alces28_79.pdf alces27_227conferenceworkshop.pdf alces27_150.pdf alces(23)_181.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces26_64.pdf alces(25)_25.pdf alces24_78.pdf alces29_235.pdf alces22_69.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces24_7.pdf alces28_123.pdf alces28_235.pdf alces27_41.pdf alces29_267.pdf alces27_111.pdf alces21_91.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces24_1.distinguishedmoosebio.pdf alces24_112.pdf rodgersar text box alces28_21.pdf alces24_143.pdf alces27_193.pdf alces26_172distinguishedmoosebio.pdf alces24_178.pdf alces22_preface.pdf alces vol. 22, 1986 alces(25)_146.pdf alces26_129.pdf alces27_8.pdf alces28_189.pdf alces(25)_58.pdf alces22_437workshopsessions.pdf alces vol. 22, 1986 alces(25)_175.pdf alces(23)_285.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces21_149.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces29_169.pdf alces21_231.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces(25)_98.pdf alces26_86.pdf alces(23)_61.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces22_377.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces22_313.pdf alces vol. 22, 1986 rodgersar text box alces 22 (1986) alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces21_321.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces21_403.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces27_161.pdf alces28_89.pdf alces26_1.pdf alces(25)_31.pdf alces29_243.pdf alces24_10.pdf alces(23)_195.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces22_83.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 temporal assessment of physical characteristics and reproductive status of moose in new hampshire daniel h. bergeron1,2, peter j. pekins 1 , and kristine rines2 1department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa; 2new hampshire fish and game department, new hampton, new hampshire 03256, usa. abstract: biological data collected from harvested moose (alces alces) were analyzed to assess whether temporal change has occurred in the physical and reproductive condition of moose from 1988–2009 in new hampshire. measurements included age and field-dressed body weight of both sexes, number of corpora lutea (cl) and ovulation rate of females, and antler beam diameter (abd) and antler spread of males. similar data were obtained from maine and vermont for comparative analysis. the only significant changes (p <0.05) occurred in the yearling age class: mean body weight of both sexes, number of cl, and abd all declined in new hampshire. the current ovulation rate (∼20%) and mean body weight (<200 kg) of yearling females in new hampshire and vermont were considered low. the declines measured in yearlings, yet relative stability in adults, are consistent with the presumption that winter ticks (dermacentor albipictus) impact the productivity of moose populations through reduced calf survival and growth and fecundity of yearlings. density-dependent factors related to habitat change are also discussed given the recent, rapid expansion of moose in the 3 states. continued monitoring of physical parameters and productivity of harvested moose, particularly the yearling cohort, is warranted to better assess the relationships among winter ticks, habitat quality, and moose populations. alces vol. 49: 39–48 (2013) key words: alces alces, body weight, moose, new england, physical characteristics, reproductive status age-specific body weight is directly related to the health and production of male and female moose (alces alces) (schwartz and hundertmark 1993), and onset of ovulation in yearlings (saether and heim 1993). antler measurements that are used routinely to estimate the health of white-tailed deer (odocoileus virginianus) populations are also used to gauge population status of moose (e.g., child et al. 2010); antler size in moose is influenced by many factors including nutritional status and health (bubenik 1997). in new hampshire, age, antler spread, antler beam diameter (abd), number of points, corpora lutea (cl) count, and field-dressed body weight of hunterharvested moose have been measured since 1988. adams and pekins (1995) found differences in body weight and number of cl in yearling cow moose relative to other age classes, but no difference within age classes from regions with different moose density. they concluded that yearling moose were useful for estimating herd health due to their substantial weight gain, change in antler characteristics, and onset of ovulation in this age class. because their data were from a relatively new and expanding moose population in the 1980–1990s, they encouraged future analyses to assess both temporal and regional trends. musante et al. (2010) found that the ovulation rate and cl count of yearling moose in new hampshire declined from 39 1988–1998 to 1999–2004, yet were un‐ changed in adults. in a comprehensive study including habitat use (scarpitti 2006) and age-specific mortality rates, they concluded that epizootics of winter ticks (dermacentor albipictus) caused periodic, annual high mortality in calves and lower fecundity in yearlings. given the relationships between certain physical characteristics and nutritional status of a moose population, periodic analysis of physical and reproductive data should reveal trends and change in the relative condition of the moose population in new hampshire. in this study we assessed temporal trends in physical characteristics and relative nutritional and reproductive status of moose in new hampshire from 1988– 2009, a period that encompassed previous studies (i.e., adams and pekins 1995, musante et al. 2010) and 5 additional years. further, we analyzed similar data from neighboring states maine and vermont to produce a regional assessment. methods study area we used data collected by new hampshire fish and game department (nhfg) personnel at mandatory harvest check stations. moose/data were from 3 northern regions that differed in moose population density (nhfg 2009) (fig. 1); the 3 regions from highest to lowest density were connecticut (ct) lakes (0.83 moose/km2), north (0.61 moose/km2), and white mountains (0.26 moose/km2), respectively (k. rines, unpubl. data, 2009). elevation in the study area ranges from ∼120–1900 m, average snow depth ranges from 0–60 cm, and ambient temperature ranges from ∼−30 to 30° c (noaa 1971–2000). the ct lakes and north regions were dominated by commercial hardwood species including sugar (acer saccharum) and red maple (a. rubrum), yellow birch (betula alleghaniensis), and american beech (fagus grandifolia). red spruce (picea rubens) and balsam fir (abies balsamea) tend to be the dominant species at higher elevations (>760 m) and in cold, wet lowland sites (degraaf et al. 1992). these regions are predominately forested and the majority of the land is privately owned and commercially harvested using various silvicultural techniques (degraaf et al. 1992); they contain ∼10% wetlands and open water, and are interspersed with trails and logging roads. the ct lakes region is hilly with few high mountains, while the north is characterized by higher forested terrain. the white mountains region contains the white mountain national forest which covers 304,050 ha and is ∼97% forested. it contains the highest elevations in new hampshire and is dominated by beech, sugar maple, and yellow birch; other common species include white ash (fraxinus americana), red maple, red spruce, and eastern hemlock (tsuga canadensis). timber harvest in this region is at smaller scale than the other regions, with maximum clear-cut size of ∼10–12 ha (degraff et al. 1992, sperduto and nichols 2004). white-tailed deer are sympatric with moose throughout the study area, and at low-moderate density (<4/km2). field measurements physical measurements of harvested moose in 1988–2009 were divided into 3 time periods (1988–1998, 1999–2004, and 2005–2009) and analyzed by region. measurements included age and fielddressed body weight for both sexes, number of cl, abd, antler spread, and number of points. a micrometer was used to measure abd on one antler at 2 perpendicular sites 2.54 cm above the pedicle; the average diameter was recorded. antler spread was the maximum distance measured between any 2 points, and an antler point was ≥2.54 cm long. ovaries were collected and stored in denatured ethyl-alcohol and sectioned later to 40 physical and reproductive status in nh – bergeron et al. alces vol. 49, 2013 visually count the number of cl (cheatum 1949). age was determined by cementum annuli counts from a lower incisor (sergeant and pimlott 1959). a subset of similar data was obtained from maine and vermont; maine data included only field-dressed body weight of cows and vermont data were from 1993–2009. data analysis new hampshire data were analyzed initially by time period and sample region, and combined statewide for comparison with maine and vermont data. analysis of variance (anova) was used to test for age-specific differences in physical parameters; age classes were 0.5, 1.5, 2.5, 3.5, 4.5, 5.5, and ≥6.5 years. a shapiro-wilk test was used to test if the data were normally distributed and a bartlett test was used to check for homogeneity of variance (zar 1999). pairwise comparisons were made with the tukey test. analyses were performed with systat v. 13. significance for all tests was assigned a priori at α = 0.05. results the analysis included measurements from >3000 and 1500 male moose, and >1500, 1300, and 2500 female moose in new hampshire, vermont, and maine, respectively. in new hampshire, sample size was >10 in the middle age classes (1.5–3.5 years) in all regions in any given time period; sample size was >20 in all age classes/time periods for state comparisons. fig. 1. location of 3 study regions with different moose density (high-low) used to evaluate temporal trends in physical and reproductive status of moose in northern new hampshire, 1988–2009. alces vol. 49, 2013 bergeron et al. – physical and reproductive status in nh 41 females statewide means for body weight and cl counts for all age classes are presented in table 1. in new hampshire the only significant differences between time periods in any region occurred in the yearling age class; albeit, body weight declined in most age classes in successive periods (table 1). body weight of yearlings declined significantly (∼25 kg) from 1988–1998 to 2005– 2009 in all regions (fig. 2): ct lakes (p = 0.033), north (p = 0.000), white mountains (p = 0.003). the number of cl in yearlings also declined from 1988–1998 to 2005–2009 in all regions (fig. 3): ct lakes (43%, p = 0.009), north (68%, p = 0.000), white mountains (76%, p = 0.003). the cl count was ∼0.20 across all regions in 2005–2009, declining from 0.60–0.80 since 1988–1998. the ovulation rate in yearling cows declined from 56 to 21% from 1988–1998 to 2005–2009 in new hampshire, and from 36 to 16% in vermont. the average body weight of yearling cows with 0 cl was 199 kg in new hampshire and 198 kg in vermont. yearling body weight declined 6% in vermont (11 kg, p = 0.001) from 1999–2004 to 2005–2009 (fig. 2), and cl counts, though not different, also declined to <0.20 (fig. 3). the cl count was lower in vermont than new hampshire in 1988– 1998 (45%, p = 0.030) and 1999–2004 (38%, p = 0.030) (fig. 3); there was no difference in 2005–2009, albeit all counts were historical lows. yearling body weight in new hampshire and vermont was not different. body weight of maine year‐ lings increased 3% from 1988–1998 to 1999–2004 (p = 0.012) (fig. 2). body weight was 6% lower in maine than new hampshire in 1988–1998 (p = 0.000), but 7% higher in 2005–2009 (p = 0.000). mean body weight of maine yearlings increased 9% from 1988–1998 to 2005–2009, and only maine had a statewide mean >200 kg table 1. statewide means (± sd) of field-dressed body weight and number of corpora lutea of harvested female moose in 3 consecutive time periods in new hampshire, 1988–2009. the only significant differences (p <0.05) occurred in the 1.5 year (yearling) age class (*); all parameters declined from 1988– 1998 to 2005–2009. age 1988–1998 1999–2004 2005–2009 body weight (kg) 0.5 110 ± 25 (74) 105 ± 20 (51) 107 ± 22 (45) 1.5 211 ± 31 (175) 203 ± 27 (206) 190 ± 29 (165)* 2.5 258 ± 34 (167) 250 ± 29 (132) 238 ± 31 (117) 3.5 255 ± 35 (102) 246 ± 29 (85) 258 ± 31 (87) 4.5 268 ± 34 (55) 263 ± 35 (68) 247 ± 43 (60) 5.5 261 ± 32 (46) 260 ± 32 (48) 246 ± 29 (40) ≥6.5 258 ± 36 (106) 263 ± 31 (133) 257 ± 36 (131) # corpora lutea 1.5 0.65 ± 0.65 (187) 0.42 ± 0.52 (200) 0.21 ± 0.42 (169)* 2.5 1.26 ± 0.66 (174) 1.09 ± 0.53 (142) 0.98 ± 0.48 (127) 3.5 1.29 ± 0.62 (102) 1.17 ± 0.60 (90) 1.08 ± 0.54 (91) 4.5 1.53 ± 0.65 (62) 1.26 ± 0.56 (72) 0.98 ± 0.61 (62) 5.5 1.37 ± 0.61 (46) 1.30 ± 0.63 (54) 1.13 ± 0.67 (48) ≥6.5 1.46 ± 0.73 (108) 1.31 ± 0.60 (151) 1.13 ± 0.66 (142) 42 physical and reproductive status in nh – bergeron et al. alces vol. 49, 2013 fig. 2. mean (± se) field-dressed body weight (kg) of harvested yearling female moose in 3 sample regions of new hampshire (1988–2009), and statewide means in new hampshire, vermont, and maine. body weight declined (p <0.05) in new hampshire and vermont from 1988–1998 to 2005–2009; conversely, body weight increased in maine. for reference, yearlings with body weight <200 kg are considered non-reproductive. 0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 ct lakes north white mt new hampshire vermont c o rp o ra l u te a sample area 1988-1998 1999-2004 2005-2009 time period fig. 3. mean (± se) number of corpora lutea (cl) in harvested yearling female moose in 3 sample regions of new hampshire (1988–2009), and statewide means in new hampshire and vermont. number of cl declined (p <0.05) in new hampshire from 1988–1998 to 2005–2009; although not different, the decline in vermont was ∼50%. alces vol. 49, 2013 bergeron et al. – physical and reproductive status in nh 43 from 1999–2009. the proportion of yearlings >200 kg in new hampshire, vermont, and maine was 44, 32, and 62%, respectively, in 2005–2009. males statewide means for body weight, abd, and antler spread are presented in table 2. in new hampshire the only significant differences between time periods in any region occurred in the yearling age class; albeit, all characteristics in table 2 declined in most age classes in successive periods. yearling body weight declined 28, 16, and 30 kg from 1988–1998 to 2005–2009 in the ct lakes (12%, p = 0.000), north (7%, p = 0.011), and white mountain (14%, p = 0.000) regions, respectively (table 3). yearling abd declined 11% in the ct lakes (p = 0.023) and 9% in the white mountains (p = 0.014) regions (table 3) from 1988– 1998 to 2005–2009. yearling antler spread declined 13, 11, and 15% from 1988–1998 to 2005-2009 in the ct lakes (p = 0.034), north (p = 0.026), and white mountains (p = 0.001) regions, respectively (table 3). as in new hampshire, vermont yearlings declined in each physical characteristic except abd; body weight declined 9% (p = 0.003) (table 3) and antler spread 7% table 2. statewide means (± sd) of field-dressed body weight, antler beam diameter (abd), and antler spread of harvested bull moose in 3 consecutive time periods in new hampshire, 1988–2009. the only significant differences (p <0.05) occurred in the 1.5 year (yearling) age class (*); all parameters declined from 1988–1998 to 2005–2009. age 1988–1998 1999–2004 2005–2009 body weight (kg) 0.5 119 ± 23 (67) 114 ± 26 (42) 115 ± 25 (46) 1.5 222 ± 39 (377) 206 ± 24 (235) 201 ± 29 (184)* 2.5 271 ± 42 (361) 262 ± 28 (246) 253 ± 30 (219) 3.5 311 ± 36 (229) 294 ± 30 (251) 284 ± 33 (214) 4.5 335 ± 40 (174) 317 ± 32 (172) 312 ± 34 (150) 5.5 350 ± 37 (108) 331 ± 32 (96) 319 ± 37 (93) ≥6.5 352 ± 37 (180) 344 ± 32 (243) 335 ± 36 (218) abd (mm) 1.5 36 ± 9 (415) 34 ± 7 (262) 34 ± 6 (199)* 2.5 45 ± 7 (391) 44 ± 5 (275) 42 ± 5 (251) 3.5 49 ± 6 (258) 47 ± 5 (291) 46 ± 4 (243) 4.5 54 ± 7 (191) 51 ± 6 (195) 50 ± 6 (162) 5.5 56 ± 8 (124) 54 ± 6 (114) 54 ± 5 (106) ≥6.5 60 ± 6 (214) 59 ± 7 (271) 58 ± 6 (236) antler spread (cm) 1.5 66 ± 11 (372) 60 ± 12 (247) 59 ± 11 (191)* 2.5 90 ± 11 (363) 85 ± 12 (275) 81 ± 11 (247) 3.5 107 ± 15 (246) 98 ± 14 (289) 96 ± 15 (242) 4.5 120 ± 16 (183) 112 ± 16 (191) 109 ± 16 (157) 5.5 126 ± 16 (114) 121 ± 12 (121) 118 ± 16 (106) ≥6.5 133 ± 15 (197) 131 ± 16 (269) 128 ± 15 (232) 44 physical and reproductive status in nh – bergeron et al. alces vol. 49, 2013 (p = 0.049) from 1988–1998 to 2005–2009 (table 3). there was no difference in body weight between new hampshire and vermont yearlings; antler spread was greater in new hampshire than vermont in 1988–1998 (9%, p = 0.031) and 2005–2009 (5%, p = 0.028), and abd was 6% larger in vermont than new hampshire in 1999–2004 (p = 0.033). discussion prior research in new hampshire (musante 2006, musante et al. 2010) indicated that new hampshire's moose population was effectively stable due to low annual growth rate (estimates = 0.95-1.07). population stability occurs despite the belief that habitat quality is high (scarpitti et al. 2005, scarpitti 2006) and adult productivity and survival are also high (musante et al. 2010). the population is presumably most influenced by winter ticks that cause periodic, high mortality of calves and reduced productivity in yearling cows (musante et al. 2010). our data indicate that body weight and cl count of yearling females have continued to decline through 2005–2009 to about 190 kg and 0.20 cl (table 1), respectively; ovulation rates of yearlings in north america average 49% (range = 0–100%, schwartz 2007). conversely, the ovulation rate of adults was not low in new hampshire or vermont (most age classes >90%, table 1); however, the cl count of adults was in decline in all age classes across the study period (table 1). yearling females <200 kg are considered non-reproductive (adams and pekins 1995), and not coincidently, mean body weight of table 3. means (± sd) of field-dressed body weight, antler beam diameter (abd), and antler spread of harvested 1.5 year-old bull moose in 3 consecutive time periods in 3 regions of new hampshire, 1988– 2009. significant declines (p <0.05) of all parameters occurred in all regions of new hampshire from 1988–1998 to 2005–2009, except abd in the north. body weight and antler spread declined (p <0.05) in vermont from 1988–1998 to 2005–2009. 1988–1998 1999–2004 2005–2009 body weight (kg) ct lakes 232 ± 43 (80) 209 ± 26 (44) 204 ± 34 (43) north 223 ± 30 (119) 212 ± 23 (80) 207 ± 27 (61) white mt. 222 ± 44 (102) 194 ± 22 (38) 192 ± 23 (36) new hampshire 222 ± 39 (377) 206 ± 25 (235) 201 ± 29 (184) vermont 216 ± 27 (58) 202 ± 28 (127) 196 ± 27 (247) abd (mm) ct lakes 38 ± 10 (85) 34 ± 6 (47) 34 ± 6 (47) north 35 ± 7 (134) 34 ± 7 (99) 34 ± 5 (65) white mt. 37 ± 9 (113) 33 ± 7 (44) 34 ± 7 (40) new hampshire 36 ± 9 (415) 34 ± 7 (262) 34 ± 6 (199) vermont 34 ± 6 (59) 36 ± 6 (128) 34 ± 7 (258) antler spread (cm) ct lakes 68 ± 21 (76) 60 ± 14 (44) 59 ± 10 (44) north 64 ± 16 (123) 59 ± 12 (96) 57 ± 11 (63) white mt. 69 ± 21 (99) 57 ± 11 (40) 59 ± 13 (39) new hampshire 66 ± 11 (372) 60 ± 12 (247) 59 ± 11 (191) vermont 60 ± 11 (54) 60 ± 12 (118) 56 ± 12 (247) alces vol. 49, 2013 bergeron et al. – physical and reproductive status in nh 45 cows with 0 cl was 199 kg in new hampshire (1988–2009) and 198 kg in vermont (1993–2009). productivity from the yearling age class in new hampshire and vermont is expectedly low based on ovulation rates ≤20% that are considerably lower (30–50%) than those measured prior to 2000. the mean cl count in new hampshire (0.22) and vermont (0.16) was equal to half the proportion of yearlings >200 kg (44 and 32%, respectively); assuming this relationship, the mean cl in maine is probably >0.30, as 62% of yearlings were >200 kg. several factors including habitat quality, weather, and disease/parasites contribute to declining trends in physical parameters of a moose population, the latter 2 typically of short-term impact. however, musante et al. (2010) believed that moose in new hampshire were mostly influenced by the annual impact, and particularly epizootics, of winter ticks. mortality of their radio-collared moose was mostly due to winter kill/parasites (41%) associated with winter tick infestations; mortality due to hunting, road-kill, poaching, predation, and weather was not considered major during the 4-year study. further, habitat was considered adequate because field-dressed weights, reproductive data, and survival of adults were not low or declining, or representative of a habitatlimited population. although our analysis identified no statistical decline in physical characteristics or ovulation rates of adults, body weight of males and females and age-specific cl counts trended downward across the ∼20-year period (tables 1 and 2). calves are most severely impacted by winter tick infestations and some mortality is likely an annual event; however, even surviving calves presumably experience lower body weight and reduced fecundity as yearlings (samuel 2004, 2007, musante et al. 2010). the declining trend in yearling condition in new hampshire and vermont from 1988–2009 suggests that average tick loads might impact moose populations through reduced fitness and fecundity of yearlings. although the field-dressed body weight of yearling cows in maine has been stable at 205 kg since 1999, it is less than the peak weight in new hampshire in 1988–1998 (217 kg, fig. 2). as a region, it is evident that productivity of yearling cows is low with cl counts probably <40% even in maine based on comparative data from new hampshire and vermont (fig. 2 and 3). new hampshire's moose population was still expanding in 1988–1998, and their physical characteristics may have peaked during this period of high resource availability related to extensive forest harvesting in the 1980s (see bontaites and gustafson 1993). their gradual decline since 1988 may reflect the combined influences of saturation of available habitat, reduced availability of preferred habitat, and gradual decline in habitat quality due to subsequent forest maturation. further, concern exists about forest regeneration in the face of dense populations in northern areas of all 3 states, and isolated examples exist (see bergeron et al. 2011); that these populations may express selflimiting impacts on habitat quality, hence fecundity, is possible. however, the steep decline in yearling body weight and that the yearling ovulation rate is well below the north american average suggests that other contributing factors exist, particularly given the relative stability of measurements in adult moose. in fact, winter ticks cause age-specific impacts because calves have higher, relative tick numbers than adults, and severe hairloss is evident on calves even in low/average tick years (samuel and barker 1979, samuel 2004, sine et al. 2009, bergeron 2011). the lack of a local epizootic of winter tick since 2002 and the declining trend in yearling physical characteristics supports the hypothesis that annual winter tick numbers affect population dynamics through reduced growth and 46 physical and reproductive status in nh – bergeron et al. alces vol. 49, 2013 fecundity of yearling moose (i.e., surviving calves). recent warmer and shorter winters that maximize spring survival and autumn questing of winter ticks presumably enhance this relationship by causing an increase in annual tick numbers, and likely increase the probability of an epizootic that produces substantial calf mortality; anecdotal reports from all 3 states suggest that a local epizootic in combination with deep snow caused high calf and yearling mortality in winter 2010–2011. the relative influences of habitat, population density, weather, and parasites on the population dynamics of moose is difficult to ascertain, and likely varies temporally. collection of long-term data sets of tick numbers and physical parameters of harvested moose in concert with annual, spring hair-loss surveys would better document the relationships between winter tick and population dynamics of moose in new hampshire. acknowledgements funding for this research was provided by the nhfg and data were available because of the dedication of nhfg biologists at harvest check stations. we are grateful to the many students from the university of new hampshire who helped at check stations through the years. we thank l. kantar of the maine department of inland fisheries and wildlife, and c. alexander of the vermont fish and wildlife department for providing comparative data. n. fortin assisted with tables and figures, and early versions of the paper. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53–59. bergeron, d. h. 2011. assessing relationships of moose populations, winter ticks, and forest regeneration in northern new hampshire. m.s. thesis, university of new hampshire, durham, new hampshire, usa. ———, p. j. pekins, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. bontaites, k., and k. a. gustafson. 1993. the history of moose and moose management in new hampshire. alces 29: 163–167. bubenik, a. b. 1997. evolution, taxonomy, and morphology. pages 77-123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. cheatum, e. l. 1949. the use of corpus lutea for determining ovulation incidence and variation in the fertility of white-tailed deer. cornell veterinarian 39: 282–291. child, k., d. a. aitken, and r. v. rea. 2010. morphometry of moose antlers in central british columbia. alces 46: 123–134. degraaf, r. m., m. yamisaki, w. b. leak, and j. w. lanier. 1992. new england wildlife: management of forested habitats. general technical report ne-144, usda forest service, northeast experiment station, radnor, pennsylvania, usa. musante, a. r. 2006. characteristics and dynamics of a moose population in northern new hampshire. m.s. thesis, university of new hampshire, durham, new hampshire, usa. ———, p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. new hampshire fish and game department (nhfg). 2009. wildlife harvest summary. new hampshire fish and game department, concord, new hampshire, usa. saether, b., and m. heim. 1993. ecological correlates of individual variation in age alces vol. 49, 2013 bergeron et al. – physical and reproductive status in nh 47 at maturity in female moose (alces alces): the effects of environmental variability. journal of animal ecology 62: 482–489. samuel, w. m. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. ———. 2007. factors affecting epizootics of winter ticks and mortality of moose. alces 43: 39–48. ———, and m. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869) on moose alces alces (l.) of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. scarpitti, d. 2006. seasonal home range, habitat use, and neonatal habitat characteristics of cow moose in northern new hampshire. m.s. thesis, university of new hampshire, durham, new hampshire, usa. ———, c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. schwartz, c. c. 2007. reproduction, natality, and growth. pages 141–171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smith‐ sonian institution press, washington, d.c., usa. ———, and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. journal of wildlife management 57: 454–468. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23: 315–321. sine, m., k. morris, and d. knupp. 2009. assessment of a line transect field method to determine winter tick abundance on moose. alces 45: 143–146. sperduto, d. d., and w. f. nichols. 2004. natural communities of new hampshire. new hampshire natural heritage bureau, concord, new hampshire, usa. zar, j. h. 1999. biostatistical analysis, 4th edition. prentice hall, inc., englewood cliffs, new jersey, usa. 48 physical and reproductive status in nh – bergeron et al. alces vol. 49, 2013 temporal assessment of physical characteristics and reproductive status of moose in new hampshire methods study area field measurements data analysis results females males discussion acknowledgements references variation in metatarsal morphology among subgroups of north american moose (alces alces) william j. silvia1, rolf o. peterson2, john a. vucetich2, william f. silvia1, and alexander w. silvia1 1department of animal and food sciences, university of kentucky, lexington, kentucky, 40546-0215; 2school of forest resources and environmental science, michigan technological university, houghton, michigan, 49931 abstract: the objectives of this study were to characterize variation in dimensional data from the metatarsus of 4 different subpopulations of north american moose (alces alces) that are known to differ in stature, and to determine if specific metatarsal width measurements (proximal, middle, distal) can be used to accurately predict metatarsal length in these subpopulations. we found that subpopulations differ in the dimensions of their metatarsal bones. alaskan moose (a. a. gigas) are significantly larger in the length and width of the metatarsus than non-alaskan moose. moose from isle royale have significantly shorter metatarsal bones than the other groups which is associated with a proportional reduction in the middle metatarsal width; the ratio of middle width:length was similar across groups in contrast to the proximal: and distal width:length ratios. these dimensions were not reduced proportionally in isle royale specimens as these ratios were greater in the isle royale moose than in other groups. predictive equations for estimating metatarsal length from each of the 3 width measurements were developed. the length could be predicted accurately from each of the width measurements if separate predictive equations were developed for specimens collected from isle royale versus the other subgroups. these data indicate that considerable variation exists in the dimensions of a single bone, the metatarsus, in subgroups of the same species. valid predictive equations developed using data sets from one subgroup may not provide accurate predictions when applied to other subgroups of the same species. alces vol. 50: 159–170 (2014) key words: alces, metatarsus, moose, morphology, variation estimating the body size of individuals is an important part of any population assessment. direct measures (e.g., shoulder height, heart girth, body weight) of large species are often difficult to obtain in the field, and estimates of body size are often made from extrapolations of other body parts. foot length is correlated with live or carcass weight in many ungulate species (bandy et al. 1956, mcewan and wood 1966, roseberry and klimstra 1975, martin et al. 2013) including moose (alces alces) (franzmann et al. 1978, lynch et al. 1995, jensen et al. 2013). for ungulates, both living and recently deceased, this is most often measured along the plantar surface from the calcaneal protuberance to the tip of the longest toe. for animal remains that are collected after significant decomposition, it may be more convenient and consistent to measure the length of the metatarsus itself, commonly referred to as the cannon bone. the length of the metatarsus is correlated with body size across mammalian species (mcmahon 1975, alexander et al. 1979). for example, the length of the metatarsus is correlated with body weight and growth rate in cattle (coble et al. 1971b), and length and width william j. silvia, department of animal and food sciences, 409 wp garrigus bldg, university of kentucky, lexington, kentucky 40546-0215 159 of the metatarsus were smaller in female than male cattle (coble et al. 1971a), a clear indication of sexual dimorphism. the length of the metatarsus is an excellent indicator of fetal age in sheep (santucci et al. 1993), the length of the metatarsus in growing lambs is directly related to maternal nutrition during gestation (pálsson and vergés 1952), and the heritability of metatarsal dimensions is relatively high (coble et al. 1971a). the length of the metatarsus has been used as an indirect measure of body size in moose (alces alces; peterson 1977). in the field, it is quite common to find metatarsal bones from moose that have been broken or damaged in such a way that an accurate length cannot be determined. however, portions of the metatarsus are often intact permitting accurate measurement of the width at some point along the length of the bone. recognizing that metatarsal dimensions are of great utility in field research with moose and that there is considerable size variation among subpopulations of moose, our first objective was to characterize the length and 3 specific width measurements of metatarsal bones collected from 4 groups of moose: 1) isle royale national park (subspecies undetermined, either a. a. americana or a. a. andersoni), 2) extant alaskan moose (subspecies a. a. gigas), 3) fossilized alaskan moose (subspecies undetermined), and 4) mainland, excluding alaska (includes subspecies a. a. americana, a. a. andersoni, a. a. shiras). our second objective was to determine if specific metatarsal widths (proximal, middle, distal) can be used to accurately predict metatarsal length of north american moose, and to determine if the relationships between length and specific widths vary among the 4 subgroups. methods quantitative measurements of metatarsal morphology of adult north american moose were made on 4 subgroups. the first subgroup consisted of 420 moose from isle royale national park (48°06’ n, 88°30’ w; peterson 1977) located in lake superior approximately 30 km from the ontario, canada coastline. the precise origin of moose on isle royale is unknown, but the founding animals could be either a. a. americana or a. a. andersoni subspecies; however, isle royale moose are morphologically different from both subspecies (peterson et al. 2011). these metatarsal specimens are currently housed at michigan technological university’s (mtu) ford center in alberta, michigan. the second group of specimens was from 170 modern alaskan moose and included specimens housed at 1) the museum of the north, university of alaska, fairbanks, alaska (collected from denali national park; 63°20’ n, 150°30’ w; n = 65), 2) the mtu ford center (collected in the kenai national wildlife refuge [knwr] at 60° 20’ n, 150°30’ w; n = 95), 3) the american museum of natural history (amnh), new york, new york (collected throughout alaska; n = 6), and 4) the field museum of natural history (fmnh), chicago, illinois (collected throughout alaska; n = 3). the third group of 49 metatarsal bones was fossil material from the late pleistocene age that was collected from several sites 10–35 km north of fairbanks, alaska (65° n, 147°40’ w) (frick 1930, wilkerson 1932) and was part of the frick collection at the amnh (n = 49); these are presumed from the subspecies a. a. gigas. the fourth set of 34 specimens, referred hereafter as mainland moose, was collected from a variety of sites in canada and the united states (excluding alaska) and included subspecies a. a. americana, a. a. andersoni, and a. a. shiras. these specimens are housed at the 1) mtu ford center (collected by the michigan and minnesota departments of natural resources (n = 20), 2) the amnh (n = 7), 3) the fmnh (n = 1), 4) brown 160 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 university (n = 1), 5) harvard university (n = 2), and 6) the university of kentucky (n = 3). all specimens were from moose either killed by hunters or vehicular collisions. specimens collected in isle royale national park, denali national park, or the knwr were obtained from animals that died of natural causes. on isle royale, the majority resulted from predation by wolves (peterson 1977); as a result, animals that were more susceptible to predation (due to age, injury, disease) may be overrepresented. the sex of specimens was determined from examination of soft tissue (when present) and morphological characteristics of the associated skull (when present). a general age was determined by the size of the remains and the complement of deciduous and permanent teeth (peterson et al. 1983). when necessary, tissue from the metatarsal bones was removed manually with a knife and/or by prolonged immersion in hot water (>80 °c). quantitative measurements of the cannon bone were made using 2 sizes of manual vernier calipers. the length was measured using a 24-inch, cen-tech aluminum caliper (harbor freight tools inc., camarillo, california, usa) that was modified by adding a vertical fence to each side, extending the height to approximately 2.2 cm (fig. 1). the width of each metatarsus was measured at the proximal end, midpoint, and distal end with a standard 5-inch manual caliper (helios, germany; fig. 2a). the width at the proximal end was measured at the widest point, typically within 1 cm of the end (fig. 2b). the width at the distal end was also measured at the widest point, but the precise location varied; in some, it was very close to the end at the lateral and medial edges of the corresponding articular condyles, and in others it was proximal to the condyles, in the approximate location of the epiphyseal plate (fig. 2c). these measurements were used to calculate width:length ratios for each width (i.e., proximal, middle, and distal). the condition of the epiphyseal plate was classified as either unfused or fused. the unfused classification included specimens in which the 2 portions of the metatarsus were separable, and specimens in which the 2 portions were not separable but a distinct suture was clearly visible (fig. 3). specimens were classified as coming from adults only if the distal epiphyseal growth plate was no longer visible. the effects of subgroup and/or sex on quantitative measurements (length and width of the metatarsus, width:length ratio) were evaluated with analysis of variance using the glm procedure of sas (1985). the relationships between the length of the cannon bone and the 3 width measurements were evaluated with linear regression using the reg procedure of sas (1985). the accuracy of the regression equations in predicting metatarsal length from width measurements was evaluated with paired t-test using the means procedure of sas (1985). fig. 1. the modified vernier caliper used to measure the length of the cannon bone. note that vertical fences were added to each of the ‘jaws’ of the caliper to extend the height. alces vol. 50, 2014 silva et al. – metatarsal dimensions in alces 161 results metatarsal length and width the length of the metatarsus was different for each of the subgroups (p < 0.01; fig. 4a). the fossil metatarsal bones from alaskan moose were the longest, followed by those of modern alaskan moose, mainland moose, and lastly isle royale moose. the width of the metatarsus at the proximal end was greater in the 2 alaskan subgroups than in the other subgroups (p < 0.01; fig. 4b); the alaskan subgroups did not differ (p = 0.10), nor did the non-alaskan subgroups (p = 0.64). the ratio of proximal metatarsal width:metatarsal length was different among groups (p < 0.01; fig. 4c). among all subgroups the ratio was highest fig. 2. dorsal view of the cannon bone with points of measurement indicated. panel a shows the measurement of length (dashed line) and proximal, middle, and distal widths (vertical arrows); 2 possible points for measure of the distal width are indicated. panel b shows a detailed view of the proximal end indicating the point of measurement more precisely. panel c shows a more detailed view of the distal end and the 2 possible points for measurement. fig. 3. the dorsal view of the distal end of the cannon bone from an adult (panel a) and a juvenile animal (panel b). although not separable, the epiphyseal plate is clearly visible on the juvenile specimen (arrow). 162 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 in specimens from isle royale (p < 0.05); the ratio among the other 3 subgroups did not differ (p > 0.70). to further examine the relationship between proximal width and length, the effect of width and subgroup on metatarsal length was determined (n = 285). proximal width had a significant effect (p < 0.01), but subgroup did not (p = 0.06). there was an interaction between proximal width and subgroup on metatarsal length (p = 0.01). since the ratio width:length appeared to be different for isle royale moose compared to the other subgroups, a second analysis was conducted without isle royale moose. again, the effect of proximal width was evident (p < 0.01), but not subgroup (p = 0.27) or the interaction term of proximal width and subgroup (p = 0.20), implying that the alaskan and mainland subgroups are similar and a different relationship exists for isle royale moose. middle (n = 226) and distal width measurements (n = 224) were not available from the fossil specimens; therefore, comparisons could only be made among the modern subgroups. the width of the metatarsal at the midpoint was greater in alaskan moose than those from non-alaskan subgroups (p < 0.01; fig. 5a); this width was similar in isle royale and non-alaskan subgroups (p > 0.30). the ratio middle width: length was not different among subgroups (p = 0.44; fig. 5b). as with the proximal metatarsal width, the distal metatarsal width was greater in the alaskan subgroup than non-alaskan subgroups (p < 0.01; fig. 5c). the ratio distal metatarsal width:metatarsal length also differed among subgroups (p < 0.01; fig. 5d). the distal width:length ratio was greater for isle royale moose than the other groups (p < 0.01); the other groups did not differ (p = 0.12). the effect of sex on metatarsal dimensions was analyzed with all specimens in which sex could be determined, which excluded the fossil subgroup. the length of the metatarsus was greater in males than females (p < 0.01; fig. 6a). as in the first analysis, the length of the metatarsus differed among subgroups (p < 0.01), and metatarsal length was longer in males than females in all subgroups. a significant interaction fig. 4. the effect of subgroup (isle royale [isro], mainland [non-isro from the lower 48 contiguous united states and canada], alaska, and fossil alaska) on metatarsal length (a), proximal metatarsal width (b), and the ratio of proximal metatarsal width to metatarsal length (c). bars with different letter superscripts are different (p < 0.05). alces vol. 50, 2014 silva et al. – metatarsal dimensions in alces 163 between sex and subgroup was also found (p < 0.01). this interaction was strongest in alaskan moose that had the longest metatarsal length and largest difference between males and females. the effect of sex on the relationship between each of the 3 width measurements and metatarsal length was examined fig. 5. the effect of subgroup (isle royale [isro], mainland [non-isro from the lower 48 contiguous united states and canada] and alaska) on middle metatarsal width (a), the ratio of proximal metatarsal width to metatarsal length (b), distal metatarsal width (c), and the ratio of distal metatarsal width to metatarsal length (d). bars with different letter superscripts are different (p < 0.05). fig. 6. the effect of sex and subgroup on metatarsal dimensions: length of the metatarsus (panel a) in which effects of sex, subgroup, and their interaction were observed (p < 0.01). proximal, middle, and distal width of the metatarsus is shown in panels b, c, d. effects of sex and subgroup on all 3 width measurements were observed (p < 0.01) but not of their interaction (p ≥ 0.09). 164 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 separately in the isle royale and non-isle royale subgroups. in both cases, the effect of proximal width was evident (p < 0.01; table 1). sex had no effect on length that was not already accounted for by proximal width (p > 0.21). the interaction term of sex with proximal width on metatarsal length was also not significant (p > 0.20) in either the isle royale or non-isle royale subgroups. similarly, there were no effects of sex or the interaction of sex and width on the relationship between middle or distal width on the length of the metatarsal (table 1). predictive equations for metatarsal length based on widths a quantitative description of the relationship between proximal width and metatarsal length was investigated with linear regression. separate regression analyses were conducted for the isle royale and non-isle royale subgroups using the following simple model: metatarsal length ¼ m � proximal metatarsal width þ b þ e ð1þ where: m = slope, b = y-intercept, and e = error term. comparison of the estimates of slope and y-intercept for the two groups (isle royale versus non-isle royale) indicated substantial difference (table 2, fig. 7). these relationships explained a high percentage of the variation in metatarsal length for isle royale (r2 = 0.47) and non-isle royale specimens (r2 = 0.66) (table 2). the accuracy of the regression lines in predicting metatarsal length from proximal width was evaluated by comparing measured lengths to estimated lengths from specimens not used to derive the regression equations; specimens from both groups were included in this test. the length of metatarsal bones from both groups was more accurately predicted using the separate regression equations derived from the respective data sets (table 3). the same analytical procedures were used to examine the relationships between middle width and metatarsal length, and distal width and metatarsal length; middle and distal widths were not available from fossil alaskan moose. there was no effect of subgroup or the width by subgroup interaction term, indicating consistency across all subgroups. subsequently, regression analysis was used and prediction equations developed with the combined subgroup data (fig. 8, 9, table 4). again, width measurements accounted for a large percentage of the variation in metatarsal length (r2 = 0.55 and 0.53 for middle and distal widths, respectively). table 1. the effect of sex on the relationship between the width of the metatarsal at 3 points of measurement (proximal, middle, and distal) and the length of the metatarsal in specimens from isle royale and non-isle royale locations including alaska (modern and fossil), canada, and the 48 contiguous united states (excluding isle royale). isle royale non-isle royale proximal width n 102 120 width p < 0.01 p < 0.01 sex p > 0.21 p > 0.78 sex × width interaction p > 0.20 p > 0.83 middle width n 65 121 width p < 0.01 p < 0.01 sex p > 0.89 p > 0.79 sex × width interaction p > 0.83 p > 0.84 distal width n 65 120 width p < 0.01 p < 0.01 sex p > 0.82 p > 0.89 sex × width interaction p > 0.86 p > 0.83 alces vol. 50, 2014 silva et al. – metatarsal dimensions in alces 165 as with proximal width measurements, a reasonably accurate estimate of metatarsal length was obtained from either middle or distal width (table 5). as expected, the length of the isle royale specimens tended to be overestimated. although not justified based on the initial analysis, a more accurate estimate of metatarsal length was developed using separate equations derived from the isle royale and non-isle royale data (table 6, fig. 8, 9). accuracy was substantially improved for the isle royale and alaskan subgroups (table 7), but not mainland moose that was a small heterogeneous group representing 3 subspecies. table 3. comparison of measured and predicted metatarsal lengths (mm) in isle royale and non-isle royale moose using separate predictive equations developed from proximal metatarsal lengths (mm). non-isle royale moose include modern and fossil specimens from alaska, and modern specimens from the 48 contiguous united states (excluding isle royale) and canada. isle royale alaska mainland sample size 6 6 6 ave. proximal metatarsal width 53.6 57.7 51.6 ave. metatarsal length 384.5 414.3 389.8 predicted metatarsal length from proximal metatarsal width (isle royale) 388.2 398.9 382.7 difference between, range, and probability that true and predicted lengths differ (isle royale) −3.7 −13.6−9.8 15.5 5.4−33.4 7.1 −0.4−18.8 p = 0.40 p = 0.02 p = 0.04 predicted metatarsal length from proximal metatarsal width (non-isle royale) 403.7 420.2 395.5 difference between, range, and probability that true and predicted lengths differ (non-isle royale) −19.2 −28.0−5.4 −5.9 −14.7−9.0 −5.6 −16.7−2.1 p = 0.01 p = 0.15 p = 0.08 table 2. the regression parameters describing the different relationship between proximal metatarsal width and metatarsal length in specimens from isle royale compared to other populations in canada and the united states including alaska. isle royale non-isle royale sample size 110 159 significance level p < 0.01 p < 0.01 adjusted r2 0.47 0.66 slope (se) 2.67 (0.27) 4.09 (0.23) y-intercept (se) 245 (14) 185 (13) fig. 7. scatter plot depicting the relationship between the proximal metatarsal width and metatarsal length for the 4 subgroups (isle royale [isro], mainland [non-isro from the lower 48 contiguous united states and canada], alaska, and fossil alaska). regression lines for the isle royale (dotted) and non-isle royale (solid line) groups are shown. 166 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 discussion the measurements of metatarsal length and width indicated that the alaskan subgroups are larger in relative size. the isle royale subgroup is different from the other subgroups with shorter metatarsal length and correspondingly larger proximal: and distal width:length ratios. the length of the metatarsus was shorter in isle royale moose than the other subgroups and may reflect the trend for large herbivores to experience a reduction in size when isolated on small islands (peterson et al. 2011), which conforms to the ‘island rule’ (van valen 1973, lomolino 2005). given this hypothesis and the short history of isle royale moose, these data demonstrate the remarkable speed at which this phenomenon can occur. the metatarsal length:width ratios also provide insight into the biological mechanism by which reduction in metatarsal size occurred on isle royale. long bones, including metatarsals, initially form in 3 parts, the proximal epiphysis (proximal articular surface), diaphysis (shaft), and distal epiphysis. growth ceases when the cartilaginous epiphyseal plates separating these portions ossify. it appears that the reduced size in isle royale specimens is limited to the diaphysis with both the length and width of the diaphysis affected proportionally. the widths at the proximal and distal epiphyses do not appear to be reduced, particularly when compared to the mainland group. thus, the shortening effect appears to be mediated solely through the diaphysis and this isolated effect may facilitate the identification of specific genes mediating such evolutionary action. the length of the metatarsus could be predicted accurately from each of the width measurements, particularly if separate predictive equations were developed for specimens from isle royale versus other subgroups. the greatest deviation between predicted and actual metatarsal length was only 4.3% using the specific equations; refinements to these predictive equations are presumably possible. for example, the distal width measurement was taken either at the distal fig. 8. scatter plot depicting the relationship between middle metatarsal width and metatarsal length for 3 subgroups (isle royale [isro], mainland [non-isro from the lower 48 contiguous united states and canada], and alaska). regression lines derived from isro specimens (dotted line), non-isro specimens (solid line), and for all specimens combined (dashed line) are shown. fig. 9. scatter plot depicting the relationship between distal metatarsal width and metatarsal length for 3subgroups (isle royale [isro], mainland [non-isro from the lower 48 contiguous united states and canada], and alaska). regression lines derived from isro specimens (dotted line), non-isro specimens (solid line), and for all specimens combined (dashed line) are shown. alces vol. 50, 2014 silva et al. – metatarsal dimensions in alces 167 epiphysis or at the distal articular condyle, whichever was wider; however, a more accurate equation might be developed with a single, consistent measurement. predictive equations based on middle and distal widths for the non-isle royale subgroups improved the accuracy of prediction for the alaskan, but not mainland group, possibly reflecting the potential heterogeneity within the mainland group. it may indicate that separate equations need to be developed for subpopulations within. finally, possible differences in the method of sample collection among data sets should be considered. the majority of mainland specimens were collected by hunters or the result of vehicular accidents, whereas specimens from isle royale, kenai national wildlife refuge, and denali national park were collected from moose presumably dying of natural causes. animals that were particularly susceptible to predation may be overrepresented in these groups. a more robust sample size reflecting consistent sampling and population variation would presumably improve the relationships presented in this paper. acknowledgements the authors would like to thank the field workers who collected metatarsal specimens on isle royale and contributed to the collection at michigan technological university. we are indebted to the officials at the united states department of the interior, national table 4. the regression parameters describing the relationship between middle metatarsal width (mm) and metatarsal length (mm), and distal metatarsal width (mm) and metatarsal length. common equations were developed using specimens from all subgroups. middle metatarsal distal metatarsal sample size 210 208 significance level p < 0.01 p < 0.01 adjusted r2 0.55 0.53 slope (se) 4.81 (0.30) 3.38 (0.22) y-intercept (se) 242 (10) 172 (15) table 5. comparison of measured metatarsal length (mm) and predicted metatarsal length (mm) in isle royale and non-isle royale moose based on middle and distal metatarsal widths (mm). common predictive equations were derived using specimens from all subgroups. isle royale alaska mainland sample size 6 6 6 ave. metatarsal length 384.5 414.3 389.8 ave. middle width of metatarsal 32.5 35.2 31.8 predicted metatarsal length from middle metatarsal width 397.8 411.2 394.7 difference between, range, and probability that true and predicted length differ (mm) −13.3 −30.0−6.2 3.2 −5.2−15.9 −4.8 −9.3−0.1 p = 0.08 p = 0.34 p = 0.01 ave. distal width of metatarsal 66.9 70.9 64.3 predicted metatarsal length from distal metatarsal width 398.3 411.7 389.7 difference between true length and predicted length (mm) and range (below) −13.8 −23.3−0.0 2.7 −13.6−11.9 0.2 −11.7−12.0 probability that the true length and the predicted length are different p = 0.04 p = 0.53 p = 0.96 168 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 parks service, and isle royale national park for granting access to the park and permitting collection of specimens. we would also like to thank mr. p. poore for the modification of the vernier calipers used to measure the length of the cannon bone. this research was supported in part by the kentucky agricultural experiment station and is published with the approval of the director (publication number 14-07-014). references alexander, r. m., a. s. jayes, g. m. o. maloiy, and e. m. wathuta. 1979. allometry of the limb bones of mammals from shrews (sorex) to elephant (loxodonta). journal of zoology 189: 305–314. bandy, p. j., i. m. cowan, w. d. kitts, and a. j. wood. 1956. a method for the assessment of the nutritional status of wild table 7. comparison of measured and predicted metatarsal lengths (mm) from middle and distal metatarsal widths (mm) in isle royale moose with those from non-isle royale moose. the alaska and mainland equations were derived from measurements from non-isle royale moose that included modern and fossil specimens from alaska, and modern specimens from the 48 contiguous united states (excluding isle royale) and canada. isle royale alaska mainland sample size 6 6 6 ave. metatarsal length 384.5 414.3 389.8 ave. middle width of metatarsal 32.5 35.2 31.8 predicted metatarsal length from middle metatarsal width 388.3 414.5 400.5 difference, range, and probability that true and predicted lengths differ −3.8 −16.2−13.7 −0.2 −7.3−14.5 −10.7 −15.5−5.5 p = 0.50 p = 0.96 p < 0.01 ave. distal width of metatarsal 66.9 70.9 64.3 predicted metatarsal length from distal metatarsal width 388.0 415.5 396.5 difference, range, and probability that true and predicted lengths differ −3.5 −13.5−11.0 −1.2 −16.6−10.4 −6.7 −16.8−6.4 p = 0.47 p = 0.78 p = 0.10 table 6. the regression parameters describing the relationship between proximal metatarsal width (mm) and metatarsal length (mm) in specimens from isle royale compared to non-isle royale moose. non-isle royale moose included modern and fossil specimens from alaska, and modern specimens from the 48 contiguous united states (excluding isle royale) and canada. middle metatarsal width distal metatarsal width isle royale non-isle royale isle royale non-isle royale sample size 71 139 71 137 significance p < 0.01 p < 0.01 p < 0.01 p < 0.01 adjusted r2 0.41 0.54 0.50 0.54 slope (se) 3.30 (0.47) 4.07 (0.32) 2.38 (.29) 2.91 (0.23) y-intercept (se) 280 (15) 270 (11) 228 (19) 209 (16) alces vol. 50, 2014 silva et al. – metatarsal dimensions in alces 169 ungulates. canadian journal of zoology 34: 48–52. coble, d. s., l. l. wilson, j. p. hitchcock, h. varela-alvarez, and m. j. simpson. 1971a. sire, sex and laterality effects on bovine metacarpal and metatarsal characters. growth 35: 65–77. ——— , ———, m. j. simpson, h. varelaalvarez, j. p. hitchcock, j. h. ziegler, j. d. sink, and j. l. watkins. 1971b. relation of bovine metacarpal and metatarsal characters to growth and carcass characters. growth 35: 79–89. franzmann, a. w., r. e. leresche, r. a. rausch, and l. l. oldemeyer. 1978. alaskan moose measurements and weights and measurement–weight relationships. canadian journal of zoology 56: 298–306. frick, c. 1930. alaska’s frozen fauna. natural history (the journal of the american museum of natural history) 30: 71–80. jensen, w. f., j. r. smith, j. j. maskey jr., j. v. mckenzie, and r. e. johnson. 2013. mass, morphology and growth rates of moose in north dakota. alces 49: 1–15. lomolino, m. v. 2005. body size evolution in insular vertebrates: generality of the island rule. journal of biogeography 32: 1683–1699. lynch, g. m., b. lajuenesse, j. willman, and e. s. telfer. 1995. moose weights and measurements from elk island national park, canada. alces 31: 199–207. martin, j. g. a., m. festa-bianchet, s. d. cote, and d. t. blumstein. 2013. detecting between-individual differences in hind-foot length in populations of wild mammals. canadian journal of zoology 91: 118–123. mcewan, e. h., and a. j. wood. 1966. growth and development of the barren ground caribou. i. heart girth, hind foot length and body weight relationships. canadian journal of zoology 44: 401–411. mcmahon, t. a. 1975. allometry and biomechanics: limb bones in adult ungulates. american naturalist 109: 547–563. palsson, h., and j. b. verges. 1952. effect of the plane of nutrition on growth and the development of carcass quality in lambs. part i. the effects of high and low planes of nutrition at different ages. journal of agricultural science 42: 1–92. peterson, r. o. 1977: wolf ecology and prey relationships on isle royale. u.s. national park service scientific monograph series 11. u.s. government printing office, washington d. c., usa. ——— , c. c. schwartz, and w. b. ballard. 1983. eruption patterns of selected teeth in three north american moose populations. journal of wildlife management 47: 884–888. ———, j. a. vucetich, d. beyer, m. schrage, and j. raikkonen. 2011. phenotypic variation in moose: the island rule and the moose of isle royale. alces 47: 125–133. roseberry, j. l., and w. d. klimstra. 1975. some morphological characteristics of the crab orchard deer herd. journal of wildlife management 39: 48–58. sas. 1985. user’s guide: statistics. sas institute inc., cary, north carolina, usa. santucci, v. l., j. a. kuller, a. f. battelli, s. a. laifer, and d. i. edelstone. 1993. fetal metatarsal length: an accurate predictor of gestational age and weight in the ovine fetus. gynecologic and obstetric investigation 35: 76–79. van valen, l. 1973. a new evolutionary law. evolutionary theory 1: 1–33. wilkerson, a. s. 1932. some frozen deposits in the goldfields of interior alaska: a study of the pleistocene deposits of alaska. american museum novitates 525: 1–22. 170 metatarsal dimensions in alces – silva et al. alces vol. 50, 2014 variation in metatarsal morphology among subgroups of north american moose (alces alces) methods results metatarsal length and width predictive equations for metatarsal length based on widths discussion acknowledgements references alces vol. 48, 2012 henningsen et al. – elaeophora in wyoming moose 35 distribution and prevalence of elaeophora schneideri in moose in wyoming john c. henningsen1, amy l. williams1,2, cynthia m. tate3, steve a. kilpatrick1,4, and w. david walter5 1wyoming game and fish department, po box 67, jackson, wyoming 83001, usa; 3wyoming game and fish departmentwildlife diseases laboratory, 1174 snowy range road, laramie, wyoming 82070, usa; 5usda/aphis/ws national wildlife research center, 4101 laporte avenue, fort collins, colorado 80521, usa. abstract: elaeophora schneideri causes disease in aberrant hosts such as moose. documented e. schneideri infections in moose are relatively rare, yet noteworthy enough that individual cases describing morbidity and mortality have been the norm for reporting. surveillance efforts for e. schneideri in wyoming moose in the 1970s found zero cases, but since 2000 several moose in wyoming discovered dead or showing clinical signs of elaeophorosis have been found infected with e. schneideri. in 2009 we searched for worms in the carotid arteries of 168 hunter-harvested moose from across wyoming to determine the prevalence and distribution of e. schneideri in moose; 82 (48.8%; 95% ci: 41.4-56.3%) were positive for e. schneideri. prevalence did not differ between sexes or among age classes but there was difference in prevalence among herd units (range = 5-82.6%). intensity of infection (range = 1-26 worms) did not differ between sexes, among age classes, or among herd units. our findings indicate that moose do not succumb to the parasite to the extent previously thought. prevalence and intensity were constant across age classes, suggesting that infected moose are surviving and an acquired, immunological resistance to further infection develops. in addition, moose might sometimes act as natural hosts to the parasite, as indicated by 1) high prevalence of infection in moose in areas where sympatric mule deer had much lower prevalence of infection, and 2) preliminary necropsy findings that revealed microfilariae in skin samples from 3 moose. however, negative impacts to moose and moose populations cannot be ruled out entirely, as this study was limited to apparently healthy hunter-harvested animals. while moose appear to often survive infection with e. schneideri, prevalence of ~50% is still cause for concern because it is unknown to what extent this parasite causes subclinical effects in moose that might impact recruitment or productivity. subsequent research on moose herds where e. schneideri occurs should consider the effects of elaeophorosis and attempt to clarify its role. alces vol. 48: 35-44 (2012) key words: alces alces, arterial worm, disease, elaeophora schneideri, elaeophorosis, moose, parasite, wyoming. elaeophora schneideri is a filarioid nematode that lives in the cephalic arteries of mule deer (odocoileus hemionus) (hibler et al. 1970) and black-tailed deer (o. hemionus columbianus) (weinmann et al. 1973). adult nematodes in the arteries of deer give birth to live young microfilariae (hibler and adcock 1971). microfilariae then migrate to the capillaries in the dermis of the host’s face and forehead where they are taken up in the blood meal of the intermediate host. horse flies (family tabanidae) of the genera hybomitra, silvius, and tabanus (hibler et al. 1970, clark and hibler 1973, espinosa 2present address: university of wyoming, department of veterinary sciences, 1174 snowy range road, laramie, wyoming 82070, usa. 4present address: conservation research center of teton science schools, 700 coyote canyon, jackson, wyoming 83001, usa. elaeophora in wyoming moose – henningsen et al. alces vol. 48, 2012 36 1983) are intermediate hosts of the parasite. transmission of the third stage infective to the vertebrate host occurs after e. schneideri larvae develop in the horse fly vector for 2-3 weeks (hibler and adcock 1971, hibler and metzger 1974, davies 1979). development of e. schneideri in the definitive host has been described previously (weinmann et al. 1973, hibler and metzger 1974). less is known about development of e. schneideri in aberrant hosts but pathogenesis usually stems from the parasite’s delayed migration through the host body or complications from circulatory impairment (i.e., elaeophorosis; adcock and hibler 1969, hibler and metzger 1974, anderson 2001). gross clinical signs of infection among aberrant hosts range from dry gangrene of nose and ear tips and antler malformations, to blindness, central nervous system damage, and death. documented aberrant hosts of e. schneideri are moose (alces alces), elk (cervus elaphus), white-tailed deer (o. virginianus), bighorn (ovis canadensis), and domestic sheep (boyce et al. 1999, anderson 2001). e. schneideri is widespread across north america occurring in mule deer in nebraska (mckown et al. 2007), south dakota (jacques et al. 2004), utah (pederson et al. 1985), texas (pence and gray 1981), colorado, and new mexico (davies 1979). it has been documented in white-tailed deer in arizona (hibler 1982), texas (waid et al. 1984), and several southeastern states (prestwood and ridgeway 1972). infected elk have been reported from oklahoma, new mexico, arizona, colorado, and wyoming (hibler 1982). the first documented e. schneideri infections in moose were in montana (worley et al. (1972). subsequently, its presence in small numbers of moose was documented in utah (jensen et al. 1982), colorado (madden et al. 1991),washington (pessier et al. 1998), and wyoming in 2000 (w. e. cook, wyoming game and fish department [wgfd], unpublished report), and oregon in 2010 (matthews, 2012). in wyoming moose both the prevalence of infection and the parasite’s geographic extent appear to have undergone a recent, notable increase. in 1973-74 worley (1975) examined 74 apparently healthy, hunter-harvested moose: 69 from teton and fremont counties in northwestern wyoming, and 5 from park and gallatin counties in southwestern montana. no wyoming moose and only 3 of 5 montana moose were infected with worms in the carotid arteries. presumably, low prevalence of e. schneideri in wyoming led hibler (1982) to believe elaeophorosis was of minimal importance to wyoming elk and moose. whereas e. schneideri has been documented consistently in small numbers of mule deer and elk in wyoming since 1967 (h. e. edwards, wgfd, unpublished data), the first infected moose was not identified until much later. within 2 weeks in january 2000, 2 moose were euthanized by wgfd field personnel in fremont county (central wyoming) because they were lethargic or walking in circles and showed signs of impaired vision; illness in each of those cases was attributed to elaeophorosis (w. e. cook, unpublished report). in 2008 a 3-yr-old cow moose in western wyoming was euthanized because of its abnormal behavior (lack of fear, blindness, and loss of motor skills). upon gross examination, a heavy load of worms (30-50) was found in the carotid arteries and its clinical signs were attributed to elaeophorosis (c. m. tate, wgfd, unpublished report). the wgfd and the wyoming state veterinary lab (wsvl) increased opportunistic surveillance of moose in 2008. animals found dead, euthanized due to illness, and road-kills were examined for e. schneideri; several were found infected with e. schneideri. most positive cases were from animals discovered dead or showing clinical signs of illness, and pathology associated with e. schneideri was implicated in several cases. in order to survey moose for e. schneideri alces vol. 48, 2012 henningsen et al. – elaeophora in wyoming moose 37 more uniformly across wyoming, a rigorous plan was developed to establish baseline data on prevalence and distribution of e. schneideri by surveying hunter-harvested moose during the 2009 hunting season. to our knowledge, this was the most comprehensive and widespread effort to date for surveillance of e. schneideri in moose. study area brimeyer and thomas (2004) described the history and status of moose in wyoming through the early 2000s. moose in wyoming occupy 3 distinct ranges: 1) bighorn mountain range in north-central wyoming, 2) the snowy range and sierra madre ranges of southeast and south-central wyoming, and 3) western wyoming among comparatively connected mountain ranges from the utah border north through yellowstone national park (fig. 1a). moose in wyoming are managed as 11 herd units (herds) comprising discrete populations for which migration among adjacent herds is thought to account for <10% of a herd population. these herds are further divided into 43 hunt areas to provide flexibility for hunting seasons; begin and end dates of hunting seasons vary among areas. the wgfd has a statewide population objective of 14,630 moose (post-hunt), yet population estimates are considered relatively unreliable or completely lacking in most herds (thomas 2008). a b fig. 1. a) moose herd units that hunter-harvested moose were collected and sampled for elaeophora schneideri in 2009 in wyoming, usa. b) intensity of elaeophora schneideri found in carotid arteries categorized as none (●), low (○), moderate (○), and high (○) in hunter-harvested moose by herd unit in 2009 in wyoming, usa. elaeophora in wyoming moose – henningsen et al. alces vol. 48, 2012 38 methods we examined hunter-harvested moose for the presence of immature or adult worms in the terminal portion of the common carotid arteries and in some instances the proximal portion of the internal maxillary arteries (hereafter field examinations). many field examinations were conducted at hunter check stations and during opportunistic field checks of successful hunters. field examinations also took place when hunters brought heads of harvested moose to wgfd regional offices, taxidermists, or meat processors. incisor teeth were collected from harvested moose for aging by cementum annuli. the specific ages obtained via cementum annuli are reported in whole years (i.e., yearling is 1). successful development and migration of e. schneideri to the carotid arteries of moose would be expected to take 5-6 months (hibler and metzger 1974), thus nematodes would not be expected in the carotids until typically december. thus field examinations as conducted in this study could not adequately diagnose infections in calves, and surveillance of calves was not included in this study. intensity of infection with worms (bush et al. 1987) was recorded as 1 of 3 categories: 1-6, 7-13, and ≥14. these categories were based on intensities observed previously in moose from wyoming and other states (worley et al. 1972, worley 1975, madden et al. 1991). in addition to examining for the presence of e. schneideri, visual signs of elaeophorosis (e.g., cropped ears, necrotized tissues, lesions, or malformed antlers) were recorded. statistical analysis prevalence was based on the total number of positive individuals; 95% confidence intervals (ci) around these proportions were calculated based on the binomial distribution (rózsa et al. 2000). fisher’s exact test was used to test for homogeneity in prevalence between sexes. chi-square was used to test for homogeneity in prevalence among age classes and herds; fisher’s exact test was used to test for differences between pairs of herds when contingency tables had ≤5 observations in ≥1 cell. because of the small number of examined animals from individual age classes, 4 combined age classes were created (1, 2-4, 5-7, ≥ 8) for statistical comparisons. likewise, some adjacent herds (jackson and targhee, lincoln and uinta) were combined to obtain adequate sample sizes for analyses. some herds were dropped from analyses because they had both small sample sizes and were too geographically separate to justify merging. because intensity was recorded as a categorical variable, chi-square was used to test for differences in intensity among sexes, age classes, and herds. statistical significance was set at p ≤0.05 for all tests. calculations were accomplished using sigmaplot 11 (systat software, san jose, ca.). results prevalence the reported harvest in fall 2009 was 548 moose (394 adult males, 135 adult females, and 19 calves); 126 males and 42 females were examined for e. schneideri from 1 september-14 november (table 1), or 31% of adult females and 32% of adult males in the harvest. e. schneideri was present in the carotid arteries of 48.8% of all moose ≥1 yr of age (95% ci: 41.4-56.3). exactly 50% of males (95% ci: 41.4-58.6%) and 45.2% of females were infected (95% ci: 31.2-60.1%); prevalence did not differ by sex (χ2 = 0.127; p = 0.72). the number of females checked in each herd was small, but prevalence did not appear to differ between sexes within any individual herd. thus, we combined sexes for statistical comparisons of prevalence among age classes and herds. ages were obtained from 151 of 168 moose: 9 were yearlings, 54 were 2-4 years old, 71 were 5-7 years old, and 17 were ≥8 years. the infection rate was 56% in yearlings (95% ci = 26.6-81.2%), 43% in 2-4 year olds alces vol. 48, 2012 henningsen et al. – elaeophora in wyoming moose 39 (95% ci = 30.3-55.8%), 56% in 5-7 year olds (95% ci = 44.8-67.3%), and 41% in those ≥8 years (95% ci = 21.6-64.0%). there were no statistical differences in prevalence of e. schneideri among age classes (χ2 = 2.420; df = 3; p = 0.490). moose from 10 herds were checked for e. schneideri (table 1). to obtain adequate sample size for statistical comparison, the adjacent jackson and targhee herds, and the lincoln and uinta herds were combined; the absaroka, dubois, and lander herds were dropped because of low sample sizes and geographic separation (table 1). prevalence was different geographically (χ2 = 27.082, df = 4, p < 0.001). the lowest prevalence occurred in the bighorn herd (5%; 95% ci = 0-25.4%) and was lower than that in the other 4 herds in the analysis. the snowy range herd had the highest prevalence (82.6%; 95% ci = 62.3-93.6%) which was higher than in the bighorns, lincoln-uinta (43.8%; 95% ci = 23.1-66.8%), and sublette herds (52.2%; 95% ci = 42.0-62.2%), but not different than in the jackson-targhee herds (61.5%; 95% ci = 35.4-82.4%). intensity of the 82 positive cases, we found 44, 26, and 11 moose with low, moderate, and high e. schneideri intensity, respectively (table 1); intensity was not recorded for 1 positive individual. the greatest number of worms counted in any moose was 26. parasite intensity was similar between sexes (χ2 = 0.564; df = 2; p = 0.754) and among age classes (χ2 = 4.177; df = 6; p = 0.653). a low-intensity worm burden was most common in all age classes, ranging from 40-71%. none of the age classes had a large number of high-intensity worm loads. high-intensity infections were not observed in the 7 infected moose in the oldest age class (≥8 years). although prevalence was high in snowy range moose, most (74%) had low-intensity infections (fig. 1b). similarly, most positive individuals in the jackson-targhee herd (88%) had low-intensity infections. the only infected moose found in the bighorns had a moderate-intensity infection (table 1). the pattern of intensity was reversed in the lincoln-uinta and sublette herds; more individuals had moderate and high intensities than low intensities. however, patterns of intensity were not different among herd units no. moose with differing intensities of infection herd no. examined no. infected % infected (prevalence) no worms low (1-6) moderate (7-13) high (≥14) absaroka 1 0 0 1 0 0 0 big horns 20 1 5 19 0 1 0 dubois 1 0 0 1 0 0 0 jackson 10 6 60 4 6 0 0 lander 4 0 0 4 0 0 0 lincoln 13 5 38.5 8 2 2 1 snowy range1 23 19 82.6 4 13 4 1 sublette 90 47 52.2 43 21 18 8 targhee 3 2 66.7 1 1 1 0 uinta 3 2 66.7 1 1 0 1 total 168 82 48.8 86 44 26 11 table 1. prevalence of elaeophora schneideri in hunter-harvested moose, wyoming, usa, 2009. 1includes 1 animal found positive for e. schneideri for which number of worms was not recorded. elaeophora in wyoming moose – henningsen et al. alces vol. 48, 2012 40 (χ2 = 11.950; df = 8; p = 0.153). clinical signs when possible, tissues were examined for gross evidence of damage as a result of infection by e. schneideri. more thorough examinations only occurred after heads had been prepared for taxidermy or when hunters donated their antlerless specimens. of the 31 infected moose that were thoroughly examined, 10 showed visual signs of elaeophorosis: 7 displayed cropped or hardened ears, 1 had antler malformation, and 2 had cropped ears and antler malformation. three of the 10 moose with visual signs had low-intensity infections, 4 had moderate-intensity, and 3 high-intensity infections. discussion prevalence of e. schneideri in wyoming moose was much higher than anticipated. documented infections in moose have been fairly rare and noteworthy enough that individual cases have been the norm for reporting (worley et al. 1972, jensen et al. 1982, madden et al. 1991, pessier et al. 1998). the prevalence reported here is probably biased low because only the main cephalic arteries were examined for worms, yet post-mortem migration of worms occurs (adcock and hibler 1969). furthermore, there was potential for false negatives because the length of the carotid artery was often short and compromised from hunter processing; there was no corresponding risk of false positives. prevalence of e. schneideri in adult mule deer has been 100% in certain local populations in the southwestern united states (hibler and adcock 1971). prevalence has been as high as 93% in elk (hibler et al. 1969, davies 1979); high prevalence in elk occurs only in areas where mule deer also have high prevalence of infection. our study focused solely on moose so we have no analogous surveillance data from deer and moose for comparison. however, opportunistic sampling indicated ~10% prevalence of e. schneideri in mule deer in a portion of the area comprising the jackson, targhee, and sublette moose herds (j. c. henningsen, unpublished data). it may be that prevalence of e. schneideri in mule deer is too low in wyoming to generate >90% prevalence in moose; however, the snowy range had 82.6% prevalence. it has long been believed that deer are the only competent definitive hosts for e. schneideri (anderson 2001). however, some researchers (worley et al. 1972, madden et al. 1991) found gravid adult female worms in moose suggesting that they may be competent hosts. histopathologic and laboratory evidence from 3 different cases in our study support the idea that moose are a competent host for e. schneideri: 1) several microfilariae associated with an adult female worm were in a cross-section of formalin-fixed carotid artery, 2) several microfilariae were in a section of formalin-fixed skin overlying the mandibular artery at the jugular notch of the mandible, and 3) one dead microfilaria was in an overnight saline soak of fresh forehead skin. although not definitive, our evidence suggests that moose are competent hosts for e. schneideri reproduction and transmission. if this is the case, prevalence in mule deer and spatial overlap with infected mule deer could be less influential in determining e. schneideri prevalence in moose. we can only speculate about the increased prevalence of e. schneideri in wyoming moose over recent decades. because elaeophorosis was perceived to have no effect on wyoming ungulate populations, there is inconsistent historical data to make inferences. numerous case reports have expanded our knowledge of the general distribution of e. schneideri, but recent reports have not attempted to describe the ecology of e. schneideri and explain observed prevalence in wildlife (e.g., davies 1979). prevalence of the disease is presumably related to the density of definitive hosts as well as the abundance of tabanid vectors. alces vol. 48, 2012 henningsen et al. – elaeophora in wyoming moose 41 tabanid populations can be highly variable among years depending on weather conditions, because temperature and precipitation influence the timing of fly emergence, seasonal longevity, and total population size (pence 1991). thus gradual climate change has been attributed with observed and predicted increases in the effects of vector-borne parasites (patz et al. 1996, hoberg et al. 2008, laaksonen and oksanen 2009). we might have either conducted our surveillance when stochastic weather conditions were temporarily conducive for high e. schneideri transmission and/or prevalence, or changing conditions over decades has lead to higher prevalence of e. schneideri. determining the vectors of e. schneideri in wyoming and subsequently confirming the impacts of temperature and precipitation on those vectors will require further research. for the same reasons prevalence varies over time, it can exhibit high spatial variability. we found higher than expected prevalence among most herds; the snowy range and bighorns stood out as having especially high and low prevalence, respectively. moose habitat use and behavior could differ across herds in ways that affect sympatry with mule deer or susceptibility to horse flies (davies 1979). domestic livestock grazing adds another layer of complexity. livestock could either increase horse fly populations and exacerbate the transmission potential among wildlife, or dilute the effect because tabanids would prey on domestic animals instead of wildlife (davies 1979). further research is needed to fully understand the spatial dynamics of elaeophorosis in moose and other species in wyoming. on a more basic level, the effects of elaeophorosis on individual moose remain unknown. the high prevalence of apparently healthy infected moose suggests elaeophorosis is often not debilitating to this host. yet e. schneideri has been implicated in morbidity or mortality in several cases (worley et al. 1972, madden et al. 1991, pessier et al. 1998). as was demonstrated in elk (adcock and hibler 1969, hibler and adcock 1971, hibler and metzger 1974), pathogenic effects of e. schneideri on moose are more complex than a simple linear or threshold response by the host to number of worms. complications from infection could arise at a number of critical stages in the life cycle of the parasite. even slightly compromised basic functions resulting from impaired blood flow such as vision, hearing, mastication, smell, and brain function could expose individuals to malnutrition, predation, and ultimately lower survival and reproduction. while the maximum number of worms found in a hunter-harvested moose in our study was 26, preliminary necropsies of symptomatic moose have sometimes revealed double that intensity (j. c. henningsen, unpublished data). additionally, while none of the ≥8-yrold moose had high-intensity infections, this may have been an artifact of low sample size. limiting surveillance to hunter-harvested moose possibly eliminates important cases from consideration. comprehensive surveillance that includes sick and dead moose with subsequent histopathologic examinations will be valuable in elucidating impacts of this parasite on individual moose. prevalence in our study was consistent across age classes. we interpret this to mean that moose of all ages are equally susceptible to infection and that infection does not affect survival differently across ages. constant intensity of e. schneideri across ages of checked moose might additionally indicate a mechanism limiting worm burdens in moose. perhaps individuals that tolerate initial infection acquire some immunity against further infection; hibler and metzger (1974) suggested as much for infected elk. immune protection has been demonstrated in other ungulate-nematode systems involving longlived adult worms. parelaphostrongylus tenuis intensities in white-tailed deer are constrained across age classes (slomke et elaeophora in wyoming moose – henningsen et al. alces vol. 48, 2012 42 al. 1995) and prestwood and nettles (1977) demonstrated white-tailed deer acquire immunity to additional p. andersoni infections. this hypothesis presumes e. schneideri are long-lived; however, it is unknown how long e. schneideri can live in moose. other filarioid nematodes live in their definitive hosts from 2->10 years (review by gems 2000). alternatively, constant e. schneideri prevalence and intensities with increasing age of moose might simply reflect new infections occurring at a rate that essentially replace those mature nematodes that die naturally. under this scenario, immune protection would not be perfect and new infections would continue through life at some rate that is tolerated by the host. on the other hand, pathologies arising from dead nematodes in the vascular system (adcock and hibler 1969) would be inconsistent with a hypothesis where moose can survive unaffected beyond the lifespan of the adult parasite. thus this hypothesis presumes moose can tolerate not only live parasites, but individuals that die within their vascular system. regardless of the mechanism, constant intensity and prevalence across age classes indicate infected moose are surviving, hence mortality caused by e. schneideri is lower than previously suggested. while moose might not overtly succumb to elaeophorosis to the extent previously thought, prevalence of 50% is still cause for concern. at high prevalence, even a moderate proportion of infected individuals suffering from subclinical effects might impact recruitment or productivity at the population level. subsequent research on moose herds where e. schneideri is present should consider the effects of elaeophorosis and attempt to clarify its role in moose population dynamics. acknowledgements the wgfd moose working group was integral in designing the survey, and numerous wgfd field personnel conducted the field examinations. t. cornish was integral in the histopathologic and laboratory work. we are grateful to c. hibler, associate editor b. mclaren, and reviewers e. addison and m. lankester for improving the manuscript. references adcock, j. l., and c. p. hibler. 1969. vascular and neuro-ophthalmic pathology of elaeophorosis in elk. pathologia veterinaria 6: 185-213. anderson, r. c. 2001. filarioid nematodes. pages 342-356 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. iowa state university, ames, iowa, usa. boyce, w., a. fisher, h. provencio, e. rominger, j. thilsted, and m. ahlm. 1999. elaeophorosis in bighorn sheep in new mexico. journal of wildlife diseases 35: 786-789. brimeyer, d. g., and t. p. thomas. 2004. history of moose management in wyoming and recent trends in jackson hole. alces 40: 133-143. bush, a. o., k. d. lafferty, j. m. lotz, and a. w. shostak. 1987. parasitology meets ecology on its own terms: margolis et al. revisited. the journal of parasitology 83: 575-583. clark, g. g., and c. p. hibler. 1973. horse flies and elaeophora schneideri in the gila national forest, new mexico. journal of wildlife diseases 9: 21-25. davies, r. b. 1979. the ecology of elaeophora schneideri in vermejo park, new mexico. ph.d. dissertation, colorado state university, fort collins, colorado, usa. espinosa, r. h. 1983. tabanid vectors of the arterial nematode, elaeophora schneideri, in southwestern montana. m.s. thesis, montana state university, bozeman, montana, usa. hibler, c. p. 1982. elaeophorosis. pages 214-218 in e. t. thorne, n. kingston, w. r. jolley, and r. c. bergstrom, edialces vol. 48, 2012 henningsen et al. – elaeophora in wyoming moose 43 tors. diseases of wildlife in wyoming. wyoming game and fish department, cheyenne, wyoming, usa. _____, and j. l. adcock. 1971. elaeophorosis. pages 263-278 in j. w. davis and r. c. anderson, editors. parasitic diseases of wild mammals. iowa state university, ames, iowa, usa. _____, _____, r. w. davis, and y. z. abdelbaki. 1969. elaeophorosis in deer and elk in the gila forest, new mexico. bulletin of the wildlife disease association 5: 27-30. _____, _____, g. h. gates, and r. white. 1970. experimental infection of domestic sheep and mule deer with elaeophora schneideri wehr and dikmans, 1935. journal of wildlife diseases 6: 110-111. _____, and c. j. metzger. 1974. morphology of the larval stages of elaeophora schneideri in the intermediate and definitive hosts with some observations on their pathogenesis in abnormal definitive hosts. journal of wildlife diseases 10: 361-369. hoberg, e. p., l. polley, e. j. jenkins, and s. j. kutz. 2008. pathogens of domestic and free-ranging ungulates: global climate change in temperate to boreal latitudes across north america. review scientifique et technique-international office of epizootics 27: 511-528. gems, d. 2000. longevity and ageing in parasitic and free-living nematodes. biogerontology 1: 289-307. jacques, c. n., j. a. jenks, d. t. nelson, t. j. zimmerman, and m. c. sterner. 2004. elaeophorosis in free-ranging mule deer in south dakota. prairie naturalist 36: 251-54. jensen, l. a., j. c. pederson, and f. l. andersen. 1982. prevalence of elaeophora schneideri and onchocerca cervipedis in mule deer from central utah. great basin naturalist 42: 351-352. laaksonen, s., and a. oksanen. 2009. status and review of the vector-borne nematode setaria tundra in finnish cervids. alces 45: 81-84. madden, d. j., t. r. spraker, and w. j. adrian. 1991. elaeophora schneideri in moose (alces alces) from colorado. journal of wildlife diseases 27: 340-341. matthews, p. e. 2012. history and status of moose in oregon. alces 48: 63-66. mckown, r. d., m. c. sterner, and d. w. oates. 2007. first observation of elaeophora schneideri wehr and dikmans, 1935 (nematoda: filariidae) in mule deer from nebraska. journal of wildlife diseases 43: 142-144. patz, j. a., p. r. epstein, t. a. burke, and j. m. balbus. 1996. global climate change and emerging infectious diseases. journal of the american medical association 275: 217-223. pederson, j. c., l. a. jensen, and f. l. andersen. 1985. prevalence and distribution of elaeophora schneideri wehr and dikmans, 1935 in mule deer in utah. journal of wildlife diseases 21: 66-67. pence, d. b. 1991. elaeophorosis in wild ruminants. bulletin of the society for vector ecology 16: 149-160. _____, and g. g. gray. 1981. elaeophorosis in barbary sheep and mule deer from the texas panhandle. journal of wildlife diseases 17: 49-56. pessier, a. p., v. t. hamilton, w. j. foreyt, s. parish, and t. l. mcelwain. 1998. probable elaeophorosis in a moose (alces alces) from eastern washington state. journal of veterinary diagnostic investigation 10: 82-84. prestwood, a. k., and v. f. nettles. 1977. repeated low-level infection of whitetailed deer with parelaphostrongylus andersoni. journal of parasitology 58: 897-902. _____, and t. r. ridgeway. 1972. elaeophorosis in white-tailed deer of the southeastern usa.: case report and distribution. jourelaeophora in wyoming moose – henningsen et al. alces vol. 48, 2012 44 nal of wildlife diseases 8: 233-236. rózsa, l., j. reiczigel, and g. majoros. 2000. quantifying parasites in samples of hosts. journal of parasitology 86: 228-232. slomke, a. m., m. w. lankester, and w. j. peterson. 1995. infrapopulation dynamics of parelaphostrongylus tenuis in white-tailed deer. journal of wildlife diseases 31: 125-135. thomas, t. p. 2008. moose population management recommendations. wyoming game and fish department, cheyenne, wyoming, usa. waid, d. d., r. j. warren, and d. b. pence. 1984. elaeophora schneideri wehr and dikmans, 1935 in white-tailed deer from the edwards plateau of texas. journal of wildlife diseases 20: 342-345. weinmann, e. j., j. r. anderson, w. m. longhurst, and g. connolly. 1973. filarial worms of columbian black-tailed deer in california 1. observations in the vertebrate host. journal of wildlife diseases 9: 213-220. worley, d. e. 1975. observations on epizootiology and distribution of elaeophora schneideri in montana ruminants. journal of wildlife diseases 11: 486-488. _____, c. k. anderson, and k. r. greer. 1972. elaeophorosis in moose from montana. journal of wildlife diseases 8: 242-244. moose status and management in montana nicholas j. decesare1, ty d. smucker2, robert a. garrott3, and justin a. gude4 1montana fish, wildlife and parks, 3201 spurgin road, missoula, montana, usa 59804; 2montana fish, wildlife and parks, 4600 giant springs road, great falls, montana, usa 59405; 3fish and wildlife ecology and management program, department of ecology, montana state university, 310 lewis hall, bozeman, montana, usa 59717; 4montana fish, wildlife and parks, 1420 east sixth avenue, helena, montana, usa 59620. abstract: moose (alces alces) are currently widespread across montana where regulated moose hunting has occurred since 1872, >140 years ago. the number of annual moose hunting permits has averaged 652 over the past 50 years. the popular permits are allocated via a random drawing, with an annual average of ∼23,000 applicants in 2008–2012 who faced a 1.9% chance of success. monitoring of moose largely occurs through annual harvest statistics collected via post-season phone surveys. recent harvest statistics indicate lower hunter success, increased effort, and lower kill per unit effort, concurrent with >50% reduction in available permits since the 1990s. aerial surveys also show decline in calf:adult ratios. in combination, these data suggest a declining trend in the statewide population, despite some ambiguity of certain data. potential limiting factors include harvest, predation, vegetative succession and degradation, parasites, and climatic conditions, which were all identified as concerns in surveys of state biologists. accordingly, montana fish, wildlife and parks will direct funds derived from moose permit auctions toward calibrating and refining statewide monitoring methods and research of population dynamics and potential limiting factors of montana moose. alces vol. 50: 35–51 (2014) key words: alces alces shirasi, elaeophora schneideri, harvest statistics, hunter success rates, kpue, montana, shiras moose, subspecies. moose (alces alces) colonized north america roughly 14,000 years ago and have since occupied much of alaska, canada, and northern portions of the contiguous united states (hundertmark et al. 2002, hundertmark and bowyer 2004). considered rare throughout the u.s. rocky mountains until the mid-1800s (karns 2007), their earlier presence in several regions of montana were documented by the lewis and clark expedition in 1805–1806, alexander ross in 1824, and others (reviewed by schladweiler 1974). widespread prevalence of moose in montana during early settlement is supported to some extent by a review of place names throughout the state, including at least 22 creeks and 6 lakes bearing “moose” in their names (schladweiler 1974). regulation of moose hunting in montana began in 1872, yet after subsequent decline brought near extirpation, hunting was closed statewide for almost 50 years from 1897–1945 (stevens 1971). in 1910, the state warden estimated a rebounding population of 300 moose as the result of “ten years of careful protection” (state of montana 1910). allowable harvest began again in 1945 with 90 permits issued. subsequently, annual permit numbers rose quickly to a maximum of 836 in 1962, and thereafter averaged 652 until 2012 (fig. 1a). the limited number of permits have been allocated via a random drawing process. in 2008– 2012, an average of ∼23,000 hunters applied annually for <600 permits, with a 1.9% chance of success. beginning in 1988, one 35 additional permit has been auctioned to the highest bidder, with revenue directly earmarked for moose management or research. additionally, since 2006 applicants can purchase unlimited numbers of chances at drawing one available moose “super-tag,” valid in any permitted hunting district. along with super-tag chances for other species, revenue from these sales is earmarked for hunting access programs and wildlife habitat conservation. moose in montana typically occur at relatively low density and are vastly outnum‐ bered by seasonally sympatric elk (cervus elaphus), white-tailed deer (odocoileus virginianus), and mule deer (o. hemionus) populations. relative ungulate densities are reflected in their harvest level; in 2012 hunters harvested ∼274 moose versus >20,000 elk, 37,000 mule deer, and 49,000 whitetailed deer. rigorous statewide abundance estimates of moose are lacking, but based on professional opinion among regional management biologists in 2006, the estimated statewide population was 4,500–5,500, albeit without estimable accuracy or precision fig. 1. statewide and regional trends of a) number of permits issued and b) hunter success rates (number harvested/number of permits issued) for moose in montana, 1945–2012. 36 moose status in montana – decesare et al. alces vol. 50, 2014 (smucker et al. 2011). moose are distributed widely across western portions of the state, with lower density extending to the east, as reflected by the current distribution of allowable harvest (fig. 2). the majority of annual permits are offered in the southwest (56% in region 3) and northwest (25% in region 1). in recent decades moose have continued to colonize, or re-colonize, portions of central and eastern montana allowing for added harvest opportunity. moose occupy forested landscapes throughout western montana ranging from regenerating areas within dense mesic forest, such as the cabinet mountains in the northwest, to areas with extensive willow fen habitat, as found within the centennial and big hole valleys in the southwest. moose in the prairie landscapes of the east inhabit wetlands, particularly along the missouri river, other riparian corridors, and areas supporting healthy willow communities. taxonomy moose within the rocky mountains of the united states have historically been classified as shiras moose (a. a. shirasi). the subspecies was first described in wyoming (nelson 1914), and subsequent morphological sampling by peterson (1952) suggested its range to extend northward through montana and into a zone of intergradation with the northwestern subspecies (a. a. andersoni) in western alberta and eastern british columbia. while genetic evaluation of subspecies designations using mitochondrial haplotypes generally upheld some level of differentiation between shiras moose in colorado and representative samples from other subspecies (hundertmark et al. 2003), such methods have not been applied to evaluate moose in montana. particular interest in subspecies distinctions has arisen recently with anecdotal evidence of immigration of moose in northern and northeastern montana from expanding populations in southern alberta and saskatchewan. for example, the boone and crockett club has traditionally used the canadian border to distinguish shiras from “canada” moose (a designation that essentially lumps northwestern and eastern [a. a. americana] subspecies into a fig. 2. number of moose permits issued by moose hunting district in montana, 2012. alces vol. 50, 2014 decesare et al. – moose status in montana 37 single category) in scoring and record keeping of trophy animals. the advent of hunting in northeastern montana’s hunting district 600 has prompted informal discussion of classifying moose harvested within northern montana and east of interstate highway i-15 as canada moose, though none have been submitted for scoring to date (personal communication, j. spring, boone and crockett club, missoula, montana). further sampling and analysis of population genetic structure of moose within and surrounding montana may be needed to evaluate and update the subspecies range extents in the region. monitoring methods and data resources have been limited for monitoring moose given their relatively low abundance and hunting opportunity compared to other montana ungulates. post-season surveys of permit holders have been used to estimate wildlife harvest since 1941 (cada 1983, lukacs et al. 2011), and in recent years phone surveys are used to collect annual harvest data. montana fish, wildlife & parks (mfwp) attempts to survey every permit holder to measure hunter success and effort, and adjusts harvest estimates according to annual hunter responses and rates. during 2005–2012, surveys yielded hunter response rates of 81–96% and statewide harvest estimates with coefficients of variation of 0.6–2.3%. these are the most consistent monitoring data through time and across the state, and are estimated distinctly for each district and permit type. though potentially less precise than more intensive aerial survey methods, hunter statistics provide a cost-effective means for monitoring moose population trend (boyce et al. 2012). generally, there are 4 statistics computed annually that provide insight into potential moose population trends: 1) number of permits issued, 2) hunter success rate, 3) days of moose hunter effort, and 4) kills per unit effort (kpue). beyond harvest statistics, mfwp biologists in most regions have made at least intermittent efforts to conduct aerial surveys, but sustained survey efforts are limited to the few areas with historically higher density. in the northwest (region 1), december helicopter surveys have been conducted annually since 1985 in a subset of moose hunting districts centered around the cabinet, purcell, salish, and whitefish mountains. moose in this densely forested region selectively use and are more visible in regenerating (15–30 years old) stands during early winter, but move into mature, closed-canopy forest as winter progresses (matchett 1985). while an explicit model with sightability covariates has not been developed for the area, an early 1990s mark-resight study with 81 neckbanded individuals produced average sightability estimates of 0.53–0.55 (brown 2006). in the southwest (region 3), fixedwing aerial surveys have been conducted during most years since the 1960s in the hunting districts of the big hole and centennial valleys. these surveys typically yield calf:adult ratios and uncorrected minimum counts, and their timing (september–may) has varied considerably by year and district. sporadic helicopter and fixed-wing aircraft surveys have occurred in other lower-density regions of the state including regions 2, 4, and 5. the mfwp is currently exploring the utility and cost-effectiveness of standardizing and coordinating survey efforts. the mfwp is also exploring the utility of cheaper monitoring methods including hunter sighting surveys at voluntary hunter check stations, and post-season phone surveys used to measure deer and elk harvests. while both the observation rate and age ratios collected from hunter sightings can be indicative of population trends (ericsson and wallin 1999, bontaities et al. 2000), there is potential to incorporate spatial and temporal attributes of sightings data into a patch occupancy modeling framework 38 moose status in montana – decesare et al. alces vol. 50, 2014 similar to recent efforts with hunter sightings of wolves (canis lupus; rich et al. 2013). additionally, the mfwp is exploring the cost-effectiveness of estimating population trends using the fates and reproductive status of marked individuals (sensu lukacs et al. 2009) which can be integrated into population models that estimate annual growth rate (decesare et al. 2012). moose harvest statistics and trend as a consequence of perceived population declines and declining population indices from harvest data in recent decades, the number of moose permits issued in montana was reduced by 53% (769 to 362) between 1995 and 2012 (fig. 1a). most reductions were in areas with traditionally the most available permits (regions 1 and 3). in contrast, the first 2 permits ever offered in northeastern montana (region 6) were added in 2008. notably, the 2010 hunting season was the first in more than 50 years when the number of statewide permits was <500 (fig. 1a). statewide hunter success is estimated as the number of moose harvested relative to the number of permits issued, averaging 78.4% during regulated moose hunting in montana (1945–2012; fig. 1b). this success rate is similar to that in adjacent idaho (61– 85%; toweill and vecellio 2004), but relatively higher than in other areas with typically more moose and moose hunters such as alberta (30–50%; boyce et al. 2012), alaska (28–37%; schmidt et al. 2005), newfoundland (25–54%; fryxell et al. 1988), and ontario (36–40%; hunt 2013). from 2008– 2012, success rates (average = 73.4%) were lower than the previous 20-year average (83.7%; t = 2.07, 23 df, p < 0.001). additionally, hunter effort, defined as the number of days spent hunting moose per hunter, increased from 6.3 in 1986 to ≥11 days/ hunter in 2010–2012 (fig. 3). similarly, kill per unit effort (kpue) that integrates hunter success and effort statistics into a metric of fig. 3. statewide annual averages of moose hunter effort (days per hunter) and moose kill per unit effort (kpue) in montana, 1986–2012. alces vol. 50, 2014 decesare et al. – moose status in montana 39 hunter efficiency, declined >50% from >0.14 to <0.07 moose killed per hunter-day over the same time period (fig. 3). the kpue for antlered bull-specific tags also varied by hunting district level (fig. 4), reflecting regional differences in moose distribution and ecotypes (e.g., more closed forests in the northwest compared to more open foothills and large riparian complexes in the southwest). in combination, lower hunter success and kpue, increased hunter effort, and a concurrent >50% reduction in available permits are indicative of a declining statewide population trend. in ontario, years with fewer permits resulted in increased hunter success rate, even after accounting for changes in underlying moose density (hunt 2013), which suggests that hunter behavior can complicate interpretation of hunter statistics (bowyer et al. 1999, schmidt et al. 2005). change in permit type over space and time (e.g., shifting between antlered bull, antlerless, or either-sex permits) can also complicate or confound interpretation of hunter statistics. for example, recent (2008–2012) increases in kpue also coincide with a prescribed reduction in the antlerless harvest that may reduce kpue by limiting the proportion of animals hunters are allowed to harvest, regardless of underlying population dynamics. thus, we cautiously interpret harvest statistics as imperfect indices. concurrent declines in available permits, success rates, and kpue may result from population decline and/or reflect other confounding factors. in addition to statewide hunter statistics, regional calf:adult ratios in areas with consistent aerial survey data indicate decline in recruitment (fig. 5). three distinct survey areas show significant (p < 0.05) overall declines in ratios since 1980, though the temporal pattern of decline may be non-linear with subsequent stability at a lower level in recent years (fig. 5). low or declining recruitment is often associated with declin‐ ing ungulate populations (e.g., decesare et al. 2012), so these data may be corroborative with harvest statistics that indicate a fig. 4. bull moose kills per unit effort (kpue; effort recorded in days) per moose hunting district by hunters carrying antlered-bull-only permits in montana, 2012. 40 moose status in montana – decesare et al. alces vol. 50, 2014 declining moose population. however, declining recruitment may also reflect an ungulate population approaching carrying capacity (gaillard et al. 1998, eberhardt 2002), so this index also does not unambiguously indicate decline. biologist interviews: local trends and management in 2010, we used structured interviews of 20 mfwp and cooperating agency biologists to assess the state of knowledge regarding moose population status, management, and factors of concern within montana (appendix a). a majority (63%) of responding biologists reported “decreasing” or “stable to decreasing” trends in their populations, with stable and increasing trends reported in some areas. these trend assessments are tempered, however, because only 10% of biologists had adequate data for making management decisions; 55 and 35% described their data as partially inadequate and inadequate, respectively. lastly, when asked about factors that potentially limit local moose populations, biologist listed predation (70%), habitat succession (45%), mfwppermitted hunter harvest (45%), parasites and/or disease (40%), native american hunter harvest (30%), and habitat loss or fragmentation (15%). potential limiting factors many factors may currently limit moose abundance and distribution including hunter harvest, predation, habitat succession, parasite and disease prevalence, and climatic conditions. the relative importance of these factors has likely changed over time. overharvest may have been responsible for decline in moose numbers in the late 1800s (stevens 1971). by the early 1970s, research fig. 5. annual moose calves per 100 adult recruitment data and associated linear regression trend lines calculated from fixed-wing and helicopter late winter aerial surveys in 3 regions of montana, 1976–2010. alces vol. 50, 2014 decesare et al. – moose status in montana 41 in southwest montana indicated that hunter harvest and nutritional inadequacies were the most important factors limiting moose populations, whereas parasites, disease, and predation had little direct effect on mortality rates (schladweiler 1974). presently there is a need to re-evaluate the relative importance of potential limiting factors in light of recent changes in many of these factors and subsequent monitoring and research in montana and elsewhere. hunter harvest the goals and objectives behind moose hunter harvest quotas vary somewhat across mfwp regional jurisdictions. managers in regions 1 and 3, where populations are largest, generally aim to sustainably maximize hunter opportunity and minimize landowner conflicts (e.g., greater numbers of permits that include either-sex or antlerless opportunities), whereas regions 2, 4, 5, and 6 manage harvest with less intent to affect moose population dynamics (e.g., bull-only hunting or low permit numbers). during the past 2 decades, numbers of antlerless permits were increased substantially in certain areas, particularly in region 3, in response to depredation complaints, perceptions that moose were unfavorably limiting vegetative growth (i.e., riparian plants), and high moose counts on aerial surveys. these prescriptive increases in moose permits were intended to induce local declines in some hunting districts. statewide, the sex ratio of harvested adult moose (i.e., excluding calves) averaged 28% female in 1971–2008, but dropped to an average of 14% in 2009–2012; female harvest is through either-sex and antlerlessonly permits. in region 1, either-sex tags were issued historically, and harvest was typically skewed heavily towards males; the 1984–2004 harvest was 78% bulls, 19% cows, and 3% calves. as of 2012, all permits in this region were changed to antlered-bull only. in region 3, permits have been typically specified as antleredor antlerlessonly, which is more restrictive to hunters but facilitates targeted management. additional moose harvest by members of the confederated salish and kootenai tribes (cskt) is permitted off-reservation by the hellgate treaty of 1855. one permit per year is allowed to each interested tribal member for hunting on primarily federal land, with mandatory reporting to cskt officials. while the sample size of animals harvested is lower than that regulated by mfwp, these harvest data provide additional opportunity for indexing population trend and are without confounding changes in permit number and type. trends in tribal harvest are similar to that of the mfwp (fig. 6); total harvest peaked in 1991 at 97 representing an additional 16.3% to the mfwp harvest of 595, and in 2012 the tribal harvest was only 18, an additional 6.6% to the mfwp harvest of 274 moose. we point out that interpretation of tribal harvest statistics with respect to the rate of population change is also not unambiguous. while some evidence exists of reduced success by tribal hunters (fig. 6), a portion of the decline can probably be attributed to fewer permit requests. also, these data do not include information about hunter effort or tribal interest in hunting other game species as allowed by treaty rights. illegal harvest of moose also occurs but has not been quantified to date. data from idaho suggest that illegal harvest can represent upwards of 31–50% of mortality (pierce et al. 1985, toweill and vecellio 2004), warranting explicit monitoring and documentation of such in montana. predation after decades of predator control in the early and mid-1900s, and subsequent recovery efforts in the late 1900s, montana currently hosts widespread populations of 42 moose status in montana – decesare et al. alces vol. 50, 2014 grizzly bears (ursus arctos), black bears (ursus americanus), wolves, mountain lions (puma concolor), and coyotes (canis latrans). while predation was not considered a concern 40 years ago (schladweiler 1974), the expanded composition and abundance of predator species may have the potential to limit local moose populations. predation was the most common concern of regional biologists relative to moose population dynamics. research on winter prey selection by recolonizing wolves in the north fork of the flathead river drainage from 1986–1996 indicated that while wolves disproportionately used areas where deer were concentrated, they preferentially killed larger moose and elk over more abundant deer. moose, particularly calves and cows, comprised a greater proportion of wolf kills as winter progressed (kunkel et al. 2004). however, annual survival of 32 adult female moose monitored concurrently in the north fork (1990–1992) was relatively high (0.9137 ± 0.0773; langley 1993), with 3 mortalities attributable to predation (1 wolf and 2 grizzly bear). in a recent dietary study of 12 wolf packs in northwest montana, moose was the most common prey item based on stable isotope analysis, constituting an average of 41% of the diet; however, these results were not supported by scat analysis from a sub-set of 4 packs in which moose averaged 18% of the diet (derbridge et al. 2012). high densities of elk and deer throughout much of the rocky mountain region may support higher predator populations and facilitate increased predation rates on sympatric moose via apparent competition (holt 1977). in such cases, a less abundant, secondary prey species can become more vulnerable to depensatory predation when faced with predator populations boosted by more numerous primary prey species (messier 1995, garrott et al. 2009). while fig. 6. moose harvest and hunter success rates by members of the confederated salish and kootenai tribes (cskt) off-reservation (primarily on federal lands in western montana), 1986–2012. alces vol. 50, 2014 decesare et al. – moose status in montana 43 moose across much of canada have been attributed with the role of a primary prey species driving predator-mediated declines in less abundant woodland caribou (rangifer tarandus caribou) populations (decesare et al. 2010), they may in fact be vulnerable themselves to such a mechanism within the elkand deer-dominated prey populations of montana. the effects of apparent competition from increased predation risk may be reduced somewhat by differential selection of winter and calving habitat among ungulates. moose in montana typically use higher elevations during winter and may accordingly spatially separate themselves from increased predation risk in some cases (jenkins and wright 1988, burcham et al. 2000, kunkel and pletscher 2001). the ultimate effect of predators on prey dynamics varies according to predation rates on different age classes (gervasi et al. 2011), as well as with differences in the nutritional quality of prey habitat (melis et al. 2009). because moose may have colonized many areas of western montana when predators were largely reduced, it is uncertain to what extent recolonized and expanding predator populations pose an additive source of mortality on local populations. in such cases, management of moose populations may require that predation rates be accounted for when deriving sustainable harvest quotas (hobbs et al. 2012). vegetative succession and degradation moose habitat requirements and preferences have been well documented (reviewed by peek 2007, shipley 2010). moose in montana use a variety of mid to high elevation forest types in summer, including closed canopy lodgepole pine (pinus contorta) and subalpine fir (abies lasiocarpa) forests, as well as aspen (populus tremuloids) and willow (salix spp.) stands, mountain parklands, and alpine meadows (knowlton 1960, peek 1962, schladweiler 1974). during winter, they often forage on willow where available, and snow depth can either restrict local use and movement (burkholder 2012) or shift use to conifer forests (tyers 2003). many studies of shiras moose in the rocky mountains have documented the importance of early successional habitats (peek 2007). large-extent wildfires in 1910, 1919, and 1929 converted much of the conifer forest in northwest montana to early-seral stages and moose populations in the state appeared to increase in response (brown 2006). while the positive association with early successional habitat following wildfires is well documented, negative impacts of the 1988 fires in yellowstone national park contradict this tenant (tyers 2006; vartanian et al. 2011). during the 1950s–1980s, timber harvest became the dominant form of disturbance shaping conifer forests in the west and was generally favorable to moose, particularly 10–30 years following harvest (eastman 1974, matchett 1985, telfer 1995). it is believed that the high amount of timber harvest combined with fire history may have set the stage for abundant moose populations through the early 1990s (brown 2006). a time-lagged decrease in early-seral forests has presumably resulted from reduced timber harvesting since the late 1980s (spoelma et al. 2004). riparian areas have been severely degraded globally by a variety of stressors (richardson et al. 2007), and in some parts of the western united states, cottonwoodwillow riparian habitats have been reduced by as much as 90–95% (johnson and carothers 1982). historically, persistent riparian habitat along rivers and streams may have provided long-term stability to moose populations and functioned as corridors to allow moose to expand into ephemeral post-fire habitats (peek 2007). in many areas of montana, habitat management has focused on restoration of riparian areas via fencing and 44 moose status in montana – decesare et al. alces vol. 50, 2014 grazing management with the goal of restoring robust willow communities. parasites moose are exposed to a suite of parasites with potential implications for population dynamics. winter ticks (dermacentor albipictus) are known to occur in moose range across much of north america south of 60° n latitude (samuel 2004), and have been detected in disparate regions and vegetation types of montana (n. decesare, unpublished data). while data are not available concerning the demographic impact of ticks on moose in montana, negative effects of ticks on moose populations have been well documented elsewhere (samuel 2007, musante et al. 2010). given that die-offs have been known to occur synchronously across various portions of moose range (delgiudice et al. 1997), impacts of tick epizootics on moose in montana seem likely. giant liver flukes (fascioloides magna) were reported as the greatest single source of mortality for a declining moose population in northwest minnesota (murray et al. 2006, lankester and foreyt 2011). such effects of flukes on moose mortality may be accentuated when individuals are malnourished (lankester and samuel 2007). both f. magna and the common liver fluke (f. hepatica) have been documented widely within montana's cattle populations (knapp et al. 1992), and multiple species of lymnaid snails, the intermediate host, are also known to occur (dunkel et al. 1996). data concerning infection rates or impacts of flukes on moose or other wild ungulates in montana are lacking. also of concern in minnesota and elsewhere in eastern north american is the meningial worm (parelaphostrongylus tenuis). prevalent in central and eastern moose populations, this parasite is carried by white-tailed deer, transmitted by terrestrial gastropod intermediate hosts, and is commonly associated with moose declines in areas of high overlap with dense deer populations (lankester 2010). while p. tenuis has not been documented in montana, detection of infected white-tailed deer in western north dakota suggest the possibility of intermittent spread into portions of montana (maskey 2008). the arterial worm (elaeophora schneideri) is a filarioid nematode found in the common carotid and internal maxillary arteries of ungulates in the west and southwestern us (henningsen et al. 2012). mule deer are definitive hosts of carotid worms, while moose and other ungulates are aberrant hosts, susceptible to blockage of blood to the optic nerve, ears, and brain and related symptoms such as blindness, ataxia, necrosis of the muzzle and nostrils, and emaciation (hibler and metzger 1974). e. schneideri was first detected in moose in montana in 1971 (worley et al. 1972), and subsequent sampling of 74 harvested moose detected carotid worms in 3 (4.0%; worley 1975). more recently, approximately 30% prevalence was detected in montana among 94 moose harvested in 2009–10 (j. ramsey, mfwp, unpublished data) and 49% prevalence (n = 165) was detected in wyoming (henningsen et al. 2012). while infection is not necessarily lethal, increasing prevalence and the potential for subclinical effects warrant further investigation. climate moose in north america occur across a great range of latitudes (40° n to 70° n), though generally are best-adapted for cold climates (renecker and hudson 1986). winter severity can affect physical condition (cederlund et al. 1991) and fecundity (solberg et al. 1999) of moose, yet recent attention has been given largely to concerns over warm temperatures. a small sample (n = 2) of captive moose in alberta exhibited metabolic and respiratory signs of heat stress alces vol. 50, 2014 decesare et al. – moose status in montana 45 at temperatures above −5°c and 14°c in winter and summer, respectively (renecker & hudson 1986). in minnesota, a heat stress index based on these thresholds explained >78% of the annual variability in moose survival (lenarz et al. 2009), and annual population growth rates decreased with increasing summer temperatures (murray et al. 2006). concerns over heat stress effects on moose are compounded by predicted patterns of future climatic warming across southern moose ranges (lenarz et al. 2010), yet much remains unclear and the relationships in minnesota were strictly correlative. it is not known whether the mechanism linking temperature to demography is a direct link between heat stress and malnutrition (murray et al. 2006) or an indirect link via parasites or other mortality agents (samuel 2007). increased mortality as a result of heat stress is likely to result in decreased abundance and a contraction in moose distribution along the southern range extent, yet local expansions of moose in other southern jurisdictions (e.g., base et al. 2006, wolfe et al. 2010, wattles and destefano 2011) and an ontario field study (lowe et al. 2010) do not directly support this hypothesis. within montana it is unclear whether any climatic variables underlie spatial variation in the productivity of local populations. research needs and future directions comprehensive review of the current status of moose and methods in practice for monitoring and management revealed 3 primary research needs in montana: 1) calibration of various trend indices to evaluate agreement and uncertainty regarding moose population trends, 2) development or refinement of monitoring programs to produce consistent data at appropriate scales to inform harvest or habitat management decisions, and 3) research into rates of adult survival and recruitment and the potential limiting factors of each. accordingly, during fiscal year 2012–2013 the mfwp began directing moose permit auction funds toward a new research program to address these research needs. generally speaking, the work aims to provide rigorous and reliable information as a foundation for understanding moose population dynamics and management practices in montana. acknowledgements funding for this work was provided by the sale of hunting and fishing licenses in montana and matching pittman-robertson grants to the montana department of fish, wildlife and parks. k. alt (retired), h. burt, g. taylor, m. thompson, and j. williams provided valuable insights on regional histories and priorities for moose management and facilitated communication with mfwp area biologists v. boccadori, r. brannon, j. cunningham, j. brown (retired), v. edwards, c. fager, a. grove, c. jourdannais, j. kolbe, b. lonner, g. olson (retired), r. rauscher, j. sika, b. sterling, s. stewart, t. thier, r. vinkey, j. vore, and a. wood. d. becker, j. cunningham, v. edwards, and j. newby were especially helpful with tracking down and interpreting moose data and reports. j. warren provided valuable insight on moose research and habitat management as well as database development. a. messer provided valuable guidance and advice on available gis data, database design, and data standardization and capture. k. smucker and j. newby provided valuable comments on previous versions of this manuscript. j. van andel provided invaluable administrative support. literature cited base, d. l., s. zender, and d. martorello. 2006. history, status, and hunter harvest 46 moose status in montana – decesare et al. alces vol. 50, 2014 of moose in washington state. alces 42: 111–114. bontaities, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69–75. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–90. boyce, m. s., p. w. j. baxter, and h. p. possingham. 2012. managing moose harvests by the seat of your pants. theoretical population biology 82: 340–347. brown, j. 2006. moose management in northwest montana: region 1 annual report. montana fish, wildlife and parks, libby, montana, usa. burcham, m., c. l. marcum, d. mccleerey, and m. thompson. 2000. final report: study of sympatric moose and elk in the garnet range of western montana, 1997–2000. university of montana, missoula, montana, usa. burkholder, b. o. 2012. seasonal distribution, winter habitat selection and willow utilization patterns of the shiras moose on the mount haggin wildlife management area. m.s. thesis, montana state university, bozeman, montana, usa. cada, j. d. 1983. evaluations of the telephone and mail survey methods of obtaining harvest data from licensed sportsmen in montana. pages 117–128 in s. l. beasom and s. f. roberson, editors. game harvest management. caesar kleberg research institute, kingsville, texas, usa. cederlund, g. n., h. k. g. sand, and å. pehrson. 1991. body mass dynamics of moose calves in relation to winter severity. journal of wildlife management 55: 675–681. decesare,n. j., m. hebblewhite, m. bradley, k. g. smith, d. hervieux, and l. neufeld. 2012. estimating ungulate recruitment and growth rates using age ratios. journal of wildlife management 76: 144–153. ———, ———, h. s. robinson, and m. musiani. 2010. endangered, apparently: the role of apparent competition in endangered species conservation. animal conservation 13: 353–362. delgiudice, g. d., r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restriction, ticks, and numbers of moose on isle royale. journal of wildlife management 61: 895–903. derbridge, j. j., p. r. krausman, and c. t. darimont. 2012. using bayesian stable isotope mixing models to estimate wolf diet in a multi-prey ecosystem. journal of wildlife management 76: 1277–1289. dunkel, a. m., m. c. rognlie, g. rob johnson, and s. e. knapp. 1996. distribution of potential intermediate hosts for fasciola hepatica and fascioloides magna in montana, usa. veterinary parasitology 62: 63–70. eastman, d. s. 1974. habitat use by moose of burns, cutovers and forests in northcentral british columbia. proceedings of the north american moose conference workshop 10: 238–256. eberhardt, l. l. 2002. a paradigm for population analysis of long-lived vertebrates. ecology 83: 841–2854. ericsson, g., and k. wallin. 1999. hunter observations as an index of moose alces alces population parameters. wildlife biology 5: 177–185. fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52: 14–21. gaillard, j. m., m. festa-bianchet, and n. g. yoccoz. 1998. population dynamics of large herbivores: variable recruitment with constant adult survival. trends in ecology & evolution 13: 58–63. garrott, r. a., p. j. white, m. s. becker, and c. n. gower. 2009. apparent alces vol. 50, 2014 decesare et al. – moose status in montana 47 competition and regulation in a wolfungulate system: interactions of life history characteristics, climate, and landscape attributes. pages 519–540 in r. a. garrott, p. j. white, and f. g. r. watson, editors. the ecology of large mammals in central yellowstone: sixteen years of integrated field studies. elsevier, san diego, california, usa. gervasi, v., e. b. nilsen, h. sand, m. panzacchi, g. r. rauset, h. c. pedersen, j. kindberg, p. wabakken, b. zimmermann, j. odden, o. liberg, j. e. swenson, and j. d. c. linnell. 2011. predicting the potential demographic impact of predators on their prey: a comparative analysis of two carnivore– ungulate systems in scandinavia. journal of animal ecology 81: 443–454. henningsen, j. c., a. l. williams, c. m. tate, s. a. kilpatrick, and w. d. walter. 2012. distribution and prevalence of elaeophora schneideri in moose in wyoming. alces 48: 35–44. hibler, c. p., and c. j. metzger. 1974. morphology of the larval stages of elaeophora schneideri in the intermediate and definitive hosts with some observations on their pathogenesis in abnormal definitive hosts. journal of wildlife diseases 10: 361–369. hobbs, n. t., h. andrén, j. persson, m. aronsson, and g. chapron. 2012. native predators reduce harvest of reindeer by sámi pastoralists. ecological applications 22: 1640–1654. holt, r. d. 1977. predation, apparent competition, and the structure of prey communities. theoretical population biology 12: 197–229. hundertmark, k. j., and r. t. bowyer. 2004. genetics, evolution, and phylogeography of moose. alces 40: 103–122. ———, ———, g. f. shields, and c. c. schwartz. 2003. mitochondrial phylogeography of moose (alces alces) in north america. journal of mammalogy 84: 718–728. ———, g. f. shields, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. schwartz. 2002. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. molecular phylogenetics and evolution 22: 375–387. hunt, l. m. 2013. using human-dimensions research to reduce implementation uncertainty for wildlife management: a case of moose (alces alces) hunting in northern ontario, canada. wildlife research 40: 61–69. jenkins, k. j., and r. g. wright. 1988. resource partitioning and competition among cervids in the northern rocky mountains. journal of applied ecology 25: 11–24. johnson, r. r., and s. w. carothers. 1982. riparian habitats and recreation: interrelationships and impacts in the southwest and rocky mountain region. usad forest service, rocky mountain forest and range experiment station, fort collins, colorado, usa. karns, p. d. 2007. population distribution, density, and trends. pages 125–140 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. knapp, s. e., a. m. dunkel, k. han, and l. a. zimmerman. 1992. epizootiology of fascioliasis in montana. veterinary parasitology 42: 241–246. knowlton, f. f. 1960. food habits, movements and populations of moose in the gravelly mountains, montana. journal of wildlife management 24: 162–170. kunkel, k. e., and d. h. pletscher. 2001. winter hunting patterns of wolves in and near glacier national park, montana. journal of wildlife management 65: 520–530. ———, ———, d. k. boyd, r. r. ream, and m. w. fairchild. 2004. factors correlated with foraging behavior of wolves in and near glacier national park, 48 moose status in montana – decesare et al. alces vol. 50, 2014 montana. journal of wildlife management 68: 167–178. langley, m. a. 1993. habitat selection, mortality and population monitoring of shiras moose in the north fork of the flathead river valley, montana. m.s. thesis, university of montana, missoula, montana, usa. lankester, m. w. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53–70. lankester, m. w., and w. j. foreyt. 2011. moose experimentally infected with giant liver fluke (fascioloides magna). alces 47: 9–15. ———, and w. m. samuel. 2007. pests, parasites, and disease. pages 479–517 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. ———, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–510. lowe, s. j., b. r. patterson, and j.a. schaefer. 2013. lack of behavioral response of moose (alces alces) to high ambient temperatures near the sourthern periphery of their range. canadian journal of zoology 88: 1032–1041. lukacs, p. m., j. a. gude, r. e. russell, and b. b. ackerman. 2011. evaluating costefficiency and accuracy of hunter harvest survey designs. wildlife society bulletin 35: 430–437. ———, g. c. white, b. e. watkins, r. h. kahn, b. a. banulis, d. j. finley, a. a. holland, j. a. martens, and j. vayhinger. 2009. separating components of variation in survival of mule deer in colorado. journal of wildlife management 73: 817–826. maskey, j. j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph. d. thesis, university of north dakota, grand forks, north dakota, usa. matchett, m. r. 1985. habitat selection by moose in the yaak river drainage, northwestern montana. alces 21: 161–190. melis, c., b. jędrzejewska, m. apollonio, k. a. bartoń, w. jędrzejewski, j. d. c. linnell, i. kojola, j. kusak, m. adamic, s. ciuti, i. delehan, i. dykyy, k. krapinec, l. mattioli, a. sagaydak, n. samchuk, k. schmidt, m. shkvyrya, v. e. sidorovich, b. zawadzka, and s. zhyla. 2009. predation has a greater impact in less productive environments: variation in roe deer, capreolus capreolus, population density across europe. global ecology and biogeography 18: 724–734. messier, f. 1995. on the functional and numerical responses of wolves to changing prey density. ecology and conservation of wolves in a changing world. canadian circumpolar institute, occasional publication 35: 187–198. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. nelson, e. w. 1914. description of a new subspecies of moose from wyoming. proceedings of the biology society of washington 27: 71–74. peek, j. m. 1962. studies of moose in the gravelly and snowcrest mountains, alces vol. 50, 2014 decesare et al. – moose status in montana 49 montana. journal of wildlife management 26: 360–365. ———. 2007. habitat relationships. pages 351–375 in in a.w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. peterson, r. l. 1952. a review of the living representatives of the genus alces. royal ontario museum. life sciences division, toronto, ontario, canada. pierce, d. j., b. w. ritchie, and l. kuck. 1985. an examination of unregulated harvest of shiras moose in idaho. alces 21: 231–252. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. rich, l. n., e. m. glenn, m. s. mitchell, j. a. gude, k. podruzny, c. a. sime, k. laudon, d. e. ausband, and j. d. nichols. 2013. estimating occupancy and predicting numbers of gray wolf packs in montana using hunter surveys. journal of wildlife management 77: 1280–1289. richardson, d. m., p. m. holmes, k. j. esler, s. m. galatowitsch, j. c. stromberg, s. p. kirkman, s. p. pysek, and r. j. hobbs. 2007. riparian vegetation: degradation, alien plant invasions, and restoration projects. diversity and distributions 13: 126–139. samuel, w. m. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. ———. 2007. factors affecting epizootics of winter ticks and mortality of moose. alces 43: 39–48. schladweiler, p. 1974. ecology of shiras moose in montana. montana department of fish and game, helena, montana, usa. schmidt, j. i., j. a. y. m. ver hoef, j. a. k. maier, and r. t. bowyer. 2005. catch per unit effort for moose: a new approach using weibull regression. journal of wildlife management 69: 1112–1124. shipley, l. 2010. fifty years of food and foraging in moose: lessons in ecology from a model herbivore. alces 46: 1–13. smucker, t., r. a. garrott, and j. a. gude. 2011. synthesizing moose management, monitoring, past research and future research needs in montana. montana fish, wildlife, and parks, helena, montana, usa. solberg, e. j., b. e. saether, o. strand, and a. loison. 1999. dynamics of a harvested moose population in a variable environment. journal of animal ecology 68: 186–204. spoelma, t. p., t. a. morgan, t. dillon, a. l. chase, c. e. keegan, and l. t. deblander. 2004. montana's forest products industry and timber harvest, 2004. resource bulletin, usda forest service, rocky mountain research station, fort collins, colorado, usa. state of montana. 1910. biennial report of the state game and fish warden of the state of montana, 1909–1910. montana fish, wildlife, and parks, helena, montana, usa. stevens, d. r. 1971. shiras moose. pages 89–95 in t. w. mussehl and f. w. howell, editors. game management in montana. montana fish, wildlife, and parks, helena, montana, usa. telfer, e. s. 1995. moose range under presettlement fire cycles and forest management regimes in the boreal forest of western canada. alces 31: 153–165. toweill, d. e., and g. vecellio. 2004. shiras moose in idaho: status and management. alces 40: 33–43. tyers, d. b. 2003. winter ecology of moose on the northern yellowstone winter range. ph. d. dissertation, montana state university, bozeman, montana, usa. ———. 2006. moose population history on the northern yellowstone winter range. alces 42: 133–149. 50 moose status in montana – decesare et al. alces vol. 50, 2014 vartanian, j. m. 2011. habitat condition and the nutritional quality of seasonal forage and diets: demographic implications for a declining moose population in northwest wyoming, usa. m.s. thesis, university of wyoming, laramie, wyoming, usa. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. wolfe, m. l., k. r. hersey, and d. c. stoner. 2010. a history of moose management in utah. alces 46: 37–52. worley, d. e. 1975. observations on epizootiology and distribution of elaeophora schneideri in montana ruminants. journal of wildlife diseases 11: 486–488. ———, c. k. anderson, and k. r. greer. 1972. elaeophorosis in moose from montana. journal of wildlife diseases 8: 242–244. appendix a: moose management survey questions provided to 20 mfwp biologists in 2010. 1. in your experience and professional judgment, what are the major concerns or limiting factors for moose in your area of responsibility (can choose more than one)? [ ] disease [ ] predation [ ] hunter harvest [ ] habitat loss/ fragmentation [ ] habitat succession [ ] other: ____________ 2. how would you describe the current status of moose within your area of responsibility? [ ] decreasing [ ] stable [ ] increasing 3. what type of moose management decisions are you typically required to make? [ ] harvest quota recommendations [ ] habitat enhancement [ ] habitat conservation [ ] large carnivore harvest recommendations 4. what information do you currently have and use for moose management (this information should be collected at the time of interview)? [ ] landowner reports [ ] hunter reports [ ] unadjusted trend counts [ ] sightability-corrected population estimates [ ] recruitment ratio counts [ ] bull: cow ratio counts [ ] harvest estimates [ ] habitat condition 5. which limiting factors have you addressed with moose management programs or decisions (this question will be accompanied by collection of past management actions: season proposals & rationales, regulations, specific habitat enhancement projects, land management plans, etc.)? [ ] disease [ ] predator harvest or control [ ] moose harvest [ ] habitat management [ ] habitat conservation [ ] other: __________ 6. how would you describe your moose survey and inventory information? [ ] adequate to make decisions for moose management [ ] adequate in some ways, not adequate in others [ ] not adequate to make moose management decisions 7. what information would most help you in your efforts to conserve and manage moose populations in your area? 8. can you list previous research projects and products from your area, and describe how results have been applied in your current management program? alces vol. 50, 2014 decesare et al. – moose status in montana 51 moose status and management in montana taxonomy monitoring methods and data moose harvest statistics and trend biologist interviews: local trends and management potential limiting factors hunter harvest predation vegetative succession and degradation parasites climate research needs and future directions acknowledgements literature cited appendix a: moose management survey questions provided to 20 mfwp biologists in 2010. space use and movements of moose in massachusetts: implications for conservation of large mammals in a fragmented environment david w. wattles1 and stephen destefano2 1massachusetts cooperative fish and wildlife research unit, department of environmental conservation, university of massachusetts, amherst, massachusetts 01003; 2u. s. geological survey, massachusetts cooperative fish and wildlife research unit, university of massachusetts, amherst, massachusetts 01003, usa. abstract: moose (alces alces) have recently re-occupied a portion of their range in the temperate deciduous forest of the northeastern united states after a >200 year absence. in southern new england, moose encounter different forest types, more human development, and higher temperatures than in other parts of their geographic range in north america. we analyzed seasonal minimum convex polygon home ranges, utilization distributions, movement rates, and home range composition of gps-collared moose in massachusetts. seasonal home range sizes were not different for males and females and were within the range reported for low latitudes elsewhere in north america. seasonal movement patterns reflected the seasonal changes in metabolic rate and the influence of the species’ reproductive cycle and weather. home ranges consisted almost entirely of forested habitat, included large amounts of conservation land, and had lower road densities as compared to the landscape as a whole, indicating that human development may be a limiting factor for moose in the region. the size and configuration of home ranges, seasonal movement patterns, and use relative to human development have implications for conservation of moose and other wide-ranging species in more highly developed portions of their ranges. alces vol. 49: 65–81 (2013) key words: alces alces, moose, home range, movements, roads, massachusetts. an animal's home range is the area where it finds the resources it needs for survival and reproduction (burt 1943); essentially it is a measure of spatial use for a given period of time. different home range estimators provide different information regarding how the animal uses space, including total area, areas of intensive use, and areas that are avoided (powell 2000, fieberg and börger 2012). animals have a cognitive map of their home range which allows them to exploit areas of concentrated resources and avoid areas that impart risks or disadvantages (powell 2000, powell and mitchell 2012, spencer 2012). thus home range size, configuration, and use can be influenced by the type, concentration, and distribution of resources, topography and other physical features, human development, and the distribution of other animals such as mates, competitors, and predators (powell and mitchell 2012). further, space use and movement patterns show seasonal changes which can be influenced by temperature and other climatic factors such as snow condition, reproductive status (börger et al. 2006, birkett et al. 2012), and for species that are affected by seasonal changes in forage quantity and quality like moose (alces alces) and other ungulates, foraging times, ruminating times, and metabolic rates (risenhoover 1986, cederlund 1989). knowledge of the size and position of an animal's home range and an individual's 65 movements and use of that area can provide insights into the distribution of resources and limiting factors in the environment (powell 2000, rettie and messier 2000, powell and mitchell 2012, spencer 2012). in areas of high human density, development of the landscape can be a major determinant of landscape use by many wildlife species (forman and deblinger 2000, lykkja et al. 2009, kertson et al. 2011). the concentration and distribution of industries and businesses, residences, roads and other infrastructure, and even the abundance of pets can affect the availability, quality, distribution, and connectivity of wildlife habitats. this is likely true for many or most taxa, but it is especially obvious for large mammals such as moose that require extensive areas to fulfill their life history needs. despite beliefs that temperature (kelsal and telfer 1974, renecker and hudson 1986, peek and morris 1998) and human development (vecellio et al. 1993, peek and morris 1998) might prevent it, moose have recently recolonized and become established in a portion of their historic range in the temperate deciduous forest of southern new england (vecellio et al. 1993, wattles and destefano 2011). this environment provides a number of potential challenges for moose, including forest types that differ from that found in most of its range (westveldt et al. 1956, degraaf and yamasaki 2001, franzmann and schwartz 2007), a thermal environment that could reduce fitness and survival (renecker and hudson 1986; boose 2001; murray et al. 2006; lenarz et al. 2009, 2010), and some of the highest densities of people in the united states (destefano et al. 2005, u. s. census bureau 2010a). habitat use, home range, and movement of moose have been studied throughout much of its range (franzmann and schwartz 2007), including elsewhere in the northeastern u. s. (leptich and gilbert 1989, garner and porter 1990, miller and litvaitis 1992, thompson et al. 1995, scarpitti et al. 2005). however, similar information has been lacking in southern new england where urban and suburban development and high road densities result in fragmentation of much of the landscape and relatively small and scattered natural areas. our objective was to determine how moose use the landscape in the humandominated and developed environment of central and western massachusetts. specifically, we wanted to quantify the seasonal home range size, space use patterns, and movement rates of moose in this recently re-established population. it is well documented that the reproductive cycle (e.g., the rut) and seasonal changes in forage affect movement patterns (belovsky 1981, risenhoover 1986, cederlund 1989, van ballenberghe and miquelle 1990), and we further predicted that movements would be influenced by weather patterns not experienced by moose elsewhere. also, due to the relatively limited number of human-moose conflicts reported in the state (wattles and destefano 2011), we predicted that moose would avoid areas with high levels of human development, locate their home ranges away from people, and that home range size and configuration would be influenced by development level. methods study area our study was conducted in central and western massachusetts, usa and adjacent portions of vermont and new hampshire, between 42° 9’ and 42° 53’ n latitude and 71° 53’ and 73° 22’ w longitude. topography is dominated by glaciated hills underlain by shallow bedrock. glacial activity created abundant small stream valleys, lakes, ponds, and wetlands whose size and nature varies with changes in beaver (castor canadensis) activity. the central and western sections of the study area are separated by the 66 home range and movements – wattles and destefano alces vol. 49, 2013 connecticut river valley which runs n-s through west-central massachusetts. elevation ranges from 100 m above sea level in the connecticut river valley, to 425 m in the hills of central massachusetts and 850 m in the berkshire hills of western massachusetts. the western two-thirds of massachusetts was >80% mixed deciduous, second, or multiple-growth forest, much of it resulting from regeneration of farm fields abandoned in the mid-late 1800s (hall et al. 2002). forest types included spruce-fir-northern hardwoods, northern hardwoods-hemlock (tsuga canadensis)-white pine (pinus strobus), transition hardwoods-white pinehemlock, and central hardwoods-hemlockwhite pine. transitions between forest types can be gradual or distinct depending on localized physiography, climate, bedrock, topography, and soil conditions, resulting in a patchwork of forest types and species groups (westveldt et al. 1956, degraaf and yamasaki 2001). dominant species included spruce (picea spp.), balsam fir (abies balsamea), american beech (fagus grandifolia), birch (betula spp.), trembling aspen (populus tremuloides), eastern hemlock, oaks (quercus spp.), hickories (carya spp.), and maples (acer spp.) depending on area and forest type. early successional habitat was created primarily through logging, and occasionally through wind and other weather events. about 1.5% of the forest was logged annually in 1984–2000, consisting of small (mean = 16.5 ha) cuts of moderate intensity (removal of 27% of timber volume) widely distributed on the landscape (kittredge et al. 2003, mcdonald et al. 2006). the pattern of forest harvest, glaciation, and transitional forest types provided a patchy mosaic of well interspersed forest types, age classes, and wetlands. july is the warmest month when mean daily temperature is 21 °c, and january the coldest when mean daily temperature is −6 °c. mean annual precipitation is 107 cm in central areas and 124 cm in western areas, with all months receiving 7–11 cm and 8–12 cm, respectively (degraaf and yamasaki 2001). the average date of last frost in the region is 15 may; the average day of first frost is 1 october and 15 september in central and western areas, respectively (degraaf and yamasaki 2001). snow depth is typically greater in western than central areas, and depths that restrict moose movement (50–70 cm) can occur in both areas (coady 1974). massachusetts is one of the most densely populated states in the u. s. (destefano et al. 2005; u. s. census bureau 2010a). development intensity is variable throughout the state, but tends to be substantially less in the uplands compared to the valley floors (<15–35 people/km2 in uplands and 35– >360/km2 in valley floors outside of major urban centers; u. s. census bureau 2010b). development in the uplands consists primarily of isolated homes and homes lining roadways within a matrix of forest; agricultural land and medium-to-large towns dominate the valleys. there is a dense road network throughout the area, consisting of state highways, paved, and unpaved municipal roads: 0.78 and 2.22 km of paved roads/km2 and 0.76 and 1.12 km of unpaved roads/km2 for uplands and valleys, respectively. study animals and gps telemetry we captured adult (>1 yr old) moose by opportunistically stalking and darting them from the ground between march 2006 and november 2009. moose were immobilized using either 5 ml of 300 mg/ml or 3 ml of 450 mg/ml xylazine hydrochloride (congaree veterinary pharmacy, cayce, sc, usa; mention of trade names does not imply endorsement by the u. s. government) administered from a 3 or 5 cc type c pneudart dart (pneudart, inc., williamport, alces vol. 49, 2013 wattles and destefano – home range and movements 67 pa, usa). we used tolazolene (100 mg/ml) at a dosage of 1.0 mg/kg as an antagonist. moose were fitted with gps collars, either ats g2000 series (advanced telemetry systems, inc., isanti, mn, usa) or telonics twg-3790 gps collars (telonics, inc., mesa, az, usa). we programmed the collars to attempt a gps fix as frequently as possible while allowing the battery life to extend for at least 1 year; depending on the collar, a gps fix was attempted every 135, 75, or 45 min. collars were also equipped with vhf transmitters, mortality sensors, and mechanisms that released the collars either at a low battery state or a pre‐ programmed date. capture and handling procedures were approved by the university of massachusetts institutional animal care and use committee, protocol numbers 25-02-15, 28-02-16, and 211-02-01. seasons we a priori defined the length and timing of seasons based on several ecological factors including vegetation phenology, weather (including temperature and snow conditions), and the moose reproductive cycle (table 1). the transition between seasons could vary by several days to several weeks depending on weather conditions and other factors. if movements were seen in the data that obviously demonstrated a change in season (e.g., a large increase in movement at the end of the winter when snow had melted or the end of summer indicating the beginning of rutting behavior), the seasons were truncated at that point and the data were included in the following season (fig. 1). home ranges and space use we used 2 methods to calculate space use by moose: minimum convex polygon (mcp) and utilization distributions (ud) by fixed kernel density estimator. we calculated100% mcp home ranges with the create minimum convex polygons tool in hawth's analysis tools (beyers 2006) and uds using the kernel density estimation tool in hrt: home range tools for arcgis (rodgers et al. 2007). all geographic information system (gis) work was performed in arcgis 9.3 (esri 2008). table 1. seasons used for calculating home-range, movements, and core-area habitat analyses for moose in massachusetts, 2006–2011. season breaks were based on phenology of vegetation, temperature, normal snow conditions, and moose reproductive activity. season dates vegetation/browse temperaturea movement moderators season length (d) spring 16 april–31 may growing season; bud-break-leaf out cool-hot potentially temperature 46 calving (females) 8–13 may–15 june growing season; bud-break-leaf out cool-hot newborn calf mobility 30 summer 1 june – 30 aug growing season; full leaf out hot temperature 92 fall 1 sept – 31 oct leaf out to leaf off hot-cool temperature and rut 61 early winter 1 nov – 31 dec dormant season; woody/evergreen warm-cold potentially metabolism 61 late winter 1 jan – 15 april dormant season; woody/evergreen cold-warm potentially snow and metabolism 107 atemperature ranges describing typical temperatures experienced during a season; cold ≤0 °c, cool >0 °c and <14 °c,warm ≥14 °c and <20 °c, hot ≥20 °c. 68 home range and movements – wattles and destefano alces vol. 49, 2013 fig. 1. the y-axis represents mean daily movement rates (m/day, thin line) for female (top; n = 5) and mature male (bottom; n = 10) moose in massachusetts, 2006–2011. the heavy line represents a 10day moving average to remove noise; the vertical dashed lines mark a priori delineated season boundaries. alces vol. 49, 2013 wattles and destefano – home range and movements 69 choice of the kernel bandwidth or smoothing factor (h) is known to have the greatest effect on the resultant utilization distribution when using kernel density estimators (worton 1989). a large h over-smooths the data resulting in a positively biased ud that encompasses unused habitats, whereas a small h under-smooths the data resulting in a fragmented ud (fieberg 2007, fieberg and börger 2012). quantitative methods of determining h can be influenced by sample size, sampling intensity, and the distribution of locations (kie et al. 2010, fieberg and börger 2012), and there is lack of agreement on the best method for calculating h (powell 2000, hemson et al. 2005, gitzen et al. 2006, fieberg 2007, kie et al. 2010, fieberg and borger 2012). we chose a 200-m bandwidth because it strikes a balance between creating a continuous polygon and over-buffering the edges of the utilization distribution. the 200-m bandwidth value merged closely separated locations into a single polygon, but did not merge widely spaced clusters. mitchell and powell (2008) noted that fragmentation of uds may be desired to identify used and unused areas in patchy and fragmented landscapes. increasing the bandwidth beyond 200 m resulted in uds with a larger buffer around all points, but failed to further merge disjointed polygons into a single polygon unless very large values of h were used. smaller values of h resulted in more fragmented uds that did not accurately represent space use. road densities in mcp home ranges and uds were calculated using the masseot (massachusetts executive office of transportation) roads layer (massachusetts office of geographic information 2005). we used a 2005 land use layer (massachusetts office of geographic information 2005) to calculate amount of forest and wetlands, and the protected and recreational open space layer (massachusetts office of geographic information 2005) to calculate amount of protected area. movements we calculated mean seasonal daily movement rates by calculating the distance between successive fixes and summing those distances for each 24-h period beginning at 0:00. mills et al. (2006) showed that decreased gps sampling intensity resulted in reduced observed movement rates in wolves (canis lupus) due to a reduction in tortuosity of the path. we corrected for the variable sampling rate in our collars (135, 75, and 45 min) by subsampling the more intensively sampled datasets (45 min), and taking every other and then every third location to simulate 90 and 135 min intervals, respectively. we saw a consistent reduction in movement rates with increasing sampling interval. therefore, we used this information to weight the movements observed in our 135(n = 23) and 45-min (n = 2) collars to the intermediate 75-min (n = 5) sampling level, making comparisons among individuals possible. statistics we used the r statistical package, version 2.12.2 (r development core team 2005) for all statistical analyses. we used mixed effect models in the r-package lme4 (bates et al. 2012) to analyze the differences in seasonal home range size and movement rates within and between sexes and seasons. we incorporated random intercept in the models to account for unequal sample sizes among sexes and seasons and to account for repeated measures on individual moose and performed post-hoc pairwise comparisons using the r-package lmerconviencefunctions (tremblay and ransijn 2012). we employed one-sample z-tests to compare road densities in the valley bottoms and uplands to home ranges. transformations failed to meet the assumption of normality; 70 home range and movements – wattles and destefano alces vol. 49, 2013 therefore, we used a nonparametric paired wilcoxon's rank-sum test to make comparisons in road density between mcp home ranges and uds. significance level for all analyses was set at 0.05. results capture and deployment of gps collars we deployed gps collars on 21 moose: 5 adult (>3 yr) females, 7 adult males, and 1 immature (<3 yr) male in central massachusetts, and 4 adult and 4 immature males in western massachusetts; 9 were recaptured to replace gps collars. we obtained 127,408 locations of the 21 moose with an overall fix rate of 85%. seasonal data for any animal were only included in the analyses if data were obtained across the entire season. the median number of locations/animal/season ranged from 402 in spring to 1,015 in late winter. the minimum number of locations was 281 for one animal in spring. home ranges and space use mean annual (mcp) home range sizes were not different for mature males (88.8 km2) and females (62.2 km2) (p = 0.28; table 2). ranges of immature males were larger in all seasons and annually (177.5 km2) than either mature males or females, except for females during summer. there were no differences in mean seasonal range sizes for mature males and females (p ≥0.22), with the exception of fall (23.0 and 59.4 km2 for females and males, respectively; p = 0.002) (table 2). seasonal home ranges for females ranged from 23.0 km2 during fall and early winter to 34.8 km2 in summer, with no difference (p ≥0.32) in seasonal home range size. seasonal home range size for mature males ranged from 17.5 km2 in late winter to 59.4 km2 during fall, with fall home ranges larger (p ≤0.01) than all other seasons. mean annual 95% ud sizes were not different between females (26.7 km2) and mature males (28.8 km2) (p ≥0.54; table 3). seasonal ud size for females did not differ among seasons (p ≥0.07; table 3). seasonal ud size for mature males ranged from 8.5 km2 in late winter to 19.6 km2 during fall, with fall larger (p ≤0.01) than summer and early and late winter; additionally, spring and summer uds were larger (p ≤ 0.01) than late winter. mature males had larger uds than central females in fall (p ≤ 0.01). seasonal uds were between 40–51% and 33–63% of seasonal mcp home ranges for females and mature males, respectively. location and composition of home ranges and utilization distributions mcp home ranges consisted of 84% (se = 0.02) forested cover types and 12% wetlands (se = 0.02), and uds were 88% (se = 0.01) forested with 9% (se = 0.01) wetlands. conservation land (state forests, table 2. seasonal and annual mean 100% minimum convex polygon home ranges (km2) for females, mature males (estimated >3 yr old), and immature male moose in massachusetts, 2006–2011. central females mature males immature males season n mean se range n mean se range n mean se range spring 5 26.9 4.2 14.1–39.0 9 28.0 3.2 14–39.0 5 61.4 25.5 15.8–158.1 summer 5 34.8 7.4 18.2–61.4 8 21.9 4.5 6.2–39.5 4 32.5 6.8 16.2–48.5 fall 5 23.0 3.7 12.8–28.8 8 59.4 15.1 31.8–161.3 5 222.6 110.0 6.6–546.8 early winter 4 23.0 2.9 14.9–29.1 10 29.6 5.4 14.3–72.9 5 50.8 11.9 14.6–83.1 late winter 5 25.8 3.8 14.3–38.0 11 17.5 2.7 5.1–31.8 5 33.2 12.8 9.3–80.4 annual 5 62.2 7.7 41.6–78.4 9 88.8 16.8 49.3–199.4 4 177.5 96.0 33.5–458.9 alces vol. 49, 2013 wattles and destefano – home range and movements 71 wildlife management areas, other protected land, and conservation easements) made up more of mcp home ranges (60%, se = 0.05, p ≤0.001) and uds (66%, se = 0.07, p ≤0.001) than was available as a whole in the central uplands (43%), and more in the mcp home ranges (59%, se = 0.1, p ≤0.004) and uds (76%, se = 0.1, p ≤0.001) than was available in the berkshire hills of western massachusetts (32%). additionally, conservation land made up a greater percentage of uds than either the overall mcp home ranges, or the area outside the ud but within the mcps (the unused portion of the mcp home range) (p ≤0.01); however, there was no difference in the amount of conservation land in mcp home ranges compared to the unused portion of the mcp (p = 0.16). all paved road types were at lower density within home ranges and uds compared to both the valley bottoms and uplands overall (p ≤0.001; table 4). additionally, all classes of paved roads (state highways, major local arteries, and local paved roads) were at lower densities within uds than either the overall mcp home ranges, or the unused portion of the mcp home range (p ≤0.04; fig. 2). state highways and local paved roads were also at greater densities in the unused portion of the mcp than in the overall mcp (p ≤0.008). seasonal movement patterns daily movement rates for female moose in central massachusetts were consistently ∼1,000–1,500 m/day in late winter (fig. 2). table 3. seasonal and annual mean 95% fixed kernel utilization distribution (km2) for females, mature males (estimated >3 yr old), and immature male moose in massachusetts, 2006–2011 (smoothing factor (h) = 200 m). females mature males immature males season n mean se range n mean se range n mean se range spring 5 10.8 0.8 8.2–12.8 9 15.5 1.4 10.4–22.0 4 19.6 3.7 13.5–28.8 summer 5 15.9 2.8 8.7–24.4 8 13.9 2.3 5.2–22.5 4 15.4 0.6 13.9–16.4 fall 4 11.4 0.8 10.0–13.6 7 19.6 2.4 10.4–30.6 4 22.2 8.1 6.5–44.8 early winter 5 11.4 1.5 8.4–15.8 9 11.5 0.8 6.3–14.5 4 19.5 1.2 16.5–20.2 late winter 5 13.1 0.7 11.6–15.7 10 8.5 1.3 4.1–15.1 4 13.5 4.4 7.4–26.2 annual 5 26.7 2.0 19.9–32.1 7 28.8 2.4 22.6–41.2 4 37.7 6.8 20.2–51.0 table 4. mean densities (km/km2) (se) of paved and unpaved roads in the valley bottoms, uplands, within maximum convex polygon (mcp) but outside utilization distributions (ud), mcp home ranges, and ud for moose in massachusetts, 2006–2011. valley bottoms uplands mcp outside ud mcp ud interstate highways 0.08 0.01 0.00 (0.00) 0.00 (0.00) 0.00 (0.00) major state highways 0.03 0.00 0.00 (0.00) 0.00 (0.00) 0.00 (0.00) state highways 0.33 0.18 0.13 (0.03) 0.11 (0.03) 0.06 (0.01) major local arteries 0.31 0.09 0.05 (0.02) 0.03 (0.01) 0.01 (0.01) local paved roads 1.48 0.50 0.40 (0.05) 0.30 (0.08) 0.14 (0.09) local unpaved/improved forest roads 0.39 0.48 0.54 (0.06) 0.49 (0.13) 0.44 (0.03) forest roads 0.73 0.28 0.33 (0.04) 0.35 (0.09) 0.38 (0.07) 72 home range and movements – wattles and destefano alces vol. 49, 2013 in spring, daily movement nearly doubled to ∼3,000 m/day prior to calving. there was a sharp decline to 500 m/day the second week of may that corresponded with the observed 8–13 may calving period. mean daily movement rates remained low for may and most of june, before peaking at ∼3,000 m/day in early july and remaining high for the remainder of the summer. movement rate declined in september to about 1,500 m/day and remained fairly consistent for the rest of the year. spring and summer seasonal movement rates for females were greater than all other seasons and calving season movement rates were lower than all other seasons (p ≤0.05; table 5). daily movement rates were lowest (1,000 m/day) for mature males from the fig. 2. road density in annual fixed kernel utilization distribution (dark gray) and minimum convex polygon home range (light gray) for a representative moose in massachusetts. heavy lines are major local roads and state highways, thin solid lines are local paved roads, and dashed lines are forest roads with limited access. alces vol. 49, 2013 wattles and destefano – home range and movements 73 beginning of february until the end of march (table 5). movements increased in early april and peaked at ∼2,500 m/day in late may and early june, before declining as summer progressed. daily movements increased to 3,000 m/day during the second week of september, indicating the start of the rut. movements increased further to a peak of nearly 8,000 m/day the last week of september and remained high through the first week of october, then declined sharply. movement rates remained relatively high at 2,000–2,500 m/day until the beginning of december when they declined to winter levels of 1,000–1,500 m/day. fall seasonal movement rates were greater than in all other seasons for mature males (p ≤0.05; table 5); additionally, spring and summer rates were greater than in late winter, and spring was greater than early winter. male daily movement rates were greater (p ≤0.05) than females during fall and lower during summer. discussion home range as a measure of resource use spatial requirements as measured by home range (second order use; johnson 1980) and uds (i.e., measuring use patterns within the home range; third order use) can provide important information about productivity of available habitat, distribution of resources and limiting factors, and how a species uses resources. this information is critical for conservation planning and habitat protection and connectivity at local and regional scales, and is particularly relevant for large mobile mammals in highly developed landscapes with fragmented patches of protected lands. harris et al. (1990) recommended using at least 2 home range estimators for all animal location data sets, including minimum convex polygon (mohr 1947) because of its prevalent use and comparability among studies. a mcp home range measures the area used by an individual to fulfill its annual or seasonal needs, but it does not describe how the area is used. alternatively, uds created by fixed kernels (worton 1989) describe the pattern and intensity of use within the mcp home range. by examining both, we can quantify areas of actual and relative intensity of use, identify important seasonal habitat patches, and delineate the area of landscape required to provide those patches comparison of uds to mcps shows that moose in southern new england used the table 5. seasonal daily movement rates (m/day) for female and mature male moose in massachusetts. mean seasonal daily movement rates and (se) in light gray, p-values for seasonal comparison between males and females in dark gray, p-values for comparisons among seasons for females above the diagonal and for males below the diagonal. female spring summer fall early winter late winter calving mean 2391 (141.0) 2464 (216.6) 1837 (81.5) 1505 (158.0) 1492 (107.9) 874 (70.6) sp 0.719 0.012 <0.001 <0.001 <0.001 spring 2019 (161.3) sp 0.22 sm 0.006 <0.001 <0.001 <0.001 summer 1731 (120.5) 0.168 sm 0.017 fl 0.112 0.097 <0.001 m a tu re m a le fall 3542 (385.2) <0.001 <0.001 fl <0.001 ew 0.951 0.008 early winter 1514 (107.0) 0.017 0.291 <0.001 ew 0.967 lw 0.009 late winter 1103 (79.8) <0.001 0.004 <0.001 0.051 lw 0.157 74 home range and movements – wattles and destefano alces vol. 49, 2013 landscape in a patchy manner; uds were typically only half the size of mcps, meaning that at any time there was a 95% probability of locating a moose within <50% of the mcp home range. additionally, uds fragmented into multiple polygons, indicating that resources were patchily distributed. maintaining connectivity of used patches within the larger landscape (mcp and larger) is essential for moose and other wide ranging species. rettie and messier (2000) argued that selection at the scale of the home range reflects attempts to reduce the effects of limiting factors. the uds measured here were located almost exclusively on the uplands of the central and western parts of the state, with limited use of valley bottoms. when valley bottoms were included in an mcp home range, they were mostly unused portions that were traversed in movements between ridge tops. overall, uds had greater amounts of forested habitat and conservation land and lower road densities than the landscape as a whole, or than the mcp home ranges. by definition moose spent 95% of their time in these less developed areas and appeared to select for more heavily forested areas away from human development. moose often crossed roads of all types in massachusetts, but seemed to show less avoidance of local residential roads with lower traffic volumes and speed limits than major highways, state highways, and major local arteries. in many instances major roads formed boundaries at the edge of an individual's home range; in other cases home ranges were bisected by highways and main roads. use of higher elevations could also be an attempt to limit thermal stress by taking advantage of reduced ambient temperatures and increased exposure to convective cooling from wind. human development and associated vehicle traffic and high temperatures that result in thermal stress may be limiting factors for moose in massachusetts. seasonal home ranges in central massachusetts, female mcp home ranges were largest during summer when energy demands were greatest because of lactation and seasonal restoration of body condition. mature male home ranges were largest during fall when they search for and attend mates during the breeding season, and smallest during late winter and summer when movements were presumably restricted by the combined effects of lower metabolism, snow conditions, and thermoregulatory constraints. despite the large number of studies on home range size (hundertmark 1997), comparisons to our results must be made with caution. most studies have used traditional vhf telemetry and home ranges were calculated with a small number of locations (e.g., <30), particularly in winter (e.g., <10), which can underestimate home range size (kernohan et al. 2001, börger et al. 2006); further, few vhf locations are collected at night when moose are often active. kernohan et al. (2001) suggested a minimum number of 30 locations, but at least 50 to calculate an accurate home range. additionally, differences in methods and the length, timing, and number of seasons used can make comparisons difficult (kernohan et al. 2001, börger et al. 2006). even with these limitations, our results fall within the range presented by hundertmark (1997) for home range sizes across north america (fig. 3). overall, home range size decreased with decreasing latitude and summer and winter home ranges in massachusetts would be expected at the low end of the scale. in the northeastern united states our results are similar to those of leptich and gilbert's (1989) in maine with >50 locations for 11 of 13 collared moose and an estimated alces vol. 49, 2013 wattles and destefano – home range and movements 75 summer mcp home range of 25 km2 for females. thompson et al. (1995) reported median summer home ranges of 32 km2 for females and 28 km2 for males in maine; their sample sizes in other seasons were too low for comparison. winter ranges were typified by concentrated use of small areas with short movements to other areas of intensive use in minnesota (van ballenberghe and peek 1971) and maine (thompson et al. 1995), a pattern similar to our observations. in northern new hampshire, scarpitti et al. (2005) observed smaller seasonal home ranges for females than our study (≤17 km2 for all seasons), with an earlier study in northern new hampshire (miller and litvaitis 1992) reporting much larger annual home ranges for females (153 km2) with the largest seasonal home ranges during fall (82 km2). garner and porter (1990) reported 36 km2 for summer and 8 km2 for winter home ranges of males in the adirondack mountains of new york. our seasonal results are the opposite of lenarz et al. (2011) who reported smaller home ranges during summer (16 km2) than in winter (33 km2) in minnesota. movements seasonal activity and movement patterns reflect changes in metabolic rate, ruminating time, and activity associated with the annual cycle of vegetation growth in temperate forests (risenhoover 1986, cederlund 1989). increased movement rates in spring corresponded with the start of the growing season and increased abundance and quality of browse. high movement rates in summer have been shown to reflect increased activity associated with more foraging bouts, lower ruminating times, and an attempt by moose to maximize foraging during the growing season (belovsky 1981, cederlund 1989, fig. 3. mean size of winter and summer home ranges in square kilometers for moose in north america relative to latitude (as reported by hundertmark 1997). data for female and male moose added as open symbols. 76 home range and movements – wattles and destefano alces vol. 49, 2013 van ballenberghe and miquelle 1990). we speculate that the periodically reduced rates in movements we observed during spring, summer, and fall were the result of thermoregulatory behavior during periods of high temperatures. the reduced movements during winter were typical of moose throughout their range (phillips et al. 1973, dussault et al. 2005, schwartz and renecker 2007). schwartz and renecker (2007) suggest that the lower winter metabolic rate of moose is an adaptation to counteract reduced forage abundance and quality and the related increased time required to digest a highly fibrous diet, resulting in fewer feeding bouts and lower activity level. movements were further reduced during periods of deep snow; however, snow depth and condition vary annually and across the state with the highest likelihood of deep snow at higher elevations in western massachusetts. when confined by deep snow, moose concentrated their habitat use into as little as 0.5 km2 for up to 3.5 months. the variability in the timing, depth, and condition of snowfall strongly influenced the variability of home range size and movements in early and late winter, as moose moved widely between suitable winter habitats until confined by snow. in addition to the influence of seasonal patterns on movements, changes in daily movement rates were greatest at times of the year corresponding to the annual reproductive cycle, i.e., calving for females and the rut for males. a final important consideration for understanding movements of moose in southern new england is the lack of their major predator, wolves (canis lupus), and the absence of moose hunting. predators and hunters can play important roles in the distribution and movements of their ungulate prey. black bears (ursus americanus) and coyotes (canis latrans) may prey on some moose calves, but in general the influence of predators or hunters on moose movements and distribution is absent in massachusetts. management implications existing distribution of vegetative communities, landscape configurations, and levels of development have allowed moose to re-colonize and establish a low density population throughout central and western massachusetts and into connecticut after 200–300 years of absence. however, southern new england is comprised of some of the most densely populated and highly developed states in the nation, and despite very active and successful conservation agencies and organizations, the trend will continue to move in the direction of more development and increased fragmentation. we have documented key elements of habitat use and movement distances and patterns by this newly re-established moose population. this information can be used to further enhance existing high priority conservation areas and identify new areas for protection and landscape connectivity. massachusetts has many well established biodiversity conservation initiatives (e.g., wildlife biomap and living waters) and planning strategies should recognize and incorporate a suitable scale to accommodate moose. if this large-scale challenge can be met, biodiversity conservation will benefit because moose use a diversity of terrestrial and wetland vegetative types (composition, size, and structure) that provide habitat for a wide array of species. acknowledgments the massachusetts division of fisheries and wildlife through the federal aid in wildlife restoration program (w-35-r) provided funding and support for this research. we appreciate the long-term involvement and support of many people, particularly r. deblinger and t. o'shea. the department of conservation and recreation, u.s. alces vol. 49, 2013 wattles and destefano – home range and movements 77 geological survey, the university of massachusetts, and safari club international provided additional funding and logistical support. capture of moose would not have been possible without the assistance of k. berger and other field assistants and volunteers. references bates, d., m. maechler, and b. bolker. 2012. lme4: linear mixed effects models using s4 classes. (accessed december 2012). belovsky, g. e. 1981. optimal activity times and habitat choice of moose. oecologia 48: 22–30. beyers, h. 2006. hawth's analysis tools for arcgis. version 3.27. (accessed december 2012). birkett, p. j., a. t. vanak, v. m. r. muggeo, s. m. ferreira, and r. slotow. 2012. animal perception of seasonal thresholds: changes in elephant movement in relation to rainfall patterns. plos one 7: 1–8. boose, e. 2001. fisher meteorological station (since 2001). harvard forest data archive: hf001. petersham, massachusetts, usa. börger, l., n. franconi, f. ferretti, f. meschi, g. de michele, a. gantz, and t. coulson. 2006. an integrated approach to indentify spatiotemporal and individual-level determinants of animal home range size. the american naturalist 168: 471–485. burt, w. h. 1943. territoriality and home range concepts as applied to mammals. journal of mammalogy 24: 346–352. cederlund, g. 1989. activity patterns in moose and roe deer in a north boreal forest. holarctic ecology 12: 39–54. coady, j. w. 1974. influence of snow on behavior of moose. naturaliste canadien 101: 417–436. degraaf, r. m., and m. yamasaki. 2001. new england wildlife: habitat, natural history, and distribution. university press of new england, hanover, new hampshire, usa. destefano, s., r. d. deblinger, and c. miller. 2005. suburban wildlife: lessons, challenges, and opportunities. urban ecosystems 8: 131–137. dussault, c., j. p. ouellet, r. coutois, j. huot, l. breton, and h. jolicoeur. 2005. linking moose habitat selection to limiting factors. ecography 28: 1–10. environmental systems research institute inc. (esri). 2008. arcgis 9.3. redlands, california, usa. fieberg, j. 2007. kernel density estimators of home range: smoothing and the autocorrelation red herring. ecology 88: 1059–1066. ———, and l. börger. 2012. could you please phrase “home range” as a question? journal of mammalogy 93: 890–902. forman, r. t. t., and r. d. deblinger. 2000. the ecological road effect zone of a massachusetts (u.s.a.) suburban highway. conservation biology 14: 36–46. franzmann, a. w., and c. c. schwartz. 2007. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. garner, d. l., and w. f. porter. 1990. movement and seasonal home ranges of bull moose in a pioneering adirondack population. alces 26: 80–85. gitzen, r. a., j. t. millspaugh, and b. j. kernohan. 2006. bandwidth selection for fixed-kernel analysis of animal utilization distributions. journal of wildlife management 70: 1334–1344. hall, b., g. motzkin, d. r. foster, m. syfert, and j. burk. 2002. three hundred years of forest and land-use in massachusetts, usa. journal of biogeography 29: 1319–1335. harris, s., w. j. cresswell, p. g. forde, w. j. trewhella, t. woollard, and s. wray. 1990. home-range analysis using 78 home range and movements – wattles and destefano alces vol. 49, 2013 http://lme4.r-forge.r-project.org http://lme4.r-forge.r-project.org http://www.spatialecology.com/htools http://www.spatialecology.com/htools radio-tracking data: a review of problems and techniques particularly as applied to the study of mammals. mammal review 20: 97–123. hemson, g., p. johnson, a. south, r. kenward, r. ripley, and d. macdonald. 2005. are kernels the mustard? data from global positioning system (gps) collars suggests problems for kernel home-range analyses with least-squares cross-validation. journal of animal ecology 74: 455–463. hundertmark, k. j. 1997. home range, dispersal, and migration. pages 303–335 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61: 65–71. kelsal, j. s., and e. s. telfer. 1974. biogeography of moose with particular reference to western north america. naturaliste canadien 101: 117–130. kernohan, b. j., r. a. gitzen, and j. j. millspaugh. 2001. analysis of animal space use and movements. pages 125–166 in j. j. millspaugh and j. m. marzluff, editors. radio tracking animal populations. academic press, san diego, california, usa. kertson, b. n., r. d. spencer, j. m. marzluff, j. hepinstall-cymerman, and c. e. grue. 2011. cougar space use and movements in the wildland-urban landscape of western washington. ecological applications 21: 2866–2881. kie, j. g., j. matthiopoulos, j. fieberg, r. a. powell, f. cagnacci, m. s. mitchell, jm. gaillard, and p. r. moorscroft. 2010. the home-range concept: are traditional estimators still relevant with modern telemetry technology. philosophical transactions of the royal society b 365: 2221–2231. kittredge, d. b., jr., a. o. finley, and d. r. foster. 2003. timber harvesting as ongoing disturbance in a landscape of diverse ownership. forest ecology and management 180: 425–442. lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–510. ———, j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. ———, r. g. wright, m. w. schrage, and a. j. edwards. 2011. compositional analysis of moose habitat in northeastern minnesota. alces 47: 135–149. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53: 880–885. lykkja, o. n., e. j. solberg, i. herfindal, j. wright, c. m. rolandsen, and m. g. hanssen. 2009. the effects of human activity on summer habitat use by moose. alces 45: 109–124. massachusetts office of geographic information. 2005. (accessed december 2012). mcdonald, r.i., g. motzkin, m. s. bank, d. b. kitteridge, j. burke, and d. l. foster. 2006. forest harvesting and landuse conversion over two decades in massachusetts. forest ecology and management 227: 31–41. miller, b. k., and j. a. litvatis. 1992. habitat segregation by moose in a boreal forest ecotone. acta theriologica 37: 41–50. mills, k. j., b. r. patterson, and d. l. murray. 2006. effect of variable sampling frequencies on gps transmitter efficiency and estimated wolf home range alces vol. 49, 2013 wattles and destefano – home range and movements 79 http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/eotroads.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/eotroads.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/eotroads.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/eotroads.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/eotroads.html size and movement distance. wildlife society bulletin 34: 1463–1469. mitchell, m. s., and r. a. powell. 2008. estimated home ranges can misrepresent habitat relationships on patchy landscapes. ecological modeling 216: 409–414. mohr, c. o. 1947. table of equivalent populations of north american small mammals. american midland naturalist 37: 223–249. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monograph 166: 1–30. peek, j. m., and k. i. morris. 1998. status of moose in the contiguous united states. alces 34: 423–434. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37: 266–278. powell, r. a. 2000. animal home ranges and territories. pages 65-110 in l. boitani and t. k. fuller, editors. research techniques in animal ecology. columbia university press, new york, new york, usa. ———, and m. s. mitchell. 2012. what is a home range? journal of mammalogy 93: 948–059. r development core team. 2005. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. (accessed december 2012). renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. rettie, w. j., and f. messier. 2000. hierarchical habitat selection by woodland caribou: its relationship to limiting factors. ecography 23: 466–478. risenhoover, k. l. 1986. winter activity patterns of moose in interior alaska. journal of wildlife management 50: 727–734. rodgers, a. r., a. p. carr, h. l. beyer, l. smith, and j. g. kie. 2007. hrt: home range tools for arcgis. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. scarpitti, d., c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. schwartz, c. c., and l. a. renecker. 2007. nutrition and energetics. pages 441–478 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. spencer, w. d. 2012. home ranges and the value of spatial information. journal of mammalogy 93: 929–947. thompson, m. e., j. r. gilbert, g. j. matula, and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31: 233–245. tremblay, a., and j. ransijn. 2012. lmerconveniencefunctions: a suite of functions to back-fit fixed and forward-fit random effects, as well as other miscellaneous functions. (accessed decem‐ ber 2012). u. s. census bureau. 2010a. census 2010. resident population data: population density (accessed february 2013). ———. 2010b. 2010 census: massachusetts profile map. (accessed february 2013). 80 home range and movements – wattles and destefano alces vol. 49, 2013 http://www.r-project.org http://www.r-project.org http://cran.r-project.org/web/packages/lmerconveniencefunctions/index.html http://cran.r-project.org/web/packages/lmerconveniencefunctions/index.html http://cran.r-project.org/web/packages/lmerconveniencefunctions/index.html http://www.census.gov/2010census/data/apportionment-dens-text.php http://www.census.gov/2010census/data/apportionment-dens-text.php http://www2.census.gov/geo/maps/dc10_thematic/2010_profile/2010_profile_map_massachusetts.pdf http://www2.census.gov/geo/maps/dc10_thematic/2010_profile/2010_profile_map_massachusetts.pdf http://www2.census.gov/geo/maps/dc10_thematic/2010_profile/2010_profile_map_massachusetts.pdf van ballenberghe, v., and d. g. miquelle. 1990. activity of moose during spring and summer in interior alaska. journal of wildlife management 54: 391–396. ———, and j. m. peek. 1971. radio telemetry studies of moose in northeastern minnesota. journal of wildlife management 35: 63–71. vecellio, g. m., r. d. deblinger, and j. e. cardoza. 1993. status and management of moose in massachusetts. alces 29: 1–7. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. westveldt, m. r., r. i. ashman, h. i. baldwin, r. p. holdsworth, r. s. johnson, j. h. lambert, h. j. lutz, l. swain, and m. standish. 1956. natural forest vegetation zones of new england. journal of forestry 54: 332–338. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70: 164–168. alces vol. 49, 2013 wattles and destefano – home range and movements 81 space use and movements of moose in massachusetts: implications for conservation of large mammals in a fragmented environment methods study area study animals and gps telemetry seasons home ranges and space use movements statistics results capture and deployment of gps collars home ranges and space use location and composition of home ranges and utilization distributions seasonal movement patterns discussion home range as a measure of resource use seasonal home ranges movements management implications acknowledgments references alces27_50.pdf alces28_243.pdf alces27_118.pdf alces21_103.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire daniel h. bergeron1 and peter j. pekins department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa. abstract: in new hampshire, winter ticks (dermacentor albipictus) probably have more influence on the moose (alces alces) population than other mortality factors, and predicting the frequency of tick epizootics is an important management consideration. weather, moose density, and habitat use influence abundance and distribution of winter ticks. we evaluated the usefulness of 3 techniques to index winter tick abundance in 3 regions with variable moose density: 1) flagging for tick larvae, 2) line-transect counts of ticks on harvested moose, and 3) roadside surveys of tick-induced hairloss on moose. although counts of tick larvae from fall flagging were not significantly different between years or regions, absolute tick abundance was measurably different (>50%) relative to moose density and years. tick abundance on harvested moose reflected annual and regional differences; in general, abundance was correlated positively with moose density and annual trends within regions were similar. tick abundance was highest for calves and lowest for cows. hair-loss surveys indicated that hair loss was generally related to moose density, and similar annual differences were reflected in all regions. we suggest measuring tick abundance on harvested moose and conducting annual roadside hair-loss surveys to create indices and threshold values useful in predicting an epizootic of winter ticks. alces vol. 50: 1–15 (2014) key words: alces alces, dermacentor albipictus, hair loss, index, moose, winter tick the winter tick (dermacentor albipictus) is a unique blood-feeding ectoparasite that periodically causes severe mortality in moose (alces alces) populations (cameron and fulton 1926–27, samuel and barker 1979). it is found in most of moose range in the united states and canada south of 60° n latitude (bishopp and trembley 1945, wilkinson 1967), but not in newfoundland or alaska, although it could presumably survive if translocated (zarnke et al. 1990, lankester and samuel 1998). winter ticks have 3 different parasitic life stages: larva, nymph, and adult. each requires a blood meal to subsequently develop to the next stage, and meals are taken from a single host throughout the course of one winter (lankester and samuel 1998). the life cycle is predictable with little annual variation (addison and mclaughlin 1988) because its reproductive cycle is dictated by environmental cues such as temperature and photoperiod (wright 1969, drew and samuel 1986). annual synchrony of the reproductive cycle is partly due to nymphal and adult diapause (glines and samuel 1984). nymphal diapause allows larvae that attach to hosts at different times to be fully developed at the same time (addison and mclaughlin 1988), and adult diapause allows for synchrony of oviposition (drew and samuel 1986). this strict cycle is probably due to the northern climate that allows only a narrow window of reproductive success (samuel 2004). 1present address: new hampshire fish and game department, hazen drive, concord, nh 03301, usa. 1 weather appears to be the most influential factor of winter tick abundance (delguidice et al. 1997, samuel 2007); however, moose population density seems to influence the distribution and abundance of winter ticks as several studies indicate that tick load increases with moose density (blyth 1995, pybus 1999, samuel 2007). a high density of moose presumably allows for higher larval attachment in autumn, yielding more adult females that produce more eggs (samuel 2004); evidence for this relationship is mostly correlative. most larvae climb vegetation in the immediate area of the hatching site and 87% of engorged adult females are found within 60 cm of moose carcasses (drew and samuel 1985, 1986), indicating that the drop site of adult female ticks is essentially the site of oviposition. therefore, distribution of winter ticks is related directly to where adult female ticks drop from moose during early spring (drew and samuel 1986, samuel 2004). moose in northern new hampshire preferentially use cut/regeneration habitat in late winter-spring (scarpitti et al. 2005). studies in canada estimated that average numbers of winter ticks on a single moose were ∼30,000 and may exceed 100,000 (samuel and barker 1979, samuel and welch 1991). high tick loads may lead to several problems including damage and loss of the winter coat, reduced visceral fat stores, restlessness, reduced growth in young moose, and death (samuel and barker 1979, mclaughlin and addison 1986, samuel 1991, addison et al. 1994). studies have found little evidence of anemia in well fed captive moose (glines and samuel 1989, addison et al. 1998); however, the authors speculated that it may occur in wild moose populations on natural diets. modeling studies conducted by samuel (2004) and musante et al. (2010) suggest anemia may have a large impact on moose if it does occur in the wild. tick induced hair-loss or alopecia is one of the most common and visual impacts of winter ticks, and rapid hair-loss occurs in march-may, coinciding with engorging by adult ticks (mclaughlin and addison 1986). berg (1975) observed high calf mortality in northwestern minnesota when calves died after 2 days of −30 °c temperatures and winds of 130 km/h; all dead calves had severe tick infestations and hair-loss. welch et al. (1990) found that tick-induced hair-loss had little impact on the metabolic rates of captive moose, possibly because of mild ambient temperatures during the study. they speculated that hypothermia is likely unimportant, as severe hair-loss rarely occurs before march, and prolonged severe cold is uncommon thereafter. studies by glines and samuel (1989) and addison et al. (1998) found only slight changes in hematologic parameters of well fed captive moose infected with winter ticks. however, the authors speculated that anemia may occur in wild animals on a natural diet. samuel (2004) and musante et al. (2007) modeled the impact of different levels of tick infestations and concluded that blood loss associated with moderate to severe infestations of winter ticks would have measurable and substantial impact on energy and protein balance, and cause anemia and possible mortality of moose calves. they predicted that calves with moderate infestations could lose 1–2 times their blood volume during the peak engorgement period; >40% loss of blood volume over a short period of time can cause death (mcguill and rowan 1989). if these models are correct, winter ticks would likely have less impact on larger adult moose that have larger blood volume and may be in better relative nutritional state in late winter. however, blood loss and/or anemia might negatively affect condition of pregnant cows and post-rut bulls, and although adult moose may be more likely to survive tick infestation, 2 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 productivity might decline, particularly in yearling females (musante et al. 2007, 2010). several techniques have been used to estimate the abundance of questing tick larvae and of adult tick loads on moose. flagging or dragging a sheet over vegetation during the questing period (piesman et al. 1986, aalangdong 1994) was used in elk island national park, alberta to measure the relative abundance of winter ticks in different habitat types to determine whether moose distribution and density in spring dictated distribution and abundance of winter tick larvae (aalangdong 1994). digestion of hide samples and subsequent counting of tick exoskeletons provide accurate estimates of tick load (addison et al. 1979), but may be impractical for managers due to time and cost (samuel 2007). in maine, sine et al. (2009) developed a useful and efficient line-transect method of counting winter ticks on hide samples from harvested moose to estimate/index tick abundance. high sampling rates are possible from harvested moose, but because harvests typically occur during the autumn questing period, onset of winter conditions that would terminate tick activity should be factored to best predict abundance of ticks in spring. the most common method of indexing winter tick abundance and impact on moose is by conducting hair-loss surveys in late winter (samuel and welch 1991, wilton and garner 1993). hair-loss on moose is highly correlated with the rate of grooming against winter ticks (mooring and samuel 1999), and annual hair-loss is correlated with the annual tick load (samuel 2004). further, years with severe hair-loss coincide with large moose die-offs (garner and wilton 1993, wilton and garner 1993). hair-loss surveys conducted since 1984 in algonquin provincial park, ontario have identified a range of hair-loss severity index values (hli) that seem to coincide with moose die-offs (steinberg 2008). in new hampshire, winter ticks probably have more influence on the moose population than disease, predation, habitat, or human-related mortality factors (musante et al. 2010), and predicting the frequency of tick epizootics is an important management consideration. this study was designed to evaluate the accuracy and potential use of 3 approaches to indexing winter tick abundance and epizootics: 1) flagging for tick larvae, 2) line-transect counts of ticks on harvested moose, and 3) roadside surveys of tick-induced hair-loss on moose. methods study area data were collected from 3 northern regions that differed in moose population density (nhfg 2009) (fig. 1); from highest to lowest density they were ct lakes (0.83 moose/km2, se = 0.23), north (0.61 moose/km2, se = 0.15), and white mountains (0.26 moose/km2, se = 0.08), respectively (k. rines, new hampshire fish and game department [nhfg], unpublished data). elevation in the study area ranges from ∼120–1900 m, average snow depth ranges from 0–60 cm, and average temperature from −13 to 19 °c (noaa 1971– 2000). the ct lakes and north regions were dominated by commercial hardwood species including sugar (acer saccharum) and red maple (a. rubrum), yellow birch (betula alleghaniensis), and american beech (fagus grandifolia). red spruce (picea rubens) and balsam fir (abies balsamea) tend to be the dominant species at higher elevations (>760 m) and in cold, wet lowland sites (degraaf et al. 1992). these regions are predominantly forested and the majority of the land is privately owned and commercially harvested using various silvicultural techniques (degraaf et al. 1992); they contain ∼10% wetlands and open water, and alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 3 are interspersed with trails and logging roads. the ct lakes region is hilly with few high mountains, while the north is characterized by high mountainous terrain. the white mountains region contains the white mountain national forest which covers 304,050 ha and is ∼97% forested. it contains the highest elevations in new hampshire and is dominated by beech, sugar maple, and yellow birch; other common species include white ash (fraxinus americana), red maple, red spruce, and eastern hemlock (tsuga canadensis). timber harvest in this region is done on a smaller scale than the other regions, with maximum clear-cut size of ∼10–12 ha (degraff et al. 1992, sperduto and nichols 2004). white-tailed deer (odocoileus virginianus) are sympatric with moose throughout the study area, and at moderate to low density (∼5/mile2; k. gustafson, nhfg, unpublished data) (fig. 1). flagging for tick larvae in each region the relative abundance of winter tick larvae was measured during fall in 10–15 clear-cuts ≥4.05 ha (10 acres) and 2–5 years old. each was sampled every 7–14 days (5–7 times) from 21 september– 12 december 2008 and 12 september–3 december 2009. winter tick larvae were collected by dragging (flagging) a 1 m2 white flannel sheet along parallel transects in each cut (aalongdong 1994). the flannel sheet was attached to a dowel with 2 hose clamps, and held to the side and dragged over the top fig. 1. three regions in northern new hampshire with different moose density; the 3 sampling techniques were used in each region to assess the influence of moose density on winter tick abundance, 2008–2010. 4 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 of vegetation. new transects were established each visit and separated by 10 m buffers to avoid repeat sampling. transects were paced to measure length (m) for calculating relative tick density. the date, time, sample site, and weather were recorded at the beginning of each sampling visit. each flannel was inspected for tick larvae at the completion of a transect, and if present, was stored in a clear plastic bag, labeled with the date, transect number, and clear-cut id, frozen within 2 days, and counted at a later date (aalongdong 1994). sampling ended in each region when prolonged cold and/or permanent snow pack occurred; such conditions cause winter tick larvae to become inactive or die (samuel et al. 2000, samuel 2007). ticks were counted by laying the flannel on a white background and recording with a tally counter (aalongdong 1994). each tick was removed from the sheet with masking tape to avoid double counts; this process was repeated on the opposite side of the sheet. the relative abundance of ticks per region (ticks/m2) was calculated by tallying the total number of ticks in each region and dividing it by the total transect length sampled. analysis of variance (anova) was used to detect differences in relative abundance between regions and between sample years. pairwise comparisons were made with tukey's test; significance level was set a priori at 0.05 for all tests. tick abundance on harvested moose winter ticks were counted on harvested moose brought to moose check stations operated by the nhfg. counts were done for the first 5 days of the moose hunt (saturdaywednesday) in 2008-2010 at the primary check station in each region: pittsburg in the ct lakes, berlin fish hatchery in the north, and twin mountain fish hatchery in the white mountain. winter ticks were counted in situ in 4, 10 × 10 cm sampling plots on a moose carcass: 1) the neck at the base of the skull, 2) the upper edge of the shoulder blade, 3) the rump midway between the hipbone and the base of the tail, and 4) the edge of the rib cage (bergeron 2011). in each plot ticks were counted on 4 parallel, 10 cm transects roughly 2 cm apart; the fur was combed/ held back and all visible ticks were counted along each transect down to the exposed hide (sine et al. 2009). only moose that had been harvested within 5 h were sampled because ticks begin leaving a carcass a few hours after death. time of death, nhfg seal number, and the relative amount of ticks leaving the carcass were recorded at the beginning of each count; biological data and sample region were identified from the seal number. a 10 × 10 cm hide sample was also cut from each of the 4 plot locations, given hunter permission. hide samples were initially cut at a larger size then trimmed to 10 × 10 cm, and ticks were then counted on 4 transects on each sample as described above. each hide sample was labeled with the date, seal number, location of the hide, check station, and then frozen in a sealed plastic bag. total tick counts were accomplished by digesting the hide samples; each was placed in a 1000 ml beaker with 800 ml of 5% potassium hydroxide solution heated to 90 °c until it was fully digested (∼2 h), leaving only the tick exoskeletons intact. the contents were filtered through a 180 µm sieve to separate undigested ticks that were counted under a lighted magnifier (addison et al. 1979). linear regression analysis was used to examine whether the transect counts and hide digestion counts were correlated. this was done to assess the accuracy of performing only transect counts in the field. anova was used to detect differences in transect counts between sample regions, year, and between bulls, cows, and calves. pairwise comparisons were made with tukey's test; alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 5 significance level was set a priori at 0.05 for all tests. roadside surveys of tick-induced hairloss on moose weekly hair-loss surveys were conducted from vehicles on predetermined routes in each of the 3 study regions to measure hair-loss on moose, 1 april–1 june 2009 and 19 april–25 may 2010. routes were chosen to survey traditional roadside salt licks that moose were known to frequent in spring and early summer. surveys coincided with the periods when nymph and adult winter ticks take blood meals and hair-loss is highest (mclaughlin and addison 1986, glines and samuel 1989); surveys should occur as late as possible because grooming against ticks continues through april (samuel 2007). the survey dates were adjusted in 2010 because few moose were observed at salt licks prior to 15 april in 2009. two single-day surveys were also conducted in 2010 to compare with the larger survey. moose were assigned to 1 of 5 categories of hair-loss: no damage to hair, slight damage (∼5–20% hair damaged/lost), moderate (∼20–40%), severe (∼40–80%), and worst case (>80%). when possible both sides of the moose were observed. however, one-sided examination should provide reliable assessment of tick induced hairloss as damage is similar on both right and left sides of moose (samuel and mcpherson 2010). moose were categorized by age and sex, gps locations, and distinguishing characteristics; digital photographs (not all moose) were also used to help distinguish individual moose to avoid double counting. repeat sightings were removed from the analysis by comparing obvious physical characteristics (e.g., antler growth) and photographs when available. other potential repeat sightings were removed by analyzing gps locations in arcgis 9.3 (environmental systems research institute, redlands, ca). buffers of 6.7 km2 were placed around each moose location because this area represents the average spring home range of moose in new hampshire (scarpitti et al. 2005). if the buffers of 2 locations overlapped and the moose was categorized as the same age, sex, and hair-loss category, it was considered a repeat sighting and removed from the analysis. an annual hair-loss severity index (hli) was calculated for each of the 3 sample regions by assigning a number to each hairloss category (1–5), multiplying the number of moose (m) in each category by that number, then dividing the sum of these numbers by the total (t) number of moose observed (wilton and garner 1993, steinberg 2008): hli ¼ m � 1ð þ þ m � 2ð þ þ m � 3ð þ þ m � 4ð þ þ m � 5ð þ t ð1þ these values were compared to trends in flagging and check station data each year, and hlis measured in algonquin provincial park, ontario. a hli was calculated for bulls, cows, and calves with combined regional data each year to identify differences in hli by sex/age. a regional calf:cow ratio was calculated from moose observed in each hair-loss survey. these were compared to ratios calculated the previous fall from moose hunter and deer hunter surveys conducted by nhfg. this exercise was done to investigate whether the proportion of calves declined from fall to spring; measureable calf loss associated with a winter tick epizootic would presumably be identified by a substantially lower calf:cow ratio in spring. results flagging for tick larvae in total, 17,036 ticks were collected on 11.7 ha of sample transect in 2008, and 6 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 11,759 ticks on 17.7 ha in 2009. ticks ranged, per flagging sheet, from 0–2,212. although there was no difference (p >0.05) among regions in the number of ticks either year or within regions between years, fewer ticks (∼40–75%) were collected in each region in 2009. the average relative density in 2008 and 2009, respectively, was 0.19 and 0.11 ticks/m2 (se = 0.09, 0.04) in the ct lakes (max = 1.30, 0.63), 0.16 and 0.07 (se = 0.05, 0.03) in the north (max = 0.62, 0.40), and 0.08 and 0.02 (se = 0.03, 0.01) in the white mountains region (max = 0.41, 0.10) (fig. 2). there was a positive correlation between moose density and tick density in both years (r2 = 0.93 and 0.99). although no significant differences were found among regions or between years, absolute differences were large. mean numbers of ticks declined 42–75% within regions between years, and the mean numbers of ticks were 58 and 82% lower in the white mountain than ct lakes regions in 2008 and 2009, respectively (fig. 2). the mean number of winter ticks collected in individual clear-cuts was below the regional mean in the majority of cuts each year (50–92%) except in the white mountain in 2008. tick abundance on harvested moose the mean number of ticks (all 4 sampling plots and transects) counted on moose ranged from 25–51 (se = 6–7), 42–101 (se = 6–13), and 14–34 (se = 5–15) in the ct lakes, north, and white mountains regions, respectively; highest counts occurred in 2010 in all regions (fig. 3). all life stages of the tick were observed on moose. the mean number of ticks for combined regional data was 53, 31, and 79 (se = 7, 4, 9). tick abundance in the ct lakes in 2010 was ∼2x higher than in 2008 (p = 0.034) and 2009 (p = 0.014) and in the north was ∼1.8x higher in 2008 (p = 0.034) and ∼2.4x in 2010 (p = 0.000) than 2009; tick abundance in the white 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 0.00 0.05 0.10 0.15 0.20 0.25 0.30 white mountainnorthct lakes m a x t ic k s /m ² m e a n t ic k s /m ² sample region 2008 2009 2008 max 2009 max fig. 2. mean (± se) and maximum number of winter tick larvae collected while flagging clear-cuts in 3 sample regions of northern new hampshire, 2008 and 2009. alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 7 mountain region was not different from other regions or between years. tick abundance in the north was ∼3x higher in 2008 (p = 0.006) and ∼2x higher in 2010 (p = 0.038) than in the ct lakes. for all regions combined in 2010, tick abundance was ∼1.5x higher than in 2008 (p = 0.032) and ∼2.5x higher than in 2009 (p = 0.000), and ∼1.7x higher in 2008 than 2009 (p = 0.024) (fig. 3). because regional calf data were minimal, statistical analysis of bull:cow:calf ratios was done using combined regional data. data were from all 4 sampling plots and transects combined. calves had more ticks than adult moose each year, and bulls had more than cows (fig. 4). in 2008, tick abundance on calves was ∼2x higher than bulls (p = 0.014) and ∼6x higher than cows (p = 0.000). tick abundance on calves was ∼4.5x higher (p = 0.004) than on cows in 2009, and tick abundance on calves and bulls was similar and >2x that on cows in 2010 (p = 0.013). a total of 148 hide samples were collected from 66 moose (26 bulls, 36 cows, 4 calves) in 2008 and 2009; 29, 45, 36, and 38 hide samples were collected from the neck, rib, rump, and shoulder, respectively. the number of ticks per transect was positively correlated with the number of ticks counted for all areas of the digested hide samples; r2 values ranged from 0.33–0.99. counts on the rib had the weakest relationship (r2 = 0.33–0.76), however, sample size was low (n = 3–9); combining regional and yearly rib samples yielded r2 = 0.70. fig. 3. mean (± se) number of winter ticks counted on harvested moose in the ct lakes, north, and white mountain sample regions, and combined regional data, in northern new hampshire, 2008–2010. means are for all 4 areas of the hide and all transects combined. numbers in columns represent sample sizes. bars with unlike letters indicate significant differences within sample regions. 8 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 combined regional and yearly data yielded similar r2 values for each area of the hide and all areas combined (r2 ≈ 0.80). roadside surveys of tick-induced hairloss on moose a total of 256 and 222 moose were surveyed in the 3 sample regions during spring 2009 and 2010, respectively: 86 and 72 in ct lakes, 96 and 77 in the north, and 74 and 73 in the white mountains. moose in each hair-loss category were observed each year. in 2009 the ct lakes had the highest hli (3.23), the north was 11% lower (2.91), and the white mountain region was 2.35 or 24% lower. in 2010, hli values were lower in every region; the north region had the highest hli (2.79), the ct lakes was 14% lower (2.44), and the white mountain region was 2.25 or 8% lower (table 1). two single-day surveys were conducted on 12 and 24 may, 2010; however, only the north region produced enough sightings to 0 20 40 60 80 100 120 140 160 180 201020092008 r e la ti v e t ic k a b u n d a n c e year bull cow calf ab b a a b a c b a 26 4 17 23 4 29 26 3 30 fig. 4. mean (± se) number of winter ticks counted on harvested bull, cow, and calf moose in northern new hampshire; data are for all sample regions combined, 2008–2010. means are for all 4 areas of the hide and all transects combined. numbers in columns represent sample sizes. bars with unlike letters indicate significant differences within sample year. table 1. hair-loss severity index (hli) values for 3 sample regions and bull, cow, and calf moose in northern, new hampshire, 2009 and 2010. single-day survey results for each region are included in parentheses (5/12/10 and 5/24/10). bull, cow, and calf data were regionally combined by sample year. region/ moose 2009 n 2010 n ct lakes 3.23 86 2.44 (2.00, 2.00) 72 (8, 3) north 2.91 96 2.79 (2.67, 2.17) 77 (51, 23) white mountain 2.35 74 2.25 (2.22, 2.38) 73 (9, 8) combined 2.86 256 2.50 222 bull 3.07 90 2.65 83 cow 2.70 111 2.45 103 calf 2.75 36 2.29 35 alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 9 make a single-day survey plausible. a total of 51 and 23 moose with corresponding hlis of 2.67 and 2.17 were observed on 12 and 24 may, values 4% and 29% lower than the regional survey. the other regions had <10 moose observations each day. the hli of bulls, cows, and calves ranged from 2.70–3.07 in 2009 and 2.29–2.65 in 2010, and varied little between sex/age of moose (2–16%); hli of bulls was always highest (table 1). calf:cow ratios calculated during spring hair-loss surveys were mid-range of the fall moose hunter and deer hunter surveys, except in the white mountains region in 2010 when it was lower than both surveys. there was little variation in calf:cow ratios among regions and between years; ratios ranged from 0.21–0.34 from moose hunter surveys, 0.33–0.41 from deer hunter surveys, and 0.30–0.43 from hair-loss surveys. the ratio from the single-day survey in the north region (0.38 both days) was mid-range of the moose and deer hunter surveys (table 1). no evidence of a winter tick epizootic or major calf mortality existed either year. discussion flagging for tick larvae abundance of tick larvae was correlated with regional moose density both years, which was consistent with trends identified in previous studies. in elk island national park the average number of ticks on moose increased as moose numbers increased, with a 1-year lag; also, many documented large die-offs of moose in the park occurred at peak moose density. although it is tempting to relate high tick densities with moose dieoffs, similar tick densities occurred in years with and without die-offs in the park (samuel 2004, 2007). clearly the relationship is not exact, and direct comparison of estimates between disparate geographic regions may be unwarranted as other factors, such as weather, likely play a role (delguidice et al. 1997, samuel 2007). the high variability in tick abundance in clear-cuts likely influenced the lack of statistical differences among regions and between years. regional means were highly influenced by a few cuts with high abundance of ticks, and the high variability among clear-cuts suggests that winter ticks are not evenly distributed even within this preferred habitat type of moose. certain clear-cuts in each region had abundance 2–7x higher than the regional mean both years; this distribution pattern may explain why individual hair-loss varies annually, and certain moose have severe hair-loss in years of overall light infestation and vice versa. conversely, local sites with high moose and tick density may explain, in part, why epizootics usually occur across wide geographic ranges that encompass variable moose population densities. a benefit of this sampling method is that it can extend through the entire questing period, which usually occurs from september until winter conditions kill remaining unattached larvae (usually november-december) (drew and samuel 1985, samuel 2004); questing usually stops at <0° c (samuel and welch 1991). because our sampling occurred from early-mid september through the first substantial snowfall, it should be representative of the relative abundance of ticks. however, because temperature and snow condition varied considerably among the adjacent study regions, tick abundance from flagging alone would not necessarily reflect regional tick abundance on moose. however, it may be possible to detect annual regional trends in tick abundance because tick numbers declined in each region from 2008 to 2009. the data also suggest that moose density influences tick abundance because relative tick density was correlated with regional moose density both years. 10 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 the flagging technique is probably not practical to index tick abundance because it is extremely labor intensive and costly. sampling occurred for ∼3 months and across a wide geographic range. two people sampled a clear-cut in ∼2 h and each needed to be visited multiple times; workdays averaged 8–10 h and counts of larvae at a later date on each flannel required 10 min–>1 h depending on the number of ticks. the relative length of the questing period is probably most easily estimated by tracking ambient temperature and snow cover, and assuming that an extended warm fall will lengthen the questing period and tick abundance. tick abundance on harvested moose tick abundance measured directly on harvested moose was highest in the north and lowest in the white mountains region each year; conversely, flagging (sampling for larvae) measurements were correlated with regional moose density. tick abundance was higher on calves than bulls and cows each year. drew and samuel (1985) suggested that bulls may have the highest absolute numbers of ticks due to their size and increased activity during the rut; however, calves have proportionally more ticks (per area) due to their smaller body size (samuel and barker 1979, samuel 2004). there was a strong relationship between transect counts and total counts from hide digestions, and the strongest relationships occurred when data from all areas of the hide were combined (r2 = 0.80). sine et al. (2009) also found strong relationships (r2 = 0.88) when combining hide samples and concluded that the total number of ticks counted on all transects was the best predictor of tick density on moose. due to the strong relationship between transect and total counts in both studies, we suggest that transect counts (easy and efficient) should suffice for use as an index of relative tick abundance on harvested moose. average time to count the 4 areas of hide was ∼5 min with a separate counter and recorder, and about twice as long if done alone (same as sine et al. 2009). further, some hunters were unwilling to donate hide samples and laboratory work was tedious and labor intensive; hide samples took ∼2 h to digest and counting tick exoskeletons varied from a few minutes to hours. the transect method identified differences between regions and years, but did not indicate a positive correlation with moose density as did the flagging method. because the moose harvest in new hampshire occurs in mid-october, this method may not translate directly to tick load and/ or related moose mortality if moose disproportionately acquire ticks in late fall. aggregations of winter tick larvae can survive into november (drew and samuel 1985), and tick larvae were collected into december during flagging. however, if the timing of the hunting season remains constant, a useful index of relative tick abundance should be evident with a few additional years of data. further, the highest tick abundance measured in fall 2010 preceded an epizootic in 2011 throughout the northeastern united states (pers. comm., l. kantar and k. rines, maine inland wildlife and fisheries and nhfg, respectively). roadside surveys of tick-induced hairloss on moose the hli values were correlated positively with regional moose density in 2009, as was the flagging method; however, although hli values declined in each region in 2010, the north region had the highest hli, the same pattern as occurred with tick abundance on harvested moose. overall, all methods indicated a reduction in tick numbers from the first to second year of the study (2008–2009: flagging and harvested moose, and 2009–2010: hair-loss) with combined data from all regions, suggesting that singly, alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 11 none is sensitive enough to detect potential differences in tick abundance among regions; however, any would probably detect large annual change in relative tick abundance. hair-loss surveys conducted in algonquin provincial park, ontario since 1984 (steinberg 2008) had hli values ranging from 1.18–3.48; hli's ≥2.50 were associated with mortality events in 1992 and 1999. however, no epizootic occurred in 1988, 2000, or 2006 with similar values. the hli values in this study were 2.20– 3.23 with the majority >2.5, but no major mortality event was evident. however, direct comparison with hli values in new hampshire are probably unwarranted because helicopter surveys are usually conducted in march in algonquin park (due to snow cover), whereas surveys occurred in aprilmay in new hampshire when hairloss is more evident (mclaughlin and addison 1986). bulls had the highest hli both years, suggesting that rutting activity of bulls during the fall questing season contributes to their tick load (drew and samuel 1985, samuel 2004). however, there was little variation overall (2–16%) in the hli of bulls, cows, and calves indicating that sex/age of moose might have little influence on survey results. a minimum of 50 moose is considered an adequate sample in algonquin provincial park (steinberg 2008), and this sample size was realized in a single day survey on 12 may in the north region (n = 51). the hli (2.67) was similar (4% lower) to that of the weekly survey (2.79), suggesting that a single-day survey should suffice given adequate sample size. in new hampshire surveys should be conducted as multiple, morning surveys, preferably condensed within a 5-day period (1 may–15 may), that are focused on the most commonly used roadside saltlicks in the highest moose density regions; the survey would be complete with ≥50 individual moose. routes within a region should be separated to ensure that the same moose is not observed at different licks by multiple observers (or use a single observer). this would reduce the duration of surveys, distance traveled, and eliminate repeat sightings. surveys should also be conducted on cool mornings with little precipitation to enhance sightings. because calves are likely most impacted by winter ticks and is the cohort most susceptible to mortality, estimates of fall and spring calf:cow ratios should indicate substantial mortality events that reduce the proportion of calves in the population. calf: cow ratios calculated from fall hunter surveys and spring hair-loss surveys were reasonably similar, and calf:cow ratios during the single-day survey in the north region (n = 51) were similar to those in the weekly survey. low sample size may be problematic for calculating such ratios in spring, and the reliability and sensitivity to detect such change is unknown because no evident dieoff occurred. conclusions although moose density and tick abundance were generally related in new hampshire, weather plays a strong role in the abundance and distribution of winter ticks (samuel and welch 1991, samuel 2007). regional weather differences that impact ticks at different life stages likely influenced regional tick abundance regardless of moose density. because major moose die-offs are usually concurrent and widespread geographically (samuel 2004), tracking regional differences in new hampshire may not be as important as obtaining adequate tick abundance samples from harvested moose and at least one regional sample of 50 moose from roadside hair-loss surveys. the combination of fall tick counts on harvested moose and spring hair-loss surveys 12 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 should prove useful in indexing winter tick abundance in northern new hampshire. both methods are time and cost-effective and capable of indicating annual change in relative tick abundance. check station counts provide an indication of transmission during the questing period; however, if weather conditions were to extend the questing period into december, check station counts may no longer be representative of actual tick loads. hair-loss surveys should help identify high tick abundance in late winter-spring caused by an extended questing period, and calf:cow ratios from the surveys could detect years of high calf mortality. combined use and comparison of these methods will increase confidence in their index value; of particular future interest is an ability to identify threshold values associated with major moose die-offs. acknowledgements funding for this research was provided by the new hampshire fish and game department through efforts of k. rines. we are grateful to the many students from the university of new hampshire who helped drag for ticks, and count ticks in the lab and on harvested moose; many nhfg biologists also counted ticks on harvested moose at check stations. c. rogers and nhfg biologists helped conduct roadside surveys. h. andreozzi assisted with tables, figures, and early versions of the paper. literature cited aalangdong, o. i. 1994. winter tick (dermacentor albipictus) ecology and transmission in elk island national park, alberta. m. s. thesis. university of alberta, edmonton, canada. addison, e. m., d. g. joachim, r. f. mclaughlin, and d. j. h. fraser. 1998. ovipositional development and fecundity of dermacentor albipictus (acari: ixodidae) from moose. alces 34: 165–172. ———, f. j. johnson, and a. fyvie. 1979. dermacentor albipictus on moose (alces alces) in ontario. journal of wildlife diseases 15: 281–284. ———, and r. f. mclaughlin. 1988. growth and development of winter tick, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670–678. ———, ———, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and unifested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469–1476. berg, w. e. 1975. management implications of natural mortality of moose in northwestern minnesota. proceedings of the north american moose conference and workshop 11: 332–342. bergeron, d. h. 2011. assessing relationships of moose populations, winter ticks, and forest regeneration in northern new hampshire. m. s. thesis. university of new hampshire, durham, new hampshire, usa. bishopp, f. c., and h. l. trembley. 1945. distribution and hosts of certain north american ticks. the journal of parasitology 31: 1–54. blyth, c. b. 1995. dynamics of ungulate populations in elk island national park. m. s. thesis. department of agricultural, food, and nutritional science, university of alberta, edmonton, alberta, canada. cameron, a. e., and j. s. fulton. 1926– 1927. a local outbreak of the winter or moose tick, dermacentor albipictus pack. (ixodoidea) in saskatchewan. bulletin of entomological research 17: 249–257. degraaf, r. m., m. yamasaki, w. b. leak, and j. w. lanier. 1992. new england wildlife: management of forested habitats. general technical report ne-144. alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 13 usda forest service, northeast experiment station, radnor, pennsylvania, usa. delgiudice, g. d., r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restrictions, ticks, and numbers of moose on isle royale. journal of wildlife management 61: 895–903. drew, m. l., and w. m. samuel. 1985. factors affecting transmission of larval winter ticks, dermacentor albipictus (packard), to moose, alces alces l., in alberta, canada. journal of wildlife diseases 21: 274–282. ———, and ———. 1986. reproduction of the winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64: 714–721. garner, d. l., and m. l. wilton. 1993. the potential role of winter tick (dermacentor albipictus) in the dynamics of a south central ontario moose population. alces 29: 169–173. glines, m. v., and w. m. samuel. 1984. the development of the winter tick, dermacentor albipictus, and its effect on the hair coat of moose, alces alces, of central alberta, canada. pages 1208–1214 in d. a. griffiths and c. e. bowman, editors. acarology, vi. ellis horwood ltd., chichester, england. ———, and ———. 1989. effects of dermacentor albipictus (acari: ixodidae) on blood composition, weight gain, and hair coat of moose, alces alces. experimental and applied acarology 6: 197–213. lankester, m. w., and w. m. samuel. 1998. pests, parasites, and diseases. pages 479–517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. mcguill, m. w., and a. n. rowan. 1989. biological effects of blood loss: implications for sampling volumes and techniques. institute for laboratory animal research news 31: 5–18. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)-induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. mooring, m. s., and w. m. samuel. 1999. premature loss of winter hair in free-ranging moose (alces alces) infested with winter ticks (dermacentor albipictus) is correlated with grooming rate. canadian journal of zoology 77: 148–156. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–111. ———, ———, and ———. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. new hampshire fish and game department (nhfg). 2009. wildlife harvest summary. new hampshire fish and game department, concord, new hampshire, usa. piesman, j., j. g. donahue, t. n. mather, and a. spielman. 1986. transovarially acquired lyme disease spirochetes (borrelia burgdorferi) in field-collected larval ixodes dammini (acari: ixodidae). journal of medical entomology 23: 219. pybus, m. j. 1999. moose and ticks in alberta: a die-off in 1998/99. occasional paper no. 20. fisheries and wildlife management division, edmonton, alberta, canada. samuel, w. m. 1991. grooming by moose (alces alces) infested with the winter tick, dermacentor albipictus (acari): a mechanism for premature loss of winter hair. canadian journal of zoology 69: 1255–1260. ———. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta. 14 assessing winter tick abundance – bergeron and pekins alces vol. 50, 2014 ———. 2007. factors affecting epizootics of winter ticks and mortality of moose. alces 43: 39–48. ———, and m. j. barker. 1979. the winter tick dermacentor albipictus (packard, 1869) on moose, alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. ———, and l. mcpherson. 2010. tick tidbits: do winter ticks cause equal damage to the haircoat of moose on right and left sides? the moose call 24: 20–22. ———, m. s. mooring, and o. i. aalangdong. 2000. adaptations of winter ticks (dermacentor albipictus) to invade moose and moose to evade ticks. alces 36: 183–195. ———, and d. a. welch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27: 169–182. scarpitti, d., c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. sine, m. e., k. morris, and d. knupp. 2009. assessment of a line transect method to determine winter tick abundance on moose. alces 45: 143–146. sperduto, d. d., and w. f. nichols. 2004. natural communities of new hampshire. new hampshire natural heritage bureau, concord, new hampshire, usa. steinberg, b. 2008. algonquin park moose hair-loss survey report 2008. ontario parks, ontario, canada. welch, d. a., w. m. samuel, and r. j. hudson. 1990. bioenergetic consequences of alopecia induced by dermacentor albipictus (acari: ixodidae) on moose. journal of medical entomology 27: 656–660. wilkinson, p. r. 1967. the distribution of dermacentor ticks in canada in relation to bioclimatic zones. canadian journal of zoology 45: 517–537. wilton, m. l., and d. l. garner. 1993. preliminary observations regarding mean april temperature as a possible predictor of tick-induced hair-loss on moose in south central ontario, canada. alces 29: 197–200. wright, j. e. 1969. hormonal termination of larval diapause in dermacentor albipictus. science 163: 390–391. zarnke, r. l., w. m. samuel, a. w. franzmann, and r. barrett. 1990. factors influencing the potential establishment of the winter tick (dermacentor albipictus) in alaska. journal of wildlife diseases 26: 412–415. alces vol. 50, 2014 bergeron and pekins – assessing winter tick abundance 15 evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire methods study area flagging for tick larvae tick abundance on harvested moose roadside surveys of tick-nduced hair-oss on moose results flagging for tick larvae tick abundance on harvested moose roadside surveys of tick-nduced hair-oss on moose discussion flagging for tick larvae tick abundance on harvested moose roadside surveys of tick-nduced hair-oss on moose conclusions acknowledgements literature cited browse removal, plant condition, and twinning rates before and after short-term changes in moose density thomas f. paragi1, c. tom seaton1, kalin a. kellie1, rodney d. boertje1,2, knut kielland3, donald d. young, jr.1, mark a. keech1,4, and stephen d. dubois5,6 1alaska department of fish and game, 1300 college road, fairbanks, alaska 99701, usa; 3institute of arctic biology, university of alaska-fairbanks, fairbanks, alaska 99775, usa; 5alaska department of fish and game, p.o. box 605, delta junction, alaska 99737, usa abstract: we monitored forage-based indices of intraspecific competition at changing moose (alces alces) densities to gauge short-term, density-dependent environmental feedback and to ultimately improve management of moose for elevated sustained yield. in 4 areas of interior alaska where moose density recently changed, we evaluated the magnitude of change among 4 browse indices: proportional offtake of current annual growth biomass (oftk), proportion of current twigs that were browsed (ptb), mean twig diameter at point of browsing (dpb), and proportion of plants with broomed architecture. in 1 area where moose density increased 100% in 6 years following effective predation control, browse removal increased 138% for oftk, 20% for ptb, and 16–42% for dpb of primary browse species, with a 44% increase in brooming. we also studied 3 areas where moose density declined 31–41% following elevated antlerless harvests of 2–4 years duration. in these areas (with intervals of 3–12 years between browse surveys) we found declines of 30–40% in oftk, 26–68% in ptb, and 11–37% in dpb, but changes in plant architecture were inconsistent. the proportion of parturient cows with neonate twins did not change between browse surveys, presumably because of a substantial lag time influenced by life history of the dominant reproductive cohorts and little change in browse nutrient content and digestibility. of the 4 browse indices studied, proportional oftk most consistently reflected the direction and magnitude of short-term changes in moose density. area-specific measures of habitat and animal conditions at high moose density provided an objective means for gauging the capacity of the respective ecosystems to support moose and maintain forage plants. we used these measures of winter forage and moose condition to justify implementing harvest strategies and to ultimately reduce high moose densities below levels of strong negative feedback. alces vol. 51: 1–21 (2015) key words: alaska, density-dependent, forage, intraspecific competition, moose, nutritional condition. moose (alces alces) management becomes increasingly challenging for populations at the extremes of the nutritional gradient. at low densities managers may consider predator control to increase abundance (gasaway et al. 1992). at high densities habitat enhancement may be an option to increase forage, or antlerless harvests could reduce abundance or population growth rates (boertje et al. 2009, young and boertje 2011). wildlife managers in alaska are often required to estimate harvestable surplus and nutritional status of wild moose populations over large areas (≤15,000 km2) of remote forested and subalpine habitats. it is difficult to estimate the 2present address: 220 blue bird, kerrville, texas 78028, usa 4present address: p.o. box 84634, fairbanks, alaska 99708, usa 6present address: p.o. box 702, delta junction, alaska 99737, usa 1 capability of habitats to support moose because of limited studies on the physiological requirements based on captive animals (reviewed in schwartz and renecker 1997) and the inherent variability in habitat and other environmental factors. biologists must either quantify forage production (kg/ha) in the context of daily food requirements for an absolute estimate of carrying capacity (e.g., wolff and zasada 1979, crete 1989, maccracken et al. 1997) or use indices to assess the relative nutritional status of the moose population and/or condition of the range. there are no standardized economical methods for assessing landscape carrying capacity in remote areas of alaska, so biologists use nutritional status of the moose population or indices related to winter forage use (boertje et al. 2007, seaton et al. 2011). these indices presumably reflect the negative feedback in nutrition from increased intraspecific competition for food resources at increasing moose density and indicate if competition is reduced as density declines. negative feedback reduces productivity and thus sustainable harvest among age and sex classes (mccullough 1984), poses a heightened risk of unsustainable forage removal, and can lead to dramatic population declines, often facilitated through winter-related mortality (gasaway et al. 1983). the most established index of nutritional status of a moose population in interior alaska is twinning rate, the proportion of parturient females with 2 neonatal calves (franzmann and schwartz 1985, keech et al. 2000, boertje et al. 2007). dressed weights of harvested calves (e.g., cederlund et al. 1991), age at first reproduction, short-yearling live mass, and browse removal rate (boertje et al. 2007) have also been used to estimate or gauge nutritional status of moose populations. however, few studies have examined how well these indices respond following intended changes in abundance through management actions. measuring animal indices can be constrained by sample size at low density, limiting their usefulness in monitoring change in abundance. it can be difficult or infeasible to observe an adequate sample of random parturient females from aircraft for estimating twinning rate in areas of low moose density (e.g., stout 2010) or in dense cover that hinders viewing of calves. conversely, browse sampling is not constrained by moose observations. seaton et al. (2011) documented an inverse correlation between proportional browse biomass removal and twinning rate across a 10-fold range in density (0.1–1.2 moose/km2) among 8 game management units of interior alaska. that study demonstrated the utility of a habitat metric for indirectly judging nutritional condition of adult female moose, which helped substantiate prior conclusions by boertje et al. (2007) based on smaller sample sizes. in this study we sought to document shortterm, landscape-level changes in browse removal rates and architecture of winter forage species following short-term, managementinduced changes in moose density. seaton (2002) reviewed methods of estimating browse removal by moose and used a modified technique to characterize “apparent” browse production (prod; kg/ha above snow) and browse offtake by moose (oftk; kg/ha), and to estimate proportional oftk (oftk/prod). the technique quantifies woody biomass through measuring twig diameter at the proximal end of current annual growth (cag) and the diameter at point of browsing (dpb) in late winter, just prior to the new growing season. earlier studies reported a correspondence between proportional oftk and the proportion of twigs browsed by moose (ptb = number of dpb > 0 divided by number of cag) (regelin et al. 1987, maccracken and viereck 1990), which suggested that the simpler twig count would be more efficient. however, seaton (2002:32) cited another 2 browse indices to moose density – paragi et al. alces vol. 51, 2015 study (k. kielland and t. osborne, unpublished data) where ptb was insensitive to a large (8-fold) change in moose abundance in western interior alaska. thus, estimating the biomass produced and removed with diameter measurements is important because moose may clip twigs at a range of diameters, and the nutritional value (e.g., digestibility and nutrient concentrations) decreases as cag diameter increases (vivås and sæther 1987, kielland and osborne 1998). the smallest diameter twigs provide the most nutrient gain per unit of mass but extend rumen fill time, whereas the largest diameter twigs provide less nutrient gain per unit mass and extend rumen processing time (gasaway and coady 1974, shipley and spalinger 1992). seaton (2002) also sought to incorporate forage plant architecture to gauge the longer-term effects of moose browsing. this information might additionally help managers and the public understand negative feedback at higher moose densities and characterize relatively less use at lower densities. in this study we followed a recommendation by seaton et al. (2011) to evaluate the utility of browse indices for detecting short-term changes in intraspecific competition following intended management actions. in 4 areas with baseline browse data and subsequent changes in moose density, we examined the magnitude of changes in browse removal and plant architecture as gauges of density-dependent feedback. we compared changes in browse metrics with changes in an established index of nutrition (twinning rates) to better understand how browse removal may or may not reflect changes in moose nutrition. we assumed that reducing moose abundance in a defined area where forage production changed relatively little over time will reduce intraspecific competition for preferred species, with an inverse response following an increase in abundance. we predicted that increased moose density would cause increased proportional oftk, increased ptb, increased mean dpb for at least the dominant or preferred browse species, and an increased proportion of plants with architecture partly or heavily affected by moose foraging. these conditions are presumed to be coincident with a decrease in moose nutrition, resulting in a lower twinning rate at higher density. where moose density decreased, we predicted the inverse responses, with one exception; that short-term reversal of trend in plant architecture as affected by moose browsing at high density (broomed → unbrowsed) would be unlikely on existing plants in the absence of widespread disturbance to regenerate young plants, such as fire or flooding. also, moose nutrition, as indexed by twinning rates, would not likely increase immediately following a decline in moose abundance unless browse quantity and quality increased substantially. intentionally changing reproductive rates is more likely a long-term proposition based on changing calf weights and eventually the life history of the dominant reproductive cohorts (females 4–10 yr old; boertje et al. 2007). study areas and moose abundance the 4 study areas (unit 19d, unit 20a central hills, unit 20a western flats, and unit 20d) were located in the boreal forest of interior alaska, usa (seaton et al. 2011, fig. 1). management actions were implemented in these areas to influence moose abundance. to illustrate the magnitude of density change in each area, we calculated moose abundance and confidence intervals for areas approximating the extent of browse sampling before and after management actions. population increase unit 19d in the remote kuskokwim valley is comprised of large floodplains and alces vol. 51, 2015 paragi et al. – browse indices to moose density 3 fig. 1. continued on next page. 4 browse indices to moose density – paragi et al. alces vol. 51, 2015 fig. 1. post-treatment sampling grids and browse plot locations before and after management actions intended to affect moose density for the 4 study areas in interior alaska. alces vol. 51, 2015 paragi et al. – browse indices to moose density 5 forested uplands within 40 km of mcgrath (62° 57′ n, 155° 36′ w). in this study area we sampled browse on 3 occasions: broadly over 10,600 km2 in 2001, narrowly in a 1368 km2 experimental area (keech et al. 2011) in 2003, and over a 2896 km2 moose survey area that included the experimental area in 2009 (keech 2012). we used the 2003 browse data for pre-treatment in unit 19d because sampling scale (paragi et al. 2008:36) was closer to that from 2009 (fig. 1a), but we provided 2001 data for additional pre-treatment context on variation in browse metrics when moose density was low (paragi et al. 2008:9). in unit 19d, keech et al. (2011) described predation control beginning in 2003 that caused a gradual increase in moose abundance over the next 6 years. in the 2009 browse sampling area, moose density estimated from fall aerial surveys doubled from 0.30/km2 in 2001 to 0.62/km2 by 2009 (keech 2012:14). population decrease units 20a and 20d are near the road system in the tanana valley near the city of fairbanks and the town of delta junction, respectively. the unit 20a central hills study area was comprised of forested uplands and subalpine shrubs 100 km south of fairbanks in the foothills of the alaska range (64° 08′ n, 147° 55′ w). this area was primarily used during fall and winter by moose that migrated to lowland flats to the north during the summer (keech et al. 2000), inclusive of when hunting and abundance surveys occurred. we sampled 600 km2 of the unit 20a central hills as part of a larger study by seaton (2002) in 2000 and sampled 790 km2 in 2012 that conformed to a harvest-reporting boundary and largely overlapped the earlier browsesampling area (fig. 1b). the unit 20a western flats (64° 26′ n, 148° 50′ w) were forested lowlands with interspersed wetlands. we sampled 1100 km2 of the western flats in 2006 and 1625 km2 in 2009; the latter survey largely overlapped the earlier survey area, but nearly half of it was influenced by large fires in 2006 and 2009 (fig. 1c). the unit 20d study area (63° 46′ n, 145° 15′ w) varied from agricultural lands and forest with several upland areas that had burned in the last 20 years near delta junction to subalpine scrub in the foothills of the alaska range 50 km south. we sampled browse in southwestern unit 20d over 3250 km2 in 2007 and 2010 (fig. 1d). in unit 20a, antlerless harvests were implemented or expanded to reduce moose abundance during 2004–07 by use of hunt zones and extended seasons, including the central hills and western flats (young and boertje 2011). we subsampled data from the larger unit 20a surveys before and after antlerless harvests to estimate post hoc moose abundance in browse survey areas in the hills and flats (kellie and delong 2006). in southwest unit 20d, antlerless harvest reduced moose abundance in fall 2007 (dubois 2008:397) and fall 2008 (dubois 2010:390). the unit 20d antlerless hunts were short duration prior to substantive snowfall, with female harvest predominantly occurring in the lowland flats north of the foothills (s. dubois, unpublished data) that were accessible by all-terrain vehicles. similar to the abundance estimates for unit 19d (keech et al. 2011, keech 2012), in the other 3 study areas we multiplied estimated abundance by a sightability correction factor (scf) for moose not observed using radiomarked individuals. we used scf = 1.21 for unit 20a (boertje et al. 2009) and scf = 1.1 for unit 20d (s. dubois, unpublished data), and the scf variance was incorporated with that of the geospatial population estimator (gspe; kellie and delong 2006) into 90% confidence limits (goodman 1960, keech et al. 2011). 6 browse indices to moose density – paragi et al. alces vol. 51, 2015 methods browse removal we used moose distribution from fall abundance surveys during shallow (<40 cm) snow as the sampling extent for browse surveys and attempted to minimize sampling bias at the landscape and vegetative stand scales. our landscape sampling design and procedures for selecting plot locations began with pre-selected random points among vegetation strata (seaton 2002) but evolved with logistical experience in the field (paragi et al. 2008:2–4). since 2006 we have primarily used rectangular gspe cells based on 2 minutes of latitude and 5 minutes of longitude (ca. 3.7 × 4.1 km) from recent moose surveys for stratified random sampling at a 3:2 ratio of high:low moose density (e.g., kellie and delong 2006:21). most plot access in remote areas was by helicopter, but we used vehicles where portions of unit 20d were near a highway or forest road (paragi et al. 2008:4). the landscape sampling protocol was developed in boreal forest, but we accommodated the linear nature of riparian browse distribution when we began sampling subalpine habitats. the helicopter flew on the designated course within gspe cells until the first patch of browse ≥0.5 m tall and above snow was encountered, at which point we selected a randomized distance (30–100 m from the nearest safe landing spot) and direction (3 tries to select a site with browse before sampling cell was skipped) for choosing the plot center in the vegetation stand. one exception to stratified sampling with gspe cells was the addition in 2009 of ad hoc systematic plots to ensure adequate sampling of the riparian zone along the kuskokwim river and takotna river in unit 19d for comparison to earlier sampling stratified by vegetation type (paragi et al. 2008). we chose a random starting point along the kuskokwim river near the eastern boundary of the sampling area and landed at the nearest willow bar every 10 km downriver (straight line by helicopter) and also every 10 km upriver on the takotna from its confluence. in 2012 we began defining study areas in unit 20a by polygon boundaries based on drainages that are used for cataloging moose harvest location from hunter reports so that inference about browse could be more directly related to changes in reported harvest. our objective was to sample at least 30 plots per study area with browse above snow to optimize precision and cost (seaton et al. 2011), so we typically selected at least 40 sample cells because some random plot locations near helicopter landing spots do not contain browse. this sample size was not achieved in unit 20a western flats in 2006, where we omitted several sites due to absence of browse near safe landing zones. however, the 15 plots achieved in this survey (table 1) are expected to accurately reflect the biomass removal level but with potentially high variance (seaton et al. 2011, fig. 3). where clumps of randomly chosen sample cells occur, sampling at least one cell in the clump provided landscape coverage if logistics became limiting (e.g., degraded flying weather or distance from fuel). we analyzed proportional oftk over the winter to describe the interaction between moose and their winter forage. the rationale for plot sampling and browse metric analysis is described elsewhere (seaton 2002, paragi et al. 2008). snow depth >70 cm can restrict access to forage, increase energetic requirements for locomotion, and influence habitat selection; snow depth >90 cm greatly restricts movement, potentially hindering adequate forage intake (coady 1974). consequently, we recorded snow depth at plots during browse surveys for a context of winter severity, particularly as a confounding factor between sampling events. we sampled only plants above snow alces vol. 51, 2015 paragi et al. – browse indices to moose density 7 with measurable cag between 0.5 and 3.0 m above ground level in a 15-m radius plot near the end of browse removal in late winter (late march or early april, before leaf emergence). we randomly selected 3 plants per species present, using plants as the sample unit for inference on browse removal at the scale of study area. plant taxonomy followed collet (2004) for willows and viereck and little (2007) for other species, with winter willow identification aided by an unpublished guide (d. simpson, alaska department of fish and game [adfg] 1986). for each randomly selected plant within a species, we randomly selected 10 twigs. for each twig we recorded to 0.1 mm precision the dpb if applicable, and cag (lyon 1970). we then counted the total number of twigs with cag on each of the 3 plants. we used the regression coefficients relating diameter to dry mass (paragi et al. 2008:40–41) and the number of twigs with cag per plant to estimate prod and oftk (telfer 1969). an exception was salix lasiandra for which 3 plants were measured on each of 2 plots in unit 20a western flats in 2006. we did not have a regression equation for s. lasiandra, so we used s. bebbiana equations for biomass analysis because these twigs have a similar morphology. we estimated oftk based on sampled twigs only (mean twig per sampled plant) with plants as the sample unit. we extrapolated prod and oftk from sampled twigs to the plot level for comparison among study areas, recognizing that this may introduce sampling bias through variation in the proportion of total plants sampled per species and variability in plant counts within plots. we used software written in r language (r development core team 2008, version 2.1.1; code and instructions available under project 5.10 at to read a microsoft® access® (version 2003) database containing plot counts, twig diameters, diameter– biomass pairs, and dry-weight conversions. we used this software to estimate the diameter–biomass relationships, prod, and oftk on the basis of plant, species, plot, and study area (paragi et al. 2008). we applied binomial 95% confidence limits (cochran 1977:58) with n as the number of plants measured, rather than twigs, to avoid table 1. sampling details and estimates of apparent browse production (sampled twigs above snow extrapolated to plot composition) by study area within game management units in interior alaska. game management unit browse sampling year sampling area (km2) browse samples (n) apparent production (kg / ha) plots plants twigs x ̄ 95% ci 19d 2001 10,600 36 251 2420 201 19 19d 2003 1368 39 298 2377 689 52 19d 2009 2896 42 278 2746 343 26 20a c. hills 2000 600 49 235 2504 745 154 20a c. hills 2012 790 37 177 1799 30 3.8 20a w. flats 2006 1100 15 109 1099 75 9.0 20a w. flats 2012 1625 44 312 2945 14 1.0 20d 2007 3250 75 437 4312 52 4.7 20d 2010 3250 57 431 4108 73 8.0 8 browse indices to moose density – paragi et al. alces vol. 51, 2015 http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat pseudo-replication (unequal proportion of plants sampled per species and per plot) and to portray a more conservative variation. where moose density changed between browse evaluation periods, we evaluated significant probability of increase or decrease in twig metrics with 1-tailed tests, where direction of change associated with removal was predicted to be the same as direction of change in moose abundance. we tested for difference between proportions of oftk and of ptb using a z-test (zar 1984:396) with the smaller number of plants for degrees of freedom in the t distribution. we tested for difference in snow depth and in dpb before and after management actions for each browse species using mann-whitney u (conover 1980) because data distributions were often non-normal (lilliefor’s test, p < 0.05). browse architecture seaton (2002:19) classified forage plants based on their history of browsing by moose and the resulting compensatory growth, termed “architecture.” in contrast to removal of cag in a specific winter, architecture describes multi-year growth history and is unidirectional (unbrowsed → browsed → broomed) unless disturbance resets plants to an unbrowsed canopy. thus, in addition to historic moose density, architecture in study areas is influenced by fire or vegetation management in prior years that stimulates young growth. three categories of plant architecture were defined from evidence of browsing prior to the current year for each plant: “unbrowsed” (no evidence of browsing prior to the current year); “browsed” (browsing in past years but <50% cag twigs between 0.5 and 3.0 m arose from lateral stems that were produced as a result of browsing); and “broomed” (>50% of cag twigs between 0.5 and 3.0 m arose as lateral stems). to reduce measurement error, architecture was classified by the first 3 authors or under their direct supervision. we used a chi-square test for independence in proportions of the plant architecture classes and portrayed variation in the proportions of plants in architecture classes with binomial confidence limits using n as the number of plants. twinning rate boertje et al. (2007) described estimation of moose twinning rates from aerial surveys shortly after peak of calving in late may. we obtained data from area or research biologists that conducted surveys annually in our 4 study areas. to evaluate trend in twinning rate, we used r script to estimate the mean rate and 95% confidence limits using a parametric bootstrap (100,000 repetitions). results moose density in the unit 20a central hills was 2.4 moose/km2 prior to antlerless harvest and was reduced 33% by the 2nd browse survey (table 2). the 2003 abundance estimate likely represented the peak density 3 years after the first browse survey in 2000 as inferred from abundance estimates in the larger unit 20a (young and boertje 2011); thus, moose density during the first browse survey may have been slightly lower. the 2012 abundance estimate was 5 years after the end of liberal antlerless harvest, a period of reduced and comparatively stable moose abundance in all of unit 20a (young 2012, table 2). in unit 20a western flats, the evidence for a 31% decline in density from 1.2 moose/km2 in 2006 was weaker because the estimate had twice the proportional variance of the other study areas (table 2). the 2006 browse survey occurred after 2 years of liberal antlerless harvest, thus likely reflected a reduced moose density (lesser expected difference in browse removal) from the peak in fall 2003 for all of unit 20a (young and boertje 2011). this population may have experienced a relatively smaller change in alces vol. 51, 2015 paragi et al. – browse indices to moose density 9 abundance than the other 3 study areas, so we expected that the magnitude of change in browse metrics and twinning rate might be ambiguous with respect to the other 3 study areas. moose density in unit 20d was 2.1 moose/km2 before antlerless harvest; and was reduced 41% by 2010. willows (salix spp.) dominated or codominated apparent browse production in most instances (fig. 2). s. alaxensis composed the majority of browse biomass in active riparian floodplains regardless of elevation (e.g., including incised drainages in subalpline), whereas s. pulchra often dominated or co-dominated production with betula neoalaskana or populus tremuloides in upland sites, particularly after recent fires or logging. prod ranged greatly among study areas and within study areas between years (14–745 kg/ha; table 1). within a study area, the greatest change was in unit 20a central hills that was possibly influenced by different browse sampling stratifications before and after antlerless harvest. sampling in the 2000 browse survey in the 20a central hills included 1 plot of extremely high production (22,148 kg/ha; paragi et al. 2008:53) that boosted mean apparent production from 329 to 745 kg/ha. lower prod in post-treatment browse surveys for units 19d and 20a western flats may also reflect slow vegetative recovery from recent fires (fig. 2a and 2c). oftk exceeded 45% for dominant willow species at higher moose densities within study areas (fig. 2): s. alaxensis in unit 19d (2009) and unit 20a central hills (2000), and s. pulchra in unit 20a central hills (2000) and unit 20a western flats (2006). mean snow depth (7–33 cm) during browse surveys differed little between table 2. estimates of moose density and browse removal (sampled twigs only) by moose reported by study area within game management units in interior alaska. moose abundance surveys were in early winter prior to the associated browse surveys unless otherwise noted. proportions of offtake and twigs browsed were predicted to change in the direction of trend in moose density. game management unit browse survey year moose density (no./km2) proportional browse offtake proportion of twigs browsed x ̄ 90% cl x ̄ 95% cla prop. 95% ci 19d 2001 0.30 0.25, 0.35 0.159 0.112, 0.195 0.110 0.039 19d 2003 0.30b 0.25, 0.35 0.171 0.144, 0.221 0.287 0.051 19d 2009 0.62 0.48, 0.72 0.405 0.332, 0.471 0.346 0.056 20a central hills 2000 2.36c 1.84, 2.87 0.433 0.394, 0.462 0.401 0.063 20a central hills 2012 1.59 1.32, 1.86 0.303 0.227, 0.357 0.130 0.050 20a western flats 2006 1.19 0.71, 1.68 0.307 0.112, 0.442 0.161 0.069 20a western flats 2012 0.82 0.50, 1.14 0.190 0.147, 0.228 0.119 0.036 20d 2007 2.11 1.63, 2.60 0.253 0.191, 0.323 0.167 0.035 20d 2010 1.24 1.00, 1.48 0.153 0.106, 0.199 0.117 0.030 abootstrapped confidence intervals may be asymmetrical. bdensity in fall 2003 assumed to be similar to that from survey in fall 2001 (keech 2012:13–14). csurvey in fall 2003; this was likely the period of maximum abundance prior to liberal antlerless harvest (young and boertje 2011, fig. 2) and the closest period with enough abundance sample units in the browse study area to permit a post hoc analysis. for comparison, the estimate with visibility correction for all of unit 20a (12,900 km2 of moose habitat <1350 m elevation) was 1.05 moose/km2 in 1999 (prior to 1st browse survey); 1.37 moose/km2 in 2003 (prior to liberal antlerless harvest); 0.98 moose/km2 in 2008 (after liberal antlerless harvest), and 0.98 moose/ km2 in 2011 (prior to 2nd browse survey) (young 2012, table 2). 10 browse indices to moose density – paragi et al. alces vol. 51, 2015 sampling periods within study areas and remained <70 cm by late winter in 3 of 4 study areas. unit 19d was the exception where mean snow depth for the winter post-treatment (107 cm) was 56 cm higher than pre-treatment (mann-whitney u = 18, p < 0.001). oftk was more consistent and precise than ptb in reflecting direction and magnitude of changes in moose density, or lack thereof. change in oftk from before to after management actions (table 2) was significant in all 4 areas and in the expected direction of change in moose density (z ≥ 2.7, p < 0.01, 1-tailed). change in ptb (table 2) was significant for unit 20a central hills (z = 6.0, p < 0.0005) and unit 20d (z = 2.1, 0.01 < p < 0.025), but not for unit 19d (z = 1.5, 0.05 < p < 0.1) or unit 20a western flats (z = 1.1, 0.2 < p < 0.5). whereas the pre-treatment moose density was similar between the 2001 and 2003 browse surveys in unit 19d (0.38 and 0.41 moose/km2 in the smaller 2003 browse study area; keech 2012:13), the lack of difference in oftk (z = 0.35, p > 0.5) was in marked contrast with the unexpected difference in ptb (z = 6.2, p < 0.001, 2-tailed; table 2). 0 100 200 300 400 k g/ ha unit 19d (a) (b) (c) (d) prod01 oftk01 prod03 oftk03 prod09 oftk09 prod00 oftk00 prod12 oftk12 prod06 oftk06 prod12 oftk12 prod07 oftk07 prod10 oftk10 0 100 200 300 400 500 600 700 k g/ ha unit 20a c hills 0 10 20 30 40 50 60 k g/ ha unit 20a w flats 0 10 20 30 k g/ ha unit 20d fig. 2. apparent production (prod) and offtake (oftk; both in kg/ha) as extrapolated from sampled twigs to plot composition in 4 study areas within interior alaska. moose abundance increased (unit 19d) or decreased (all others) coincident with intended outcome of management actions. years of prod and oftk are represented by last 2 digits of year starting with 2000. note differences in y-axis scale among areas; error bars are 95% confidence limits. an additional year of pre-treatment data (2001) was available for comparison in unit 19d. species codes: bene (betula neoalaskana; formerly b. papyrifera), cost (cornus stolonifera), poba (populus balsamifera), potr (p. tremuloides), saal (salix alaxensis), saar (s. arbusculoides), sabe (s. bebbiana), sagl (s. glauca), sain (s. interior), sapu (s. pulchra), and sari (s. richardsonii). alces vol. 51, 2015 paragi et al. – browse indices to moose density 11 further, ptb was highly variable for various degrees of oftk whether a moose population increased or decreased (fig. 3). direction of species-level change in dpb generally corresponded with the direction of change in both oftk and ptb. the strongest correspondence existed for those species composing the dominant (or codominant) biomass in a study area (table 3). among the 4 study areas, proportional magnitude of change in dpb (p < 0.05 by species) was 11–35% (x̄ = 22%, n = 6) in the predicted direction for each study area. the relative importance of dpb as a component of oftk is evident in scaling among browse metrics during the 100% increase in moose in unit 19d; 138% increase in oftk corresponded to increases of 20% in ptb (all species combined) and 16–42% in dpb of the primary browse species. where moose density decreased 31–41% in the other 3 study areas, decreases were documented in oftk (30% in unit 20a central hills to 40% in unit 20d), ptb (26% in unit 20a western flats to 68% in unit 20a central hills), and dpb of the primary browse species (11–37%, both extremes in unit 20d). plant architecture also responded to changes in moose density. there was a 44% increase in broomed plants following a moose population increase in unit 19d (fig. 4). in the absence of vegetative disturbance such as fire, the proportion of unbrowsed plants increased 6-fold (2% to 16%) in the unit 20a central hills after the moose population declined from the highest density in our 4 study areas; this was also the longest period between end of liberal antlerless harvest (presumed greatest point of density reduction; table 2) and the post-treatment browse survey (5 years later). there was an 83% increase in proportion of unbrowsed plants in unit 20a western flats following a reduction in moose density, where fires created new unbrowsed plants during and after antlerless hunts. however, we found no change in plant architecture in unit 20d (fig. 4) despite a 41% decline in moose density over 3 years. when compared with the 4 browse metrics, twinning rate showed no changes or trend in these moose populations during the intervening period of reduced moose density between browse surveys or soon thereafter (fig. 5). (a) (b) 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 devo mer ssa moibfo noitr oporp propor�on of twigs browsed unit 19d 2003 2009 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 devo mer ssa moibfo noitroporp propor�on of twigs browsed unit 20a c hills 2000 2012 fig. 3. examples of greater range in proportion of twigs browsed for a given range of offtake where moose populations had increased (a) and decreased (b) in interior alaska. the dashed lines illustrate a 1:1 relationship for comparison to the yearspecific plot data. 12 browse indices to moose density – paragi et al. alces vol. 51, 2015 discussion proportional oftk was a more comprehensive metric than ptb or dpb for detecting short-term, landscape-level changes in intraspecific competition for winter forage and potential effects on plants following management actions intended to affect moose density. proportional oftk consistently reflected change in moose density despite substantial variation in prod (total and among species) before and after management actions. we infer this relationship of oftk and density change as evidence that proportional oftk is unbiased, likely because moose distribution reflects browse distribution and oftk reflects available prod. our estimates of prod and oftk were complicated by 4 factors: 1) a change in landscape sampling design between preand post-treatment for unit 19d and unit 20a central hills, 2) by differences in size of some preand post-treatment study areas because of changing management issues, 3) by measuring a relatively limited number of plots over large diverse landscapes that met precision objectives for proportional oftk (seaton et al. 2011) but increased chance of sampling error, and 4) by measuring a limited number of forage plants without regard to nutrition or digestibility. we attribute the relatively high variation in the relationship between oftk and ptb (fig. 3) to condensing species with different twig diameters to a simple count of browsed table 3. change in mean diameter (mm) at point of browsing (dpb) by moose on the primary winter forage species in areas where moose populations increased (unit 19d, 2003–09) or decreased (all others) in accordance with intended outcome of management actions, interior alaska. dpb was predicted to change in the direction of trend in moose abundance. bold text represents dominant species (>50% of total estimated apparent production extrapolated from sampled plants to plot). trend in dpb (positive or negative) is inferred from the mann-whitney u statistic (p < 0.05). an additional year of pre-treatment data (2001) in unit 19d is shown for comparison. species codes: bene (betula neoalaskana), saal (salix alaxensis), sabe (s. bebbiana), and sapu (s. pulchra). area year bene saal sabe sapu unit 19d 2001 2.7 3.9 3.0 2.1 unit 19d 2003 2.8 4.3 3.4 3.2 unit 19d 2009 3.0 5.0 3.2 3.1 trend none pos none none p 0.43 <0.001 0.11 0.7 unit 20a central hills 2000 2.9 4.3 3.3 3.2 unit 20a central hills 2012 2.8 4.7 3.2 2.5 trend none none none neg p 0.67 0.28 0.2 <0.001 unit 20a western flats 2006 3.1 5.2 2.8 3.2 unit 20a western flats 2012 2.8 3.4 2.6 2.2 trend none neg none neg p 0.06 0.03 0.25 <0.001 unit 20d 2007 2.5 4.5 2.4 2.8 unit 20d 2010 2.5 3.8 2.6 2.5 trend none neg none neg p 0.58 <0.001 0.32 <0.001 alces vol. 51, 2015 paragi et al. – browse indices to moose density 13 twigs of all species combined. the effect of seemingly small change in dpb is relatively more important than changes in ptb in explaining change in oftk because the species-specific mass-diameter relationship is based on non-linear twig geometry that 296 277 235 177 109 310 431 295 0 0.2 0.4 0.6 0.8 1 19d 2003 19d 2009 20a c hills 2000 20a c hills 2012 20a w flats 2006 20a w flats 2012 20d 2007 20d 2010 unbrowsed browsed broomed χ2 = 1.8, p = 0.40 χ2 = 6.2, p = 0.045 χ2 = 26.1, p <0.001 χ2 = 29.9, p <0.001 fig. 4. the proportional changes in categories of browse plant architecture where the moose populations increased (unit 19d) or decreased (all others) coincident with the intended outcome of management actions in interior alaska. the binomial confidence interval (95%) and sample size are shown above bars. (b) 0.00 0.10 0.20 0.30 0.40 1998 2002 2006 2010 tw in ni ng ra te unit 20a c hills(a) 0.00 0.20 0.40 0.60 0.80 2001 2003 2005 2007 2009 2011 2013 tw in ni ng ra te unit 19d (d)(c) 0.00 0.10 0.20 0.30 0.40 2005 2008 2011 tw in ni ng ra te unit 20a w flats 0.00 0.10 0.20 0.30 0.40 2005 2008 2011 tw in ni ng ra te unit 20d fig. 5. moose twinning rates in late spring for (a) 1 moose population that increased and (b-d) 3 populations that decreased coincident with intended outcome of management actions in interior alaska. open circles indicate years of browse surveys; confidence intervals were 95% bootstrapped. 14 browse indices to moose density – paragi et al. alces vol. 51, 2015 does not scale equally with the linear proportion of twigs browsed. our data may be the first to report how changes in moose foraging behavior, as shown by changes in stem cropping diameter (dpb), is influenced by moose density across a broad geographic range. these observational and experimental data strongly suggest that variation in dpb reflects variation in competition which translates into variation in demographic processes (e.g., twinning rates). however, dpb of several species is not readily condensed to a single numeric score, and using dpb in isolation could complicate comparisons of browse removal before and after management actions if preferences for forage species change over time. oftk incorporates both ptb and dpb, thus reduces potential confounding of either component alone. over relatively short periods of change in moose density (2 years in unit 20d to 6 years in unit 19d), we expected change in architecture reflecting plant life histories to lag behind (or be of lesser magnitude) than change in the 3 metrics based on cag and dpb. however, magnitude of changes in architecture reflecting increases in moose density (more broomed plants) or decreases in moose density (more unbrowsed plants) were often of equal magnitude to change in cag and dpb metrics in as few as 5 years after change in density (unit 20a western flats). this rate of change is comparable to recovery of broomed willows after elk (cervus elaphus) reductions in wyoming (singer and zeigenfuss 2003:80–81). the lack of architectural changes in unit 20d despite a 41% decline in moose density may be explained by mismatch in scale of abundance surveys and distribution of antlerless harvest, where the latter primarily occurred in the flat and relatively accessible northern portion of the study area (s. dubois, unpubl. data). we had stratified unit 20d for browse surveys into flats and hills (sampling design before liberal antlerless harvest described in paragi et al. [2008]) and found no change in browse plant architecture in the flats. however, we noted a significant increase in brooming in the hills (t. paragi, unpublished data), indicating the continued effect of high browsing pressure where moose density had not been reduced. this experience highlighted the importance of scaling the browse sampling appropriately to the extent of abundance estimates and management actions. browse removal rate at the moose population level is not an absolute measure of carrying capacity for a given moose density, nor a meaningful demographic parameter linked in real time to changes in birth and death rates. browse removal rate is best used in concert with nutritional indices such as twinning rate that are also demographic parameters conveying population status relative to carrying capacity. singer and zeigenfuss (2003:67–70) studied consumption of willows as a product of percent leader use and percent twig use and found that moose density alone had little value in describing willow consumption where moose (x̄ = 1.9/km2) shared winter range with elk (x̄ = 16.3/km2) in wyoming. similarly, månsson (2009) found no positive relationship between betula spp. biomass removal and moose density where biomass removal of the dominant browse species, scots pine (pinus sylvestris), was related to moose density in sweden. however, we found that oftk of all browse species combined correlated with the direction and magnitude of moose density change. the highest biomass removal for all species combined approached 45% of cag where moose density was highest (unit 20a central hills). we have focused on how browse removal relates to animal condition, but high levels of offtake warrant consideration of sustainable plant health or productivity as another trigger for habitat enhancement or prudent reduction in herbivore density. alces vol. 51, 2015 paragi et al. – browse indices to moose density 15 low to moderate levels of browse removal can stimulate browse production (suter 1992), but removal beyond a threshold causes decline in production (danell et al. 1985, persson et al. 2005a). persson et al. (2007) found that production response of betula spp. to simulated browsing in northern sweden was highest at 25–40% biomass removal (representing ca. 3 moose/km2 on the winter range) on sites with moderate to high soil nutrients but lower on sites with low soil nutrients. singer and zeigenfuss (2003:70) observed a range (0–47%) in willow removal among study sites, with the highest growth response occurring at moderate (ca. 21%) consumption levels. they surmised that repeated consumption >30% is likely detrimental to plants and >45% removal is exceptionally high. we observed proportional oftk >45% for browse species (fig. 2) when moose in units 20a and 20d were at relatively high density (>2/km2). the poor nutritional condition indicated by low twinning rate of these populations (fig. 5; also boertje et al. 2007) suggests that such levels of browsing intensity have a negative effect on browse production. despite uncertainty in the threshold of sustainable browse removal, the recent antlerless harvests in units 20a and 20d intended to reduce relatively high moose densities and negative effects on the winter forage base seemed prudent regardless of whether proportional oftk is considered at the level of study area or plant species. we acknowledge that factors independent of change in moose density may influence estimates of browse metrics. we caution that our estimates of “apparent” production were for twigs above late winter snow depth in a limited number of plots, rather than a rigorous landscape estimate of total biomass production. the actual biomass available to moose might vary over time independent of moose abundance because of plant density in response to disturbance, plant structural change since disturbance, compensatory growth over a life history of exposure to browsing, growing season limits to cag (e.g., drought or insect defoliation), and competing herbivory by hares. diet quality can multiply effects on intake rate to produce greater removal on better quality forages (white 1983, mcart et al. 2009). study areas should remain large enough so that browse sampling allows inference at the population level and reduces annual variation in estimates of production and removal inherent at smaller scales (ahlén 1975:111, 134–135, 165; mackie 1976, månsson et al. 2007). managers wishing to monitor plant health should consider sampling designs focused on total production during the snow-free period and stratifying by vegetation type in addition to moose density. finally, we urge others to replicate our evaluation of management actions in a more rigorous and balanced experimental design, recognizing the risk that public regulatory bodies can interrupt the duration or type of management “treatment” once baseline data are collected. our only case study of deep snow (107 cm in unit 19d, 2009) was also our only case study of population increase. we do not know if unit 19d would have had removal levels equally as high in 2009 had snow depth been <90 cm. deep snow likely contributed to the increased 2009 browse removal in unit 19d by concentrating moose and exacerbating the effect of increased density (k. kellie, unpublished data). testa (2001) noted that the proportional number of s. alaxensis twigs browsed on 2 important wintering areas in the nelchina basin of southcentral alaska was 60–82% during winters with snow >70 cm compared with 12–35% during winters <70 cm. collins (2002:11) further defined a positive relationship between snow depth and biomass removal for this population. repeating surveys among winters of markedly different 16 browse indices to moose density – paragi et al. alces vol. 51, 2015 snow depth for a relatively stable moose population would be instructive as to the effect of snow-depth on spatial distribution and degree of browse removal. twinning rates did not respond in an expected fashion during the intervening period between the end of management actions causing moose density to change and subsequent browse surveys (2–6 yr); presumably, immediate consequences to reproduction were small. we surmise that increased body weight of the youngest, non-reproductive cohorts may be a better short-term index to improved nutrition following reduced intraspecific competition. lag in nutritional condition of wild ruminants has been documented following density reduction (blood 1974, albon et al. 1987, boertje et al. 2007). reduced twinning rate following prolonged high density might persist until enough more robust female calves born during periods of lower food competition enter the breeding population and affect the birth rate (e.g., solberg et al. 2004). in 3 of our 4 study areas, estimates of apparent production also decreased, possibly reflecting a decline in per capita forage that could dampen reproductive responses (solberg et al. 2012). our relevant case study in interior alaska occurred in unit 20a moose, where twinning rate took 12 years to increase despite a dramatic decline from peak abundance in the 1960s (fig. 5 in boertje et al. 2007). the lower twinning rate in the 1990s in unit 20a, despite lower moose density than in the 1960s, may have been evidence of degraded range capacity from having moose at prolonged high density in the 1960s and less extensive wildfires after the 1960s (less forage per capita). unfortunately, we do not have historic data on forage abundance or quality to evaluate this speculation. time lag in reproductive response may differ among periods, or among populations, in part due to differences in vegetation recovery rate (sand et al. 1996:242) that is potentially influenced by negative effects on soil fertility from prolonged high biomass removal (persson et al. 2005b). monitoring systems that quantify densitydependent responses of ungulates to their habitat as a correlate of population density have existed for decades (e.g., aldous 1945), but few jurisdictions manage ungulate abundance primarily based on monitoring indices of habitat (e.g., keigley and fager 2006) or nutrition. dubois (2008:388) first proposed use of twinning rate thresholds developed by boertje et al. (2007) rather than a population objective to recommend management of moose population trend in unit 20d. our study demonstrated the value of browse biomass offtake for corroborating intended reductions in intraspecific competition and gauging relationship of moose abundance to carrying capacity at higher densities. offtake and twinning rate provide managers with objective means to recommend and monitor effectiveness of forage enhancement, or timely reduction in moose density through harvest across age and sex classes to reduce forage competition and avoid prolonged negative effects of high density (boertje et al. 2007, young and boertje 2011). we urge managers and public regulatory bodies to utilize an empirical monitoring and decision framework for moose population management that incorporates measures of plant and animal condition in addition to population objectives. when reporting metrics on forage or animal condition, managers need to clearly identify that maintaining higher moose densities incurs an increased risk of strong negative feedback after severe winters (gasaway et al. 1983, boertje et al. 2009). when the public desires high moose densities, we encourage managers to discuss the risks associated with various management options and acceptable means for achieving proposed harvest alces vol. 51, 2015 paragi et al. – browse indices to moose density 17 objectives (adfg 2011, young and boertje 2011). acknowledgements our work was funded by federal aid in wildlife restoration and the alaska legislature. several adfg employees (particularly l. parrett), other agency personnel, and volunteers helped with field work. c. maurer, q. slade, a. shapiro, r. swisher, and m. terwilliger safely and efficiently piloted helicopters. b. taras estimated bootstrap confidence limits for twinning rates and provided formulas for incorporating visual detection variance on moose population estimates. we thank j. benson, s. brainerd, t. hanley, and d. james for constructive comments. references adfg (alaska department of fish and game). 2011. intensive management protocol. division of wildlife conservation. juneau, alaska, usa. ahlén, i. 1975. winter habitats of moose and deer in relation to land use in scandinavia. viltrevy 9: 45–192. albon, s. d., t. h. clutton-brock, and f. e. guinness. 1987. early development and population dynamics in red deer. ii. density-independent effects and cohort variation. journal of animal ecology 56: 69–81. aldous, c. f. 1945. a winter study of mule deer in nevada. journal of wildlife management 9: 145–151. doi: 10.2307/ 3795893. blood, d. a. 1974. variation in reproduction and productivity of an enclosed herd of moose (alces alces). international congress of game biologists 11: 59–66. boertje, r. d., m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314–327. doi: 10.2193/ 2007-591. ———, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. doi: 10.2193/2006-159. cederlund, g. n., h. k. g. sand, and a. pehrson. 1991. body mass dynamics of moose calves in relation to winter severity. journal of wildlife management 55: 675–681. doi: 10.2307/38 09517. coady, j. w. 1974. influence of snow on behavior of moose. naturaliste canadien 101: 417–436. cochran, w. g. 1977. sampling techniques. third edition. john wiley and sons, new york, ny, usa. collet, d. m. 2004. willows of interior alaska. u.s. fish and wildlife service, yukon flats national wildlilfe refuge, fairbanks, alaska, usa. collins, w. b. 2002. interrelationship of forage and moose in game management unit 13. federal aid in wildlfe restoration, research final performance report, grants w-24-4 to w-27-4, study 1.50. alaska department of fish and game, juneau, alaska, usa. conover, w. j. 1980. practical nonparametric statistics. second edition. john wiley and sons, new york, ny, usa. crete, m. 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67: 373– 380. doi: 10.1139/z89-055. danell, k., k. huss-danell, and r. bergström. 1985. interactions between browsing moose and two species of birch in sweden. ecology 66: 1867–1878. doi: 10.2307/2937382. dubois, s. d. 2008. unit 20d moose. pages 386–423 in p. harper, editor. moose management report of survey and inventory activities, 1 july 2005 through 30 june 2007. project 1.0. alaska 18 browse indices to moose density – paragi et al. alces vol. 51, 2015 department of fish and game, juneau, alaska, usa. ———. 2010. unit 20d moose. pages 380–409 in p. harper, editor. moose management report of survey and inventory activities, 1 july 2007 through 30 june 2009. project 1.0. alaska department of fish and game, juneau, alaska, usa. franzmann, a. w., and c. c. schwartz. 1985. moose twinning rates: a possible population condition assessment. journal of wildlife management 49: 394–396. doi: 10.2307/3801540. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. ———, and j. w. coady 1974. review of energy requirements and rumen fermentation in moose and other ruminants. naturaliste canadien 101: 227–262. ———, r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. goodman, l. a. 1960. on the exact variance of products. journal of the american statistical association 55: 708–713. doi: 10.1080/01621459.1960.10483369. keech, m. a. 2012. response of moose and their predators to wolf reduction and short-term bear removal in a portion of unit 19d. federal aid in wildlife restoration, final wildlife research report. project 1.62. adf&g/dwc/ wrr-2012-7, grants w-33-4 to w-3310. alaska department of fish and game, fairbanks, alaska, usa. ———, r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450–462. doi: 10.2307/3803243. ———, m. s. lindberg, r. d. boertje, p. valkenburg, b. d. taras, t. a. boudreau, and k. b. beckmen. 2011. effects of predator treatments, individual traits, and environment on moose survival in alaska. journal of wildlife management 75: 1361–1380. doi: 10.1002/jwmg.188. keigley, r. b., and c. w. fager. 2006. habitat-based adaptive management at mount haggin wildlife management area. alces 42: 49–54. kellie, k. a., and r. a. delong. 2006. geospatial survey operations manual. alaska department of fish and game, fairbanks, alaska, usa. kielland, k., and t. osborne. 1998. moose browsing on feltleaf willow: optimal foraging in relation to plant morphology and chemistry. alces 34: 149–155. lyon, j. l. 1970. lengthand weight-diameter relations of serviceberry twigs. journal of wildlife management 34: 456–460. doi: 10.2307/3799033. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. wildlife monographs 136. ———, and l. a. viereck 1990. browse regrowth and use by moose after fire in interior alaska. northwest science 64: 11–18. mackie, r. j. 1976. evaluation of range survey methods, concepts, and criteria (effectiveness of the key browse survey method). job progress report. federal aid project w-120-r-6, study 28.01, job bg-2.01, 1 july 1974-30 june 1975. montana department of fish and game, helena, montana, usa. månsson, j. 2009. environmental variation and moose alces alces density as determinants of spatio-temporal heterogeneity in browsing. ecography 32: 601–612. doi: 10.1139/z07-015. alces vol. 51, 2015 paragi et al. – browse indices to moose density 19 ———, h. andren, a. pehrson, and r. bergström. 2007. moose browsing and forage availability: a scale-dependent relationship? canadian journal of zoology 85: 372–380. mcart, s. h., d. e. spalinger, w. b. collins, e. r. schoen, t. stevenson, and m. bucho. 2009. summer dietary nitrogen availability as a potential bottom-up constraint on moose in south-central alaska. ecology 90: 1400–1411. doi: 19537559doi: 10.1890/08-1435.1. mccullough, d. r. 1984. lessons from the george reserve, michigan. pages 211– 242 in l. k. halls, editor. white-tailed deer ecology and management. wildlife management institute, washington, dc, usa. paragi, t. f., c. t. seaton, and k. a. kellie. 2008. identifying and evaluating techniques for wildlife habitat management in interior alaska: moose range assessment. federal aid in wildlife restoration, project 5.10. final research technical report, grants w-33-4 through w-33-7. alaska department of fish and game, juneau, alaska, usa. persson, i.-l., r. bergström, and k. danell. 2007. browse biomass production and regrowth capacity after biomass loss in deciduous and coniferous trees: response to moose browsing along a productivity gradient. oikos 116: 1639–1650. doi: 10.1111/j.0030-1299.2007.15946.x. ———, k. danell, and r. bergström. 2005a. different moose densities and accompanied changes in tree morphology and browse production. ecological applications 15: 1296–1305. doi: 10.1890/04-0499. ———, ———, j. pastor, j. danell, and r. bergström. 2005b. impact of moose population density on the production and composition of litter in boreal forests. oikos 108: 297–306. doi: 10.1111/ j.0030-1299.2005.13844.x. regelin, w. l., c. c schwartz, and a. w. franzmann. 1987. effects of forest succession on nutritional dynamics of moose forage. swedish wildlife research supplement 1: 247–262. sand, h., r. bergström, g. cederlund, m. östergren, and f. stålfelt. 1996. density-dependent variation in reproduction and body mass in female moose alces alces. wildlife biology 2: 233–245. schwartz, c. c., and l. a. renecker. 1997. nutrition and energetics. pages 441–478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. seaton, c. t. 2002. winter foraging ecology of moose in the tanana flats and alaska range foothills. m.s. thesis, university of alaska-fairbanks, fairbanks, alaska, usa. ———, t. f. paragi, r. d. boertje, k. kielland, s. dubois, and c. l. fleener. 2011. browse biomass removal and nutritional condition of moose, alces alces. wildlife biology 17: 55–66. doi: 10.2981/10-010. shipley, l. a., and d. e. spalinger. 1992. mechanics of browsing in dense food patches: effects of plant and animal morphology on intake rate. canadian journal of zoology 70: 1743–1752. doi: 10.1139/ z92-242. singer, f. j., and l. c. zeigenfuss. 2003. part ii. a survey of willow communities, willow stature and production, and correlations to ungulate consumption and density in the jackson valley and the national elk refuge. pages 58–86 in l. c. zeigenfuss and f. j. singer, editors. ecology of native ungulates in the jackson valley: habitat selection, interactions with domestic livestock, and effects of herbivory on grassland and willow communities. final report. natural resource preservation progress project #00-03, and interagency agreement nos. 1460-0013 and 1460-01-005 20 browse indices to moose density – paragi et al. alces vol. 51, 2015 between grand teton national park and the u.s. geolgical survey. united sates geological survey, fort collins, colorado, usa. solberg, e. j., a. loison, j.-m. gaillard, and m. heim. 2004. lasting effects of conditions at birth on moose body mass. ecography 27: 677–687. doi: 10.1111/ j.0906-7590.2004.03864.x. ———, o. strand, v. veiberg, r. andersen, m. heim, c. m. rolandsen, r. langvatn, f. holmstrøm, m. i. solem, r. eriksen, r. astrup, and m. ueno. 2012. moose, red deer and reindeer – results from the monitoring program for wild cervids, 1991–2011 (abstract in english). nina report 885. nina, trondheim, norway. stout, g. w. 2010. unit 24b moose. pages 572–610 in p. harper, editor. moose management report of survey and inventory activities, project 1.0, 1 july 2007–30 june 2009. alaska department of fish and game, juneau, alaska, usa. suter, s. m. 1992. the morphology and chemistry of two willow species in relation to moose winter browsing. m.s. thesis, university of alaska-fairbanks, fairbanks, alaska, usa. telfer, e. s. 1969. twig weight-diameter relationships for browse species. journal of wildlife management 33: 917–921. doi: 10.2307/3799325. testa, j. w. 2001. population dynamics of moose and predators in game management unit 13. federal aid in wildlife restoration, research final performance report, study 1.49, grants w-24-3 to w-27-3. alaska department of fish and game, juneau, alaska, usa. viereck, l. a., and e. l. little, jr. 2007. alaska trees and shrubs. second edition. university of alaska press, fairbanks, alaska, usa. vivås, h. j., and b.-e. sæther. 1987. interactions between a generalist herbivore, the moose (alces alces) and its food resources: an experimental study of winter foraging behavior in relation to browse availability. journal of animal ecology 56: 509–520. white, r. g. 1983. foraging patterns and their multiplier effects on productivity of northern ungulates. oikos 40: 377– 384. doi: 10.2307/3544310. wolff, j. o., and j. c. zasada. 1979. moose habitat and forest succession on the tanana river floodplain and yukontanana upland. alces 15: 213–244. young, d. d., jr. 2012. unit 20a moose. project 1.0. moose management report of survey and inventory activities, 1 july 2009–30 june 2011. alaska department of fish and game, juneau, alaska, usa. ———, and r. d. boertje. 2011. prudent and imprudent use of antlerless moose harvests in interior alaska. alces 47: 91–100. zar, j. h. 1984. biostatistical analysis. second edition. prentice-hall, upper saddle river, new jersey, usa. alces vol. 51, 2015 paragi et al. – browse indices to moose density 21 browse removal, plant condition, and twinning rates before and after short-erm changes in moose density study areas and moose abundance population increase population decrease methods browse removal browse architecture twinning rate results discussion acknowledgements references the changing role of hunting in sweden—from subsistence to ecosystem stewardship? sara lindqvist1,3, camilla sandström1,2, therese bjärstig2, and emma kvastegård1 1swedish university of agricultural sciences, department of wildlife, fish, and environmental sciences, skogsmarksgränd, 901 83 umeå, sweden; 2umeå university, department of political sciences, umeå universitet, 901 87 umeå, sweden. abstract: although hunting served traditionally to supply game meat, and that is still important in sweden, recreation is the most common reason for hunting moose (alces alces) today. hunting also serves an important management purpose in regulating moose populations to control crop and forest damage. this study used semi-structured interviews with key stakeholders and officials involved in the recently implemented ecosystem-based, adaptive local moose management system where hunters and landowners become environmental stewards responsible for managing moose in context with forest damage, vehicular collisions, large carnivores, and biodiversity. our study found that participation and collaboration in reaching management objectives was perceived as positive by stakeholders, although their stewardship is jeopardized if specific management responsibilities are not clarified regarding monitoring. further, it is important to find long-term funding solutions for monitoring activities that are critical for adequate data collection and to support the stakeholder role as steward. the importance of monitoring must be communicated to individual hunters and landowners to achieve an ecosystem-based moose management system that effectively incorporates both social and ecological values. alces vol. 50: 53–66 (2014) key words: adaptive, alces alces, biodiversity, knowledge-based, management areas, monitoring, moose, sweden. historically, hunting was primarily associated with human subsistence and livelihood, whereas today it is mostly associated with recreation (hendee 1974, barnard 2004), and has gradually developed into a tool to meet sustainable wildlife and ecosystem management objectives (holsman 2000, fischer et al. 2012). with this shift, hunters and other stakeholders should be directly or indirectly involved in environmental stewardship; i.e., the responsible use and protection of the natural environment through conservation and sustainable practices (leopold 1950, holsman 2000, chapin et al. 2009). moose (alces alces) hunting in sweden embodies this development where utility and leisure lately have become closely intertwined with management objective such as game population control (holsman 2000, council of europe 2007). a new moose management system was implemented in sweden in january 2012 that emphasizes the stewardship role of hunters and landowners. it provides a unique opportunity to analyze the extent that stakeholders support this institutional change and whether the new system offers the resources necessary for hunters and landowners to exercise ecological stewardship. the swedish moose management system has evolved several times in past decades to balance the interest of hunters (i.e., high moose populations) with other societal 3present adress: sjövägen 96, 834 34 brunflo, sweden 53 interests, most notably the commercial forestry sector concerned with browsing on economically valuable tree species. an increasing number of moose-vehicle collisions (mvc), negative effects on agriculture, and biodiversity are also of high concern (e.g., angelstam et al. 2000, edenius et al. 2002, lavsund et al. 2003, seiler 2003). because these incremental changes did not resolve these conflicts, the swedish parliament decided, in one step, to move from a top-down administrative system consisting of a patchwork of organizational management units, into an ecosystem-based, adaptive local management system incorporating moose biology components like home range size and seasonal migration in combination with stakeholder engagement (swedish government 2010; see also broman 2003, wennberg-digasper 2008). to avoid repeating previous management failures, the main objective of this reform was to establish a knowledge-based and adaptive moose management system with the capacity to balance different interests (sandström et al. 2013). in particular, monitoring that was performed irregularly and incoherently across sweden, is now considered a key focus in the new system to integrate knowledge for establishing and evaluating management objectives. accordingly, the extent to which the new management system will succeed is dependent on 3 central elements of ecological stewardship: the will among hunters and landowners to 1) support wildlife management program goals designed to balance social and ecological values, 2) support and participate in the development of institutions for defining and implementing stewardship goals, and 3) participate in monitoring activities related to meeting social or ecological objectives (holsman 2000, chapin et al. 2009). our objective was to study the implementation phase to assess stakeholder acceptance and capacity to handle the objectives of ecosystem-based institutions and monitoring as part of knowledge-based management. background compared to previous management systems, the current system has a pronounced national objective for long-term balance of the moose population with forest resources and societal interests (swedish official investigation 2009). to reach this objective, an official investigation identified the need to overcome institutional deficiencies, primarily the lack of collaboration between key stakeholders and an ecosystembased approach, but also the lack of systematic monitoring of moose and forest resources (swedish official investigation 2009). the institutional change included some redistribution of tasks between the management levels, but also adding a new management body at the ecosystem level (i.e., moose management areas) that covers the equivalent of a moose population (at least 50,000 ha in the south and 100,000 ha in the north). this approach should bridge the regional level with moose management units and license areas at the local level (table 1). the moose management areas are governed by a moose management group consisting of 3 landowners and 3 hunters. this group is responsible for 1) making an ecosystembased and adaptive moose management plan for their respective area (stretching over maximum 3 years), 2) advising hunters and landowners in creating local management plans within the moose management units, and 3) coordinating and evaluating monitoring activities (swedish official investigation 2009). the institutional amendment of the moose management system resulted in redistribution of funding as well, from primary administration use by the county administrative boards to include all monitoring of moose and forest resources (swedish government 2010). 54 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 the lack of systematic monitoring is addressed by science-based recommendation to focus on 4 accepted monitoring methods to provide annual information about the moose population including harvest statistics, hunter observation rates, pellet-group counts, and calf weights (bergström et al. 2011, danell et al. 2011, ericsson and kindberg 2011, kindberg et al. 2011). jointly, these methods require long-term, standardized implementation to function as reliable indices either singly or in combination (bergström et al. 2011, månsson et al. 2011). to further meet the broader goals of the ecosystem-based system, information about large carnivore populations and mvcs will be evaluated, and assessment of forage and browsing damage are also necessary to address the primary management objective (swedish government 2010). moose damage survey methods and forage forecasts are suggested as the basis of monitoring and are preferably conducted every 3rd and 5th year, respectively (kalén and bergquist 2011, rolander et al. 2011). the damage survey methods estimate the proportion of old and fresh stem damage within a given height interval and area (rolander et al. 2011), and the forage forecasts estimate the availability and quality of food resources through combination of satellite mapping of clearcuts and field sampling (kalén and bergquist 2011). study area we conducted our study in 5 counties (västerbotten, dalarna, södermanland, västra götaland, and kronoberg) distributed across sweden: 62°00' n, 15°00' e (fig. 1). these counties were selected because of their differences in ecology, landownership, and use patterns that present varied challenges to fully implement the new ecosystem-based, adaptive local management system. the counties cover all swedish vegetation types including alpine, boreal, boreonemoral, and nemoral zones. forests are dominated by commercially valuable scots pine (pinus sylvestris) and norway spruce (picea abies), and by deciduous tree table 1. the institutional framework of the old and new swedish moose management systems. the swedish environmental protection agency has the ultimate responsibility at the national level with the swedish forest agency functioning primarily as an advisory and supporting authority. at the regional level the county administrative boards have authoritative responsibility for moose management. at the regional level in the new system, wildlife management delegations with members from all interest groups are included. at the ecosystem level in the new system are moose management areas consisting of moose management groups with stakeholder representatives (3 hunters and 3 landowners). the different categories of license areas in the old system are removed, and license areas or moose management units exist only at the local level in the new system. moose management units also include stakeholders at the local level. old system new system national level swedish environmental protection agency /swedish forest agency regional level county administrative boards county administrative boards including wildlife management delegations ecosystem level (50,000–100,000 ha) moose management areas led by a moose management group with 3 landowners and 3 hunters local level (10,000–15,000 ha) license areas (a, b, c, and e) wildlife management units moose management units (hunters and landowners) moose management units (hunters and landowners) license areas (in exceptional cases) alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 55 species such as birch (betula spp.), aspen (populus tremula), and rowan (sorbus aucuparia), as well as broadleaved trees like oak (quercus spp.) in the south. the 5 counties also differ in several other ways that could affect local and regional moose management, including moose population density (hörnberg 2001) and health; for example, moose tend to be larger in the north (sand et al. 1995). predation by brown bears (ursus arctos) is mostly in västerbotten and dalarna, whereas wolves (canis lupus) occur in dalarna and västra götaland. the effect of predation on a local moose population varies among and within these counties depending on the composition and number of predators (sand et al. 2011). while roe (capreolus capreolus) and red deer (cervus elaphus) exist in all 5 counties, some of sweden’s densest fallow deer (dama dama) populations exist in västra götaland and södermanland; södermanland also has a small population of mouflon sheep (ovis aries orientalis). only these 2 counties were considered with potential for interspecific competition between moose and other ungulates. land ownership among the counties also differ with more commercial forest companies in the north, and more private land in the south; forest ownership in sweden is ∼51% private and 42% commercial (bergman and åkerberg 2006). methods representative officials and stakeholders within moose management areas were interviewed in each county. in total, 29 semi-structured, qualitative interviews were conducted in october–december 2012. phone interviews, except 2 face-to-face, were used to ensure a high response rate. an interview manual with a vast spectrum of qualitative questions regarding the moose management system guided the interviews, and all respondents were asked identical questions. the recorded interviews lasted 45–120 min and were transcribed in full. respondents were given the opportunity to read their transcribed interview, and to clarify and/or alter any content to ensure the information was as valid as possible. the first interview in each county was with the wildlife manager of the county administrative board, which is the regional authority responsible for wildlife management issues, including moose management. given their local knowledge, each was asked to suggest a typical management area that was neither more collaborative, nor more conflicted and turbulent, than any other management area in the county. the stakeholders (3 landowner and 3 hunters) in this management area were contacted for further fig. 1. the 5 counties that served as the study area; from the north, västerbotten, dalarna, södermanland, västra götaland and kronoberg counties. the figure in each county represents the number of moose management areas in that county. 56 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 interviews. in addition, the swedish forest agency (sfa) forest manager who was responsible for wildlife issues in each county was interviewed, since the sfa is an important advisory authority for the stakeholders. there were 40 potential interviews in the study: 5 wildlife managers, 5 forest managers, and 6 stakeholders per county (n = 30) of which 19 were interviewed (9 landowners and 10 hunters) with 1–2 nonrespondents (unreachable or unwilling) in each county. interviews were conducted with at least one hunter and one landowner in each management group; therefore, the data were regarded as reasonably balanced and useful to analyze the new adaptive moose management system. the interviews were thoroughly read and all information regarding or related to monitoring was extracted from the material. specific quotes that strengthen, clarify, or illustrate general (or divergent) responses are provided in the results. to ensure integrity, respondents are anonymous and we only describe their stakeholder group or agency and county. the interviews were conducted in swedish and we present interview quotes based upon our translation to english. results willingness and capacity to support nationwide objectives we asked respondents about their attitudes towards the overarching objective of the new moose management system, the need for collaboration, and the ecosystembased approach. the management of moose in larger geographical areas, as opposed to smaller management units, and the comprehensive ecosystem approach were considered positive in all counties. all but one hunter and landowner felt they would be more able to actively influence the management process. the single hunter from västerbotten claimed that “local decisions have been moved even further away”. hunters and landowners also emphasized the increased collaboration between them as important in fostering management legitimacy and trust among all participants. there was a strong conviction among officials and stakeholders that the new system would enable fact-based rather than assumption-based management, and that information derived from moose monitoring, forage forecasts, and browsing pressure estimates would allow more local, detailed, and nuanced management decisions. two example quotes were: “an advantage is that we will be able to get a clear view of the moose population, and that we, with determined effort, will achieve a high-quality moose population” (hunter from dalarna), and “we will have a moose population that is adjusted to the forage production—that’s what’s important, that we do not have too many moose, but we use the resources we have in our forests” (landowner in kronoberg). the stakeholders were developing moose management area plans in all counties. the plan was considered an important tool towards realizing management objectives and was regarded as a living document which could be altered if conditions changed. the perceptions of system resilience and how quickly management might respond to change differed among respondents; some believed that a yearly revision was reasonable, while others thought it possible to make immediate amendments during a hunting season. rapid changes in the moose population would primarily be based on hunter observations, and to some degree on harvest statistics, but this assumes that reporting is relatively fast and that management groups meet frequently to assess the situation. how such reports would be used differed among respondents; some claimed that if hunter observations differed from specified sexratios (e.g., equal male:female ratio), then restrictions might be implemented. others alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 57 observed that reporting is generally quite slow, and that the hunting season would likely be over before information could reach the moose management groups and subsequently to hunting teams. willingness and capacity to implement stewardship goals we asked respondents to describe the ongoing implementation process to understand the degree of support and willingness to participate in the development of institutions for defining and implementing stewardship goals. despite pervasive belief in the new moose management system and a willingness among stakeholders and officials to participate in the implementation process, several obstacles were identified in fully implementing the system. the first obstacle was the short time period between the political decision in april 2011 and when the moose management groups would initiate their work (january 2012); both officials and stakeholders claimed they had inadequate time. specifically, more time was needed to acquire monitoring information, to fully understand the function and implementation of the new moose management system, and to adjust work arrangements before the first 3-year plan was due. a landowner in kronoberg summarized this with: “the process has been too fast, we do not have enough knowledge, we have too little knowledge regarding our game populations, we have no monitoring methods we all agree on applying, we have poor knowledge of browsing damage, we have no good overview of forage forecasts either, we rule and believe that we can manage the moose population, and we almost become overconfident and imagine we can calculate, down to nearest decimal, how many moose we can harvest.” the second obstacle was the lack of available and sufficient knowledge from monitoring to define management plans and objectives. the need for better data and the desire to develop a better overview of resource status was apparent in all counties. a new database (algdata.se) providing easy access to monitoring information and statistics important for developing their management plans was an identified need. certain regional county data requirements differed. västra götaland respondents stressed the lack of resolution in forage forecasts, moose damage surveys, predator densities, and mvcs. they also preferred that numbers and data be available for each management unit or even each hunting team, rather than an average number at the county or management area level. in södermanland, the importance of including other ungulate species was highlighted with regard to species-specific browsing damage surveys. the third obstacle was the vagueness of management responsibilities regarding data collection, statistics, and monitoring. despite the high degree of awareness of the need for monitoring, the responsibilities for such activities were especially perceived as confusing among the respondents. the main emphasis of responsibility for implementation, interpretation, and evaluation of monitoring lies with the moose management groups in their respective areas; the moose management units are obliged by regulations to participate in monitoring (environmental protection agency 2011). yet, there was confusion concerning role and authority among both officials and stakeholders. a forest manager in västerbotten stated: “we should have a locally based management. and what does it mean? does it mean that the management unit or maybe even the level below [hunting team] holds the steering capacity? or does it mean that if the moose management group does not approve the unit’s management plan then you have to do it all over again?”. the county administrative board, the agency with actual decision-making power to demand participation 58 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 http://www.algdata.se in monitoring (swedish official investigation 2009), referred to the management groups and claimed the task was theirs to solve. they in turn claimed a lack of instruments and authority of decision-making and cannot demand monitoring participation from the management units. it is unclear if the management unit participation refers to all specified monitoring activities or if they can choose among suggested actions. stakeholders also claimed that hunters had the actual capability for control. license areas that cover parts of many moose management areas in all counties were another issue because they are not obligated by law to follow any management plan or participate in monitoring activities; consequently, they risk counteracting or interfering with plans in management units or management areas. respondents were apprehensive that the vagueness about management responsibility might undermine the purpose and role of management areas and the crucial monitoring required for ecosystem-based management. there were also uncertainties in all counties regarding high costs associated with decisions, prioritization, and the interval of monitoring activities to provide management with sufficient information. the lack of financial resources to support the moose management system was the fourth obstacle. originally, the moose management system was intended to be economically self-sufficient through harvest fees, including the cost of monitoring programs (environmental protection agency 2011). both officials and stakeholders found it unrealistic that, in accordance with the decision made by the swedish parliament, the entire moose management system (administration, management group members, and monitoring) should be funded by harvest fees alone; cost of monitoring was of greatest concern. most funding during the implementation phase was used for administrative work with little left for monitoring. most monitoring methods listed here are relatively inexpensive and require minimal effort of hunters, with pellet-group counts the exception as field work should occur twice annually (bergström et al. 2011). to save money, an official and a stakeholder suggested that monitoring be conducted in 4–5 year intervals rather than annually. however, this approach would reduce the ability to detect trends in data and recent information would not be unavailable during a 3-year plan undermining the adaptive advantage of the management plan. neither wildlife or forest managers could financially support monitoring activities and higher harvest fees were not considered a viable solution; in södermanland and västra götaland, this might have a contrary effect where hunters would refocus efforts on other game. one perception was that monitoring expenditures should be shared among stakeholders, versus exclusively funded by hunters, by having landowners responsible for forest resources and hunters responsible for moose. stakeholders also indicated that the sfa should be responsible for funding and providing forage forecasts and moose damage surveys. in all counties but västerbotten, another problem was how to fairly subdivide monitoring costs among many small landowners in the fragmented ownership common in sweden. one solution was that all monitoring be conducted by volunteers; however, certain deficiencies were identified relative to voluntary monitoring including lack of trust among some stakeholders regarding the credibility of monitoring data, and the time that stakeholders were willing to spend on voluntary activities. circumventing such concerns requires hiring professionals for monitoring which would increase costs substantially. the money available from harvest fees in each specific county differs because of variable harvest fees, but more alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 59 importantly, on the annual moose harvest in each county; for example, harvest is about ∼1,000 moose in södermanland versus 6,000 in dalarna. raising harvest fees to increase management finances was not considered as an option, especially in södermanland and västra götaland where increasing costs for hunters might have an unintended effect. rather than investing money and efforts in moose hunting, officials and hunters suggested that other game species might instead become increasingly important to the hunters. willingness and capacity to participate in monitoring activities we asked respondents to describe monitoring methods used previously, new methods to be implemented in the adaptive system (fig. 2), and their general knowl‐ edge about different monitoring methods. information about local moose populations during implementation of the new system was mostly obtained from harvest statistics and hunter observations, but data from all base-monitoring methods were used to some extent when moose management areas defined their initial plans. throughout sweden in 2012, ∼50% of management areas used hunter observation rates of moose during the first week of moose hunting. rates within our management areas were higher than the national average (unpublished data, swedish hunting association), although none of the 5 counties considered this sufficient information to manage their moose population. pellet-group counts were conducted in all counties; however, they did not provide complete data for any county or moose management area due to their fragmented application. only södermanland collected data on calf weight but it was regarded as a monitoring method for future use. hunters still recorded calf weights and future use of these data might be facilitated by a fully developed database (algdata.se). this fig. 2. the monitoring methods planned to be used in each county (cab) and selected moose management areas (mma). additional monitoring will be conducted in mmas than at the county level in västerbotten, dalarna and södermanland. 60 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 database will gather data from all monitoring effort and other areas of interest, and will presumably serve as an important management tool in facilitating stewardship. forage forecasts were provided by the sfa in all counties but a comprehensive forage forecast should include both satellite mapping and field visits; only satellite images were available in kronoberg and södermanland. the quality of forecasts in other counties could not be ascertained from responses. a countywide moose damage survey was conducted only in västerbotten, the only county with a long-term tradition of monitoring browsing damage. browsing was monitored on a smaller scale in dalarna. the need and desire to collect more detailed and comprehensive information was evident in the planned monitoring activities identified by officials and stakeholders in all counties. the necessity for more information and data about forest resources was specifically highlighted by a landowner in västerbotten: “in order to achieve a functioning adaptive management based on more facts and local knowledge that consider both the quality of the moose population, but also take moose damages into account”. there was confusion and opinions were divided about the applicability of monitoring methods and related validity of derived estimates. respondents in västra götaland desired estimates at the management unit level or lower, believing estimates at the management area level were too coarse. officials and stakeholders indicated their intention to address mvcs in management plans. they generally believed that mvcs were not only of societal importance, but reflected the relative size and trend of the moose population. stakeholders in västra götaland claimed that moose impacts on crops must be assessed, since moose damage, especially to oat (avena sativa) fields, is a recurring issue within their area. all respondents identified the need to include prominent county-specific conditions such as other ungulate species, moose migration patterns and predation, and habitat changes in the moose management system. in the 3 counties with established populations of large carnivores (i.e., brown bears in västerbotten and dalarna, wolves in dalarna and västra götaland), information about predator populations and predation rates was considered an important com‐ ponent of moose management plans. res‐ pondents in södermanland stressed the importance of co-managing other deer and ungulate species, rather than focus on moose singly, arguing that moose and deer interact and share several resources (e.g., habitat and food). they also questioned why deer and wild boar (sus scrofa) were not already considered in the ecosystem-based management system. in västerbotten, seasonal migration patterns of moose were reflected in the management plan. in kronoberg, habitat changes resulting from the 2005 hurricane gudrun received special attention, as it felled 75 million cubic meters of forest in southern sweden (swedish hydrological and meteorological institute 2013), creating beneficial habitat for moose and roe deer. discussion the respondents generally perceived the increased participation and collaboration between hunters and landowners as positive. despite some concern that the new mana‐ gement approach might remove decisionmaking from the local to ecosystem level and lose legitimacy with individual hun‐ ters and landowners, the majority of officials and stakeholders supported the program goals and believed the opposite. most envisioned a decentralized influence from the county to ecosystem level or down to the local level, and believed that enhanced alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 61 stakeholder stewardship would consolidate management legitimacy. despite the positive spirit and increased collaboration, most respondents found the specific management responsibilities to be unclear. the term “local” seemed to create confusion about division of power, at what scale, and where power resided to influence decision-making. lack of funding and the actual time provided for implementation of the new system were other obstacles identified by respondents. while time constraints would gradually be remedied, the funding issue risks undermining the entire management system since it reduces the possibilities for stakeholders to contribute to ecosystem stewardship. stakeholders were aware of the need to maximize information to successfully balance moose and forest resources. all acknowledged the need for reliable monitoring and appropriate sampling as part of adaptive management, and that the purpose of monitoring might be lost without using appropriate techniques. however, stakeholder expectations and scientific recommendations regarding data acquisition and monitoring differed. the suggested monitoring methods were most useful for moose management areas, and less so for smaller local units where the goal is to use accurate estimates (ericsson 2011). pellet-group counts are an exception, but using this method in a smaller management unit would not be as useful when managing a large population, a primary goal for achieving ecosystem-based management (swedish official investigation 2009). understanding the limitations of monitoring techniques is crucial to achieve transition from an assumption-based to a knowledge-based management system. the forest monitoring methods are applicable at both the management area and local scales (kalén and bergquist 2011, rolander et al. 2011). however, local damage surveys are often only roughly correlated with population density (rolander et al. 2011), and applying such data with moose population estimates at larger scale risks error in interpretation within the larger management area. further, data collected at different resolutions probably poses risk to the overall assessment that includes various monitoring techniques, potentially complicating stakeholder stewardship further. the ability to modify the management plan was perceived as a core and essential feature of the moose management system; however, perceptions varied as to whether it was possible to alter management decisions during an ongoing hunting season. indeed, invoking change given new information is essential in adaptive management (allen et al. 2011). the question remains as to the extent of adaptability in the new moose management system. the regulations (environmental protection agency 2011) provide for changes during a hunting season, yet this would require continual monitoring and rapid response by stakeholders. unless reporting efforts improve, it will be difficult to adjust during a hunting season. even if these components function flawlessly, most moose in sweden are harvested in the first weeks of the hunting season (ball et al. 1999); therefore, from a practical standpoint, non-emergency changes and deviations would likely occur the following year. accounting for mvcs and predation was relatively well developed in all counties. predation, for example, could be estimated precisely from wolf monitoring (sand et al. 2011). conversely, with respect to other ungulates—such as in södermanland where red deer, roe deer, fallow deer, and wild boar (sus scrofa) are abundant—current monitoring of populations, damage, and forage forecasts were considered highly insufficient; neither wild boar nor deer are systematically monitored in sweden (apollonio et al. 2010). deer populations could 62 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 be assessed with pellet-group counts, but density estimates are dependent upon species-specific identification of pellet groups and defecation rates (neff 1968). it might also be possible to expand hunter observations to include other deer with the methods of kindberg et al. (2009), where effort-corrected observations of brown bears was suggested as a useful monitoring technique. adequate evaluation would be required before implementing this approach into moose management plans. despite progress with genetic monitoring of browsing damage (spong 2011), information about different ungulate forage selection and food overlap is generally lacking. overall, our results suggest that critical knowledge-gaps exist with both hunters and landowners that preclude their effective participation in, and use of many monitoring techniques. conclusions we found strong stakeholder support for the moose management goals to balance social and ecological values. although the willingness to embrace ecosystem stewardship was pervasive among stakeholders, the moose management program faces several challenges in fully implementing an ecosystem-based, adaptive moose management system. this includes the need to clarify primary concepts like “ecosystem-based” and “local”, and to have all stakeholders clarify and agree about definitions within the plan and the role and responsibility at each level; such an approach should help mitigate conflicts and avoid further confusion. several obstacles were identified concerning monitoring, the key tool enabling environmental stewardship. specifically, unclear responsibilities and inadequate funding threaten local and regional data collection, both of which could jeopardize stakeholder stewardship. the importance and benefits of monitoring and reporting, as emphasized by the respondents, must be communicated to the grass-root level to enhance future participation; i.e., the responsibility of individual stewardship. demonstrating that monitoring is a worthwhile effort of individual hunters or landowners is perhaps the most difficult challenge. it is important to create a sustainable moose management system in balance with forage resources, traffic safety, and large carnivore populations, as well as other land uses and biodiversity objectives (moller et al. 2004). unless these challenges are resolved, the primary objective of the new moose management system risks the same failure as with previous plans. to address the monitoring issues, an addition or revision of the regulations (environmental protection agency 2011) regarding the purpose of monitoring is suggested; this approach would be positive for all stakeholders and should facilitate their commitment to the system. the issue of monitoring costs, specifically regarding forest resources, needs to be addressed directly with full support of stakeholders to achieve adequate participation. lastly, information regarding coexistence and co-management of ungulates in management areas needs to be incorporated into the system, an approach similar to that for large carnivores in sweden (andrén et al. 2011). acknowledgements this research was funded through future forests, a multi-disciplinary research program supported by mistra (the foundation for strategic environmental research), the swedish forestry industry, the swedish university of agricultural sciences, umeå university, and the forestry research institute of sweden. we thank the two anonymous reviewers for valuable comments on an earlier draft of the manuscript. alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 63 references allen, c. r., j. j. fontaine, k. l. pope, and a. s. garmestani. 2011. adaptive management for a turbulent future. journal of environmental management 92: 1339–1345. andren, h., a. jarnemo, h. sand, j. mansson, l. edenius, and p. kjellander. 2011. ekosystemaspekter på älgförvaltning med stora rovdjur. ecosystem aspects on moose management with large carnivores. fakta skog 12/2011. swedish university of agriculture, umeå, sweden. (in swedish). angelstam, p., p. e. wikberg, p. danilova, w. e. faber, k. nygren, w. b. ballard, and a. r. rodgers. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland, and russian karelia. alces 36: 133–145. apollonio, m., r. andersen, and r. putman. 2010. european ungulates and their management in the 21st century. cambridge university press, new york, new york, usa. ball, j. p., g. ericsson, and k. wallin. 1999. climate changes, moose and their human predators. ecological bulletins 47: 178–187. barnard, a. 2004. hunter-gatherers in history, archaeology and anthropology. berg publishers, oxford, united kingdom. bergman, m., and s. åkerberg. 2006. moose hunting, forestry, and wolves in sweden. alces 42: 13–23. bergstrom, r., j. mansson, j. kindberg, å. pehrson, g. ericssonr, and k. danell. 2011. spillningsinventering för älg. pellet-group count for moose. fakta skog 12/2011. swedish university of agriculture, umeå, sweden. (in swedish). broman, e. 2003. environment and moose population dynamics. ph.d. thesis. department of environmental sciences and conservation, göteborg university, göteborg, sweden. chapin, s., g. kofinas, and c. folke. 2009. principles of ecosystem stewardship: resilience-based natural resource management in a changing world. springer science and business media, new york, new york, usa. council of europe. 2007. european charter on hunting and biodiversity. adopted by the standing committee of the bern convention at its 27th meeting in strasbourg, 26–29 november, 2007. (accessed april 2013). danell, k., j. p. ball, r. bergstrom, g. ericsson, j. kindberg, and h. sand. 2011. älgkalvvikter – ett konditionsmått. moose calves’ weights – an index of fitness. fakta skog 13/2011. swedish university of agriculture, umeå, sweden. (in swedish). edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forests. silva fennica 36: 57–67. environmental protection agency. 2011. nfs 2011:7. naturvårdsverkets föreskrifter och allmänna råd om jakt efter älg och kronhjort. environmental protection agency’s regulations and general guidelines on hunting for moose and red deer. (in swedish). ericsson, g. 2011. inventering för adaptiv älgförvaltning i älgförvaltningsområden (äfo). monitoring for an adaptive moose management within moose management areas (mma). inventeringsmanualens förord. manual preface. swedish university of agriculture, umeå, sweden. (in swedish). ———, and j. kindberg. 2011. älgobservationer (älg-obs). hunter observations. fakta skog 11/2011. swedish university of agriculture, umeå, sweden. (in swedish). fischer, a., c. sandstrom, m. delibesmateos, b. arroto, d. tadie, d. randall, f. hailu, a. lowassa, m. msuha, 64 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 http://www.cicwildlife.org/uploads/media/hunting_charter_en.pdf http://www.cicwildlife.org/uploads/media/hunting_charter_en.pdf http://www.cicwildlife.org/uploads/media/hunting_charter_en.pdf v. kerezi, s. reljic, j. linnel, and a. majic. 2012. on the multifunctionality of hunting – an institutional analysis of eight cases from europe and africa. journal of environmental planning and management 56: 531–552. hendee, j. c. 1974. a multiple-satisfaction approach to game management. wildlife society bulletin 2: 104–113. holsman, r. h. 2000. goodwill hunting. exploring the role of hunters as ecosystem stewards. wildlife society bulletin 28: 808–816. hornberg, s. 2001. the relationship between moose (alces alces) browsing utilization and the occurrence of different forage species in sweden. forest ecology and management 149: 91–102. kalen, c., and j. bergquist. 2011. fodpro, foderprognoser. skogliga inventeringsmetoder i en kunskapsbaserad älgförvaltning. fodpro, forage forecasts, moose damage survey methods in a knowledge-based moose management. version 1.0. skogsstyrelsen (swedish forest agency). (in swedish). kindberg, j., g. ericsson, r. bergstrom, and k. danell. 2011. avskjutningsstatistik för älg. harvest statistics for moose. fakta skog 10/2011. swedish university of agriculture, umeå, sweden. (in swedish). ———, ———, and j. e. swenson. 2009. monitoring rare or elusive large mammals using effort-corrected voluntary observers. biological conservation 142: 159–165. lavsund, s., t. nygren, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109–130. leopold, a. 1950. a sand county almanac and sketches here and there, illustrated by charles w. swartz. oxford university press, new york, new york, usa. mansson, j., c. e. hauser, h. andren, and h. p. possingham. 2011. survey method choice for wildlife management: the case of moose alces alces in sweden. wildlife biology 17: 176–190. moller, h., f. berkes, p. o. b. lyver, and m. kislalioglu. 2004. combining science and traditional ecological knowledge: monitoring populations for comanagement. ecology and society 9(3): 2. http://www.ecologyandsociety.org/vol 9/iss3/art2/ (accessed april 2103). neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. the journal of wildlife management 32: 597–614. rolander, m., c. kalen, and j. bergquist. 2011. äbin, skogliga inventeringsmetoder i en kunskapsbaserad älgförvaltning (moose damage survey methods in a knowledge-based moose management). älgbetesinventering (äbin) version 1.0. skogsstyrelsen (swedish forest agency). (in swedish). sand, h., h. andren, j.e. swenson, and j. kindberg. 2011. flera jägare på älgpopulationen – predationsmönster hos varg och björn. several hunters of the moose population – predation patterns of wolf and brown bear. fakta skog 25/2011. swedish university of agriculture, umeå, sweden. (in swedish). ———, g. cederlund, and k. danell. 1995. geographical and latitudinal variation in growth patterns and adult body size of swedish moose (alces alces). oecologica 102: 433–442. sandstrom, c., s. wennberg-digasper, and k. öhman. 2013. conflict resolution through ecosystem-based management: the case of swedish moose management. international journal of the commons 7: 549–570. seiler, a. 2003. the toll of the automobile. ph. d. thesis. swedish university of agricultural sciences, umeå, sweden. spong, g. 2011. dna-analyser och viltövervakning. dna analysis and wildlife surveillance. fakta skog 17/2011. swedish university of agriculture, umeå, sweden. (in swedish). alces vol. 50, 2014 lindqvist et al. – hunting role in sweden 65 http://www.ecologyandsociety.org/vol9/iss3/art2/ http://www.ecologyandsociety.org/vol9/iss3/art2/ swedish government. 2010. proposition 2009/10:239. älgförvaltningen. 2010. proposition 2009/10:239 moose management. (in swedish). swedish hydrological and meteorological institute. 2013. (accessed march 2013). swedish official investigation. 2009. 2009:54. uthållig älgförvaltning i samverkan. 17th june 2009. sustainable moose management in collaboration. fritzes. stockholm, sweden. (in swedish). wennberg-digasper, s. 2008. natural resource management in an institutional disorder: the development of adaptive comanagement systems of moose in sweden. ph. d. thesis. division of political science, department of business administration and social sciences, luleå university of technology, sweden. 66 hunting role in sweden – lindqvist et al. alces vol. 50, 2014 http://www.smhi.se/kunskapsbanken/meteorologi/gudrun-januaristormen-2005-1.5300 http://www.smhi.se/kunskapsbanken/meteorologi/gudrun-januaristormen-2005-1.5300 http://www.smhi.se/kunskapsbanken/meteorologi/gudrun-januaristormen-2005-1.5300 the changing role of hunting in sweden—from subsistence to ecosystem stewardship? background study area methods results willingness and capacity to support nationwide objectives willingness and capacity to implement stewardship goals willingness and capacity to participate in monitoring activities discussion conclusions acknowledgements references alces28_31.pdf alces(23)_1.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces24_118.pdf alces24_150.pdf alces vol. 48, 2012 oehlers et al. visibility of moose 89 visibility of moose in a temperate rainforest susan a. oehlers1,2, r. terry bowyer3, falk huettmann2, david k. person4, and winifred b. kessler5 1yakutat ranger district, tongass national forest, 712 ocean cape drive, yakutat, alaska 99689 usa; 2ewhale lab, institute of arctic biology, and department of biology and wildlife, university of alaska fairbanks, alaska 99775 usa; 3department of biological sciences, 921 south 8th avenue, stop 8007, idaho state university, pocatello, idaho 83209 usa; 4division of wildlife conservation, alaska department of fish and game, 2030 sea level drive, ketchikan, alaska 99901 usa; 526700 west fork road, prince george, british columbia, v2k 5l6 canada. abstract: aerial surveys are the principal methods used to estimate populations of moose (alces alces gigas) in alaska. accounting for missed animals during aerial surveys is problematical, especially in forested habitats; incorporation of a visibility correction factor to account for the proportion of animals missed is known to improve accuracy of population estimates. our purpose was to study factors affecting visibility of radio-collared moose during aerial surveys in a temperate rainforest on the yakutat foreland, alaska, usa. wildlife managers in the area typically assume they observe only 50% of moose during surveys regardless of widely varying conditions. we used logistic regression to examine factors that influenced visibility including vegetation, light conditions, snow cover, and sex, age, and group size of moose. we then used logistic regression to develop a simpler model that only contained variables easily measured during aerial surveys: forest cover, snow cover, light, open versus vegetated habitat, and group size. we used that model to estimate a visibility correction factor. the mean correction factor was 1.304, ranging from1.005-2.138, yielding a population estimate of 699 (90% ci = 671-724) moose from a survey count of 595 animals. our correction factor was within the range reported for other populations of moose, and lower than the correction factor (2.0) currently used in this area. we conclude that application of site and time-specific visibility models is critical when estimating populations of large ungulates, especially in forested habitats. alces vol. 48: 89-104 (2012) key words: aerial surveys, alaska, alces alces gigas, gis, moose, population estimate, radio-telemetry, visibility. population estimates of ungulates based on aerial surveys are subject to error associated with the inability to detect animals that are present (visibility bias; timmerman 1993, anderson and lindzey 1996). environmental factors such as rugged terrain or dense cover may obscure visibility of animals, and differences in habitat selection and morphology by sex and age groups may make some animals more difficult to observe, thereby biasing their visibility (peek et al. 1974, thompson and veukelich 1981, bowyer et al. 2002, bowyer 2004). grouping behavior, activity of individuals (i.e., lying or standing), weather, and ground conditions (e.g., snow cover) can measurably affect visibility of animals. many of these problems are manifest in aerial surveys of moose (alces alces gigas) in temperate rainforests on the yakutat foreland of southeast alaska, usa where snow conditions that facilitate detecting moose can be intermittent, weather conditions for flying are frequently poor, and forest cover is dense and widespread. ideally, population surveys should be conducted during the mating season when moose are more active and sexes aggregated (miquelle et al. 1992, oehlers et al. 2011). because yakutat does not generally receive sufficient snowfall to enhance visibility before sexes spatially segregate after mating visibility of moose oehlers et al. alces vol. 48, 2012 90 and males cast antlers, identification of sex is difficult. sightability (also referred to as detectability or visibility) is the probability that an animal within the field of search for an observer will be seen by that observer (caughley 1974). that probability can be expressed as a scalar, or correction factor for visibility bias, which is then multiplied by the number of moose observed to obtain a more accurate population estimate than an uncorrected count (steinhorst and samuel 1989). correction factors for visibility bias (commonly referred to as sightability correction factors or scfs, and hereafter referred to as correction factors) that account for the proportion of animals undetected during aerial surveys are known to improve the accuracy of population estimates (timmerman 1993), particularly for areas with extensive forest cover and variable weather conditions that occur on the yakutat foreland. survey precision incorporates both the variance of total moose sighted and the variance of the correction factor (timmerman 1993). logistic regression is commonly used to develop correction factors for ungulates (mccorquodale 2001, quayle et al. 2001, mcintosh et al. 2009); this method is designed for use with binomial dependent variables (observed or not), and can accommodate continuous and categorical independent variables (hosmer and lemeshow 2000). we studied factors affecting visibility of moose on the yakutat foreland to improve population estimates from aerial surveys. we derived a series of models predicting correction factors using data from visibility trials from aerial surveys involving radio-collared moose. we examined the influence of temporal and weather-related variables such as month, time of day, cloud cover, light intensity, precipitation, and wind speed on visibility. we considered effects of environmental variables such as snow, forest, and vegetation cover on visibility of moose. in addition, we investigated the influence of sex, age, group size, sex and age composition of groups, activity, and intensity of site use on visibility. logically, we expected that forest cover and lack of snow cover would reduce the probability of moose being observed, and that visibility would increase with increasing snow cover. we also hypothesized that visibility would decline with smaller group size or if moose were bedded. further, we postulated that age or sex would affect visibility, because of morphological differences or if age groups and sexes used different habitats. we derived a model containing all of the covariates that we determined were important predictors of visibility, and a second model that included only those variables for which information could be obtained from routine aerial surveys. the full model was needed to consider all variables, including life-history characteristics such as sex and age and their potential influence on visibility, and would be useful in areas where sex and age composition is known or could be determined during surveys. we considered the second model to be more appropriate for management purposes in our study area, because it did not require data that could only be obtained reliably from radio-collared animals, and is more appropriate for late-winter surveys when sex cannot be accurately determined. finally, we applied the management model to a sample data set to estimate the density of moose within our study area. ours is one of few studies to examine factors influencing visibility of ungulates in a northern temperate rainforest, and our results should be useful to biologists managing ungulates throughout the northern coastal forests of the pacific northwest. study area we conducted research on the yakutat foreland of the tongass national forest, located along the southeast coastline of alaska (fig. 1). our study area of approximately 1,280 km2 encompassed most of the foreland, and included ~80 km of coastline extending from alces vol. 48, 2012 oehlers et al. visibility of moose 91 yakutat bay to dry bay. distance between the coast and mountain ranges varies from 8-24 km. there are several large rivers as well as numerous smaller streams distributed throughout the study area (fig 1). the yakutat foreland (lat. 59°20’ n, long. 139°0’ w) falls within the humid temperate domain, characterized by year-round cloudy, cool, and wet conditions (shephard 1995). the mean annual temperature was 4.1° c and the mean total precipitation was 381 cm (combined snow and rain) from 19712000 (noaa 2005). the mean temperature during this same time period was -3.4° c during january (the coldest month) and 12° c during july (the warmest month). total snowfall during the study was 345 cm; mean daily snowfall was 3.0 cm and the mean snow depth was 20 cm. other than a few rolling bedrock hills, most of the foreland is of low relief (average elevation 20 m; shephard 1995), and is a mosaic of forests, wetlands, and shrublands (shephard 1995). forested areas are dominated by sitka spruce (picea sitchensis), and a small percentage of the upper canopy is composed of black cottonwood (populus trichocarpa), western hemlock (tsuga heterophylla), and mountain hemlock (t. mertensiana). shephard (1995) documented 20 different forest communities on the foreland, with canopy cover ranging from 1-80% and averaging 60% for the common forest communities, with stand heights ranging from 15-47 m. nonforested areas include wetlands and shrublands composed primarily of graminoids, forbs, and shrubs including several species of tall and low willow (salix spp.) ranging from 1-6 m in height, and sitka alder (alnus sinuata) up to 4 m. nonforested areas dominate the coastal areas on the western half of the study area, with patches of spruce dispersed on the heaths and adjacent fig. 1. study area for developing a visibility model for moose on the yakutat foreland, alaska, usa, 2003-2004. visibility of moose oehlers et al. alces vol. 48, 2012 92 to some riparian zones; contiguous forested stands predominate the remainder. the most recent aerial surveys (2002) conducted in the forelands by alaska department of fish and game (adfg) estimated a density of 0.5 moose/km2 with a composition ratio of 19 males:100 females:14 young (n. l. barten, adfg, pers. comm.). total count surveys by parallel transects set approximately 0.4-0.5 km apart are conducted in the nonforested portions of the foreland by adfg in late autumn as soon as snow covers most of the ground, but often those conditions do not occur until well into winter. the adfg assumes that 50% of moose along transects are detected; consequently, the observed number of moose is doubled to estimate population size. that correction factor, however, has never been empirically evaluated or assessed. in addition, adfg does not survey forested portions of the forelands because of low (unknown) visibility, which constitutes about one-half of the study area. other large mammals that occur on the forelands include brown bear (ursus arctos), black bear (u. americanus), and gray wolves (canis lupus). sitka black-tailed deer (odocoileus hemionus sitkensis) occupy some of the islands offshore but are uncommon on the mainland. in addition, moose are an important part of the subsistence economy (ballew et al. 2006, schmidt et al. 2007). methods capture and handling twenty-two female and 16 male moose were darted from a helicopter by adfg personnel with palmer cap-chur equipment with the immobilizing drugs carfentanil and xylazine (roffe et al. 2001) during march and november 2002, and march and december 2003. dosages ranged from 3.0-5.0 mg of carfentanil and 100-130 mg of xylazine depending on time of year, sex, and animal condition. all capture and handling methods followed guidelines established by the american society of mammalogists animal care and use committee (1998) for research on wild mammals. our protocols were approved by independent institutional animal care and use committees at the university of alaska fairbanks (protocol # 04-26) and the adfg (protocol # 03-0001). we fitted moose with gps radio-collars (model 4000, lotek wireless, ontario, canada) that recorded locations 4 times daily, or standard vhf radio-collars (model mp2-mpp4, avm, colfax, california and model 600nh, telonics, mesa, arizona). we programmed both types of collars to release remotely relative to time of deployment (typically 1.5 yr). a lower incisor was removed from each moose to determine age from cementum annuli (gasaway et al. 1978). naltrexone (350-1300 mg) and tolazoline (400-800 mg) were subsequently administered and moose were monitored until they recovered from the immediate effects of immobilization. we also monitored each moose by aerial survey for 1 month post-capture to assess capture-related mortality. three females died or their collars malfunctioned within 1 month of capture, and were not included in the visibility trials. visibility trials we flew surveys to locate collared moose between 24 november 2003 and 18 march 2004 using a cessna® 185 fixed-wing aircraft. timing of sampling and type of aircraft were the same used by adfg when conducting moose surveys. we defined a visibility trial as the effort by the survey crew to count all moose within a 5 km2 sampling quadrat (square survey block) that included a radio-collared moose on a particular day. the aerial-survey crew was composed of the pilot, the primary observer in the front seat, and a secondary observer in the back seat behind the pilot. we attempted to control as many factors as possible, such as using the same aircraft, pilot, and primary observers for all trials. one pilot and 2 primary observers with >150 h of moose alces vol. 48, 2012 oehlers et al. visibility of moose 93 survey experience were used in the trials, with 8 secondary observers ranging in initial experience level of 6-8 (40-150 h of moose survey experience) on a lickart scale of 1-10. our trial procedure was similar to that of quayle et al. (2001). trials to locate individual radio-collared moose were separated by ≥3 days to reduce autocorrelation among locations. the extremely large home ranges of moose on the foreland (mean seasonal home ranges varied from 24.3-86.3 km2; oehlers et al. 2011) made this interval a reasonable choice for attempting to achieve independence among locations. frequencies for the subset of moose to be sampled during a flight were programmed into a receiver (model r4000, ats, isanti, minnesota) and scanned while flying at an altitude of 245-300 m above ground level. once a signal was received, the primary observer obtained the general position of the moose without identifying an exact location. we used a laptop computer equipped with baker geolink sketchmapping software (michael baker corporation, moon township, pa) to record our location and flight path so that the telemetry operator (primary observer) could identify the approximate location of the collared moose on the map without viewing the ground, thereby minimizing observer bias. the pilot and secondary observer also avoided scanning the ground in the immediate survey area to prevent detection of the target animal before beginning the survey. the survey crew noted if the collared moose was accidentally spotted by either observer while obtaining the general location; those observations were eliminated from analyses. the primary observer then delineated a 5 km2 (2.23 km x 2.23 km) quadrat (quayle et al. 2001) centered around the general location of the identified moose on the laptop computer using a 0.4 km grid overlay on the screen. because the location of the moose was inexact, the actual location of the moose was not centered within the quadrat. consequently, observer bias was minimized because none of the observers knew where in the quadrat to expect to find the radio-collared moose. the pilot then flew over the quadrat along transects spaced 0.4 km apart, which were delineated by the grid overlaid on the screen. the laptop screen displayed our flight path, allowing the pilot to navigate and follow the specified transect lines. the pilot flew the aircraft at an altitude of approximately 185 m and speed of about 130 km/h, resulting in a search intensity of approximately 1.0 min/ km2. we circled the location of each moose sighted in the quadrat to identify and record information on all of the variables included in the appendix, and recorded the location of the moose using the sketchmapper software. if the targeted moose was not sighted during the survey, we located that animal via telemetry and recorded the same information. forest cover was measured at 2 scales and recorded as “0” if the predominant vegetation within both a 10 m and 250 m radius of the radio-collared moose was nonforested, and “1” if this same area was predominantly forested (including a range of canopy covers). vegetation cover was defined as “0” if the predominant vegetation was open habitat such as muskeg, meadow, sand, or gravel bar, or “1” if there was vegetation such as tall shrubs or forest that could obscure visibility of the moose. percent vegetation was recorded as a categorical variable (1-3) representing percentage of vegetative cover (shrubs or trees) within a 10 m and 250 m radius of the observed moose that could obscure visibility of that moose. we defined a “group” as 1 or more moose within 50 m of each other (siegfried 1979, molvar and bowyer 1994, bowyer et al. 2001) to encompass the complete range of sociality for this species (monteith et al. 2007). we categorized age of non-collared animals as young (<1 year) or adult (≥1 years old) through visual observation. we expected a high pregnancy rate of yearlings (boer 1992), because preliminary data indicated a predator-limited visibility of moose oehlers et al. alces vol. 48, 2012 94 population (bowyer et al. 2005, oehlers et al. 2011). consequently, we considered yearling females as adults (monteith et al. 2007). moreover, distinguishing between yearlings and adults during aerial surveys in winter was difficult, and further distinguishing of ages beyond yearling or adult was not possible during aerial surveys. we used arcview 3.2 geographic information software (esri, redlands, ca, usa) to plot gps locations for moose and determine elevation and distance to the coast for each moose. elevation was extracted from a raster data layer provided by the u.s. forest service (usfs), which was based on usgs digital elevation model with 20-m resolution. distance from shore was calculated with the usfs shoreline polygon layer for the study area. statistical analyses detection of a radio-collared moose during visibility trials was coded 1 if detected and 0 if not observed. we used sas 9.1 (sas institute, cary, nc) for all statistical tests, and adopted an α = 0.05. we used multivariate logistic regression to model visibility. our suite of potential predictors of detection included parameters such as sex, age, group size, forest cover, snow cover, light conditions, aircraft speed, and experience of observers (appendix). group size was squared because the untransformed covariate was not linear in the logit. we included the identification of individual moose as a coded variable to control for making repeated measures of individual moose. we reduced potential multicollinearity among independent variables by testing for strong correlations between pairs of covariates (│r│≥0.7) and preventing their simultaneous inclusion in logistic regression models. during initial model screening, we also examined variance inflation factors (vif) and tolerance (tol) of independent continuous and discrete variables to identify intercorrelated variables. values of vif <10 and tol >0.40 were considered acceptable (neter et al. 1996, allison 2001). we ultimately considered 16 variables from the initial set of 27 candidate predictor variables. we then screened these remaining covariates using forward step-wise logistic regression (proc logistic; agresti 1990) with an alpha to enter of 0.15 (hosmer and lemeshow 2000, p. 118) and alpha to remove of 0.3, and backward logistic regression with alpha to remove of 0.3, to define a broad initial set of candidate models. we restricted the number of covariates within any candidate model to ≤8, because our sample size of visibility trials was 88; our sample size precluded a global model. our sample size also precluded an all possible regressions approach. we used hosmer and lemeshow tests for goodness-of-fit (hosmer and lemeshow 2000) to determine the appropriateness of the logistic models. once we had established a large set of candidate models, we used akaike’s information criterion (aicc) (burnham and anderson 2002) to select model variables. age and sex were included in most of the top candidate models. classifying moose into discrete age classes (i.e., beyond yearling or adult) is not possible from aerial surveys, and correct classification of sex is difficult once males have cast their antlers, so we repeated this same process omitting age and sex to allow development of models that did not rely on data from captured moose. accordingly, we developed overall explanatory models that included life-history characteristics, as well as management models which included variables that could be measured easily during aerial surveys alone. we used model-averaging procedures to derive composite explanatory and management models (burnham and anderson 2002, giudice et al. 2012). we only considered candidate models with aicc δ values ≤4 for inclusion in composite models. we calculated relative effects (risk ratios) for covariates included in our composite models (farmer et al. 2006). relative effects estimate the change in relative probability of detection alces vol. 48, 2012 oehlers et al. visibility of moose 95 for an incremental change in magnitude of a predictor variable (riggs and pollock 1992). we evaluated relative effects to determine the comparative importance of independent variables in affecting the probability of detection. in general, relative effects >2.0 or <0.5 indicated large effects of covariates on detectability (riggs and pollock 1992). for demonstrative purposes, we applied our composite management model to existing surveys of the moose population that were conducted by adfg on the yakutat foreland from 30 november-4 december 2005 using their survey methodology previously described in study area. model variables were assessed for each individual or group of moose observed during these surveys, and then the corresponding correction factor was calculated for each observation and multiplied by the number of animals in that observation. these corrected estimates were then totaled to derive a mean population estimate and the range of population estimates using the upper and lower correction factor based on the 90% ci (becker and reed 1990, anderson and lindzey 1996, white 2005). these data included 262 observations of single moose or groups and 595 total moose observed. results the median age for both females (n = 22, range = 3-13 yr) and males (n = 16, range = 1-10yr) was 6 years. we conducted 88 trials involving 55 radio-collared females and 33 males; each was surveyed 1-4 times (x = 2.3, sd = 0.70). snow conditions were generally adequate for aerial surveys from novemberjanuary and during the last 20 surveys conducted in march, but comparatively poor during february. we observed 254 groups of moose. radio-collared animals were sighted in 71% of the surveys; males were observed in 76% and females in 66% of the trials. radiocollared animals were detected in 82% of trials in nonforested areas, and in 27% of trials in forested cover. animals 1-3, 4-6, 7-10, and 11-13 years old were detected in 89, 55, 75, and 100% of trials, respectively. mean (± se) group size of collared animals was 3.7 ± 0.4. radio-collared animals were observed in open (31%), shrub (52%), and forested (17%) habitat during the trials. the location of females and males in nonforested and forested habitat was similar; 82 and 85% and 18 and 15%, respectively. logistic regressions forest cover, vegetation cover, and percent cover were each correlated (│r│≥0.7) between the 2 scales of measurement (10 m and 250 m). we considered the 10-m scale more easily estimated and likely to be consistent between observers; consequently, we chose to include the 10-m scale for each of these variables for consideration in our models. following tests for collinearity, variance inflation factors, and tolerance, candidate models for overall visibility included the parameters age, group size, forest cover, light, snow cover, experience secondary, and wind speed start (table 1). age, group size2, forest cover, and snow cover were included in each of the top 3 candidate models. visibility increased by 38% for each additional year of the moose aged, and by 75% for each additional (increasing) experience level of the secondary observer (table 2). overcast skies (versus sun) increased visibility by 175%. visibility increased with group size2 and speed of the plane (flight speed ranged from 129-145 km/h), but effects were small. visibility declined under forested cover (94%), snow cover of 0-33% (76%) or 34-66% (82%), and for females (23%). candidate models derived for management purposes (omitting sex and age) included group size2, forest cover, snow cover, light conditions, and vegetation cover (table 3). similar to the overall model, detectability increased with group size, nonforested and open habitat, overcast skies, and higher snow cover in the composite management model (table visibility of moose oehlers et al. alces vol. 48, 2012 96 4). application of the composite management model to our sample data yielded a range of correction factors from 1.005-2.138 for each observation. the mean correction factor was 1.304, and mean upper and lower (90% ci) correction factors were 1.215 and 1.390, yielding a population estimate of 671-724 animals (x = 699 moose) from an uncorrected count of 595 animals. discussion both our overall and management models included group size, forest cover, and snow cover as covariates of visibility. lack of snow cover strongly reduced visibility of moose and confirmed our hypothesis that visibility would be higher as snow cover increased. nonetheless, that relationship was not linear because visibility was similar between snow cover of 0-33% and 34-66% (57% and 54%, respectively). we believe that snow cover of 34-66% did not improve visibility because snow was still sufficiently patchy to obscure many moose against a dark background. we hypothesize that no snow cover actually may be preferable to patchy snow because patchy snow conditions may fatigue observers more quickly than uniform coverage. forest cover has been included in visibility models for both north american elk (cervus elaphus; samuel et al. 1987, bleich et al. 2001) and moose (peterson and page 1993, anderson and lindzey 1996, drummer and aho 1998, quayle et al. 2001). in our study area, coniferous tree species predominate in the forested areas, obstructing visibility of moose year-round, whereas vegetation in nonforested areas included alders and willows that model k parameters aicc∆i aiccwi a 5 age, group2a, forest cover, light, snow 0.000 0.4428 b 7 age, sex, observer2b, speedc, group2, forest cover, snow 0.7520 0.3041 c 6 age, sex, observer2, group2, forest cover, snow 1.8083 0.1793 d 4 group2, forest cover, snow, light 3.5851 0.0738 e 16 saturatedd 18.8200 0.0000 table 1. number of model parameters (k), differences in akaike’s information criterion (aicc) scores (∆) and aicc weights (wi) for candidate visibility models for moose on the yakutat foreland, alaska, 2003-2004. agroup size2. bexperience secondary. cwind speed start. dincludes survey start time, temperature, group, sex, age, experience primary, experience secondary, wind speed start, flight speed, group size2, forest cover, vegetation cover, percent cover, activity, light, snow cover, and elevation. variable β se rr rr 90% ci intercept -14.284 16.844 n/a n/a age 0.325 0.169 1.384 1.047-1.829 group size2 0.074 0.075 1.077 0.951-1.219 forest cover -2.849 0.945 0.058 0.0120.275 light 1.010 1.157 2.746 0.407-18.524 snow cover 1 (0-33%)a -1.419 1.201 0.242 0.033-1.755 snow cover 2 (34-66%)a -1.661 1.059 0.190 0.033-1.090 sexb -0.256 0.352 0.774 0.433-1.384 flight speed 0.063 0.075 1.065 0.941-1.205 experience secondary 0.562 0.693 1.754 0.559-5.504 table 2. regression coefficients and risk ratios (rr) for selected composite overall explanatory model for visibility of moose on the yakutat foreland, alaska, 2003-2004. confidence intervals did not overlap 1 in the individual models. asnow cover is relative to the reference variable of level 3, 67-100%. bsex is relative to the reference variable of male. alces vol. 48, 2012 oehlers et al. visibility of moose 97 do not retain leaves during winter and have less effect on visibility. group size was less influential on visibility than either forest or snow cover. group size affects visibility of elk (samuel et al. 1987, bleich et al. 2001, mccorquodale 2001), feral horses (equus caballus; ransom 2012), and mule deer (odocoileus hemionus; ackerman 1988); logically, larger groups are generally more visible. moose tend to aggregate in open areas in alaska during rut (miquelle et al. 1992, molvar and bowyer 1994); therefore, if snow conditions are adequate, visibility would be highest during the peak of rut. visibility did not differ when moose were standing or bedded. light condition also was an important predictor of visibility as moose were more visible in overcast conditions when glare and shadows were minimized. fox (1977) noted similar issues with glare from snowfields during mountain goat (oreamnos americana) surveys conducted in clear weather in southeast alaska. visibility increased with increasing age of moose, and was higher for males than for females, although the relative effect of sex was small. greater visibility of males could distort male:female ratios and result in the underestimation of the female population, unless a correction for differential visibility is incorporated. although several other studies of ungulate visibility reported that sex or group composition was accounted for in multivariate models because of correlation with other covariates such as group size or vegetation (anderson and lindzey 1996, bleich et al. 2001, mccorquodale 2001), sex in our model was not correlated with any other variable. the effect of sex on visibility probably occurred because of physical differences between the sexes; larger body size, darker color, and presence of antlers in early winter likely explain the higher visibility of males. solberg et al. (2010) also reported that male moose were observed by hunters with a 1.26 higher probability than females during the hunting season, and suggested that this difference was reflective of fundamental differences in antipredator behavior, including risk taking (such as use of open habitat), activity level, and space use. although age was not model k parameters aicc∆i aiccwi a 4 group2a, forest cover, snow, light 0.0000 0.5404 b 5 group2, forest cover, vegetation cover, snow, light 1.0609 0.3180 c 3 group2, forest cover, snow 2.6844 0.1442 d 14 saturatedb 14.4800 0.0004 table 3. number of model parameters (k), differences in akaike’s information criterion (aicc) scores (∆), and aicc weights (wi) for candidate visibility management models for moose on the yakutat foreland, alaska, 2003-2004. agroup size2. bincludes survey start time, temperature, group, experience primary, experience secondary, wind speed start, flight speed, group size2, forest cover, vegetation cover, percent cover, activity, light, snow cover, and elevation. variable β se rr rr 90% ci intercept 0.048 0.905 n/a n/a group size2 0.070 0.038 1.073 1.007-1.142 forest cover -2.551 2.190 0.078 0.002-2.894 light 1.441 0.920 4.225 0.926-19.279 snow cover 1 (0-33%)a -1.028 0.935 0.358 0.076-1.673 snow cover 2 (34-66%)a -1.377 0.897 0.252 0.057-1.109 vegetation cover -0.284 0.383 0.753 0.400-1.416 table 4. regression coefficients and risk ratios (rr) for selected composite management model for visibility of moose on the yakutat foreland, alaska, 2003-2004. confidence intervals did not overlap 1 in the individual models. asnow cover is relative to the reference variable of level 3, 67-100%. visibility of moose oehlers et al. alces vol. 48, 2012 98 significantly correlated with other variables, all observations of the oldest animals were in large groups in non-forested habitat; therefore, other covariates besides age were likely more influential on visibility. we did not detect an influence of age and sex composition of groups on visibility. we attempted to standardize flight speed during surveys, and weather conditions resulted in a minimal range of speeds (130145 km/h). remarkably, visibility of moose increased with speed of the plane. nonetheless, flight speed was in only 1 of the top 4 candidate overall explanatory models, its relative effect was small, and the 90% ci included 0; within the range of speeds we flew, this variable was likely of minimal importance. experience level (1-10) of the second observer increased visibility by 75% in the explanatory model; however, the effect was highly variable, and was not included in the management model. although experienced observers have developed a search image, and therefore may be more likely to observe moose, observer experience is difficult to quantify, and experience level changes over the course of visibility trials. previous studies have documented differences in visibility related to observer experience (leresche and rausch 1974, caughley et al. 1976); however, recent studies have noted little effect on visibility when observers were experienced (ackerman 1998) or when observer experience correlated with other variables in the model (samuel et al. 1987, anderson and lindzey 1996). all second observers in our study were experienced in moose surveys (i.e., 40-150 h of moose survey experience); consequently, our model will be most effectively applied when using experienced observers, a conclusion also reached by quayle et al. (2001). our overall visibility of moose was 70.5% and similar to that in quebec (crête et al.1986; 73%), alberta (rolley and keith 1980; 64%), and isle royale, michigan (peterson and page 1993; 78%), and higher than in minnesota (giudice et al. 2012; 38-56%), michigan (drummer and aho 1998; 39%), wyoming (anderson and lindzey 1996; 59%), and alaska (leresche and rausch 1974; 43-68%). correction factors for moose range from 1.03-3.2 (oosenberg and ferguson 1992, timmerman and buss 1998) and are generally higher in areas of denser cover and higher moose density (gasaway et al. 1986, peterson and page 1993). comparisons of visibility rates may be tenuous, however, because of differences in aircraft type (crête et al. 1986), number of observers, search intensity, and habitat (anderson and lindzey 1996). our results are within the range of correction factors reported for moose, but emphasize the variability in visibility and the need to develop correction factors specific to a particular area and time frame. the use of a dynamic correction factor, such as that developed with a visibility model, is superior to the use of a static correction factor. our modeled correction factor is offered as an alternative to the use of both a calculated scf (scfc) and an observed scf (scfo) as described by gasaway et al. (1986). observed scfs must be calculated for each survey (preferable daily), and are cost prohibitive in areas dominated by dense coniferous forests and areas of low moose density (gasaway et al. 1986), both of which occur in our study area. our results confirm that visibility of moose from aircraft varies with environmental factors and group size. therefore, application of the visibility model, combined with an appropriate sampling strategy, and with sophisticated analytical methods such as machine learning (‘non-linear statistics’; breiman 2001), may improve the accuracy and precision of population estimates over the use of a static correction factor. our method could be extended to other areas of similar environmental conditions such as the remainder of coastal alaska and british columbia (and could be tested for applicability to interior alaska) if protocols associated with alces vol. 48, 2012 oehlers et al. visibility of moose 99 the chosen model are followed (mccorquodale 2001). because visibility may differ among types of aircraft used (crête et al 1986), surveys should be conducted using a cessna® 185 or similar fixed-wing aircraft at approximately 185 m above ground elevation, as used in model development (samuel et al. 1987, anderson and lindzey 1996). additionally, observers should be experienced and their observation skills constantly calibrated in aerial surveys of moose. conducting surveys when moose are likely to be most visible (i.e., with nearly continuous snow cover and overcast light conditions) will provide the most precise population estimates. improved population estimates will allow for more knowledgebased and effective management decisions by state and federal managers. acknowledgements this research was funded primarily by the u.s. forest service, with additional funding from the department of interior bureau of indian affairs, and in-kind support from the adfg. the institute of arctic biology, department of biology and wildlife, and the alaska fish and wildlife cooperative research unit of the university of alaska fairbanks were all instrumental in the educational and funding portion of the project. w. eastland contributed technical support. many thanks to t. o’connor, c. grove, and e. campbell of the u.s. forest service for project support. special thanks to adfg biologists j. crouse, s. jenkins, n. barten, and k. white for their support in capture and handling of moose. pilots d. russel, l. hartley, b. bingham, and j. liston contributed to captures of moose, and our aerial survey efforts. thanks also to helicopter pilots of temsco helicopter and u.s. forest service helicopter managers a. stearns, j. schlee, and d. andreason. we acknowledge u.s. forest service personnel n. catterson, k. schaberg, b. lucey, d. gillikin, s. mehalick, m. moran, and c. wiseman for field and logistical support. references ackerman, b. r. 1988. visibility bias of mule deer aerial census procedures in southeast idaho. ph. d. dissertation, university of idaho, moscow, usa. agresti, a. 1990. categorical data analysis. john wiley & sons, new york, new york, usa. allison, p. d. 2001. logistic regression using the sas system: theory and application. sas institute inc., cary, north carolina, usa. anderson, c. r., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24: 247-259. animal care and use committee. 1998. guidelines for the capture, handling, and care of mammals as approved by the american society of mammalogists. journal of mammalogy 74: 1416-1431. ballew, c., a. r.tzilkowski, k. hamrick, and e .d. nobmann. 2006. the contribution of subsistence foods to the total diet of alaska natives in 13 rural communities. ecology of food and nutrition 45: 1-26. becker, e. f., and d. j. reed. 1990. a modification of a moose population estimator. alces 26: 73-79. bleich, v. c., c. s. y chun, r. w. anthes, t. e. evans, and j. k. fischer. 2001. visibility bias and development of a sightability model for tule elk. alces 37: 315-327. boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1: 1-10. bowyer, r. t. 2004. sexual segregation in ruminants: definitions, hypotheses, and implications for conservation and management. journal of mammalogy 85: 1039-1052. _____, d. r. mccullough, and g. e. belovsky. 2001. causes and consequences of sociality in mule deer. alces 37: 371402. visibility of moose oehlers et al. alces vol. 48, 2012 100 _____, d. k. person, and b. m. pierce. 2005. detecting top-down versus bottom-up regulation of ungulates by large carnivores: implications for conservation of biodiversity. pages 342-261 in j. c. ray, k. h. redford, r. s. steneck, and j. berger, editors. large carnivores and biodiversity conservation. island press, covelo, california, usa. _____, k. m. stewart, s. a. wolfe, g. m. blundell, k. l. lehmkuhl, p. j. joy, t. j. mcdonough, and j. g. kie. 2002. assessing sexual segregation in deer. journal of wildlife management 66: 536-544. breiman, l. 2001. statistical modeling: the two cultures. statistical science 16: 199-231. burnham, k. p., and d. r. anderson 2002. model selection and multimodel inference: a practical information-theoretic approach. second edition. springer-verlag inc., new york, new york, usa. caughley, g. 1974. bias in aerial survey. journal of wildlife management 38: 921-933. _____, r. sinclair, and d. scott-kemmis. 1976. experiments in aerial survey. journal of wildlife management 40: 290-300. crête, m., l. rivest, h. jolicoeur, j. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose with observations made from fixed-wing aircraft in southern quebec. journal of applied ecology 23: 751-761. drummer, t. d., and r. w. aho. 1998. a sightability model for moose in upper michigan. alces 34: 15-19. farmer, c. f., d. k. person, and r. t. bowyer. 2006. risk factors and survivorship of black-tailed deer in a managed forest landscape. journal of wildlife management 70: 1403-1415. fox, j. l. 1977. summer mountain goat activity and habitat preference in coastal alaska as a basis for the assessment of survey technique. pages 190-199 in w. samuel and w. g. macgregor, editors. proceedings first international mountain goat symposium, 19 february, 1977, kalispell, montana, usa. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological paper of the university of alaska fairbanks 22: 1-108. _____, d. b. harkness, and r. a. rausch. 1978. accuracy of moose age determinants from incisor cementum layers. journal of wildlife management 42: 558-563. guidice, j. h., j. r. fieberg, and m. s. lenarz. 2012. spending degrees of freedom in a poor economy: a case study of building a sightability model for moose in northeastern minnesota. journal of wildlife management 76: 75-87. hosmer, d. w., and s. lemeshow. 2000. applied logistic regression. second edition. john wiley and sons, new york, new york, usa. leresche, r. e., and r. a. rausch. 1974. accuracy and precision of aerial moose censusing. journal of wildlife management 38: 175-182. mccorquodale, s. m. 2001. sex-specific bias in helicopter surveys of elk: sightability and dispersion effects. journal of wildlife management 65: 216-225. mcintosh, t. e., r. c. rosatte, j. hamr, and d. l. murray. 2009. development of a sightability model for low-density elk populations in ontario, canada. journal of wildlife management 73: 580-585. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monographs 122: 1-57. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75: 621-630. alces vol. 48, 2012 oehlers et al. visibility of moose 101 monteith, k. l., c. l. sexton, j. a. jenks, and r. t. bowyer. 2007. evaluation of techniques for categorizing group membership of white-tailed deer. journal of wildlife management 71: 1712-1716. national oceanic and atmospheric administration (noaa). 2005. local climatological data. (accessed may 2007). neter, j. w., m. h. kutner, c. j. nachtsheim, and w. wasserman. 1996. applied linear statistical models. fourth edition. the mcgraw-hill companies, inc., new york, new york, usa. oehlers, s. a., r. t. bowyer, f. huettmann, d. k. person, and w. b. kessler. 2011. sex and scale: implications for habitat selection by alaska moose. wildlife biology 17: 67-84. oosenberg, s. m., and s. h. ferguson. 1992. moose mark-recapture survey in newfoundland. alces 28: 21-29. peek, j. m., r. e. leresche, and d. r. stevens. 1974. dynamics of moose aggregations in alaska, minnesota, and montana. journal of mammalogy 55: 126-137. peterson, r. o., and r. e. page. 1993. detection of moose in midwinter from fixedwing aircraft over dense forest cover. wildlife society bulletin 21: 80-86. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43-54. ransom, j. i. 2012. detection probability in aerial surveys of feral horses. journal of wildlife management 76: 299-307. riggs, m. r., and k. h. pollock. 1992. a risk ratio approach to multivariable analysis of survival in longitudinal studies of wildlife populations. pages 74-89 in d. r. mccullough and r. h. barrett, editors. wildlife 2001: populations. elsevier applied science, new york, new york, usa. roffe, t. j., k. coffin, and j. berger. 2001. survival and immobilizing moose with carfentanil and xylazine. wildlife society bulletin 29: 1140-1146. rolley, r. e., and l. b. keith. 1980. moose population dynamics and winter habitat use at rochester, alberta, 1965-1979. canadian field-naturalist 94: 9-18. samuel, m. d., e. o. garton, m. w. schlegel, and r. g. carson. 1987. visibility bias during aerial surveys of elk in northcentral idaho. journal of wildlife management 51: 622-630. schmidt, j. i., j. m. ver hoef, and r. t. bowyer. 2007. antler size of alaskan moose: effects of population density, harvest intensity, and use of guides. wildlife biology 13: 53-65. shephard, m. e. 1995. plant community ecology and classification of the yakutat foreland, alaska. us forest service, r10-tp-56. alaska region, juneau, alaska, usa. siegfried, w. r. 1979. vigilance and group size in springbok. madoqua 12: 151154. solberg, e. j., c. m. rolandsen, m. heim, j. d. c. linnell, i. herfindal, and b. seather. 2010. age and sex-specific variation in detectability of moose (alces alces) during the hunting season: implications for population monitoring. european journal of wildlife research 56: 871-881. steinhorst, r. k., and m. d. samuel. 1989. sightability adjustment methods for aerial surveys of wildlife populations. biometrics 45: 414-425. thompson, i. d., and m. f. veukelich. 1981. use of logged habitats in winter by moose with calves in northeastern ontario. canadian journal of zoology 59: 2103-2114. timmerman, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations a review. alces 29: 35-46. _____ , and m. e. buss. 1998. population and harvest management. pages 559-615 visibility of moose oehlers et al. alces vol. 48, 2012 102 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d. c., usa. white, g. c. 2005. correcting wildlife counts using detection probabilities. wildlife research 32: 211-216. alces vol. 48, 2012 oehlers et al. visibility of moose 103 variable type description method/time of collection month discrete month of visibility trial aerial survey age discrete age of collared moose capture sex discrete sex of collared moose capture group indicator 0 = single moose, 1 = ≥ 2 moose aerial survey group size discrete total number of moose seen within 50 m of collared moose aerial survey composition indicator 0 = single-sex group; 1 = both sexes in group aerial survey males discrete number of adult males in group aerial survey females discrete number of adult females in group aerial survey calves discrete number of calves in group aerial survey unknown sex discrete number of unknown sex adults in group aerial survey forest cover 10 m indicator 0 = nonforested, 1 = forested, within 10 m of moose aerial survey forest cover 250 m indicator 0 = nonforested, 1 = forested, within 250 m of moose aerial survey vegetation cover 10 m indicator 0 = open habitat such as muskeg, 1= shrub or forested habitat within 10 m of moose aerial survey vegetation cover 250 m indicator 0 = open habitat such as muskeg, 1= shrub or forested habitat within 250 m of moose aerial survey percent vegetation 10 m indicator 1 = 0-33%, 2 = 34-66%, 3 = 67-100% vegetative cover within 10 m of moose aerial survey percent vegetation 250 m indicator 1 = 0-33%, 2 = 34-66%, 3 = 67-100% vegetative cover within 250 m of moose aerial survey elevation continuous elevation above sea level in meters gis distance from coast continuous straight-line distance from coastline to center of moose group in meters gis activity indicator 0 = bedded, 1 = active (any moose in group) aerial survey site use indicator 0 = no beds, few tracks, 1 = beds and multiple tracks aerial survey cloud cover indicator 0 = clear, 1 = partly cloudy, 2 = overcast aerial survey precipitation indicator 0 = none, 1 = mist, 2 = light rain, 3 = hard rain, 4 = snow aerial survey snow cover indicator 1 = 0-33%, 2 = 34-66% ,3 = 67-100% aerial survey wind speed start continuous wind speed (km/h) at beginning of survey aerial survey wind speed end continuous wind speed (km/h) at end of survey aerial survey appendix candidate predictor variables considered during initial modeling for visibility of moose on the yakutat foreland, alaska, 2003-2004. visibility of moose oehlers et al. alces vol. 48, 2012 104 flight speed continuous average flight speed (km/h) during survey (excludes circling) aerial survey (plane instrumentation) temperature continuous average temperature (celsius) during survey aerial survey start time discrete survey start time; military time rounded to hour aerial survey light indicator 0 = sunny, 2 = flat light/even shadows aerial survey experience primary continuous previous experience level of primary observer, scale of 1-10 collected from each surveyor prior to visibility trials number flights primary discrete number of previous visibility trials by primary observer collected from each surveyor prior to visibility trials experience secondary continuous previous experience level of secondary observer, scale of 1-10 collected from each surveyor prior to visibility trials number flights secondary discrete number of previous visibility trials by secondary observer tabulated throughout visibility trials alces27_208.pdf alces24_188.pdf alces vol. 48, 2012 broders et al. moose response to temperature 53 ecothermic responses of moose (alces alces) to thermoregulatory stress on mainland nova scotia hugh g. broders1, andrea b. coombs1,2, and j. r. mccarron1 1department of biology, saint mary’s university, halifax, nova scotia, canada b3h3c3; 2current address: department of natural resources, wildlife division, 136 exhibition street, kentville, nova scotia, canada b4n 4e5. abstract: the size of the mainland nova scotia moose (alces alces) population has declined precipitously over the last several decades and their current distribution is discontinuous. in recognition of the state of its moose population, nova scotia declared moose as ‘endangered’ under nova scotia’s endangered species act in 2003. a variety of factors have been attributed to the decline, and the goal of this project was to determine whether thermoregulatory stress may be impacting the viability of the moose population. location and temperature information were collected from gps-collared moose to test predictions related to whether moose behaviour changes in response to high temperatures. overall, our results suggest that moose exhibit behaviours (i.e., ectothermy) that are consistent with thermoregulatory stress, but the actual impacts of this, if any, on population productivity requires further study. the greatest response occurred in the summer during both day and night, when moose moved to areas of lower ambient temperature. further, overall movements were significantly reduced during periods of high temperatures. alces vol. 48: 53-61 (2012) key words: alces alces, ectothermy, endangered, moose, nova scotia, temperature. the incidence of population declines are increasing due to a variety of factors, most notably overexploitation, habitat destruction, and food chain disruption (campbell and reece 2002). large mammals are highly vulnerable to human exploitation and it has been estimated that less than 21% of the earth’s terrestrial surface contains all the large mammals it once did (morrison et al. 2007). in the northern hemisphere, populations along the southern extent of their species range are particularly vulnerable to climate change (renecker and schwartz 1997, lenarz et al. 2009) and may eventually shift north in response to warming temperatures. these and a suite of other threats have been identified as negatively impacting north american moose (alces alces) populations. although some north american moose populations are stable or increasing, population declines have occurred in alaska (timmerman 2003), minnesota (murray et al. 2006), manitoba (v. crichton, manitoba natural resources, pers. comm.), and nova scotia (pulsifer and nette 1995) which has had closures of hunting seasons as a result (parker 2003). the southern range limit of moose may be determined by thermoregulatory stress (renecker and hudson 1986) and links between declining populations and increased ambient temperature associated with climate change have been suggested (murray et al. 2006, lenarz et al. 2009, 2010; but see lankester 2010). marai and haebb (2010) define heat stress as “the state at which mechanisms activate to maintain an animal’s body thermal balance, when exposed to untolerable (uncomfortable) elevated temperatures.” although the initial response may be physiological, behavioural modifications can reduce these physiological stressors (e.g., movement to cooler areas in response to heat stress). of all the extant boreal ungulate species, moose are the most likely candidate to suffer moose response to temperature broders et al. alces vol. 48, 2012 54 from heat stress due to their relatively low, upper critical temperature limit (-5° c in winter and 14° c in summer; karns 1997, renecker and hudson 1986). in addition to panting to ameliorate thermal stress (renecker and hudson 1986), moose often use aquatic areas (renecker and schwartz 1997) or forest stands that buffer from extremes in temperature, a form of ectothermy. such areas of thermal cover (dussault et al. 2004, mysterud and ostbye 2008) could provide conditions that may be as much as 7° c cooler than forest edges (chen et al. 1995). demarchi and bunnell (1995) and dussault et al. (2004) found that nocturnal activity of moose increased in summer and fall as ambient temperature increased, and use of thermal cover was lower at night suggesting that activity of moose may be inversely related to temperature and/or exposure to solar radiation. in late winter heat stress strongly influenced cover selection as moose tend to avoid areas where the temperature exceeds 8° c (the temperature when panting begins to dissipate heat from the body in the late winter months; schwab and pitt 1991). leblond et al. (2010) suggested that when temperatures are cooler, moose chose areas that had less thermal cover and higher amounts of solar energy. contrary to these studies however, lowe et al. (2010) did not find a behavioral response by moose to temperatures in ontario. they suggested that, within their study area, there were no obvious thermal refugia and that moose either were not impacted by the temperature range to which they were exposed, or the resolution of their measurements were not fine enough to detect a response. prior to european settlement of nova scotia, it was believed that the local moose population was large (≈15,000; parker 2003). approximately 100 years ago, the population on cape breton island was extirpated and the current population was founded by the introduction of moose from alberta (pulsifer and nette 1995). today, the remnant native population on the mainland has a discontinuous distribution with a crude population estimate of 1000 (parker 2003). on the mainland, the most significant populations occur in cumberland-colchester counties, pictou-antigonish counties, and in the tobeatic wilderness area. the last hunting season for mainland moose was held in 1981 (parker 2003), and in 2003 the population was classified as endangered under the nova scotia endangered species act. currently, there are no reliable demographic estimates or other data that could be used to justify management decisions. some of the factors believed to be affecting population growth include: parasites such as parelaphostrongylus tenuis, deterioration in the quantity and quality of moose habitat, poaching, predation, and thermal stress (brannen 2004, beazley et al. 2006). the goal of this study was to determine if there is any evidence that moose on mainland nova scotia exhibit signs of heat stress. specifically, our hypothesis was that moose would alter their behavior to reduce physiological stress in response to high temperatures. to assess this hypothesis, we tested 2 emergent predictions: 1) during periods of high temperature moose would select cooler areas, and 2) during periods of high temperature movement would be reduced relative to times when it was cooler. to test these predictions we used data from gps-collars deployed on moose on the mainland of nova scotia, 2002-2006. if our hypothesis is supported, our results may lend justification for further study to quantitatively characterize the population level impacts of increasing temperatures on moose, including population recovery in nova scotia. materials and methods location and temperature data from 12 gps-collared adult moose from cumberlandcolchester counties (n = 5), antigonish county (n = 1), and the chebucto peninsula of halifax county (n = 6) were provided by the nova scotia department of natural resources (nsdnr). the gps collars (lotek alces vol. 48, 2012 broders et al. moose response to temperature 55 gps 2200l; lotek wireless inc., newmarket, ont., canada) were programmed to acquire and store location and temperature data every 2-4 h. although gps collars have the advantage of greater location accuracy and resolution in movement dynamics data relative to conventional telemetry (girard et al. 2002), the ability to record a location at pre-determined times is reduced if the animal is under dense forest cover (rempel et al. 1995, moen et al. 1996, rodgers et al. 1997, dussault et al. 1999) or on steep slopes (gamo and rumble 2000); hence, there is potential for location bias. regardless, location accuracy was expected to be within 10 m under most conditions and times. location data were imported into arcgis geographic information system (gis), version 9.1 (esri, redlands, california) and each location was assigned to 1 of 4 cover types using land-use and forest resource inventory data for the region (interpreted from 1:10,000 aerial photos; nsdnr). the 4 cover types included softwood (75% softwood species by basal area), mixedwood (26-74% softwood species by basal area), hardwood (<25% softwood species by basal area), and other (water and all other land use types that were not forest cover types). for each prediction we controlled for time of day and season effects by conducting separate analysis during the day and night for each of 3 seasons (summer: 15 june-15 september; early winter: 16 november-14 january; late winter: 15 january-15 april). further, because individuals were likely to respond to temperature variations in different ways (e.g., due to differences in age, body condition, location, gender, and reproductive status), data collected from each individual were analyzed separately, but global inferences were based on the results from all individuals. to test our first prediction regarding whether moose selected cooler stands when temperatures were high, we determined the magnitude of the difference between the temperature recorded by the collar and the temperature recorded by the nearest environment canada weather station (collar temperature minus weather station temperature; called ∆t) for each record. positive or negative ∆t values indicated that the animal was in a location warmer or cooler than the temperature recorded at the weather station, respectively. however, complicating this measure was the fact that it was possible that the temperature as recorded on the collar was affected by radiant heat and variation in the degree of shading from the animal. we predicted that the impact of shading might vary between day and night, regardless of season but be relatively consistent all year. further, we expected the impact of radiant heat would be consistent for these endothermic homeotherms across the range of temperatures experienced by the animal in any one season. however, the impact of radiant heat should be greater in the summer due to the lower insulative potential offered by the summer coat. therefore, independent analyses were conducted for day and night as well as during the summer, early winter, and late winter to minimize bias. because of radiant heat, we expected that the collar temperature would be a positively biased measure of local temperature and therefore the power to detect selection of cooler areas based on temperature is reduced. therefore, we expected our results to be conservative. we did not analyze spring and fall data because of our inability to control for variation in the growth or shedding of the winter coat (samuel et al. 1986). to be further conservative and minimize the impacts of equipment malfunction, within each dataset (e.g., summer day data) we first sorted data by ∆t and deleted 5% of the data on each extreme of the continuum so that we only worked with 90% of the data. for prediction 1, we regressed ∆t on the temperature data from the nearest environment canada weather station recorded at the same time, or within an hour, for each individual moose. we predicted that if moose were selecting cooler areas during periods of warmth that moose response to temperature broders et al. alces vol. 48, 2012 56 there would be a significantly negative slope. assumptions of regression were confirmed for all analysis via an examination of residuals for normality and homoscedasticity (sokal and rohlf 1995). to test prediction 2 that moose would move less during periods of high temperatures, we divided the range of environmental temperature values for each of the 3 seasons and for day and night into 2 groups with a 7º c range (arbitrarily chosen based on the range of values in the dataset). we did not use the extremes of the temperature continuum because of low sample sizes and we omitted the records with temperature values in the middle of the distribution to ensure there was opportunity to detect variation between the 2 groups, if indeed there was meaningful variation. during the summer, the temperature ranges used were 10 to 16º c for low temperature and 20 to 26º c for high temperature. during early and late winter the temperature ranges used were -11 to -5º c for low temperature and 0 to 6º c for high temperature. when temperatures were within either the low or high range, for each animal, we calculated the average movement distance (i.e., straight-line distance between successive locations) over all 2 h periods for which we had location data. a one tailed ttest was used to test if the distance travelled during periods of high temperature was less than during periods of low temperature (for that season); we used α = 0.05 for decisionmaking criteria. where we found evidence of ectothermic response to temperature, we characterized the types of sites where moose were located during these times to better understand site type selection. results we recorded 29,964 locations for 12 moose from 2002-2006. summer we examined the temperature response of 7 moose (3 males from halifax county and 2 females from each of halifax and cumberland county) but we only had 2 h movement data from 5 of these animals (all from halifax county), as 2 individuals had collars programmed to record data at 4 h intervals. during the daytime, the slope of the regression lines varied among animals but all were negative and different from zero (all ps < 0.001). there was also a trend in the response to environmental temperature such that the male response (combined results: ∆ t = 8.0 – 0.28 ec temp; p <0.001, df = 1732, β0 se = 0.39, β1 se = 0.02) was greater than that of females (combined results: ∆ t = 4.8 – 0.20 ec temp; p <0.001, df = 1972, β0 se = 0.36, β1 se = 0.02). during the day all 5 moose moved less (all ps <0.001) during periods of high temperature than low temperature. the average movement distance during low temperatures was 2.2 x further than during high temperatures (range = 1.8-2.6 x). during periods of high temperature there was a greater proportion of locations in softwood, and a smaller proportion in mixed wood and open areas than during periods of cooler temperatures (table 1). at night the slopes of the regression lines for all 7 moose were negative and different from zero (all ps < 0.001). as in daytime, there was also a trend in the response to environmental temperature by gender, with the response by males (combined results: ∆ t = 6.6 – 0.39 ec temp; p <0.001, df = 12552, β0 se = 0.47, β1 se = 0.03) greater than that of females (combined results: ∆ t = 2.9 – 0.24 ec temp; p <0.001, df = 1629, β0 se = 0.38 β1 se = 0.02). there were far fewer instances (only 5-16 % as many per individual) of high temperature records than low temperature records. there was no difference (p < 0.05) in the movement distance between periods of high and low temperature for 3 of the 5 moose; 2 others (a male and a female from halifax county) moved more during high temperatures. alces vol. 48, 2012 broders et al. moose response to temperature 57 early winter we examined the temperature response of 5 moose (3 males from halifax county and 1 female from each of halifax and cumberland county). during this time we only had 2 h interval location data from 4 moose (all from halifax county) because the collar for the moose in cumberland county was programmed to record at 4 h intervals. during the day the slopes of the regression lines for all 5 moose were negative, but only 4 of these slopes (range of -0.11 to -0.23) were different from zero (p <0.001; the exception was a halifax county male). one of the individuals for which we have movement data had only 6 records for ‘low’ temperature, therefore analysis was only conducted for 3 individuals (2 males and 1 female), each with ≥20 records for each of ‘high’ and ‘low’ temperature; there was no difference in movement distance between periods of high and low temperature (all ps >0.05). at night there was no consistent trend in temperature response among the 5 moose tracked. regression lines for 2 of the 5 moose (both males from halifax county) were not different from zero (p >0.05), whereas another male and the female from the same area had positive relationships (p <0.05); a female from cumberland county had a negative relationship (p <0.05). one of the 4 moose (a halifax county female) moved more (p <0.05) during high temperature periods; movement distance was not different for the other 3 moose. late winter we had data to examine the temperature response of 11 moose (3 males and 3 females from halifax county, 4 females from cumberland county, and 1 female from antigonish county). however, only 6 animals wore collars programmed to record locations at 2 h intervals; therefore, movement analysis was conducted only with these 6 (2 males and 3 females from halifax county and 1 female from antigonish county). during the day there appeared to be minimal effects of gender in temperature response, but there was a trend with geography. of the 6 moose (3 males and 3 females) from halifax county, only 1 (male) had a regression line different (p <0.05) from zero (it was negative). however, each of the other 5 moose (all female; 4 from cumberland county and 1 from antigonish) had negative regression lines that were all similar to one another (combined results of 5 non-halifax county moose: ∆ t = 0.51 – 0.17 ec temp; p <0.001, df = 892, β0 se = 0.12, β1 se = 0.02). during the day we only had 2 h interval location data from moose in halifax county and one animal in antigonish. although 5 of the 6 moose had average movement distances that were less in high temperature periods than low temperature periods, only 3 were different (p <0.05), the 2 females from halifax and 1 from antigonish. there was a geographic trend in temperature response at night in that each of the 5 non-halifax county moose (all female; 4 from cumberland county and 1 from antigonish) had negative regression lines (combined results of 5 non-halifax county moose: ∆ t = -0.50 – 0.145 ec temp; p <0.001, df = 914, β0 se = 0.132, β1 se = 0.016). of the 6 halifax county moose, 3 had regression temperature site type 20-26 °c 10-16 °c difference softwood 47.6 39.5 8 mixedwood 26.1 30.3 -4.2 open 14.7 19.5 -4.8 water 7 5.7 1.3 hardwood 4.5 4.6 -0.1 other 0.2 0.4 -0.2 # locations 1331 1359 table 1. proportion (%) of the total locations of 11 moose on mainland nova scotia in each of 6 site-types when the temperature recorded at the nearest environment canada weather station was between 20-26 °c and 10-16 °c, 15 june-15 september, 2002-2006. moose response to temperature broders et al. alces vol. 48, 2012 58 lines not different from 0 (p >0.05) and 3 had positive regression lines (p <0.05). during the night there was no difference in the movement distance by 5 of the 6 moose during periods of high and low temperature. one moose (halifax county female) moved less (p <0.05) during high temperature periods relative to low temperature periods. discussion annual movement patterns, home ranges, daily, seasonal, and annual temperature regimes and a number of anthropogenic factors affect moose behavior (andersen 1991, schwab and pitt 1991, courtois et al. 2002, dussault et al. 2004). in this paper we present evidence from 2 predictions that support the hypothesis that moose on mainland nova scotia alter their behavior to reduce physiological stress in response to high temperatures. the extent to which this behavior is able to ameliorate the impacts of heat stress is unknown and would require further investigation. this finding is consistent with dussault et al. (2004) who suggested that moose spend more time under cover during hot days but come out to feed during cool nights. contrary to our results, research along the southern extent of moose range in ontario did not support our hypothesis as moose in their study did not seem to exhibit any behavioral response to increased temperatures (lowe et al. 2010). they found that there was little variation in temperature trends among site types such that there was minimal, if any, thermal advantage to selecting one site type over another. instead, they found that animal movement was negatively related to snow depth. in nova scotia, moose did not always display ectothermy by moving to thermal cover at the same temperature threshold or at the metabolic heat stress temperature threshold identified by renecker and hudson (1986). for example, during summer nights, based on individual regressions, moose moved to cooler areas when temperatures reached 14o c, which is consistent with their findings. however, during summer days there was more inter-individual variation and many moose did not move to cooler areas until around 24o c. during early winter days moose tended to seek cooler areas as day temperatures still generally exceeded their upper critical temperature. this pattern is not reflected during early winter nights or in late winter, possibly due to the lack of microclimate variation, indicating that temperature may be similar in all cover types. dussault et al. (2004) noted that some moose used open, deciduous, or mixed areas even when air temperatures are warm because thermal cover often offers low food availability. given the complexity of an animal’s thermal environment, factors such as wind and solar radiation in combination with ambient temperature presumably influence habitat use and movement. the temperature patterns were not as distinct in winter as they were in the summer, but we found that moose in winter had a greater than expected use of softwood cover based on availability of this cover type in their home range during times of expected heat stress. this finding agrees with other studies which suggest that during some seasons moose will use certain types of forest cover in disproportion to availability (cook et al. 2004). there is an assumption that closed canopy forests such as the softwood stands chosen by moose might provide areas with low snowfall amounts as well as thermal cover which makes them an even more attractive choice (jung et al. 2009). the moose population on the mainland of nova scotia is near the southern periphery of the species range in north america and is listed as ‘endangered’ due to its low population size. the behavioral patterns we document herein are suggestive of moose responding to increased temperature, and although may indicate heat stress, such thermoregulatory behaviour is not unexpected. given the concern for certain moose populations at the alces vol. 48, 2012 broders et al. moose response to temperature 59 southern fringe of their range, it would be prudent for further investigation into the extent to which behavioural thermoregulation (i.e., ectothermy) impacts population productivity. to be cautious, forest managers should be cognizant of, and explicitly address the need to maintain appropriate thermal cover on the landscape that allows moose to use ectothermy to ameliorate the effects of high temperatures at critical times of the year. further study may also be required to quantitatively characterize the relative ability of different cover types to buffer extremes of temperature. acknowledgements arcgis technical assistance was provided by greg baker, research tools technician for the mp_sparc lab. funding for this project was provided by an nserc undergraduate student research award and the nova scotia department of natural resources. this manuscript was improved based on comments provided by b. patterson and 2 reviewers. references andersen, r. 1991. habitat changes in moose ranges: effects on migratory behavior, site fidelity and size of summer home-range. alces 27: 85-92. beazley, k., m. ball, l. isaacman, s. mcburney, p. wilson, and t. nette. 2006. complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada. alces 42: 89-109. brannen, d. c. 2004. population parameters and multivariate modeling of winter habitat for moose (alces alces) on mainland nova scotia. m. sc. thesis, department of biology, acadia university, wolfville, nova scotia, canada. campbell, n. a., and j. reece. 2002. biology. sixth edition. benjamin cummings, san franscisco, california, usa. chen, j. q., j. f. franklin, and t. a. spies. 1995. growing-season microclimatic gradients from clear-cut edges into oldgrowth douglas-fir forests. ecological applications 5: 74-86. cook, j. g., l. l. irwin, l. d. bryant, r. a. riggs, and j. w. thomas. 2004. thermal cover needs of large ungulates: a review of hypothesis tests. transactions of the 69th north american wildlife and natural resources conference: 708-726. courtois, r., c. dussault, f. potvin, and g. daigle. 2002. habitat selection by moose (alces alces) in clear-cut landscapes. alces 38: 177-192. demarchi, m. w., and f. l. bunnell. 1995. forest cover selection and activity of cow moose in summer. acta theriologica 40: 23-36. dussault, c., r. courtois,j.-p. ouellet, and j. huot. 1999. evaluation of gps telemetry collar performance for habitat studies in the boreal forest. wildlife society bulletin 27: 965-972. _____, j.-p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioural responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321-328. gamo, r. s., and m. a. rumble. 2000. gps radio collar 3d performance as influenced by forest structure and topography. biotelemetry 15: proceedings of the 15th international symposium on biotelemetry 15: 464-473. www.fs.fed.us/rm/pubs_other/ rmrs_2000_gamo_r001.pdf . girard, i., j. p. ouellet, r. courtois, c. dussault, and l. breton. 2002. effects of sampling effort based on gps telemetry on home-range size estimations. journal of wildlife management 66: 1290-1300. jung, t. s., t. e. chubbs, c. g. jones, f. r. phillips, and r. d. otto. 2009. winter habitat associations of a low-density moose (alces americanus) population in central labrador. northeastern naturalist 16: 471-480. moose response to temperature broders et al. alces vol. 48, 2012 60 karns, p. d. 1997. population distribution, density and trends. pages 125-140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. lankester, m. w. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53-70. leblond m, c. dussault, and j. p. ouellet. 2010. what drives fine-scale movements of large herbivores? a case study using moose. ecography 33: 1102-1112. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013-1023. _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. lowe, s. j., b. r. patterson, and j. a. schaefer. 2010. lack of behavioral responses of moose (alces alces) to high ambient termperatures near the southern periphery of their range. canadian journal of zoology 88: 1032-1041. marai, i. f. m., and a. a. m. haeeb. 2010. buffalo’s biological functions as affected by heat stress a review. livestock science 127: 89-109 moen, r., j. pastor, y. cohen, and c. c. schwartz. 1996. effects of moose movement and habitat use on gps collar performance. journal of wildlife management 60: 659-668. morrison, j. c., w. sechrest, e. dinerstein, d. s. wilcove, and j. f. lamoreux. 2007. persistence of large mammal faunas as indicators of global human impacts. journal of mammalogy 88: 1363-1380. murray, d. l., e. w. cox, w. b. ballard, h. a.whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1-30. mysterud, a., and e. ostbye. 2008. cover as a habitat element for temperate ungulates: effects on habitat selection and demography. wildlife society bulletin 27: 385-394. parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. available from www.gov.ns.ca/natr/wildlife/largemammals/pdf/statusreportmoosens.pdf . pulsifer, m. d., and t. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. rempel, r. s., a. r. rodgers, and k. f. abraham. 1995. performance of a gps animal location system under boreal forest canopy. journal of wildlife management 59: 543-551. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322-327. _____, and c. c. schwartz. 1997. food habits and feeding behaviour. pages 403-440 in a.w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. rodgers, a. r., r. s. rempel, r. moen, j. paczkowski, c. c. schwartz, e. j. lawson, and m. j. gluck. 1997. gps collars for moose telemetry studies: a workshop. alces 33: 203-209. samuel, w. m., d. a. welch, and m. l. drew. 1986. shedding of the juvenile and winter hair coats of moose (alces alces) with emphasis on the influence of the winter tick, dermacentor albipictus. alces 22: 345-359. schwab, f. e., and m. d. pitt. 1991. moose alces vol. 48, 2012 broders et al. moose response to temperature 61 selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071-3077. sokal, r. r., and f. rohlf. 1995. biometry. third edition. w. h. freeman, new york, new york, usa. timmerman, h. r. 2003. the status and management of moose in north america circa 2000-01. alces 39: 131-151. alces(25)_167.pdf alces(25)_1.pdf alces21_17.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces28_203.pdf alces26_142.pdf alces22_439_nbmoosemodel.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces(25)_178.pdf alces27_12.pdf alces(23)_301workshopsessions.pdf alces vol. 23, 1987 alces21_253.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces(25)_63.pdf alces(23)_89.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces29_175.pdf alces(25)_104.pdf alces26_91.pdf alces21_161.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces28_95.pdf alces27_169.pdf alces21_339.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces22_395.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces26_9.pdf alces29_249.pdf alces21_419.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces24_14.pdf alces(23)_221.pdf alces vol. 23, 1987 rodgersar text box alces vol. 23, 1987 alces vol. 23, 1987 alces(25)_36.pdf estimating moose abundance in linear subarctic habitats in low snow conditions with distance sampling and a kernel estimator eric j. wald1,3,4 and ryan m. nielson2 1us fish and wildlife service, yukon delta nwr, po box 346, bethel, alaska 99559; 2western ecosystems technology, inc, 415 w. 17th st., suite 200, cheyenne, wyoming 82001; 3university of wyoming, department of ecosystem science and management, dept. 3354, 1000 e. university ave., laramie, wyoming 82071-3354 abstract: moose (alces alces) are colonizing previously unoccupied habitat along the tributaries of the lower kuskokwim river within the yukon delta national wildlife refuge (ydnwr) of western alaska. we delineated a new survey area to encompass these narrow (0.7–4.3 km) riparian corridors that are bounded by open tundra and routinely experience winter conditions that limit snow cover and depth necessary for traditional moose surveys. we tested a line-transect distance sampling approach as an alternative to estimate moose abundance in this region. additionally, we compared standard semi-parametric detection functions available in the program distance to a nonparametric kernel-based estimator not previously used for moose distance data. a double-observer technique was used to verify that the probability of detection at the minimum sighting distance was 1.0 (standard assumption). average moose group size was 2.03 and not correlated with distance from the transect line. the top semi-parametric model in the program distance was a hazard-rate key function with no expansion terms. this model estimated average probability of detection as 0.70 with an estimated abundance of 352 moose (95% ci = 237–540). the cv for the semi-parametric model was 20% and had an estimated bias of 1.4%. the nonparametric kernel-based model had an average probability of detection of 0.73 and an estimated abundance of 340 (95% ci = 238–472) moose. the cv for the kernel method was 18% and the estimated bias was <0.001%. line-transect distance sampling with a helicopter worked well in the narrow riparian corridors with low snow conditions, and survey costs were similar to traditional surveys with fixed-wing aircraft. the kernel estimator also performed well compared to the standard semi-parametric models used in program distance. our technique provides a viable approach for surveying moose in similar areas that have restrictive conditions for standard aerial surveys. alces vol. 50: 133–158 (2014) key words: alaska, alces alces gigas, distance sampling, kernel, line-transect, moose, population estimate, survey, y-k delta the yukon delta national wildlife refuge (ydnwr) is divided into 4 primary moose (alces alces) survey units along the yukon and kuskokwim rivers. surveys in these units are typically conducted using the geospatial population estimator (gspe) technique (ver hoef 2002, delong 2006, kellie and delong 2006, ver hoef 2008), which is the preferred method adopted by the alaska department of fish and game (adfg) and several other federal agencies including other national wildlife refuges in alaska. only one survey unit is on the lower kuskokwim river within ydnwr, and encompasses nearly 2250 km2 of contiguous habitat along the 4present address: arctic national wildlife refuge, 101 12th ave, rm# 236, fairbanks, alaska 99701 133 relatively wide riparian corridor. the gspe technique overlays a grid of sample blocks on the study area where each block is stratified into high or low moose density based on a previous stratification flight. a random selection of survey blocks in each strata are surveyed using a fixed-wing aircraft to completely search each selected block. the analysis uses the block’s spatial correlation to increase the estimate’s precision based on finite population block kriging (ver hoef 2002). complete and adequate depth of snow cover is essential for this type of survey. ideally surveys are conducted approximately every 3–5 years to monitor trends in moose abundance. however, the yukon-kuskokwim delta and other coastal areas of western alaska experience moderating climatic effects from the bering sea and have frequent thawrefreeze events (1–9 events/winter; wilson et al. 2013). as a result, weather and snow conditions may preclude survey initiation or completion, extending the typical period between surveys. despite adequate habitat to sustain a higher moose population, the lower kuskokwim river has historically had a low moose density (0.03 moose/km2 in 2004; perry 2010) because of extensive hunting pressure (coady 1980). therefore, a moratorium was implemented on the lower kuskokwim river watershed between 2004 and 2009 when the population increased substantially (0.23 moose/km2 in 2008; perry 2010) and expanded into previously unoccupied, or occasionally occupied habitat making it necessary to create an expanded survey unit. the new survey unit was developed to include the major tributaries of the lower kuskokwim river within the ydnwr, which are narrow riparian corridors that originate from the adjacent mountains (fig. 1). these tributaries can support a substantial population and are important wildlife corridors to other parts of ydnwr and neighboring conservation units (i.e., togiak national wildlife refuge; aderman and woolington 2006). the kuskokwim tributary survey unit was first proposed, designed, and partially surveyed in the winters of 2009 and 2010 using the gspe technique; weather and lack of snow cover prevented completion of both surveys. environmental conditions such as snow cover are among the most influential variables that affect survey quality (leresche and rausch 1974, gasaway et al. 1986, quayle et al. 2001, oehlers et al. 2012). the gspe technique recommends that surveys occur after fresh or moderately fresh snow with complete ground coverage (gasaway et al. 1986, kellie and delong 2006), typically ≥20 cm in this area. retrospectively, we questioned the suitability of the gspe technique for these tributaries because of the time and cost required to conduct the survey given the unreliable weather and snow conditions. in addition, we sought to evaluate whether this technique is ideal for use in the narrow linear habitats given that large portions of many survey blocks (∼3.7 km � 4.5 km) included non-moose habitat (i.e., open tundra). the stratified random block design of the gspe is better suited for larger and more contiguous blocks of similar habitat (kellie and delong 2006). a minimum count (termed complete count, a non-sampling approach) survey is used in adjacent areas (aderman 2008) with a fixed-wing aircraft flown throughout the entire area, counting all moose observed; the count is the population estimate (lancia et al. 2005). this method requires more flying time to search all areas completely and the minimum count has neither an estimate of precision (i.e., confidence interval) nor typically a sightability correction factor (gasaway and dubois 1987). simple aerial strip-transect surveys require less flying than minimum counts and incorporate an estimate of precision (timmermann 1974, timmermann and buss 2007, jung et al. 134 estimating moose abundance – wald and nielson alces vol. 50, 2014 2009); however, this method assumes equal detection of animals from the centerline out to the designated strip width (burnham and anderson 1984). typically there is no estimate of sightability with strip transect sampling (evans et al. 1966, timmermann fig. 1. the yukon delta national wildlife refuge encompasses the yukon-kuskokwim delta in western alaska. bethel is the main community along the kuskokwim river. the four main tributaries of the lower kuskokwim river include the tuluksak, kisaralik, kwethluk, and eek rivers which form the study area. these rivers are characterized by narrow riparian corridors bounded by open tundra. alces vol. 50, 2014 wald and nielson – estimating moose abundance 135 1993), although sightability could be estimated with marked animals (anderson and lindsey 1996), or more intensive flying at an increased cost (gasaway et al. 1986, gasaway and dubois 1987). we determined that a viable survey method was line-transect distance sampling (burnham et al. 1985, buckland et al. 2001) using a helicopter for several reasons: 1) the area tends to have marginal snow cover each year making it difficult to complete a gspe, 2) a helicopter can fly lower and more slowly with better visibility than a fixed-wing aircraft, helping to compensate for minimal snow cover, 3) line-transects can “fit” in the narrow riparian corridors better than gspe blocks that typically encompass large portions of non-moose habitat, 4) distance sampling incorporates sightability corrections (e.g., weather, lighting, snow conditions, observer experience) provided that probability of detections at some distance is known or assumed and, 5) we expected time, logistics, and costs may be similar compared to a fixed-wing gspe survey in the same region. thompson (1979) initially applied a distance sampling approach to estimate moose abundance in ontario, canada, where it was later improved upon by dalton (1990). thompson (1979) had problems fitting detection functions, and both surveys had difficulties meeting some of the sampling assumptions (e.g., exact distance measurements, movement of animals before detection, sightings not always independent) and were limited to the technological and statistical challenges of that period (gasaway and dubois 1987, pollock and kendall 1987, dalton 1990). significant advances in distance sampling methodology and statistical analysis have been recognized over the last 30 years (buckland et al. 2001, 2004; thomas et al. 2010), and these improvements led to the development of distance sampling protocol for moose in interior alaska (nielson et al. 2006). although distance sampling has been used successfully to estimate moose abundance in relatively large contiguous blocks of boreal forest habitat with adequate snow conditions, no study has demonstrated that this technique works in subarctic tundra along small, narrow riparian corridors. distance sampling analyses typically involve estimation of semi-parametric detection functions (buckland et al. 2001, thomas et al. 2010). during early development and analyses of line-transect distance data, burnham et al. (1980) suggested that other nonparametric methods such as kernel estimators or splines might prove “fruitful” for estimating probability of detection, and mack and quang (1998) further suggested that kernel methods could be a viable technique in wildlife distance sampling. the nonparametric kernel density estimator does not assume an underlying distribution for the detection function, and thus has more flexibility by allowing the data to “speak for themselves” or dictate the shape of the detection function (silverman 1986, wand and jones 1995). kernel estimators are considered a robust alternative to other density function estimators (chen 1996a) and are computationally more efficient than polynomials (buckland 1992). both kernel and semi-parametric methods are robust against changing detection functions during a survey (gerard and schucany 2002) and are resilient to changing survey conditions such as snow depth/coverage, sun angle and overcast skies, and wind or other environmental conditions that could change during a survey over time and space (burnham et al. 1980, chen 1996b); it is assumed that no correlation exists between moose density and these changing conditions. popular computer programs such as distance 6.0 do not include a kernel-based detection function (thomas et al. 2010) for use in analysis of linetransect data, although the kernel method 136 estimating moose abundance – wald and nielson alces vol. 50, 2014 has been used for distance data in other types of surveys (buckland 1992, chen 1996a, mack and quang 1998, gerard and schucany 2002, jang and loh 2010, nielson et al. 2013, nielson et al. 2014), but not for moose. the objectives of this study were twofold: 1) evaluate helicopter-based aerial line-transect distance surveys with a doubleobserver modification to obtain an estimate of moose abundance within narrow riparian corridors during a low snow year, and 2) compare the nonparametric kernel-based detection function to the more traditional semi-parametric models in the program distance (thomas et al. 2010). we investigated what we presumed was a viable alternative to traditional moose survey methods for areas with environmental conditions that preclude traditional surveys. study area the yukon delta nwr is located in western alaska and encompasses the delta formed by the yukon and kuskokwim rivers which empty into the bering sea (fig. 1). the kuskokwim tributary survey unit includes parts of 4 main lower kuskokwim river tributaries originating from the mountains to the south and east. these tributaries include the eek, kwethluk, kisaralik, and tuluksak rivers and are characterized by narrow (0.7–4.3 km wide) riparian corridors (fig. 2) running through the foothills and tundra flats that drain the northwest sides of the eek and kilbuck mountains. fig. 2. tributary rivers within the study area are characterized by narrow riparian corridors bounded by open tundra. the relatively open forest and shrub habitat is conducive to sighting moose during a line-transect survey with a helicopter. this corridor is a part of the kwethluk river and is approximately 800 m wide. alces vol. 50, 2014 wald and nielson – estimating moose abundance 137 the eek and upper kwethluk rivers are represented by open riparian shrubs (salix spp. and alnus spp.) and scattered clumps of balsam popular (populus balsamifera), whereas the lower kwethluk river transitions to open forests that include sporadic mixing of spruce (picea glauca), balsam popular, and birch (betula papyrifera) as the overstory with an understory of open willow and alder. the tuluksak river riparian zone is primarily a narrow corridor of spruce and birch with an understory of willow and alder. the kisaralik river riparian zone is mostly mixed coniferous open woodland which exhibits a moderate transition between the kwethluk and tuluksak riparian habitats. all 4 tributary habitats are bounded by tundra and include variously sized open meadows, old river channels, and beaver ponds. weather conditions are highly variable across the survey area. average temperatures and snow depth at bethel, alaska airport (2000–2010; noaa 2011) during the primary survey months were −21 °c (range −36 to 4 °c) with 23 cm (0–56 cm) of snow in january, −10 °c (−37 to 5 °c) and 23 cm (0–64 cm) of snow in february, and −9 °c (−27 to 4 °c) with 20 cm (0–56 cm) of snow in march. in many years there are freeze-thaw events (wilson et al. 2013) throughout the winter which ultimately limit total snow accumulation and duration. our study period (2010) was an el niño year which affected the winter weather pattern on the yukon-kuskokwim delta from june 2009 to march 2010 (noaa 2013). repeated high pressure systems over the delta kept numerous low pressure systems south and subsequently pushed easterly, resulting in unusually clear and dry conditions with periods of colder temperatures and little snowfall over the study area during winter 2009–2010. a portion of the survey unit experiences a “banana belt” effect, especially along the foothills between the kwethluk and kisaralik rivers. this area is usually affected by a warming trend that typically melts snow more frequently and rapidly than other parts of the area, perhaps resulting from an inversion. these conditions can limit gspe surveys along the lower kwethluk and kisaralik rivers in any given year. nearly 3 cm of new snow accumulated 4 days prior to the survey. total snow depth was 5 cm (bethel airport) but was 8–10 cm at 2 snow stakes in the study area (kwethluk river) during the survey. snow coverage ranged from 85–100% throughout the survey area with meadow grasses and short vegetation protruding and snow melted off stumps and root wads. weather conditions during this survey were mostly clear, 9–37 kph winds, and −12 to 2 °c. day length was nearly 12 h with sunrise at 0900 hr and shadows becoming long at about 1500 hr. survey times were typically between 0900 and 1700 hr each day and flying conditions were generally favorable during the entire survey. methods field survey aerial line-transect distance sampling protocol for moose followed nielson et al. (2006). the survey area (i.e., sampling universe) was limited to the riparian corridor for the rivers of interest. polygons were created around rivers to encompass riparian vegetated areas within the floodplain and between the tundra benches on each side of the river (arcmap 9.2, environmental systems research institute, redlands, california, usa; fig. 1). satellite imagery was used to facilitate creation of survey areas which encompassed nearly 730 km2. survey transects were created within river corridors and varied by length and number along each river (fig. 1). multiple transects were placed in areas wide enough to allow equidistance spacing of 700 m, 138 estimating moose abundance – wald and nielson alces vol. 50, 2014 providing a maximum 350 m search area on each side of a transect centerline. transect length varied according to stretches of river that allowed straight transects with sections changing direction in a saw-toothed manner as the river corridor meandered (nielson et al. 2006). some riparian corridors were sufficiently narrow to allow only a single transect which had a random start point contingent on allowing the minimum half transect width (350 m) to be in moose habitat. the centerline could not be on the edge of the habitat (i.e., one side having a hard boundary of no moose habitat or open tundra and the other side all moose habitat) to avoid extreme asymmetry of g(y), although this source of bias is minimal in most studies (buckland et al. 2001). areas with systematic parallel transects had a random start point for the first transect. a total of 46 transects were delineated with a combined length of 698 km (fig. 1). we used a robinson (r-44) helicopter with bubble windows to survey moose during 16–17 march 2010. helicopters provide a better platform for observing moose because of better sightability, smaller variances, and often comparable cost with fixed-wing surveys (smits et al. 1994, gosse et al. 2002). protocol recommends a flight altitude of 122 m above ground level (agl) which results in good visibility and minimal disturbance of moose (nielson et al. 2006). however, snow conditions were poor, so flight altitude was adjusted to 100 m agl to increase visibility while remaining high enough to minimally affect moose and prevent ground flash, the visual effect of ground zooming by too fast when flying at a lower altitude (becker and quang 2009). our target ground speed was 64 kph (40 mph) depending on terrain and wind. four people were onboard during this survey. the pilot was responsible for maintaining desired altitude, speed, and heading on transect centerlines using a preprogramed gps (garmin 695). two observers were seated in the back (one on each side) and were the primary observers for the survey. their responsibilities were to sight moose groups, count and classify each group, and measure the perpendicular distance from the transect centerline to each group centerpoint. the data recorder sat in the front-left seat and was responsible for recording all data including gps locations, performing as a double-observer, frequently measuring helicopter agl, and overall survey coordination. the front observer recorded survey data while flying off transect in order to not interfere with double-observer duties while flying on transect. the double-observer method was used in conjunction with the line-transect survey to test the assumption that detection was 100% on or near the transect centerline, or at the minimum available sighting distance (buckland et al. 2001, laake and borchers 2004, borchers et al. 2006). this assumption is sometimes violated (chen 1999, 2000) and information regarding heterogeneity in observer bias should be modeled (graham and bell 1989) because it can produce negatively biased estimates. the data recorder in the front-left seat was paired with the rear left observer to conduct the double-observer sampling, which is essentially a mark-recapture method (borchers et al. 2006). the data recorder focused on or near the centerline to detect moose, but recorded all moose observations at any distance. double-observer data are used to estimate detection rate on or near the centerline by the rear seat observers. this requires that the front and rear seat observers operate independently of each other (buckland et al. 2010). data were recorded on the number of moose groups detected by both observers, and groups detected by the front and not the rear observer. to account for observer bias, the 2 rear observers rotated sides each day to be paired with the front-left observer alces vol. 50, 2014 wald and nielson – estimating moose abundance 139 to incorporate biases from both observers into the model (cook and jacobson 1979). thus, we considered the probability of detection estimated for the back-left observer based on the mark-recapture data to be relevant for both backseat observers during analyses. the recorder also worked with the pilot to keep flight speed and altitude within the range of survey protocol. a laser rangefinder (nikon forestry 550 hypsometer) was frequently used to measure true vertical distance from the ground to helicopter every 2–3 min to check flight altitude and recommended adjustments as needed. moose groups were defined as one or more moose within a 50 m radius (molvar and bowyer 1994). the distance of groups perpendicular to the transect centerline was measured by the rear observer with a laser rangefinder (leupold rx-1000 tbr) with a built-in clinometer that had maximum range of ∼900 m; clinometers allow for accurate horizontal measurements regardless of survey altitude. a group that was hard to laser-range (e.g., trees, helicopter movement, animal movement) required flying back over the group and marking its location with a gps (marques et al. 2006). distance was measured to the center of a group using the laser rangefinder directed at their feet to avoid over-estimating the distance; this measurement was associated with the location of the group when first observed. if moose moved before a distance was acquired, the observer ranged the location of the initial observation. doubling back to mark gps locations worked well, but some moose moved after detection because of the aircraft hovering directly overhead; tracks in the snow proved reliable as reference points for these cases. the distances measured with the gps method (∼90% of all groups observed) were calculated in a gis. additional moose observed “off transect” while doubling back to obtain gps locations were not included in any analysis. observers determined group size, composition (i.e., adults and calves), and classified percent habitat cover for ∼50 m radius around each group. data analysis standard distance sampling theory assumes all individuals (objects) available to be detected on the centerline, or the minimal available sighting distance, are observed, and that the probability of detection is a function of perpendicular distance from the centerline. there are 3 essential assumptions for accurately estimating density using distance sampling. in order of importance these are: 1) objects at the minimal available sighting distance are detected with certainty, that is g(w1) = 1.0, or can be estimated, 2) objects are detected prior to any movement in response to the survey, and 3) perpendicular measurements to the object are accurate (buckland and turnock 1992, buckland et al. 2001). other design/ analysis assumptions exist but are less stringent, including accurate measurement of group size and that object density is independent of the placement of transects (i.e., uniform distance distributions; fewster et al. 2008). fulfilling these assumptions allow for an accurate density estimate using: d̂ ¼ nêðsþ 2ðw2 � w1þlp̂ ð1þ where n is the number of observed groups, êðsþ is the expected (or average) group size, w1 and w2 are the minimum and maximum search distances from a transect, respectively, l is the total length of transects flown, and p̂ is the estimated average probability of detection within the area searched (buckland et al. 2001). we tested the assumption g(w1) = 1.0, where g(w1) is the minimum sighting 140 estimating moose abundance – wald and nielson alces vol. 50, 2014 distance with the double-observer technique (chen 2000, borchers et al. 2006, buckland et al. 2010) using observations from the left side of the helicopter. observations collected independently by individual observers on the left side were used to estimate the probability of detecting a moose group at the minimum available sighting distance. this probability was used to adjust the estimated detection curve starting at that distance (laake and borchers 2004). we used logistic regression (mccullagh and nelder 1989) in the mark-recapture analysis to estimate the probability of detecting a moose group by the back-left observer given detection by the front-left observer. we considered 3 models that 1) treated the probability that a group was detected by the back-left observer as constant across all distances from the transect line (intercept only model), 2) treated the probability of detection as a function of distance from the transect line, and 3) included both linear and quadratic terms for distance from the transect line. we used akaike’s information criterion for small sample sizes (aicc; burnham and anderson 2002) to identify the best model for estimating probability of detection by the backseat observers based on the markrecapture data. the aicc was calculated as: aicc ¼ �2logðlikelihoodþ þ 2kn/ðn�k�1þ ð2þ where k was the number of parameters in the model (including intercept term), n was the number of observations used to fit the model, likelihood was the value of the logistic likelihood evaluated at the maximum likelihood estimates, and ‘log’ was the natural logarithm. the logistic regression model was fit using the program r (r development core team 2010). we designed transect centerlines to be a minimum of 700 m apart to ensure that moose groups were not counted more than once if they moved during the survey. to meet this assumption, we set the maximum search width, w2, equal to the maximum distance a moose group was observed within 300 m of the transect centerline. since the backseat observers had a blind spot underneath the helicopter, we used a laser rangefinder (hypsometer) to determine the minimum sighting distance for the backseat observers. to determine the width of the blind spot at the survey altitude, the backseat observer laser-ranged through the bubble window along their line-of-sight to the ground, just clear of the helicopter skid. the minimum available sighting distance for backseat observers was ∼43 m from the centerline. moose were visible to the frontleft observer from 0–43 m through the front helicopter window, but lumping these data into a single distance of “zero”, and the fact that the front-right observer (pilot) was not focused on observing moose on that side of the line, precluded using these data in the analyses; therefore, the front-left observer observations in the 0–43 m range were not used. thus, the minimum available sighting distance, w1, was set at the minimum distance at which a moose group was detected by a backseat observer. we evaluated whether correlation existed between expected group size and detection distance because group size can influence detectability, especially at longer distances; larger groups may have a higher probability of detection than smaller groups further from the transect line (drummer and mcdonald 1987, drummer et al. 1990). we used a pearson’s correlation analysis to estimate the correlation (r) between group size and distance from the transect line, and calculated a 95% confidence interval (ci) for the statistic (zar 1999). we determined no relationship existed between group size and detection distance if the 95% ci included 0.0. in this situation, we used the average alces vol. 50, 2014 wald and nielson – estimating moose abundance 141 of all observed group sizes for êðsþ (equation 1). if a correlation was detected, we used the regression method (buckland et al. 2001) to estimate expected group size. we examined the habitat covariate, percent cover as a potential influence on the probability of detection of groups (i.e., higher percent cover may reduce probability of detection; anderson and lindzey 1996, oehlers et al. 2012). the underpinning of distance sampling is the detection function g(y) which expresses the probability of detecting a group given that the group was observed at distance (y) from a random transect, and that the assumption g(w1) = 1.0, or can be estimated, holds true (buckland et al. 2001). there are many models that can be fitted to distance data in order to estimate the shape of a detection function, and we used the computer program distance 6.0 release 2 (thomas et al. 2010) to model semi-parametric detection functions for moose groups. we considered robust key functions with expansion terms as outlined in buckland et al. (2001), including the half-normal with hermite polynomial and cosine expansion terms, the hazard-rate with a cosine expansion, and the uniform with simple polynomial and cosine expansion terms. additionally, the half-normal or hazard-rate key functions allow for predictor variables to help model the detection function. we used the half-normal (including hermite polynomial and cosine expansion terms) and hazard-rate (including a cosine expansion term) key functions for modeling the detection function while incorporating the percent cover variable. the number of expansion terms for each key function was allowed to vary from 0–5; the aicc was used to select the number of expansion terms among the various models. the model with the lowest aicc value was selected as the best model to describe the detection function (burnham and anderson 2002). use of parametric or semi-parametric detection functions may not always be the best approach to fit probability detection curves (burnham and anderson 1976, buckland 1992); instead, a nonparametric kernel density estimator without an assumed probability density function may provide a better fit to the data. we fit a nonparametric kernel estimator (silverman 1986, wand and jones 1995) to our group observations, and used the general univariate kernel density estimator described in wand and jones (1995): f̂ xð þ ¼ nhð þ�1 xn i¼1 k x � xi h � � ð3þ where x is a perpendicular distance within the range of observed distances, xi is one of the n observed distances, h is a smoothing parameter, or ‘bandwidth’, and k is a kernel function satisfying the conditionð k xð þdx ¼ 1. since the bandwidth (h) governs the function smoothness (chen 1996a, gerard and schucany 1999), the choice of bandwidth is more crucial than the choice of kernel (mack and quang 1998, jang and loh 2010). we used a gaussian kernel function (silverman 1986, chen 1996a) and the direct plug-in bandwidth selection method (sheather and jones 1991, wand and jones 1995, sheather 2004) to develop the detection function for groups. the direct plug-in method objectively fits the bandwidth and is considered to be the best compromise between bias and variance among the available methods (sheather and jones 1991, wand and jones 1995, venables and ripley 2002, sheather 2004). kernel estimators inherently do not perform well near sharp boundaries (jang and loh 2010). a boundary bias is created, as in our case, when distance observations are not distinguished from the right or left side of the transect line and where all values are 142 estimating moose abundance – wald and nielson alces vol. 50, 2014 non-negative (buckland 1992, jang and loh 2010). in order to model the distances with a kernel estimator, chen (1996a, 1996b) and silverman (1986) recommended reflecting the observed distances to both sides of the transect line in order for the kernel density estimator to perform properly. after shifting all observed distances by the left-truncation distance (w1), we multiplied (reflected) the observed distances by (−1) and added them to the dataset (buckland 1992, chen 1996a, 1996b). once the kernel density function was created from the expanded dataset, the detection function to the right of the zero line (positive) was used for the density estimate. the kernel estimator was fit using the program r (r development core team 2010) and the mass package in r (venables and ripley 2002). we used bootstrapping to estimate ses and 95% cis for final estimates of moose density and abundance within the sampled region (efron 1981a, 1981b, quang 1990, diciccio and efron 1996). estimates derived from the program distance 6.0 were bootstrapped within the program, which uses a default of 999 bootstrap re-samplings with replacement (thomas et al. 2009). standard errors and confidence intervals for the kernel estimates were derived from bootstrapping 999 resamples (with replacement) to be consistent with program distance and because the bootstrapped estimates (se, ci) usually become stable and asymptotic between 500 and 1000 resamples (efron and tibshirani 1994). we bootstrapped the 46 line-transects surveyed in which the bootstrap would rerun the analysis for all parameter estimates, including the shape of the detection function and the average probability of detection during each iteration. additionally, we evaluated bias and precision of the density estimate using the bootstrap (efron and tibshirani 1986). we used the percentile method (efron 1981b, 1982) for calculating the 95% confidence intervals using the 2.5th and 97.5th percentiles of the 1,000 estimates (999 bootstrap estimates + original estimate). the percentile method is the preferred method for calculation of cis when bootstrapping, because using the standard formula (i.e., estimate ± 1.96[se]) requires the additional assumption that the bootstrap estimates generally follow a normal distribution (buckland 1984, efron and tibshirani 1994). we estimated relative percent bias of the density estimates as: %bias ¼ dboot � dorig � � dorig � � � 100 ð4þ where dboot is the average density estimate from the bootstrap and dorig was the original density estimate. we measured dispersion or the extent of variability in relation to the final density estimate by calculating the coefficient of variation (cv) as cv ¼ ðse=d̂þ � 100%. the standard deviation (sd) of the 1,000 estimates was used as the estimated se. in order to estimate the total length (l) of transects needed in future surveys to achieve a certain level of precision (i.e., cv value), we used the formula from buckland et al. (2001): l ¼ l0 cvðdþf g2 . cvtðdþf g2 ð5þ where l0 is the total length of transects surveyed, cv (d) is the coefficient of variation of the density estimate from this study and cvt (d) is the desired target coefficient of variation. results a total of 162 moose (112 adults, 48 calves) in 78 groups were detected on 46 transects along 698 km within a series of polygons encompassing 730 km2 of riparian moose habitat. there were 37 cow moose with calves including 26 singletons and 11 alces vol. 50, 2014 wald and nielson – estimating moose abundance 143 sets of twins (30% twinning rate). group size ranged from 1–5 with 73% of groups comprised of 1–2 moose and only 6% with 4–5 moose. because the mark-recapture portion of this study was intended to occur only on the left side of the aircraft, a single group observation detected by the front-left observer that was 244 m to the right of the transect line (and not detected by the back-right observer) was excluded from the analysis. ten of the 78 groups were detected only by the front-left observer and were recorded as seen directly on the transect line (i.e., perpendicular distance = 0). since these groups could not be seen by the back-left observer and ‘lumping’ of the perpendicular distances occurred during data recording (i.e., these moose were likely somewhere ± 43 m from the transect line and not all directly on the line), these observations were also excluded from the analysis. of the remaining observations, the minimum observed distance of a moose group by the backseat observers was 46 m from the transect line, thus w1 was set to 46 m. we truncated our data to 300 m, which corresponded to a reduction of approximately 8% of the farthest distance observations; buckland et al. (2001) recommend truncating the farthest 5–10% of distance observations. the maximum observed distance of a group within 300 m of a transect line was 299 m, so w2 was set at 299 m. analysis of moose observations within the defined search width (46–299 m from a transect line) indicated that group size was not correlated with distance from the transect line (r = 0.048, 95%, ci = −0.20–0.29). the average group size by the backseat observers within the search strip was 2.03 (95% ci from 1.78–2.32) and was used in the density estimate. the mark-recapture trials used 34 observations to fit logistic regression equations to estimate the probability of detection by the back-left observer given detection by the front-left observer. only 3 groups were missed by the back-left observer and these were deleted for the density estimate. the logistic equation with linear and quadratic terms for distance from the transect line had the lowest aicc value (19.8 versus 22.4 and 24 for the intercept only and linear distance function models, respectively). the final estimated logistic regression model was: e yi½ � ¼ exp 51:461 � 0:583distancei þ 0:0012distance2i � 1 þ exp 51:461 � 0:583distancei þ 0:0012distance2i � ð6þ where e [yi] was the expected probability of detection for mark-recapture observation i. based on this final model, the predicted probability of detection of moose at the minimum sighting distance by the back-left observer was 1.0. therefore, the estimated probability of detection curve was not scaled by a correction factor prior to integration and estimation of p̂, and only observations by the rear seat observers were used to estimate moose density. comparison of models using aicc values requires that the competing models are all estimated using the same number of observations and the same response (y) values. percent cover was not recorded for one observation so this record was not initially included during estimation of the probability of detection curve. in addition, due to the distribution of percent cover values for observations (10–70% in 10% increments) and few observations at the extremes, the original values were collapsed into 3 categories: 1) 10–30, 2) 30–50, and 3) 60–70%. comparison of the models with and without the covariate for percent cover indicated that a hazard-rate key function with no expansion terms was the best fit to the data (table 1). because the models 144 estimating moose abundance – wald and nielson alces vol. 50, 2014 containing the covariate for percent cover ranked last according to aicc values, we refit the models without the predictor variable using all the observations from the rear seat observers within 46–299 m of a transect line, including the single observation where percent cover was not recorded. the results of this analysis were similar to the analysis, minus the missing observation, in that the top model was a hazard-rate key function with no expansion terms (table 2). we used the goodness of fit (gof) test statistic to determine if the top model fit the data well (buckland et al. 2001). there was no evidence of lack of fit for the top model (gof test; χ2 = 6.07, df = 4, p = 0.194). based on this final model using 59 groups, the estimated average probability of detection was 0.70 (fig. 3), and the estimated density was 0.48 moose/km2 or a population of 352 moose (95% ci = 237–540); this model had a cv of 20% and an estimated bias of ∼1.4%. the detection function calculated using the kernel density estimator (without covariates) also had a good fit (gof test; χ2 = 6.42, df = 5, p = 0.73) with an estimated average probability of detection of 0.73 (fig. 3). the estimated density was 0.47 moose/km2 or a population of 340 moose (95% ci = table 1. estimated semi-parametric detection functions fit to 58 moose group observations using the program distance (thomas et al. 2010), including the number of expansion terms, whether the covariate for percent cover was included in the model, the number of parameters (k), aicc value, and estimated average probability of detection (p̂), estimated moose density (d̂), and coefficient of variation (cv) for each model, alaska, 2010. key function expansion expansion terms k % cover (yes/no) aicc p̂ d̂ % cv hazard-rate cosine 0 2 n 627.85 0.71 0.480 19.1 uniform cosine 2 2 n 629.26 0.67 0.505 25.3 uniform simple polynomial 1 1 n 629.47 0.70 0.485 17.3 half-normal cosine/hermite polynomial* 0 1 n 629.48 0.62 0.547 21.0 half-normal cosine 1 4 y 631.40 0.85 0.398 18.2 hazard-rate cosine 0 4 y 631.42 0.68 0.498 18.4 half-normal hermite polynomial 1 4 y 631.68 0.82 0.413 18.4 *no expansion terms were selected using aicc values. table 2. estimated semi-parametric detection functions fit to 59 moose group observations using the program distance (thomas et al. 2010), including the number of expansion terms, whether the covariate for percent cover was included in the model, the number of parameters (k), aicc value, and estimated average probability of detection p̂ , estimated moose density d̂, and coefficient of variation (cv) for each model, alaska, 2010. key function expansion expansion terms k aicc p̂ d̂ % cv hazard-rate cosine 0 2 643.25 0.70 0.482 20.0 uniform cosine 1 1 643.53 0.60 0.562 19.1 half-normal cosine/hermite polynomial* 0 1 643.67 0.64 0.533 20.4 uniform simple polynomial 1 1 644.47 0.72 0.472 17.4 *no expansion terms were selected using aicc values. alces vol. 50, 2014 wald and nielson – estimating moose abundance 145 238–472); the corresponding cv was 18% and the estimated bias based on bootstrapping was <0.001%. based on an encounter rate of 0.085 moose groups/km (59 groups/698 km) and the cv from the kernel estimator (18%), we calculated the total length of transects needed to achieve a targeted cv of 20, 15, and 10%. this analysis indicated that transect lengths of 566, 1006, and 2262 km, respectively, were required to meet these targeted cvs assuming that the encounter rate remains constant. discussion others have utilized distance sampling with varying degrees of success in model fitting and achieving adequate levels of estimate precision under adequate snow conditions within boreal transition forest of westcentral alaska (nielson et al. 2006) and central canadian boreal forest habitats (thompson 1979, dalton 1990, thiessen 2010, peters et al. 2014). we evaluated this method as an alternative technique for surveying moose in a subarctic tundra ecosystem with variable and often minimal snow conditions, and present an alternative technique for analyzing distance data using a nonparametric kernel density estimator to fit the detection function. overall, distance sampling proved to be a viable technique to monitor moose along the kuskokwim fig. 3. histogram of the 59 moose group distance observations with corresponding detection functions superimposed. the final hazardrate detection function was fit using the program distance (thomas et al. 2010). the non-parametric kernel-based detection function was fit using the program r (r development core team 2010). perpendicular distances were shifted left by 46 m prior to analysis, but shifted back for graphing visual clarity. 146 estimating moose abundance – wald and nielson alces vol. 50, 2014 tributary rivers in the ydnwr in southwestern alaska. assumptions distance sampling depends on 3 main assumptions that need to be met, or nearly so, in order to produce unbiased and reliable estimates (buckland et al. 2001). although the assumptions can be relaxed to some degree in certain circumstances and still provide dependable estimates (buckland et al. 2001), we designed our study in an attempt to meet all assumptions or to estimate our biases if we failed to adequately meet one. assumption (1), that objects at the minimum available sighting distance were detected with certainty, is typically addressed by developing a sightability correction factor (scf) through modeling covariates against mark-recapture data of collared animals (gasaway et al. 1986, samuel et al. 1987, anderson and lindzey 1996). distance sampling inherently accounts for, or corrects for, visibility (perception) biases provided that all objects are detected on the transect centerline or at the minimum sighting distance (buckland et al. 2001, marques and buckland 2004). although many distance sampling studies do not address assumption (1) (bachler and liechti 2007), we utilized the doubleobserver method (graham and bell 1989) to test the primary assumption that detection of moose groups on the centerline or at the minimum sighting distance was certain. the mark-recapture logistic regression analysis indicated that we met the assumption with 100% probability of the backseat observer detecting a moose group that was detected by the front-left observer at the minimum sighting distance (g(w1)). detection certainty along the transect line in our study was enhanced by several factors including that the survey area was within narrow riparian habitat with a relatively open canopy (range of percent cover data was 10–70% with an average of 39%). additionally, detection on the centerline was further increased by the fact that we used a helicopter flying at 100 m agl at a relatively slow speed of 48–88 kph; although snow was shallow, visibility on the centerline was excellent. moose in our study area were assumed to be available for detection because of the relatively open habitat (39% canopy closure). one concern was that root-wads of wind-fallen trees could hide an adjacent, bedded moose. however, if we flew directly over a root-wad we detected the moose on the transect line (g(w1) = 1.0), and it would be accounted for in the detection function at distances off the transect line (laake et al. 2008). availability bias could possibly be removed or reduced if the area was surveyed at different times (e.g., hours apart) to allow animals to become available; this is likely dependent on species (laake and borchers 2004) and unlikely of concern with moose in most conditions. assumption (2), that objects are detected at their initial location, is sometimes difficult to assess (fewster et al. 2008), and in our case, related to disturbed moose “pushed” by the helicopter some distance before detection. thompson (1979) reported that moose were not disturbed by circling fixed-wing aircraft and did not move as the airplane approached; however, cumberland (2012) reported that moose usually initiated some movement from a larger turbine helicopter (bell 206) at a lower survey altitude (60 m agl). random movement is acceptable as long as it is not caused by the observer (buckland et al. 2001), but failure to meet assumption (2) would bias the density estimate low. we investigated the validity of this assumption, in part, by reviewing the distance data histogram and identifying whether a bump or peak is evident some distance from the center line. figure 3 has a slight alces vol. 50, 2014 wald and nielson – estimating moose abundance 147 bump in frequency at ∼150 m which might reflect recording bias or movement from the helicopter prior to detection (dalton 1990). our front observer was focused on the centerline for the double-observer method, and looking forward of the helicopter to identify any pre-detection movement (fewster et al. 2008). few moose were observed moving prior to when the backseat observers detected moose; movement was mainly in response to the helicopter directly overhead. because these few moose did not move into a zone of detection for the backseat observer, movements were considered moot in our study. further, observations of moose within the effective search width (46–299 m) did not indicate movement prior to detection by the backseat observer. a possible explanation for the bump in our data was that the backseat observers had a comfortable scanning level or sight line (i.e., distance) while sitting in the helicopter. observer fatigue increases as a survey progresses (briggs et al. 1985, schroeder and murphy 1999), possibly contributing to the desire to scan (subconsciously) at a less strained position. for example, jang and loh (2010) graphed the classic wooden stake data outline in burnham et al. (1980) and showed that their histogram had a large bump or spike of detections that clearly was not associated with movement. we believe that we met assumption (2) because the front observer scanned forward of the flight path and there was no change in observation rate moving from the center line (fig. 3). flying at a higher altitude (e.g., 122 m instead of 100 m agl). as suggested by nielson et al. (2006), would presumably further address assumption (2). to meet assumption (3), that perpendicular distance measurements are exact, we utilized both rangefinders with built-in clinometers and gps units to determine perpendicular distances to moose groups. although both methods can be accurate and efficient, it was difficult to range groups at times, especially when close to the helicopter which required a quick response, and when in cover. higher quality or industrial-type laser rangefinders should reduce this problem. we eventually adopted the technique used by marques et al. (2006) and used gps locations to measure distances. this required more flying time and effort to fly off transect to identify the group location. because this process caused some moose to move prior to marking the location, we followed tracks in the snow to ascertain the initial location. increasing flight altitude while off transect may alleviate or reduce disturbance while marking locations. we are confident that movements were minimal and measured adequately, and that groups were not double counted on subsequent transects. other assumptions that are not generally discussed in literature, such as the uniformity of the distance distribution and independence of group observations, are typically addressed during the survey design process. the uniformity assumption is addressed by randomly distributing transect lines across the study area, or systematically arranging transects with a random start point, as we did (fewster et al. 2008, jang and loh 2010). the assumption that observations are independent is addressed in the same manner as distance uniformity (buckland et al. 2001) provided that moose are not clustered together in one part of the survey area. estimates of density are robust to the independence assumption especially when bootstrapping to obtain confidence intervals (thomas et al. 2002). additionally, we did not incorporate “dependent” moose groups observed while flying off transect when obtaining gps locations for groups that moved. detection functions survey design and protocol are paramount for meeting the 3 primary assumptions 148 estimating moose abundance – wald and nielson alces vol. 50, 2014 of distance sampling in order to model detection functions reliably (thomas et al. 2010). our survey transects were systematically distributed throughout the riparian corridors with random start points allowing for statistical inference (fewster et al. 2009). although transects were spaced 700 m apart and a maximum search width of 350 m could have been used in the analysis, there were few observations beyond 300 m. restricting the analysis to observations within 300 m reduced the possibility that moose groups were counted more than once. we dropped nearly 8% of the farthest observations which was within the recommended 5–10% range suggested because outliers may have undue influence on the shape and scale of the detection function (buckland et al. 2001). our effective search width (w1 and w2) was 253 m (46–299 m) which was narrower than the 700 (dalton 1990) and 800 m (thiessen 2010) widths used in canada, and similar to the 250 m search width used by thompson (1979). our search width was narrow because of the narrow corridor of habitat, we flew relatively low, and the relatively poor snow conditions which presumably reduces visibility. we recommend using narrow transect search widths during low or poor snow condition years to increase effectiveness of the survey. we measured percent cover because covariates can improve model precision by accounting for heterogeneity in the data (buckland et al. 2004, marques and buckland 2004), but at an added cost of sample size (giudice et al. 2012). however, it did not improve the detection function based on the higher aic values for the models including this covariate. this might reflect the relatively narrow range of values (range = 10–70%, median = 40%) that were lumped into 3 categories. visual obstruction by vegetation influenced detection of moose in minnesota where percent cover was higher (range = 0–95%, median = 60%) (giudice et al. 2012). buckland et al. (2001) recommend 60–80 observations and at least 10– 20 replicate transects to obtain reliable estimates with relatively good precision; we had 58 group observations with percent cover measurements. seddon et al. (2003) improved their survey precision (i.e., decreased cv values) by increasing observations. in an analysis of several surveys, thiessen (2010) found a strong relationship between the number of observations and cvs for those surveys; surveys with 60 observations had a cvof ∼20%. our survey (59 observations) corroborates this relationship as our model without covariates had a cv of 20% (using the hazard-rate key function). adding covariates or stratifications would require considerably more observations to ensure reliable estimates with a cv of 20%. we examined the possibility of stratifying our study area by each tributary river but moose density was too low to acquire sufficient observations for reliable estimates for all rivers, with the possible exception of the kwethluk river (∼60–80 observation in each strata; buckland et al. 2001). group size influences detection at distance (drummer et al. 1990), and can be included as a covariate in the program distance (laake et al. 2008). we investigated whether a correlation existed between group size and detection distance prior to “penalizing” our analysis with an additional covariate (giudice et al. 2012). our analysis indicated that group size was not correlated with detection distance in this study which most likely reflects the group composition in the study area. most groups were relatively small (73% had 1–2 moose) and most observations were cows with calves. since there was no correlation between distance and group size, and the composition of groups had a narrow range of sizes (no major outliers), we used the average group size as equivalent to the expected group alces vol. 50, 2014 wald and nielson – estimating moose abundance 149 size in the density estimate (buckland et al. 2001). our model choices were based on recommendations of buckland et al. (2001) and past experience (nielson et al. 2006) in order to prevent a “shotgun” approach to modeling. models that included the percent cover covariate were ranked last and did not contribute to or improve the model according to aicc; we subsequently removed the covariate and analyzed the data without it. our top model was the hazard-rate key function with no expansion terms. as in our study, several other ungulate studies reported the hazard rate key function with various expansions to be the top models (focardi et al. 2002, shorrocks et al. 2008, young et al. 2010, schmidt et al. 2011). other studies with ungulates found the half normal key function with various expansions terms to perform best (trenkel et al. 1997, jathanna et al. 2003, peters 2010, thiessen 2010). the kernel estimator used for the probability of detection curve does not use maximum likelihood methods, so aicc values are not available for comparison with models with semi-parametric detection functions in the program distance. however, the kernel-based estimated probability of detection, animal density, and cv were similar to those obtained from the hazard-rate model in distance, although the ci was much narrower for the kernel estimator. based on bootstrapping, all the semi-parametric models estimated that the probability of detection biased low (i.e., lower than the kernel’s p̂ ¼ 0:73), thus estimated densities were biased high. the estimated bias in the hazard-rate model estimate (1.4%) was higher than the estimated bias for the kernelbased estimate (<0.001%). an advantage of the nonparametric estimator is that it is free of parametric assumptions on the detection function. additionally, using a kernel-based model does not require that detection is a monotonically decreasing function of distance away from the transect centerline, unlike semi-parametric models (cassey and mcardle 1999). one limitation is that it requires an adequate sample size to produce a reasonable estimate (chen 2000), but sample size was not a problem in our univariate analysis for the level of precision we achieved. the hermite polynomial and kernel estimates are very similar with the kernel estimate less intensive to compute (buckland 1992). survey effectiveness the study area is characterized by marginal snow conditions during any given year with the most reliable conditions in february (fig. 4). the daily average snow depth clearly demonstrates that the area does not accumulate deep snow and daily variation is high because of periods of warming and rain causing snowmelt. we considered ∼20 cm of snow accumulation as moderate to good conditions, required for the standard gspe survey method. conversely, dalton (1990) considered 20 cm “shallow” for a moose survey in ontario, canada. comparison of the helicopter linetransect method with the gspe method considers time, logistics, cost, and the estimate of precision. the gspe method requires a minimum of 60 units surveyed between 2 moose density strata (30 low and 30 high strata), with preferably more units in high density areas assuming greater variation within that strata (kellie and delong 2006). gspe survey areas are a minimum of 777 km2 because smaller areas have insufficient sample units to generate estimates. these survey units are approximately 16.6 km2 and require a minimum search intensity of 40 min/block with cub-like or tandem style fixed-winged aircraft (kellie and delong 2006). the time required to fly the minimum intensity and number of units would be ∼40 h. two aircraft for a minimum 150 estimating moose abundance – wald and nielson alces vol. 50, 2014 of 5 days would be required to complete the survey given continuous optimal snow conditions, a situation unlikely in the study area. the helicopter line-transect survey was considerably more efficient in terms of area sampled and flight time than a fixed-wing gspe. we flew a total of 16 h in one helicopter including about 2.5 h of training prior to the actual survey (training is highly recommended to prepare the survey team for duties and search patterns) and 13.5 h during the actual survey, including ferry times from the base of operation. we accomplished the survey in 2 days which improves the chance of continuous, optimal snow conditions. cost based solely on flight hours was similar; a fixed-wing survey for the minimum sampling under the gspe method in this area would be approximately 6% less than the helicopter line-transect technique, a reasonable compromise given the reduction in survey days (2 versus 5). our helicopter survey had a cv of 18%, a difficult level of precision to obtain with the minimum number of gspe units required to address potential stratification errors, sample sizes, and high variability of observations between units (kellie and delong 2006). management application the recent expansion and establishment of moose in the lower kuskokwim river tributaries prompted our survey efforts. our survey indicated that moose density is 0.47 moose/km2 or twice that of the adjacent lower kuskokwim river survey unit in 2008 (0.23 moose/km2 without scf; perry 2010). the difference can arguably be attributed to better habitat along the kuskokwim tributaries compared to the main channel of the lower kuskokwim river, and the 5-year moose hunting moratorium in the lower kuskokwim drainage that allowed moose to establish a viable population and disperse into unoccupied habitat. the exploitation of underutilized habitat was expressed in population production and is emphasized by the 30% rate of twinning during our march survey, which is high for that time of year. fig. 4. average daily snow depth at bethel, alaska airport (2000–2010) during typical moose survey months. in this area the minimum snow depth for gspe type surveys is ∼20 cm (dashed line). the variability in snow depth is due to periodic and rapid warming trends (noaa 2011). alces vol. 50, 2014 wald and nielson – estimating moose abundance 151 comparative twinning rate data are collected in may during the peak calving period and can be used in conjunction with other variables to determine the nutritional status of moose in an area. density dependence occurs at high moose densities in interior alaska when may twinning rates are 4– 21% indicating nutritional stress, whereas in years of low moose densities and recovery of vegetation, rates are 30–47% (boertje et al. 2007). the may twinning rate in our study area has recently been estimated as 47–67% (concurrent study; unpublished survey data, ydnwr), corroborating our assumption of a population at high nutritional status. this moose population continues to grow and is an important subsistence resource for local people and a monitoring program is essential for sound management regarding appropriate harvest levels to maintain a healthy, sustainable population. we recommend that surveys occur every 3–5 years to assess population trends and inform management decisions. future surveys along the lower kuskokwim tributaries should follow the same protocol used here (and potentially the same transects; buckland et al. 2001), with the exception of how moose locations under the aircraft were recorded. perpendicular distances of moose detected by the front-left observer under the helicopter (i.e., moose groups approximately ± 43 m of the center line) should be estimated and recorded in future surveys. the precision of our density estimates would have increased if we did not lump and remove from analyses the 10 observations directly under the helicopter. the kernel density estimator was fairly precise (cv = 18%); however, additional transects are required to increase precision in future surveys. attaining a cv closer to 15% would require an additional 307 km of transects; a precision with cv = 10% would require an additional 1564 km, given the current group encounter rate. given the narrow riparian corridors, it would be difficult to add transects without greatly increasing the chance of double counting groups. thus, improved precision is limited by increasing the encounter rate; however, if a cv of 20% is acceptable, transect length could be reduced by ∼133 km. another consideration is pooling data across years to obtain more robust, and potentially more precise estimates of detection probability (burnham et al. 1980, burnham et al. 2004, fewster et al. 2005). distance sampling is pooling robust and a common practice in a single study area because each transect typically has too few observations to calculate separate detection functions (gerard and schucany 2002). pooling by year to increase sample size (observations) and to account for various survey conditions (e.g., snow conditions) could improve the global detection function for the area if repeated surveys are in the same area and preferably along the same transects (nielson et al. 2014). if the cv ranges from 13–19%, managers should be able to detect at least a 38% change in abundance using a 90% ci with 80% statistical power. given our density estimate of 0.47 moose/km2 (340 moose), we should detect a change in density if the population changed by ∼0.18 moose/ km2 (129 moose; 38%). furthermore, if there is a 5-year period between surveys, this would require a finite rate of change (λ = e{(ln nt − ln n0) / t}, where n0 is the starting abundance estimate, nt is the abundance estimate at time t, and t is the time period between surveys; skalski et al. 2005: 295) equal to λ = 0.909 annually for a decreasing population, or λ = 1.066 for an increasing population. this is a realistic change for moose in this area since the lower kuskokwim survey unit showed an extreme growth rate of λ = 1.647 over a 4-year period (from 70 to 515 moose; perry 2010). our research provides a viable alternative method to survey moose in a subarctic 152 estimating moose abundance – wald and nielson alces vol. 50, 2014 tundra ecosystem with marginal snow conditions. presumably this technique could be applied elsewhere in areas with larger contiguous habitat and variable snow conditions such as portions of moose range within subarctic alaska, canada, scandinavia, and russia. areas with dense canopy cover may require accounting for vegetation covariates affecting sightability which would require a larger number of group observations. nevertheless, as climate change increases the disruption of prevailing weather patterns and causes more atypical and uncertain scenarios such as freeze-thaw or rain-on-snow events, our research and recommendations provide wildlife managers an accurate, efficient, and cost-effective option for surveying moose populations. acknowledgements we would like to thank observers m. gabrielson and v. anvil, and pilot s. hermens (hermen helicopters) for his expert flying. we also thank g. walters for his aerial fixed-wing support and t. doolittle for encouraging this effort. we are grateful to drs. l. munn, r. mccormick, j. beck, m. murphy, and s. miller, as well as the alces editors and 2 anonymous reviewers for providing helpful comments. funding for this project was through the us fish and wildlife service, yukon delta national wildlife refuge, bethel, alaska, usa. the use of trade names of commercial products in this manuscript does not constitute endorsement or recommendation for use by the federal government. references aderman, a. r. 2008. demographics and home ranges of moose at togiak national wildlife refuge, southwest alaska, 1998 – 2007. progress report. u.s. fish and wildlife service, togiak national wildlife refuge, dillingham, alaska, usa. ———, and j. woolington. 2006. demographics and home ranges of moose at togiak national wildlife refuge, southwest alaska: march 1998–march 2005. progress report. u.s. fish and wildlife service, togiak national wildlife refuge, dillingham, alaska, usa. anderson, c. r., jr., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24: 47–259. bachler, e., and f. liechti. 2007. on the importance of g(0) for estimating bird population densities with standard distance-sampling: implications from a telemetry study and a literature review. ibis 149: 693–700. becker, e. f., and p. x. quang. 2009. a gamma-shaped detection function for line-transect surveys with double-count and covariate data. journal of agricultural, biological, and environmental statistics 14: 207–223. boertje, r. d., k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. borchers, d. l., j. l. laake, c. southwell, and c. g. m. paxton. 2006. accommodating unmodeled heterogeneity in double-observer distance sampling surveys. biometrics 62: 372–378. briggs, k. t., w. b. tyler, and d. b. lewis. 1985. comparison of ship and aerial surveys of birds at sea. journal of wildlife management 49: 405–411. buckland, s. t. 1984. monte carlo confidence intervals. biometrics 40: 811–817. ———. 1992. fitting density functions with polynomials. journal of the royal statistical society, series c-applied statistics 41: 63–76. ———, d. r. anderson, k. p. burnham, j. l. laake, d. l. borchers, and l. thomas. 2001. introduction to distance alces vol. 50, 2014 wald and nielson – estimating moose abundance 153 sampling. oxford university press, new york, usa. ———, ———, ———, ———, ———, and ———, editors. 2004. advanced distance sampling: estimating abundance of biological populations. oxford university press, new york, new york, usa. ———, j. l. laake, and d. l. borchers. 2010. double-observer line transect methods: levels of independence. biometrics 66: 169–177. ———, and b. j. turnock. 1992. a robust line transect method. biometrics 48: 901–909. burnham, k. p., and d. r. anderson. 1976. mathematical models for nonparametric inferences from line transect data. biometrics 32: 325–336. ———, and ———. 1984. the need for distance data in transect counts. journal of wildlife management 48: 1248–1254. ———, and ———. 2002. model selection and multimodel inference: a practical information-theoretic approach. second edition. springer, new york, new york, usa. ———, ———, and j. l. laake. 1980. estimation of density from line transect sampling of biological populations. wildlife monographs 72: 1–202. ———, ———, and ———. 1985. efficiency and bias in strip and line transect sampling. journal of wildlife management 49: 1012–1018. ———, s. t. buckland, j. l. laake, d. l. borchers, t. a. marques, j. r. b. bishop, and l. thomas. 2004. further topics in distance sampling. pages 307– 392 in s. t. buckland, d. r. anderson, k. p. burnham, j. l. laake, d. l. borchers, and l. thomas, editors. advanced distance sampling: estimating abundance of biological populations. oxford university press, new york, new york, usa. cassey, p., and b. h. mcardle. 1999. an assessment of distance sampling techniques for estimating animal abundance. environmetrics 10: 261–278. chen, s. x. 1996a. a kernel estimate for the density of a biological population by using line transect sampling. applied statistics 45: 135–150. ———. 1996b. studying school size effects in line transect sampling using the kernel method. biometrics 52: 1283–1294. ———. 1999. estimation in independent observer line transect surveys for clustered populations. biometrics 55: 754–759. ———. 2000. animal abundance estimation in independent observer line transect surveys. environmental and ecological statistics 7: 285–299. coady, j. w. 1980. history of moose in northern alaska and adjacent regions. the canadian field naturalist 94: 61–68. cook, r. d., and j. o. jacobson. 1979. a design for estimating visibility bias in aerial surveys. biometrics 35: 735–742. cumberland, r. e. 2012. potvin doublecount aerial surveys in new brunswick: are results reliable for moose? alces 48: 67–77. dalton, w. j. 1990. moose density estimation with line transect survey. alces 26: 129–141. delong, r. a. 2006. geospatial population estimator software user’s guide. alaska department of fish and game. fairbanks, alaska, usa. diciccio, t. j., and b. efron. 1996. bootstrap confidence intervals. statistical science 3: 189–228. drummer, t. d., and l. l. mcdonald. 1987. size bias in line transect sampling. biometrics 43: 13–21. ———, a. r. degange, l. l. pank, and l. l. mcdonald. 1990. adjusting for group size influence in line transect sampling. journal of wildlife management 54: 511–514. efron, b. 1981a. nonparametric estimates of standard error: the jackknife, the 154 estimating moose abundance – wald and nielson alces vol. 50, 2014 bootstrap and other methods. biometrika 68: 589–599. ———. 1981b. nonparametric standard errors and confidence intervals. the canadian journal of statistics 9: 139–172. ———. 1982. the jackknife, the bootstrap and other resampling plans. cbms-nsf regional conference series in applied mathematics (vol. 38). society for industrial and applied mathematics, philadelphia, pennsylvania, usa. ———, and r. tibshirani. 1986. bootstrap methods for standard errors, confidence intervals, and other measures of statistical accuracy. statistical science 1: 54–77. ———, and ———. 1994. an introduction to the bootstrap. chapman and hall/ crc, boca raton, florida, usa. evans, c. d., w. a. troyer, and c. j. lensink. 1966. aerial census of moose by quadrat sampling units. journal of wildlife management 30: 767–776. fewster, r. m., s. t. buckland, k. p. burnham, d. l. borchers, p. e. jupp, j. l. laake, and l. thomas. 2009. estimating the encounter rate variance in distance sampling. biometrics 65: 225–236. ———, j. l. laake, and s. t. buckland. 2005. line transect sampling in small and large regions. biometrics 61: 856–859. ———, c. southwell, d. l. borchers, s. t. buckland, and a. r. pople. 2008. the influence of animal mobility on the assumption of uniform distances in aerial line-transect surveys. wildlife research 35: 275–288. focardi, s., r. isotti, and a. tinelli. 2002. line transect estimates of ungulate populations in a mediterranean forest. journal of wildlife management 66: 48–58. gasaway, w. c., and s. d. dubois. 1987. estimating moose population parameters. swedish wildlife research supplement 1: 603–617. ———, ———, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, no. 22, institute of arctic biology, fairbanks, alaska, usa. gerard, p. d., and w. r. schucany. 1999. local bandwidth selection for kernel estimation of population densities with line transect sampling. biometrics 55: 769–773. ———, and ———. 2002. combining population density estimates in line transect sampling using the kernel method. journal of agricultural, biological, and environmental statistics 7: 233–242. giudice, j. h., j. r. fieberg, and m. s. lenarz. 2012. spending degrees of freedom in a poor economy: a case study of building a sightability model for moose in northeastern minnesota. journal of wildlife mangement 76: 75–87. gosse, j., b. mclaren, and e. eberhardt. 2002. comparison of fixed-wing and helicopter searches for moose in a midwinter habitat-based survey. alces 38: 47–53. graham, a., and r. bell. 1989. investigating observer bias in aerial surveys by simultaneous double-counts. journal of wildlife management 53: 1009–1016. jang, w., and j. m. loh. 2010. density estimation for grouped data with application to line transect sampling. the annals of applied statistics 4: 893–915. jathanna, d., k. u. karanth, and a. j. t. johnsingh. 2003. estimation of large herbivore densities in the tropical forests of southern india using distance sampling. journal of zoology (london) 261: 285–290. jung, t. s., t. e. chubbs, c. g. jones, f. r. phillips, and r. d. otto. 2009. winter habitat associations of a low-density moose (alces americanus) population in central labrador. northeastern naturalist 16: 471–480. kellie, k. a., and r. a. delong. 2006. geospatial survey operations manual. alces vol. 50, 2014 wald and nielson – estimating moose abundance 155 alaska department of fish and game. fairbanks, alaska, usa. laake, j. l., and d. l. borchers. 2004. methods for incomplete detection at distance zero. pages 108–189 in s. t. buckland, d. r. anderson, k. p. burnham, j. l. laake, d. l. borchers, and l. thomas, editors. advanced distance sampling: estimating abundance of biological populations. oxford university press, new york, new york, usa. ———, m. j. dawson, and j. hone. 2008. visibility bias in aerial survey: markrecapture, line-transect or both? wildlife research 35: 299–309. lancia, r. a., w. l. kendall, k. h. pollock, and j. d. nichols. 2005. estimating the number of animals in wildlife populations. pages 106–153 in c. e. braun, editor. techniques for wildlife investigations and management. sixth edition, the wildlife society, bethesda, maryland, usa. leresche, r. e., and r. a. rausch. 1974. accuracy and precision of aerial moose censusing. journal of wildlife management 38: 175–182. mack, y. p., and p. x. quang. 1998. kernel methods in line and point transect sampling. biometrics 54: 606–619. marques, f. f. c. and s. t. buckland. 2004. covariate models for the detection function. pages 31–47 in s. t. buckland, d. r. anderson, k. p. burnham, j. l. laake, d. l. borchers, and l. thomas, editors. advanced distance sampling. oxford university press, new york, usa. ———, m. andersen, s. christensen-dalsgaard, s. belikov, a. boltunov, o. wiig, s. t. buckland, and j. aars. 2006. the use of global positioning systems to record distances in a helicopter linetransect survey. wildlife society bulletin 34: 759–763. mccullagh, p., and j. a. nelder. 1989. generalized linear models. second edition. chapman and hall, london, united kingdom. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75: 621–630. nielson, r. m., t. j. evans, and m. b. stahl. 2013. investigating the potential use of aerial line transect surveys for estimating polar bear abundance in sea ice habitats: a case study for the chukchi sea. marine mammal science 29: 389–406. ———, l. l. mcdonald, and s. d. kovach. 2006. aerial line transect survey protocols and data analysis methods to monitor moose (alces alces) abundance as applied on the innoko national wildlife refuge, alaska. technical report prepared for us fish and wildlife service. west, inc., cheyenne, wyoming, usa. (accessed december 2009). ———, l. mcmanus, t. rintz, l. l. mcdonald, r. k. murphy, b. howe, and r. e. good. 2014. monitoring abundance of golden eagles in the western united states. journal of wildlife management 77: 1436–1448. noaa (national oceanic and atmospheric administration). 2011. climate data online. national climate data center, asheville, north carolina, usa. (accessed april 2011). ———. 2013. el niño -southern oscillation (enso). climate prediction center, national weather service, camp springs, maryland, usa. (accessed january 2013). oehlers, s.a., r. t. bowyer, f. huettmann, d. k. person, and w. b. kessler. 2012. visibility of moose in a temperate rainforest. alces 48: 89–104. perry, p. 2010. unit 18 moose management report. pages 271–285 in p. harper, 156 estimating moose abundance – wald and nielson alces vol. 50, 2014 http://www.west-inc.com/reports/moose_survey.pdf http://www.west-inc.com/reports/moose_survey.pdf http://www.ncdc.noaa.gov/cdo-web/ http://www.ncdc.noaa.gov/cdo-web/ http://www.cpc.ncep.noaa.gov/products/precip/cwlink/mjo/enso.shtml http://www.cpc.ncep.noaa.gov/products/precip/cwlink/mjo/enso.shtml http://www.cpc.ncep.noaa.gov/products/precip/cwlink/mjo/enso.shtml editor. moose management report of survey and inventory activities 1 july 2007–30 june 2009. alaska department of fish and game. project 1.0. juneau, alaska, usa. peters, w. e. b. 2010. resource selection and abundance estimation of moose: implications for caribou recovery in a human altered landscape. m.s. thesis, university of montana, missoula, montana, usa. (accessed january 2012). peters, w., m. hebblewhite, k. g. smith, s. m. webb, n. webb, m. russell, c. stambaugh, and r. b. anderson. 2014. contrasting aerial moose population estimation methods and evaluating sightability in west-central alberta, canada. wildlife society bulletin: doi: 10.1002/wsb.433. first published online 3 may 2014. pollock, k. h., and w. l. kendall. 1987. visibility bias in aerial surveys: a review of estimation procedures. journal of wildlife management 51: 502–510. quang, p. x. 1990. confidence intervals for densities in line transect sampling. biometrics 46: 459–472. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–54. r development core team. 2010. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. samuel, m. d., e. o. garton, m. w. schlegel, and r. g. carson. 1987. visibility bias during aerial surveys of elk in northcentral idaho. journal of wildlife management 51: 622–630. schmidt, j. h., k. l. rattenbury, j. p. lawler, and m. c. maccluskie. 2011. using distance sampling and hierarchical models to improve estimates of dall’s sheep abundance. journal of wildlife management 76: 317–327. schroeder, b., and s. murphy. 1999. population surveys (ground and aerial) on nesting beaches. pages 45–55 in k. l. eckert, k. a. bjorndal, f. a. abreugrobois, and m. donnelly, editors. research and management techniques for the conservation of sea turtles. iucn/ssc marine turtle specialist group publication no. 4. seddon, p. j., k. ismail, m. shobrak, s. ostrowski, and c. magin. 2003. a comparison of derived population estimate, mark-resighting and distance sampling methods to determine the population size of a desert ungulate, the arabian oryx. oryx 37: 286–294. sheather, s. j. 2004. density estimation. statistical science 19: 588–597. ———, and m. c. jones. 1991. a reliable data-based bandwidth selection method for kernel density estimation. journal of the royal statistical society series b 53: 683–690. shorrocks, b., b. cristescu, and s. magane. 2008. estimating density of kirk’s dikdik (madoqua kirkii gunther), impala (aepyceros melampus lichtenstein) and common zebra (equus burchelli gray) at mpala, laikipia district, kenya. african journal of ecology 46: 612–619. silverman, b. w. 1986. density estimation for statistics and data analysis. chapman and hall, london, united kingdom. skalski, j. r., k. e. ryding, and j. j. millspaugh. 2005. wildlife demography: analysis of sex, age, and count data. elsevier academic press, burlington, massachusetts, usa. smits, c. m. m., r. m. p. ward, and d. g. larsen. 1994. helicopter or fixed-wing aircraft: a cost-benefit analysis for moose surveys in yukon territory. alces 30: 45–50. thiessen, c. 2010. horn river basin moose inventory january/february 2010. bc ministry of environment, fort st. john, british columbia, canada. alces vol. 50, 2014 wald and nielson – estimating moose abundance 157 http://etd.lib.umt.edu/theses/available/etd-01102011-145105/ http://etd.lib.umt.edu/theses/available/etd-01102011-145105/ (accessed march 2013). thomas, l., s.t. buckland, k. p. burnham, d. r. anderson, j. l. laake, l. borchers, and s. strindberg. 2002. distance sampling. pages 544–552 in a. h. el-shaarawi and w. w. piegorsch, editors. encyclopedia of environmetrics, vol. 1. ———, ———, e. a. rexstad, j. l. laake, s. strindberg, s. l. hedley, j. r. b. bishop, t. a. marques, and k. p. burnham. 2010. distance software: design and analysis of distance sampling surveys for estimating population size. journal of applied ecology 47: 5–14. ———, j. l. laake, e. rexstad, s. strindberg, f. f. c. marques, s. t. buckland, d. l. borchers, d. r. anderson, k. p. burnham, m. l. burt, s. l. hedley, j. h. pollard, j. r. b. bishop, and t. a. marques. 2009. user’s guide: distance 6.0. release 2. research unit for wildlife population assessment, university of st. andrews, united kingdom. (accessed january 2013). thompson, i. d. 1979. a method of correcting population and sex and age estimates from aerial transect surveys for moose. proceedings of the north american moose conference and workshop 15: 148–168. timmermann, h. r. 1974. moose inventory methods: a review. naturaliste canadien 101: 615–629. ———. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29: 35–46. ———, and m. e. buss. 2007. population and harvest management. pages 559–615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. second edition. university of colorado press, boulder, colorado, usa. trenkel, v. m., s. t. buckland, c. mclean, and d. a. elston. 1997. evaluation of aerial line transect methodology for estimating red deer (cervus elaphus) abundance in scotland. journal of environmental management 50: 39–50. venables, w. n., and b. d. ripley. 2002. modern applied statistics with s. fourth edition. springer, new york, new york, usa. ver hoef, j. m. 2002. sampling and geostatistics for spatial data. ecoscience 9: 152–161. ———. 2008. spatial methods for plot-based sampling of wildlife populations. environmental and ecological statistic 15: 3–13. wand, m. p., and c. m. jones. 1995. kernel smoothing. fourth edition. crc press, london, united kingdom. wilson, r. r., a. bartsch, k. joly, j. h. reynolds, a. orlando, and w. m. loya. 2013. frequency, timing, extent, and size of winter thaw-refreeze events in alaska 2001–2008 detected by remotely sensed microwave backscatter data. polar biology 36: 419–426. young, j. k., k. m. murray, s. strindberg, b. buuveibaatar, and j. berger. 2010. population estimates of endangered mongolian saiga saiga tatarica mongolica: implications for effective monitoring and population recovery. oryx 44: 285–292. zar, j. h. 1999. biostatistical analysis. fourth edition. prentice-hall, inc., upper saddle river, new jersey, usa. 158 estimating moose abundance – wald and nielson alces vol. 50, 2014 http://scek.ca/documents/scek/final_reports/hrb%20moose%20inventory%202010%20(phase%201)_final%20report.pdf http://scek.ca/documents/scek/final_reports/hrb%20moose%20inventory%202010%20(phase%201)_final%20report.pdf http://scek.ca/documents/scek/final_reports/hrb%20moose%20inventory%202010%20(phase%201)_final%20report.pdf http://scek.ca/documents/scek/final_reports/hrb%20moose%20inventory%202010%20(phase%201)_final%20report.pdf http://www.ruwpa.st-and.ac.uk/distance/ http://www.ruwpa.st-and.ac.uk/distance/ estimating moose abundance in linear subarctic habitats in low snow conditions with distance sampling and a kernel estimator study area methods field survey data analysis results discussion assumptions detection functions survey effectiveness management application acknowledgements references alces29_201.pdf alces22_245.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces24_48.pdf alces26_30.pdf alces21_501distmoosebio.pdf alces vol. 21, 1985 alces26_66.pdf alces24_90.pdf alces27_65.pdf alces28_249conferenceworkshops.pdf alces28_35.pdf alces(23)_33.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 61 a possible source of brain abscesses in bull moose vince crichton1 and rick wowchuk2 11046 mcivor ave., winnipeg, manitoba r2g 2j9, canada; 2box 2217, swan river, manitoba r0l 1z0, canada abstract: the presence of cranial infections and abscessations is well documented in males of multiple cervids in north america. the preponderance of such infections is related directly to antlers and all processes from antler growth, fighting, and through to casting. one proposed infection pathway is through an open wound at the pedicle formed at casting. moose generally do not cast antlers in synchrony, and we propose that males irritated by the imbalance of a remaining antler are more likely to actively remove that antler by striking trees. this behavior is a possible explanation for the occurrence of cast antlers with attached bone and that antlers from bulls of all ages can have substantial amounts of parietal bone attached. the force of this activity may cause breakage of the parietal bone leaving either an opening to the meninges in the cranial vault or a significant depression in the bone. we propose that shed antlers with measurable parietal bone attached, estimated as high as 10% of cast moose antlers, would create abnormally large wounds and possibly an enhanced route of cranial infection and subsequent abscessations. alces vol. 55: 61–65 (2019) key words: abscesses, antlers, brain, cervids, males, moose. antlers are principally a secondary sexual characteristic critical in the reproductive behavior of cervids. they occasionally have structural abnormalities associated with either injury during the growth period or in response to a previous physical injury. another type of abnormality is associated with the annual process of casting antlers. specifically, multiple reports document cast white-tailed deer (odocoileus virginianus), elk (cervus elaphus), and moose (alces alces) antlers, and to a lesser extent caribou (rangifer tarandus) antlers, with measurable parietal bone attached (fig. 1). surprisingly, during filming of deer engaged in antler sparring for an outdoor program, the host displayed a broken-off antler with part of the skull attached. of consequence is that in this or a similar outcome after casting, a possible infection route into the cranial cavity and meninges may occur in open wound sites at the pedicle (w. samuel, university of alberta, retired). intracranial abscessations in male white-tailed deer are well documented in north america, seasonally focused in september-april, and presumed to be associated with breeding behaviors involving antlers (i.e., sparring, rubbing, and casting) (davidson et al. 1990, baumann et al. 2001, cohen et al. 2015). these studies report a frequency of <10%, but karns et al. (2009) documented a 35% rate of abscessations in a high-density deer population in maryland; albeit, their sample size was less robust than the other studies. infections are principally associated with arcanobacterium pyogenes (davidson et al. 1990, baumann et al. 2001, karns et al. 2009), a common bacterium that invades superficial wounds of ungulates (zulty and montali 1988). the infection brain abscesses in moose – crichton and wowchuk alces vol. 55, 2019 62 pathway is presumed to be through the open wound associated with either normal casting or abnormal antler breakage at the pedicle. in an examination of 4953 male deer from georgia, cohen et al. (2015) found 91 abscesses (1.8%) and higher probability of an abscess with increasing age; no cranial abscesses were found in 2562 females. although they looked at site-specific variables, none were strongly associated with observations of infection. van ballenberghe (1982) noted an infection at the pedicle of an alaskan moose that had prematurely cast an antler in early september prior to the peak of the rutting period. although documented reports are rare with moose, maccracken et al. (1994) found a high frequency of cast moose antlers with attached pedicle bone in the copper river delta in alaska; however, the authors attributed this to genetic and/or local geographic causes. alternate hypotheses for this anomaly include trauma associated with antler rubbing, male confrontations during the rut, behavior and activity during the casting process (davidson et al. 1990), and physiological stress associated with relative nutritional condition following harsh winters (landete-castillejos et al. 2010). regardless of origin, these wounds and skull fractures presumably open a pathway for intracranial infection by arcanobacterium pyogenes that could prove fatal. it is suggested that the skeletal fractures, abscessations, and infection generally confined to male cervids are directly related to antlers. the origin of injuries to the pedicle area is not clear other than the fact that parietal bone is observed on some cast antlers. both authors (biologist and licensed antler dealer) have observed multiple cast antlers (>1000; range – 1–10% annually) from moose, elk, and deer with a substantial piece of the skull attached (fig. 1 and 2). we find this more common in moose and deer than elk, and although we have examined fewer, have documented erosion of the parietal bone in a caribou skull (fig. 3). we and others searching fig. 1. cast antler from a bull moose showing a portion of the parietal bone attached to the cast antler. alces vol. 55, 2019 brain abscesses in moose – crichton and wowchuk 63 for cast antlers occasionally locate moose antlers at the base of trees which is suggestive of using trees to physically remove antlers. the process as to how breakage occurs is not entirely clear given the multiple behavioral and nutritional explanations. for example, although breakage could occur from blunt force during rutting activity (sparring and fighting), antlers are not necessarily shed in synchrony when cast, fig. 2 . antler from mature bull moose showing pedicle area totally covered with skull bone. fig. 3. caribou skull showing erosion of the parietal bone at the point of antler attachment. brain abscesses in moose – crichton and wowchuk alces vol. 55, 2019 64 as single-antlered moose are frequently observed during the casting period. we suggest that some moose (1.5 years and older) facilitate casting of the remaining antler, which may be an irritant due to imbalanced weight distribution, by striking or rubbing the attached antler against trees or a rigid object in an attempt to dislodge it. newsom (1937) observed that moose may facilitate casting by knocking their antlers against trees and such behavior may explain why some cast antlers are occasionally found at the base of trees. in moose, the torque force caused by the massive, horizontal antlers with a center of gravity far from the skull likely exceeds that in other cervids (nygren et al. 1992). although we have observed this phenomenon in all-aged bulls, larger antlers from animals > 2.5 years old generally have more attached parietal bone than those from younger animals. we did not examine antlers for mineral content as done by maccracken et al. (1994) and recognize that rutting behavior may cause weakening of or hairline fractures at the pedicle (davidson et al. 1990), and that age (hindelang and peterson 1996) and environmental factors (landetecastillejos et al. 2010) influence antler growth, relative bone strength, and presence of osteolytic lesions on the skull. nevertheless, we propose that a contributing factor to substantial pieces of parietal bone on cast moose antlers may be the physical force used to shed the remaining contralateral antler when striking it against solid objects, most often trees. in turn, associated larger wounds at the pedicle provide a possible route of infection for arcanobacterium pyogenes to the meninges resulting in intracranial abscessations. will an antler grow in subsequent years if the pedicle is damaged? field observations suggest that antlers may not develop normally from a damaged pedicle the following year, and bulls with a single antler are observed occasionally in summer (pers. observation, crichton). however, permanent damage or failure to grow antlers is considered rare in advanced cervid species such as moose in which pedicle wounds have multiple months to heal between casting and regrowth (bubenik 1982, goss 1983). regardless, reduced health and survival of moose, elk, deer, and caribou can occur from a cranial infection, and physically-forced antler casting may enhance that possibility. consideration of such is warranted with regard to management strategies aimed at regulating or banning the use of antler traps that ensnare or knock off antlers forcibly. references baumann, c. d., w. r. davidson, d. e. roscoe, and k. beheler-amass. 2001. intracranial abscessation in whitetailed deer of north america. journal of wildlife diseases 37: 661–670. doi:10.7589/0090-3558-37.4.661 bubenik, g. a. 1982. the endocrine regulation of the antler cycle. pages 73–107 in r. d. brown, editor. antler development in cervidae. caesar kleberg wildlife research institute, kingsville, texas, usa. cohen, s., e. h. belser, c. h. killmaster, j. w. bowers, b. j. irwin, m. j. yabsley, and k. v. miller. 2015. epizootiology of cranial abscess disease in whitetailed deer (odocoileus virginianus) of georgia. journal of wildlife diseases 51: 671–679. doi:10.7589/2014-05-129 davidson, w. r., v. f. nettles, l. e. hayes, e. w. howerth, and c. e. couvillion. 1990. epidemiologic features of an intracranial abscessation/suppurative meningoencephalitis complex in white-tailed deer. journal of wildlife diseases 26: 460–467. doi:10.7589/0090-3558-26.4.460 goss, r. j. 1983. deer antlers: regeneration, function, and evolution. academic press, new york, new york, usa. alces vol. 55, 2019 brain abscesses in moose – crichton and wowchuk 65 hindelang, m., and r. o. peterson. 1996. osteoporotic skull lesions in moose at isle royale national park. journal of wildlife diseases 32: 105–108. doi:10.7589/0090-3558-32.1.105 karns, g. r., r. a. lancia, c. s. deperno, m. c. conner, and m. k. stoskopf. 2009. intracranial abscessation as a natural mortality factor for adult male deer (odocoileus virginianus) in kent county, maryland, usa. journal of wildlife diseases 45: 196–200. doi:10.7589/ 0090-3558-45.1.196 landete-castillejos, t., j. d. currey, j. a. estevez, y. fierro, a. calatayud, f. ceacero, a. j. garcia, and l. gallego. 2010. do drastic weather effects on diet influence changes in chemical composition, mechanical properties and structure in deer antlers? bone 47: 815–825. doi:10.1016/j.bone.2010. 07.021 maccracken, j. g., t. r. stephenson, and v. van ballenberghe. 1994. peculiar antler cast by moose on the copper river delta, alaska. alces 30: 13–19. newsom, w. m. 1937. winter notes on the moose. journal of mammalogy 18: 347–349. doi:10.2307/1374210 nygren, k., r. silvennoinen, and m. karna. 1992. antler stress in the nasal bone region of moose. alces supplement 1: 84–90. van ballenberghe, v. 1982. growth and development of moose antlers in alaska. pages 37–48 in r. d. brown, editor. antler development in cervidae. caesar kleberg wildlife research institute, kingsville, texas, usa. zulty, j. c., and r. j. montali. 1988. actinomyces pyogenes infection in exotic bovidae and cervidae: 17 cases (1976–1986). journal of zoo animal medicine 19: 30–32. doi:10.2307/ 20094849 screenposition alces27_220.pdf alces24_195.pdf alces24_159.pdf alces28_137.pdf alces29_281conferenceworkshop.pdf alces(25)_75.pdf alces27_24.pdf alces(23)_303distinguishedmoosebio.pdf alces vol. 23, 1987 alces(25)_182distinguishedmoosebio.pdf alces22_443_moose&forestmgmt.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces(25)_168.pdf alces21_37.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 45, 2009 sipko and kholodova – eurasian moose population 25 fragmentation of eurasian moose populations during periods of population depression taras p. sipko and marina v. kholodova institute of ecology and evolution, russian academy of sciences, 33 leninsky prospect, moscow 119071, russia abstract: changes in the distribution of eurasian moose (alces alces) populations during the pleistocene and holocene eras were analyzed from historical and contemporary literature. we focused on how range boundaries varied, suitable habitat was fragmented, and how local and regional populations were isolated, especially during periods of population depression. we discuss how the occurrence and duration of isolation of local populations likely influenced the genetic structure of eurasian moose. we question the geographic division of certain subspecies, and suggest that our analysis be used to reinterpret and revise genetic structure of eurasian moose populations. alces vol. 45: 25-34 (2009) key words: alces alces, eurasia, fragmentation, genetic structure, history, isolation, moose, population depression, range, subspecies. the eurasian moose (alces alces) popula-population returned to most of its original range during the 20th century. historically it experienced numerous range reductions and fragmentations that were followed by restoration and dispersal into new areas. recent research (danilkin 1999) suggests that species differentiation in moose could be greater than believed previously. in order to best understand and interpret the genetic structure of eurasian moose (i.e., its phylogeography and polymorphism), it is necessary to identify dispersal centres where genetic diversity was presumably highest. assuming that mitochondria are inherited maternally and females disperse shorter distance than males, we expect that the geographic distribution of mitochondrial dna haplotypes was fairly stable and should reflect past migration routes of moose. we collected and analyzed extensive archaeological and paleontological data to trace moose range during periods of substantial population depression. we regard our results as preliminary because these data are incomplete, and distribution of moose populations varied in space and time and was not always documented accurately. distribution and taxonomoy subspecies of moose in eurasia are considered to have no distinct differences; variable morphological characteristics relative to geographic location represent a cline (markov and danilkin 1996, danilkin 1999). moose constitute one macropopulation in europe (rozhkov et al. 2002, davydov et al. 2004) with european moose (alces alces alces) inhabiting europe, altai, and western siberia up to the yenisei river (danilkin 1999). in summer and winter the yenisei river is a stem of a large river system rich in valleys that are preferred moose habitat with no hindrance to moose migration. we suspect that the subspecies boundary lies either westward along the ob and yenisei rivers divide, or eastward along the divide of the yenisei and lena river basins; further research is required to better delineate this boundary. caucasian moose (a. a. caucasicus) in the caucasus area were eliminated by the beginning of the 20th century. repopulation of the north caucasus region by this subspecies in the 20th century indicates that caucasian and european moose habitats were probably well eurasian moose population – sipko and kholodova alces vol. 45, 2009 26 connected in the past, and calls into question whether to identify caucasian moose as a separate subspecies (danilkin 1999). the subspecies status seems more appropriate for the moose population in the westernmost part of europe. yakut moose (a. a. pfizenmayeri) inhabit the area to the east of the yenisei river up to the stanovoi ridge, khakassia (tikhonov 1990), northern mongolia to the south, and the tchersky or verkhoyansky ridges to the northeast. kolyma moose (a. a. buturlini) are distributed in northeast siberia to the east of the tchersky ridge (zheleznov 1990). we think that the boundary of this subspecies runs along an arc formed by the verkhoyansky ridge and subtar-khayata ridge; further genetic research is needed to delineate this boundary. ussuri moose (a. a. cameloides) are restricted to the southwestern part of siberia and the amur basin; the distribution of this subspecies requires further study. it’s highly possible that moose populations inhabiting the area east of baikal lake (ditsevich 1990) and further east in steppe valleys of the selenga and orkhon basins, and in northeast china also belong to this subspecies. the northern boundary runs along the stanovoy ridge (kutcherenko 1975) and the lena river divide (danilkin 1999). biological parameters in russia moose live mainly in forest habitats but also occupy the forest steppe and forest tundra. moose were documented in tundra on the lena river estuary along the coastline of the arctic sea (nasimovich 1955), and moose remnants were found on new siberian islands in the arctic sea (vereshchagin 1967). they were also documented in the desert near aral lake, >500 km from their original ecotopes (heptner and nasimowitsch 1974). moose have also been documented in the alpine zone to 2500 m elevation (semenov-tjan-shansky 1948, heptner et al. 1961). although typically sedentary, certain populations migrate and individuals may disperse long distances. it took them several decades to disperse >1500 km to reach the caucasus foothills (yasan 1966), but they easily cross waterways 10-15 km wide (timofeeva 1974). their size, mobility, and relatively high reproductive rate aid them in repopulating vacated areas. moose in the pleistocene era the earliest evidence of moose remains dates back to the mid-pleistocene era (vereshchagin 1967) that had no less than 3 periods of glaciation (gerasimov and markov 1939) that dramatically altered landscapes and moose range (fig. 1). according to vereshchagin (1967), there is scarce evidence of moose in the ice age, whereas moose remains are considerable in the holocene era. active morphogenetic processes apparently occurred in the late pleistocene era (boeskorov 2001). moose distribution during the ice age can be described only in general terms. because the greatest part of europe was covered by glacial sheets and lakes, moose probably inhabited a rather narrow territory between the ocean and glacial lakes adjoining the alpine and nordic ice sheets. this was possibly a connecting link between the moose population in southwest europe overgrown with forests very much like those of modern scandinavia (woillard 1979), and that inhabiting vast areas of the south ural mountains where forest refugia existed (panov 1999). the european moose population in this area seems to be the most abundant and genetically diverse, and as the glacial sheet retreated northward it covered eastern europe. much later, about 10,000 years ago when the southern part of scandinavia was free of ice, european moose with red deer (cervus elaphus) and wisent (bison bonasus) found their way from west europe to fennoscandia using denmark as a land bridge (filonov 1983), and later dispersed from the east through the karelian isthmus. in west siberia when alces vol. 45, 2009 sipko and kholodova – eurasian moose population 27 the last glacial period was at its utmost, the glacier blocked the yenisei and ob rivers forming mansi lake, a large reservoir twice as large as the black sea; surplus water ran via the turgai channel to the caspian sea (groswald 1983). during this period western and eastern siberian moose appeared to be totally isolated, which explains the chromosomal differences between the eastern and western populations. the altai-sayan mountain region had no solid glacial sheet during the ice age (gerasimov and markov 1939), although a chain of mountain glaciers factored into the isolation of northern and southern moose populations in this area. there was no solid glacial sheet in east siberia during the pleistocene (groswald 1998, 1999) and moose occupied all suitable areas. southeast siberia is limited by lake baikal in the west and by the stanovoi highlands, then the stanovoi range, and part of the dzhugdzhur range in the east. although moose currently cross these mountain systems, these mountains were covered with vast glaciers in the pleistocene era (preobrazhenskiy 1960, groswald 1984) and were impassable for some period isolating the so-called ussuri moose from the rest of the population. northeastern asia is a huge amphitheatre sloping towards the arctic ocean that is characterized by strong orographic contrasts; though subdued mountains prevail, they are joined with highlands and plains. the verkhoyansk mountains present an orographic barrier of the area in the west. to the south of the verkhoyansk range, the sette-daban and the yudom range stretch divided by the yudom-mai highlands, and further along the okhotsk sea coastline lies the dzhugdzhur range. the tchersky range stretches 1800 fig. 1. distribution of ice sheets, mountain glaciers, and ice-dammed lakes in eurasia during the midpleistocene era (according to groswald 1984) that influenced the distribution and range of moose. major seas include the aral (as), black (bs), and caspian (cs). glacial sheets include the chukchi (ch), east siberian (es), karskii (ka), ohotskii (oh), and scandinavian (sc). mountain glaciers include the altai (as), baikal (ba), central asian (ca), tibetan (ti), and verkhoyansk (ve). other features include lake mansijskoe (ml), amur river (am), and the turgaiskii trench (tu). eurasian moose population – sipko and kholodova alces vol. 45, 2009 28 km northwest to the east of the verkhoyansk mountains. glaciers developed in the mountains to various extant and during glacial maximums reached highland valleys which had several glacial and interglacial periods (groswald and kotljakov 1989). highlands occupy the inner part of the area, and lowlands lie along the coast and narrow stretches penetrate between the mountains to the south. these valleys formed refugia where isolated moose populations survived. four moose populations that formed during the ice age are documented in the area (safronov 2008). further, during the sartan ice age there was a forest refugium in the middle reach of the anadyr (kozhevnikov and zheleznovchokotskij 1995) where a moose population most closely related to alaskan moose might have existed. range recession in eurasia moose were distributed across most of europe during the early holocene era (heptner et al. 1961, vereshchagin 1967). later the range retreated eastward; the last moose was killed in saxony in 1777 and in galicia in west ukraine in 1769 (gebel 1879). by the end of the 18th century moose were eliminated in belovezhskaya pusha (sablina 1955). in the beginning of the 20th century moose were still in east prussia (now kaliningrad region, russia; obermeier 1913) but were never encountered after world war i. thus, the european moose was preserved only in russia and the nordic countries by the mid20th century. european part of russia during the period of utmost population depression, the southern boundary of moose range retreated 450-1000 km northward, the northern boundary 500-600 km southward (danilkin 1999), and the range was fragmented (fig. 2). the northern boundary corresponded to the northern extent of the forest zone and reached 65° n in the ural mountains (sokolov 1959). the southern boundary coincided with the latitudinal flow of major rivers such as the volga, kama, and belaya rivers (filonov 1983). figure 2 depicts the location of 20 isolated moose populations; in the following text each population is described both temporally and geographically with an accompanying number [#] identified on figure 2. the main area was divided roughly into western [1] and eastern parts [2] along the vologda-arkhangelsk railroad and the white sea-baltic canal; there is no information regarding the duration of this fragmentation. the western part was characterized by irregular moose distribution of variable configuration and included the leningrad, pskov, novgorod, and tver regions, the western part of the vologda region, the northern part of the smolensk region, and north of byelorussia (serzhanin 1961). there is good reason to believe that a small breeding population survived in the area of the pripyat and pinsk [3] marshes (serzhanin 1961, galaka 1964). a small population of moose also survived in the bryansk forests [4] along the left bank area of the desna river (fedosov and nikitin 1951) and moose were also documented in the sumy region (galaka 1964). moose also survived extirpation in the meshcherskaya lowland [5], the boggy interfluve of the volga and oka rivers (severtsev 1854, kulagin 1932). one other isolated population existed in the area between the tsna and sura rivers [6] in the mordovia, penza, and tambov regions (filonov 1983). scandinavian peninsula current norwegian and swedish moose populations are abundant, yet in the beginning of the 19th century only small, isolated groups survived in the southwest of the scandinavian peninsula [7] (markgren1974, danell and bergström 2008). the same situation occurred in finland where moose disappeared by the mid-19th century (markgren1974, nygrén et al. 2008). single animals migrated gradually alces vol. 45, 2009 sipko and kholodova – eurasian moose population 29 from adjoining regions of south karelia [8] where the moose population was not abundant (vereshchagin and rusakov 1979). data on the kola peninsula [9] is rather contradictory. some authors (semenov-tjan-shanskij 1948, kirikov 1966) state that the species disappeared from the area and later repopulated it from adjoining territories. others (vereshchagin and rusakov 1979) quote data supporting that a moose population survived on the peninsula. west siberia the moose population between the ural mountains and the yenisei river in west siberia [10] was at its minimum by the beginning of the 1920s. the northern boundary arched southward reaching 63° n, and the range was a strip of land about 450 km wide (laptev 1958), with a 250-500 km wide gap in the region of the ob river dividing the area into eastern and western parts (laptev 1958, yurlov 1965). it is suggested that migration could occur between these areas, but insufficient evidence exists to support or refute this idea. this gap existed for an unknown period, but we believe it lasted no less than 100 years. south siberia the moose population of the altai-sayan mountains was at its minimum at the end of the 19th century (sobansky 1975). moose were also eliminated in the adjoining regions lying to the north and northwest in the kemerovo region in kuznetsk alatau (sobansky 1992), in the south altai, and the adjoining kazakhstan regions (sludsky 1953). moose were considered fully extirpated in altai (filonov 1983), nevertheless, a small population may have possibly survived in the upper reaches of the abakan, biya [12] (dmitriev 1938), katun, and tchuya rivers [11] (sobansky 1975). repopulation of the altai region resulted from migration from both the east and west (filonov 1983). moose was never abundant in the sayan fig. 2. the southern boundary of moose range retreated 450-1000 km northward and the northern boundary 500-600 km southward during the period of severest population depression, this figure identifies the location of 20 isolated moose populations during this period; each population is described both temporally and geographically in the text. eurasian moose population – sipko and kholodova alces vol. 45, 2009 30 region. in the beginning of the 20th century they disappeared from the khakas-minusinsk basin (skalon et al. 1941), as well as from the west sayan where only single animals were encountered migrating from the west. moose survived only in the east sayan, east of tuva, and in the eastern part of the tannu-olu range [13]; the latter were connected with a mongolian population (yanushevich 1952). the mongolian population occupied the area south of the altai-sayan region and was limited to mountainous taiga regions. moving west to east, moose left in the mongolian altai were encountered only at the beginning of the eastern part of the tanu-ola mountains. further east moose occupied an area around habsugul lake, the khangai mountains up to the khabgai-nuru range in the south, and the upper reaches of the onon, kerulen, and tola rivers and other rivers in the khantae uplands (bannikov 1954). east siberia this region was a northern siberian upland between the yenisei and lena rivers where the distribution of moose has changed considerably (michurin and mironenko 1967). during the period of lowest population at the end of the 19th-early 20th centuries, the range boundary crossed the yenisei river at 590 30’ n and went northeast crossing the podkamennaya tunguska river. it followed the verkhnyaya tunguska and podkamennaya tunguska divide to 1000 e, where it turned north to the upper reach of the kotui river basin, went along the kotui to 700 n, and then east to the mid-reach of the anabar river [14] (naumov 1934, heptner et al. 1961). according to middendorf (1869), in the mid-19th century moose were occasionally documented along the banks of the nizhnyaya tunguska river. however, maak (1887), whose expedition in 1854–55 visited the vilyuy river starting not far from the nizhnyaya tunguska river that flows east to the lena river, reported that moose were absent in the area and local people had no knowledge of them. moose were also absent in central yakutia (tavrovsky et al. 1971). we suggest that northern [14] and southern [15] populations in this region were isolated for an extended period. southeast siberia moose range has changed little in the region of the trans-baikal and amur basin [16]. the southern boundary along the sea of japan retreated northward to 440 45’ n (kaplanov 1948), and moose abandoned the ussuri and part of the amur bottomlands [18] (zhitkov 1914, rakov 1965). more apparent changes occurred in southwestern china where at the beginning of the 20th century moose were distributed to the north of the chita-kharbinvladivostock railroad (oshanin 1934, zhen zuoxin 1956). they inhabited the great and little khingan mountains and were in direct contact with moose on the opposite bank, often crossing the amur river (rakov 1964). thus, this chinese population and the adjacent russian population should be considered the same. one other population survived in east manchuria at the interfluve of the ussuri and sungari rivers [17] (abramov 1949); it was isolated long enough to develop different antler morphology than moose in the sikhote alin (rakov 1965). further, in an east [18]-west [16] direction along the chita–khabarovsk transect, average body weight increases stepwise to the west boundary of the khabarovsk territory (rozhkov et al. 2001) where the little khingan meets the bureinsky range. it is possible that these mountain ranges divide genetically and taxonomically different moose populations. northeast siberia moose disappeared before the beginning of the 18th century in the kamchatka peninsula (vereshchagin and nikolaev 1979) and sakhalin island (kozyrev 1960, alekseev 1974) alces vol. 45, 2009 sipko and kholodova – eurasian moose population 31 where none were documented by the first russian explorers. in the 1820s moose were rare in the whole region (mensbir 1878) and were not documented in the koryak district (fil and demyanyuk 1972). the range boundary along the pacific coast retreated westward to the kolyma tributary of the omolon river [19] (zheleznov 1990, sipko et al. 2004). moose did not inhabit the central axial zone of the mountain ranges of the verkhoyansk uplands (tavrovsky et al. 1971), though some survived in mountain valleys [20] forming 4 isolated populations (safronov 2009). repopulation of the kamchatka peninsula occurred from recent transplants (sipko et al. 2004, 2008). conclusion moose range in russia underwent repeated changes during the pleistocene and holocene eras in response to climatic and landscape changes. this complicated geographic history caused periods of isolation for particular local moose populations. some isolated populations were depressed severely and probably passed through a genetic bottleneck that influenced their local genetic structure. because certain populations demonstrate distinct morphological and physical differences, we suggest that the taxonomic and phylogenetic structure of moose range in eurasia is more complex than considered previously. further genetic studies of distinct populations would help elucidate such genetic relationships and differences. references abramov, k. g. 1949. the distribution, ecology, and economy of moose in priamurie. the nature branch of biology 54 (1): 1728. moscow, russia. (in russian). alekseev, a. i. 1974. the amur expedition, 1849-1855. publishing house of nauk academy, leningrad, russia. (in russian). bannikov, a. g. 1954. the mammals of the mongolian people’s republic. publishing house of the academy of sciences of ussr, moscow, russia. (in russian). boeskorov, g. g. 2001. systematization and origin of modern moose. siberian branch of the russian academy of science, novosibirsk, russia. (in russian). danell, k., and r. bergström. 2008. the history of moose management in sweden an 800 year perspective. 6th international moose symposium, yakutsk, russia, 1420 august, 2008. abstract only. danilkin, a. a. 1999. the deer (cervidae), mammals of russia and adjacent regions. moscow, russia. (in russian). davydov, a. v., o. p. piskunov, a. v. pronjaev, and j. i. rozhkov. 2004. spatial differentiation moose of eurasian moose by results of an estimation of the hunting trophies. bulletin of game biologists 1: 36-41. moscow, russia. (in russian). ditsevich, v. n. 1990. systematization and morphology of east siberian moose. 3rd international moose symposium. syktyvkar, russia, 27 august-5 september. abstract only. dmitriev, v. v. 1938. ungulates of the altaisky nature reserve and surroundings (east altai and west sayan). the altay reserve 1: 171-262. (in russian). fedosov, l. v., and k. n. nikitin. 1951. pages 6-8 in fauna of the bryansk area. bryansk publishing house, bryansk, russia. (in russian). fil, v. i., and v. p. demjanjuk. 1972. distribution and number of moose in the kamchatka area. russian scientific research institute of hunting and fur farming 34: 32-36. kirov, russia. (in russian). filonov, k. p. 1983. moose. lesnaya promyshlennost, moscow, russia. (in russian). galaka, b. a. 1964. expansion of moose range in ukraine. pages 35-43 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). gebel, g. 1879. influence of the person and its culture on an empire of animals. nature eurasian moose population – sipko and kholodova alces vol. 45, 2009 32 and hunting 1: 95-117. moscow, russia. (in russian). gerasimov, i. p., and k. k. markov. 1939. the glacial age in territory of the ussr. works of institute of geography, vol. 33, academy of sciences, moscow– leningrad, russia. (in russian). groswald, m. g. 1983. integumentary glaciers of continental shelves. science, moscow, russia. (in russian). _____. 1984. the euro-asian glacial cover. pages 130-131 in glaciology the dictionary gidrometeoizdat. leningrad, russia. (in russian). _____. 1998. materials glaciology. 84: 121129. moscow, russia. (in russian). _____. 1999. euro-asian hydrosphere accidents and a congelation of arctic regions. moscow, russia. (in russian). _____, and v. m. kotljakov. 1989. great glacial systems and drain of northern eurasia and its significance for inter-regional correlations (quaternary period). paleogeography and litologia 5-13. kishinev, shtiincza, ussr. (in russian). heptner, w. g., and a. a. nasimowitsch. 1974. der elch (the moose). ziesmanverlag, wittenburg lutherstadt, germany. (in german). _____, a. a. nasimovich, and a. g. bannikov. 1961. mammals of the soviet union. artiodactyls and non-artiodactyls. moscow, russia. (in russian). kaplanov, l. g. 1948. tiger, red deer, and moose. otdelenia zoologia no. 14 (29). moscow, russia. (in russian). kirikov, s. v. 1966. hunting animals, natural environment, and man. nauka, moscow, russia. (in russian). kozhevnikov, j. p., and n. k. zheleznovchukotskij. 1995. beringia: history and evolution. moscow, russia. (in russian). kozyrev, r. v. 1960. the ancient past of sakhalin. yuzhno-sakhalinsk, russia. (in russian). kulagin, n. m. 1932. moose of the ussr. publishing house of nauk academy, leningrad, russia. (in russian). kutcherenko, s. p. 1975. the distribution, number, and use of ussurijskogo moose. pages 108-109 in ungulate fauna of the ussr. мoscow, russia. (in russian). laptev, i. p. 1958. mammals of a taiga zone of western siberia. tomsk university publishing house, tomsk, russia. (in russian). maak, r. k. 1887. viljujsky district of the yakut area. st. petersburg, russia in 1994 edition: viljujsky district of the yakut area. yana publishing house, moscow, russia. (in russian). markgren, g. 1974. the moose in fennoscandia. canadian naturalist 101: 185-194. markov, g. g., and a. a. danilkin. 1996. kraniometricheskaja in eurasia (the characteristics of moose in eurasia). russian academy of sciences, biological series p: 244-247. moscow, russia. (in russian). menzbir, m. 1878. characteristics of distribution of animals. nature and hunting 10: 32-34. moscow, russia. (in russian). michurin, l. n., and o. n. mironenko. 1967. about moose in the putorano mountains. pages 72-75 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). middendorf, a. 1869. travel in northern and eastern siberia by siberian fauna. 2(5) st. petersburg, russia. (in russian). nasimovich, a. a. 1955. the role of snow cover in the lives of ungulates in the ussr. publishing house of the academy of sciences of ussr, moscow, russia. (in russian). naumov, н. п. 1934. mammals of the tungus district. work of the polar commission. publishing house of the academy of sciences of ussr, moscow, russia. (in russian). nygrén, k., p. danilov, and t. nugren. 2008. alces vol. 45, 2009 sipko and kholodova – eurasian moose population 33 the dynamics and management of north european moose. 6th international moose symposium. yakutsk, russia, 14-20 august, 2008. abstract only. obermeier, g. 1913. the prehistoric person. translated from german, p.j. schmidt. edition of brockhaus and efron, st. petersburg, russia. (in russian). oshanin, a. 1934. trade and sports hunting from northern manchuria. collection of works 7: 12-24. moscow, russia. (in russian). panov, n. k. 1999. history of development of vegetation of a mountain part of the southern urals mountains in the late pleistocene and holocene, on paleontological data. pages 144-158 in n. g., smirnova, editor. historical ecology of animals in the southern ural mountains. sverdlovsk, russia. (in russian). preobrazhenskiy, v. s. 1960. kodarskiy glacial region (trans-baikal). ninth scetion of igy program (glaciology). no. 4. publishing house of the academy of sciences of ussr, moscow, russia. (in russian). rakov, n. v. 1964. geography and ecology of moose populations in amur-ussurijskij kraj. pages 44-100 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). _____. 1965. distribution and ecology of moose in amur-ussurijskij kraj. pages 28-65 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). rozhkov, j. i., a. v. pronjaev, o. v. piskunov, n. e. ovsjukova, a. v. davydov, and l. v. rozhkova. 2001. moose. the population-biological analysis of license information. hunting russian animals. issue 4. tsentrohotkontrol, moscow, russia. (in russian). _____, _____, _____, _____, _____, and _____. 2002. moose of the european part of the russian federation. population division, selection, seasonal changes, morphological, and ethological differentiation (by results of the analysis of the license information). pages 153-158 in questions for modern game biologists. tsentrohotkontrol, moscow, russia. (in russian). sablina, t. b. 1955. ungulate belovezhskoi pushi. page 15 in a. n. severtsova, editor. works of the institute of morphology and ecology of animals. moscow, russia. (in russian). safronov, v. m. 2009. regional populations and migration of moose in northern yakutia, russia. alces 45: 17-20. semenov-tjan-shanskij, o. i. 1948. moose on the kola peninsula. works of the lappish reserve 2: 91-162. moscow, russia. (in russian). serzhanin, i. n. 1961. mammals of belarus. minsk publishing house, izdatelstvo akademii nauk, belorusskoi ssr, minsk, russia. (in russian). severtsev, n. a. 1854. moose or sohatii. bulletin of natural sciences 19: 290-299. sipko t. p., v. i. fil, and a. r. gruzdev. 2004. resettlement of moose. the siberian zoological conference, novosibirsk, russia, 15-22 september, 2004. abstract only. _____, _____, and _____. 2008. moving moose on kamchatka. 6th international moose symposium, yakutsk, russia, 1420 august, 2008. abstract only. skalon, v. n., i. scherbakov, and m. bazikin. 1941. nature and socialist facilities. volume 8. moscow, russia. (in russian). sludsky, a. a. 1953. eviction of “taiga” animals in the forest-steppe and steppes of western siberia and kazakhstan. the nature branch of biology 53 (2): 14-32. moscow, russia. (in russian). sobansky, g. g. 1975. territorial accommodation of ungulates in the altay mountains. pages 122-123 in ungulate fauna of the ussr. moscow, russia. (in russian). eurasian moose population – sipko and kholodova alces vol. 45, 2009 34 _____. 1992. ungulates in the altai mountains. novosibirsk, russia. (in russian). sokolov, i. i. 1959. ungulates (perissodactyla and artiodactyla). in fauna of the ussr. mammals. nauka publishing house, moscow, russia. (in russian). tavrovsky, v. a., o. v. egorov, and v. g. krivosheev. 1971. mammals of yakutia. nauka, moscow, russia. (in russian). tikhonov, a. n. 1990. history and systematization moose. 3rd international moose symposium. syktyvkar, russia, 27 august-5 september. abstract only. timofeeva, e. k. 1974. moose: ecology, distribution, and economic importance. leningrad university publishing house, leningrad, russia. (in russian). vereshchagin, n. k. 1967. the geological history of moose and its relationship to man. pages 3-37 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). _____, and a. nikolaev и. 1979. trade animals of neolith tribes of kamchatka. the nature branch of biology 84 (5): 40-44. moscow, russia. (in russian). _____, and o. s. rusakov. 1979. ungulates from the north-western part of the ussr. nauka, leningrad, russia. (in russian). woillard, g. 1979. abrupt end of the last interglacial in north east france. nature 281 no. 5732: 558-562. yanushevich, a. i. 1952. vertebrate fauna of the tuva area. zap-sib publishing house, novosibirsk, russia. (in russian). yazan, j. p. 1966. moose in the northern caucasus. the nature branch of biology 71 (4): 123-125. moscow, russia. (in russian). yurlov, k. t. 1965. change in moose range in the southern west-siberian lowlands. pages 17-27 in biology and economy of the moose. rosselhozizdat, moscow, russia. (in russian). zheleznov, n. k. 1990. wild ungulates in northeastern ussr. vladivostok, russia. (in russian). zhen zuoxin. 1956. animals and birds of china. the nature journal 10: 3-18. moscow, russia. (in russian). zhitkov, b. m. 1914. moose in ussurijskogo kraj. zoological branch of natural sciences: a new series 2 (3): 5-14. st. petersburg, russia. (in russian). diurnal defecation rate of moose in southwest finland juho matala1 and antti uotila2 1finnish forest research institute metla, p.o. box 68, fi-80101 joensuu, finland; 2university of helsinki hyytiälä forest station, hyytiäläntie 124, fi-35500 korkeakoski, finland. abstract: an accurate measure of defecation rate is essential for application of pellet group counts in moose (alces alces) population estimates. we measured the wintertime, diurnal defecation rate of moose by tracking 7 gps-collared and 22 uncollared moose in southwest finland. the mean defecation rate was 23.5 ± 4.2 pellet groups/d, one of the highest values reported. the mean defecation rate did not differ between the tracking methods (gps vs. uncollared moose); limited sample size precluded conclusions about sex and age differences. the defecation rate was not correlated with calendar week, length of accumulation period, or number of diurnal beds. our results are appropriate for use in southwest finland when using the pellet group method to assess moose population density. alces vol. 49: 155–161 (2013) key words: alces alces, finland, defecation rate, moose, pellet group, tracking. counting fecal pellet groups of moose (alces alces) has been widely used to estimate habitat utilization, feeding behavior, and population trends and density (see franzmann et al. 1976b, forbes and theberge 1993, härkönen and heikkilä 1999, rönnegård et al. 2008, månsson 2009, månsson et al. 2011a, b). reliable moose population estimates are not always realized from pellet group counts (rönnegård et al. 2008), but the method's usefulness has been noted (neff 1968, lautenschlager and jordan 1993, månsson et al. 2011b), despite some uncertainty (neff 1968). however, to successfully estimate moose population density using a pellet group count, it is critical to use an accurate defecation rate in the survey area and time period (timmermann 1974). earlier studies have used 2 main methods to estimate the defecation rates of moose: 1) track moose in snow-covered terrain (joyal and ricard 1986, andersen et al. 1992), and 2) estimate the number of pellet groups in a closed area or island where the number of moose is known (jordan et al. 1993). defecation rates have occasionally been estimated by comparing the aerial censuses and pellet group counts in specific areas (rönnegård et al. 2008). moose enclosures could also be utilized, but results from domestic moose can be affected by food quality and behaviour that are dissimilar to natural conditions. gps radio-collars enable intensive and accurate tracking by identifying specific individuals and the beginning and end points of their specific tracks, providing ideal conditions to measure defecation rates of free-ranging moose. moose defecation rates vary by age, sex, habitat, food quality, season, and year (desmeules 1968, franzmann et al. 1976a, oldemeyer and franzmann 1981, joyal and ricard 1986, andersen et al. 1992, månsson et al. 2011b). large variations in defecation rate have been reported in different areas; for example, in north america variation was 9.6–32.2 pellet groups/d (timmermann 1974), and in northern europe rates varied corresponding author: juho matala, finnish forest research institute metla, p.o.box 68, fi-80101 joensuu, finland. +358 40 801 5275. juho.matala@metla.fi 155 from 14–26.9 pellet groups/d (andersen et al. 1992, remm and luud 2003, rönnegård et al. 2008). these data emphasize the importance of using area-specific rates when using the pellet group method to estimate moose population density. our main objective was to formulate a general estimate of the wintertime, diurnal defecation rate of moose in southwest finland by tracking both gps-collared and uncollared moose in snowy terrain. we also compared these 2 tracking methods and looked for differences in defecation rates between sex and age. study area moose were tracked in 2 separate areas approximately 100 km apart in southwest finland (fig. 1). uncollared moose were tracked in the orivesi-kangasala area (∼wgs84 61°36′ n, 24°22′ e) and gpscollared moose in the loppi-hyvinkää area (∼wgs84 60°38′ n, 24°35′ e). both areas are located in the southern boreal vegetation zone with scots pine (pinus sylvestris) and norway spruce (picea abies) as the dominant tree species. forest cover was 78% of the total land area in orivesi-kangasala and 71% in loppi-hyvinkää (metla 2012). methods tracking of uncollared moose we actively searched for uncollared moose in their known habitats during fresh snow conditions between december and april, 1999–2003. the accumulation period for pellet counts began when moose were seen or flushed, enabling data collectors to accurately time their count by locating fresh resting places, pellet groups, or tracks in fig. 1. study area locations and starting points of moose tracking periods in southern finland. in the orivesi-kangasala area some of the starting point coordinates of uncollared moose are rough estimates and overlap because they could not be separated at the map scale. 156 diurnal defecation rate of moose – matala and uotila alces vol. 49, 2013 the snow. the pellet groups were counted the following day, until moose fled from the counter. the end of the accumulation period was determined in a similar manner as the starting point. the accumulation periods varied from 8–31 h. moose were classified as either adults or calves by visually observing them and their pellets, bedding places, and behavior. their sex could be determined without visual contact by analyzing their urination methods, as bulls urinate in front of the hind hooves. we measured the diurnal defecation rate of 5 bulls, 3 cows, and 3 calves. additionally we measured 3 cows with twin calves and 1 cow with a single calf, without separating the pellet groups of the cows and calves. in total, we were able to record the diurnal defecation rates of 22 uncollared moose and count the number of beds of 7 individuals. tracking of gps-collared moose the finnish game and fisheries research institute (fgfri) provided location data for the gps-collared moose. the fgfri implemented gps-collaring procedures in accordance with finnish legislation, with permission from the national animal experiment board of finland. an individual gps-collared cow was tracked once and another on 5 separate occasions, while 1 cow with twin calves was tracked once, and one cow with a single calf twice; altogether, 7 individuals were tracked. the gps-collared moose were tracked a few days after fresh snowfall in december– march 2010. the counter went to the most recent location of moose tracks which were usually ∼12 h old. the pellet groups were counted along the moose track by following it against the original course of the moose; the coordinates of the pellet groups, beds, and urination sites were located with handheld gps devices. the accumulation period finished when an individual track became mixed with others or sunset precluded tracking. the duration of the accumulation period was determined by identifying the time associated with the closest location of the tracks with gps collar data. the accumulation periods ranged from 6–47 h. data analysis at least 20 pellets were required to make a pellet group. we processed the pellet group data of 29 moose (7 gps-collared, 22 uncollared) to calculate the diurnal defecation rate (number of pellet groups produced per individual in 24 h) from the accumulation periods of individual bulls, cows, and calves. mean values were calculated for the group of cows and calves when it was impossible to identify calf from cow pellet groups; analyses of 25 separate cases were used to calculate the diurnal defecation rate. for comparison, we sampled the mean values of diurnal defecation rates of 3 sex and age classes: 1) bulls, 2) cows, and 3) calves and cow-calves. the last class was required because we were only able to track 3 individual calves, which was insufficient for any reasonable analysis. we also compared the mean defecation rates of the uncollared and gps-collared moose. due to limitations in the linearity and homogeneity of variances in the data, we used the kruskal-wallis test for multiple samples and the mann-whitney test for paired sample comparisons. furthermore, we searched for possible correlations between the diurnal defecation rate and 1) the calendar week of the accumulation period, 2) the duration of the accumulation period, and 3) the number of diurnal beds for part of the samples. all statistical calculations were performed using the statistical package for the social sciences (spss) 17.0 software. data relating to home range of moose during the accumulation period (calculated using the home range tools for arcgis® version 1.1 with the minimum convex polygon method), length of the moose track alces vol. 49, 2013 matala and uotila – diurnal defecation rate of moose 157 during the accumulation period, and the number of diurnal urinations were also collected from the gps-collared moose, but insufficient sample sizes precluded their utilization in the analysis. results the diurnal defecation rate ranged from 12.2–32 pellet groups with an overall mean of 23.5 ± 4.2 (sd; table 1, fig. 2). the mean values of bull, cow, and calf/cow-calf groups were different (kruskal-wallis test: χ2 = 9.9, df = 2, p = 0.007; table 1). the calf and cow-calf group had the highest mean rate, but was statistically different only from the cow group (table 1). the bull group had the lowest mean rate, but also the lowest and highest absolute values (i.e., the widest range; table 1). the cow and bull group rates and the rates of the uncollared and gps-collared moose were not different (table 1). the diurnal defecation rate was not related to the calendar week of the accumulation period (fig. 2, table 2). the gpscollared individual which was tracked 5 times between 17 january and 23 february 2010 showed no trends during this period. the defecation rate did not correlate to the duration of the accumulation period or to the number of diurnal beds (table 2). the mean values for the other variables were: diurnal number of beds = 8.0 ± 4.2 (sd, n = 15), the diurnal urination rate = 1.0 ± 1.0 (sd, n = 9), the area of home range during the accumulation period = 95,267 m2 ± 122,399 (sd, n = 9), and length of track during the accumulation period = 955 m ± 832 (sd, n = 6). discussion the mean defecation rate was considered high (23.5 ± 4.2 pellet groups/d) but similar to the average rate measured in relatively good moose habitat in southern norway (22.9 pellet groups/d; andersen et al. 1992). lower defecation rates were measured in southern sweden (14 pellet groups/d; rönneberg et al. 2008) and on isle royale, north america (20.9; jordan et al. 1993). many studies have reported lower defecation rates in alaska and canada (see desmeules 1968, franzmann et al. 1976a, oldemeyer and franzmann 1981, joyal and ricard 1986). the high values reported in our study area are not unreasonable when comparing the status of the moose population in finland to other nordic countries. the finnish moose population is lower and of a higher productive state compared to those in table 1. mean values of the wintertime, diurnal defecation rate and comparisons between moose groups in southern finland. pellet groups (#/ind/24 h) mann-whitney test mean min max n sd test against u-value p grouping by moose type: bulls 20.1 12.2 32.0 5 7.3 cow 12 0.152 cows 22.3 18.9 24.0 9 2.2 calf + groups 10 0.002 calf and cow-calf groups 25.9 23.0 31.3 11 2.2 bull 11 0.061 grouping by tracking method: gps-collared 23.5 20 26.7 9 1.8 uncollared 64 0.647 uncollared 23.4 12.2 32.0 16 5.2 all moose 23.5 12.2 32.0 25 4.2 158 diurnal defecation rate of moose – matala and uotila alces vol. 49, 2013 sweden and norway (lavsund et al. 2003, tiilikainen et al. 2012). these differences presumably imply better foraging habitat, higher nutritional condition, and a resultant higher defecation rate in finland. because defecation rates are influenced by food quality and availability and show large variation (andersen et al. 1992), it follows that areaspecific defecation rates should be measured when using pellet group counts to estimate population density. the extremes in defecation rate varied largely in our study, but most observations, especially of the gps-collared individuals, were similar to the mean indicating the general reliability of these data and their application to estimate population density. the extreme values could result from local foraging conditions or from longer movements associated with unintended disturbances while tracking the uncollared moose. because the defecation rate can correlate with age and sex, these relationships should be taken into account if a change in population structure occurs (franzmann et al. 1976a). our data indicate that the calf and cow-calf group had slightly higher defecation rates than individual cows as also reported by desmeules (1968), but opposite that of joyal and ricard (1986). bulls and cows were not different in our analyses; however, general conclusions about sex and age differences cannot be made due to the limited sample size. using an accurate defecation rate is critical when applying pellet group counts in moose population density estimates. defecation rates have not been published previously in finland, and given the considerable variation in similar data from northern europe, we consider our results the best available for local use. furthermore, similar defecation rates were obtained for both regions in our study, making the mean values applicable to all of southern finland. the lack of correlation between calendar week and defecation rate indicates a stable accumulation rate throughout winter, suggesting that these rates might be applicable for springtime pellet group counts. variations in the absolute data and means should be taken into account when calculating confidence limits for the final moose density estimate. acknowledgements we thank mr. j. sillanpää for tracking the gps-collared moose and mr. h. mattila 50 fig. 2. the diurnal defecation rate of moose in southern finland relative to the time of the tracking period (calendar weeks). moose types are classified as: 1) individual bulls, calves, and cows, and 2) cow-calf groups (when individuals could not be separated during tracking); cow1 = cow with single calf, cow2 = cow with 2 calves. table 2. pearson correlations between the diurnal defecation rate (pellet groups/moose/24 h) and temporal variables in southern finland; no differences were found. calendar week accumulation period (h) beds/ moose/ 24 h coefficient −0.144 −0.086 −0.403 p-value 0.492 0.682 0.136 n 25 25 15 alces vol. 49, 2013 matala and uotila – diurnal defecation rate of moose 159 for tracking in the kangasala area. dr. j. pusenius from the finnish game and fisheries research institute is greatly acknowledged for organizing the gps-collaring and thus enabling data availability on exact moose locations for our study. we also thank the finnish forest research institute, the university of helsinki, and the ministry of agriculture and forestry for providing financing for our work. m.sc. m. melin is acknowledged for producing the maps in fig. 1 and ms. s.thompson for checking the language of this manuscript. references andersen, r., o. hjeljord, and b-e. saether. 1992. moose defecation rates in relation to habitat quality. alces 28: 95–100. desmeules, p. 1968. détermination du nombre de tas de crottins rejetés et du nombre de reposées établies, par jour, par l'original (alces alces), en hiver. le naturaliste canadien 95: 1153–1157. (in french with english abstract). franzmann, a. w., p. d. arneson, and j. l. oldemeyer. 1976a. daily winter pellet groups and beds of alaskan moose. journal of wildlife management 40: 374–375. ———, j. l. oldemeyer, p. d. arneson, and r. k. seemel. 1976b. pellet-group count evaluation for census and habitat use of alaskan moose. proceedings of north american moose conference workshop 12: 127–142. forbes, g. j., and j. b. theberge. 1993. multiple landscape scales and winter distribution of moose, alces alces, in a forest ecotone. canadian field-naturalist 107: 201–207. härkönen, s., and r. heikkilä. 1999. use of pellet-group counts in determining density and habitat use of moose alces alces in finland. wildlife biology 5: 233–239. jordan, p. a., r. o. peterson, p. campbell, and b. mclaren. 1993. comparison of pellet-group counts and aerial counts for estimating density of moose at isle royale: a progress report. alces 29: 267–278. joyal, r., and j-g. ricard. 1986. winter defecation output and bedding frequency of wild, free ranging moose. journal of wildlife management 50: 734–736. lautenschlager, r. a., and p. a. jordan. 1993. potential use of track-pellet group counts for moose censusing. alces 29: 175–179. lavsund, s., t. nygrén, and e. j. solberg. 2003. status of moose populations and challenges to moose management in scandinavia. alces 39: 109–130. månsson, j. 2009. environmental variation and moose alces alces density as determinants of spatio-temporal heterogeneity in browsing. ecography 32: 601–612. ———, h. andrén, and h. sand. 2011a. can pellet counts be used to accurately describe winter habitat selection by moose alces alces. european journal of wildlife research 57: 1017–1023. ———, c. e. hauser, h. andrén, and h. p. possingham. 2011b. survey method choice for wildlife management: the case of moose alces alces in sweden. wildlife biology 17: 176–190. metla. 2012. finnish forest research institute, national forest inventory. http:// www.metla.fi/metinfo/vmi/ . neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. journal of wildlife management 32: 597–614. oldemeyer, j. l., and a. w. franzmann. 1981. estimating winter defecation rates of moose, alces alces. canadian fieldnaturalist 95: 208–209. remm, k., and a. luud. 2003. regression and point pattern models of moose distribution in relation to habitat distribution and human influence in ida-viru county, estonia. journal for nature conservation 11: 197–211. 160 diurnal defecation rate of moose – matala and uotila alces vol. 49, 2013 http://www.metla.fi/metinfo/vmi/ http://www.metla.fi/metinfo/vmi/ rönnegård, l., h. sand, h. andrén, j. månsson, and å. pehrson. 2008. evaluation of four methods used to estimate population density of moose alces alces. wildlife biology 14: 358–371. tiilikainen, r., e. j. solberg, t. nygrén, and j. pusenius. 2012. spatio-temporal relationship between calf body mass and population productivity in fennoscandian moose alces alces. wildlife biology 18: 304–317. timmermann, h. r. 1974. moose inventory methods: a review. naturaliste canadien 101: 615–629. alces vol. 49, 2013 matala and uotila – diurnal defecation rate of moose 161 diurnal defecation rate of moose in southwest finland study area methods tracking of uncollared moose tracking of gps-ollared moose data analysis results discussion acknowledgements references impact of moose browsing on forest regeneration in northeast vermont haley a. andreozzi1, peter j. pekins1, and matt l. langlais2 1department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa; 2department of forests, parks & recreation, st. johnsbury, vt 05819, usa abstract: moose (alces alces) play an important role in the ecological and economic resources of northern new england, a landscape dominated by commercial forests. this study measured the impact of moose browsing on forest regeneration in wildlife management unit e1 in northeastern vermont where moose density was considered high in the 1990–2000s. we surveyed 37 clearcuts categorized into 4 age classes (3–5, 6–10, 11–15, and 16–20 years old). the stocking rate (stems/plot) of commercial species ranged from 74–76% in the 3–5, 6–10, and 11–15 year age classes, increasing to 86% in the 16–20 year age class. the proportion of plots containing a commercial tree without severe damage was above the accepted threshold stocking level of 40–60% in all age classes. the proportion of plots containing a commercial hardwood stem declined with increasing age class; the opposite occurred with softwood stems indicating a possible shift from hardwoodto softwood-dominated stands from selective browsing pressure. height of 11–20 year old stems was less than in new hampshire, indicating that growth was possibly suppressed in vermont due to higher moose density. overall, browsing was not considered a major problem based upon stocking rates. further study is warranted to evaluate whether compensatory growth occurs in response to reduced browsing as forests age and/or moose population density declines. alces vol. 50: 67–79 (2014) key words: alces alces, browsing, clearcut, damage, moose, new england, regeneration, stocking. moose (alces alces) populations have experienced a regional increase in northern new england over the last several decades, making them an increasingly valuable wildlife resource. they play an important role ecologically and economically in vermont, with 78% of the state open to regulated moose hunting and 406 hunting permits issued statewide in 2011 (vtfw 2008, 2011). with forests also covering 78% of vermont's landscape, the state generates over $1.5 billion annually from forest-based manufacturing and forest-related recreation and tourism (nefa 2007). the majority of forestland, 4 million acres, is owned privately or by timber investment management organizations; local, state, and fed‐ eral government owns ∼19% (919,440 acres) (nefa 2007). forest and wildlife management aimed at sustainable forest production is critical for the long-term stability of both vermont's economy and moose population. with adult moose weighing 300–600 kg, substantial browse is required to maintain such large body size (bubenik 1997), estimated at daily dry matter intake of 2.8 kg/ moose/day in january (pruss and pekins 1992). moose have the ability to substantially alter plant communities and are capable of damaging woody plants (renecker and schwartz 1997); repeated browsing can suppress height growth and recruitment of saplings into the canopy (risenhoover and maass 1987). moose browsing has the capability to affect the structure and dynamics of forest ecosystems over the long-term (mcinnes et al. 1992), which has important 67 implications for the management of forests where moose populations are regulated. moose show preference for forage in clearcut and early-successional habitat that is typical of the commercially managed forests of the northeast (thompson et al. 1995, scarpitti et al. 2005). for example, productive moose habitat in new hampshire was linked directly to the early successional forage created by commercial forest harvesting and early-successional browse is a dietary component year-round (scarpitti et al. 2005, scarpitti 2006). clearcuts 5–20 years old provide suitable early winter habitat, as regenerating hardwood and softwood species provide both browse and cover for moose (thompson and stewart 1997). while the impact of moose browsing on forest regeneration has received substantial attention elsewhere, little attention has been paid to the potential and actual effects in northern new england (pruss and pekins 1992, scarpitti 2006, bergeron et al. 2011). in order to manage moose and forest resources with respect to moose density and damage to regeneration, it is important to have extensive ecological knowledge of the relationships among moose, the ecosystems they inhabit, the plants they use as forage (edenius et al. 2002), and the associated impacts on forest production such as timber quality impairment. as moose populations have increased in northern new england, land managers have implied that a relationship exists between high population density and reduced forest regeneration in clearcuts. on isle royale, mcinnes et al. (1992) found that moose browsing affected the structure and dynamics of forest ecosystems on a long-term scale; however, in larger landscapes such impacts are usually more localized and often relate to high seasonal density. in northern new hampshire, bergeron et al. (2011) evaluated the impact of browsing on the regeneration of commercial tree species in 3 regions with different moose population density (0.26–0.83 moose/km2). while regeneration of commercial trees was not considered a regional problem at any density, specific clearcut sites with low regeneration were found adjacent to traditional moose wintering areas. it was predicted that such sites could change from hardwood to softwood dominance over time (bergeron et al. 2011). by the early 2000s, there was anecdotal evidence that the moose population in northeastern vermont, specifically wildlife management unit (wmu) e, was causing measurable damage to forest regeneration; moose densities in wmu e were thought to be well over 1.5 moose/km2 (4 moose/ mile2) (c. alexander, vtfw wildlife biologist, pers. comm.). to achieve the desired population level, hunting permit numbers were dramatically increased by the vermont department of fish and wildlife (vtfw) from 440 to 833 permits in 2004, when it was believed moose had approached their biological carrying capacity (vtfw 2008). the number of hunting permits rose to 1046 in 2005 and continued to increase until 2009, when 1223 permits were issued statewide in an effort to accelerate population reductions to protect forest habitat. by 2008, the population density was approaching the goal set by the 10-year moose management plan (0.7 moose/km2 [1.75 moose/mile2]) and the number of permits was reduced to 765 in 2010 and 405 in 2011. in response, this study was designed to evaluate the impact of moose browsing on the regeneration of commercial tree species in wmu e1 in northeast vermont by conducting qualitative assessments of damage in clearcuts between 3–20 years of age. methods study area the study area was located in north‐ east vermont and encompassed all of 68 browsing and forest regeneration – andreozzi et al. alces vol. 50 vtfw wmu e1, covering an area of 682 km2 bordered by new hampshire and quebec (fig. 1). elevation ranges from ∼250–1,130 m, and it is dominated by maple (acer saccharum, a. pensylvanicum, a. rubrum) and birch (betula alleghaniensis, b. papyrifera) hardwoods, and coniferous stands of balsam fir (abies balsamea) and red spruce (picea rubens.) while heavily forested, timber harvesting is common throughout as the majority of the land is privately owned and commercially harvested (nefa 2007). the 2011 moose density was estimated at 0.77 moose/km2 (1.96 moose/ mi2) based on a rolling 3-year average of moose sightings by early winter (november) deer hunters, and was previously estimated in 2010 as 0.93 moose/km2 (2.41 moose/ mile2) based on aerial surveys (millette et al. 2011). field measurements regeneration surveys were performed in june 2012 to measure the impact of moose browsing on forest regeneration in clearcuts 3–20 years old (leak 2007, bergeron et al. 2011). clear-cuts were separated into 4 age classes (3–5, 6–10, 11–15 and 16–20 years old) in order to assess temporal changes during both the period of typical browsing (0–10 years) and at least 10 years post-browsing (11–20 years). in each age class, 8–11 clear-cuts were located using aerial photography; each was a minimum of 4.1 ha (10 acres) and a maximum of 16.2 ha (40 acres) in size to reflect the typical range in size of clear-cuts in the region (m. langlais, vermont department of forests, parks & recreation county forester, pers. comm.). in certain cases, clear-cuts >16.2 ha were used to achieve appropriate sample sizes within an age class; a section ≤16.2 ha was surveyed. small plot surveys (milacre, ∼2.3 m diameter circle) were evenly spaced on equidistant transects throughout each clear-cut (fig. 2). in each milacre plot, the dominant stem (tallest tree) was recorded as a commercial or non-commercial tree species. if the dominant stem was non-commercial, the plot was searched for the presence of commercial species; commercial species included yellow and white birch, sugar and red maple, american beech (fagus grandifolia), aspen (populus spp.), black cherry (prunus serotina), balsam fir, red and black spruce (picea mariana), and tamarack (larix laricina). stem damage was assessed on a qualitative basis as fork, broom, or crook (fig. 3). the height of the damage above or below breast height (approximately 1.4 m) was recorded, as well as the number of forks and crooks, and the severity of crooks based on angle. light crooks were those ≤30°, moderate crooks were those 30–60°, and severe crooks were those ≥60° from the dominant stem. the relative height of the fig. 1. the location of the study area in vermont used to assess the impact of moose browsing on forest regeneration, 2012. the area included all of wmu e1 in northeast vermont. alces vol. 50 andreozzi et al. – browsing and forest regeneration 69 dominant stem was estimated to the nearest foot when <3.05 m (10 ft), or as ≥3.05 m. data analysis broomed stems and multiple forks above breast height were considered browse defects indicative of a severely damaged tree; otherwise, damage was considered light or moderate. trees with lesser damage are expected to recover during future growth (switzenberg et al. 1955, carvell 1967, trimble 1968, jacobs 1969). stems with single forks above breast height, or multiple forks either above or below breast height were considered to have moderate damage. stems with a single fork below breast height or crooks were considered to have light damage. a fully stocked stand (average density for undisturbed stand) at 80 years was assumed if a minimum of 40–60% of plots (threshold) contained a dominant commercial stem without severe damage (leak et al. 1987). to evaluate relative height between age classes and further assess browse impact, comparisons were made of the proportion of plots containing a dominant commercial stem ≥3.05 m height without severe damage, as vegetation ≥3.05 m was presumed to be above the typical height of moose browsing (bergström and danell 1986). temporal comparisons were made to assess if younger age classes with high initial browse pressure recover to fully stocked stands after 10–15 years. analysis of variance (anova) and pairwise tukey's test were used to look for differences in browse damage between clear-cuts and age classes. analyses were performed with systat v. 13. significance for all tests was assigned a fig. 2. example of the sampling design used to measure browse damage in clearcuts in northeast vermont, summer 2012. equidistant transects were established upon which 100– 400, 2.3 m diameter plots were established to measure the presence of dominant commercial stems, stem quality, and relative height; modeled after bergeron et al. 2011. fig. 3. the 3 qualitative browse categories used to describe browsing damage of dominant stems in milacre sample plots (bergeron et al. 2011). 70 browsing and forest regeneration – andreozzi et al. alces vol. 50 priori at α = 0.05. results are presented throughout as ± se. results a total of 37 clearcuts were surveyed: 11, 8, 8, and 10 in the 3–5, 6–10, 11–15, and 16–20 year age classes, respectively. there were 1709, 1291, 1442, and 1585 milacre plots surveyed in the 4 age classes, respectively. stocking rate of commercial trees (stems/plot) was high in all age classes, and increased with age class (table 1); it ranged from 74–76% in the 3–5, 6–10, and 11– 15 year age classes, increasing to 86% in the 16–20 year age class. the proportion of commercial trees with severe damage was low overall, with <10% damaged severely in all age classes except in the 16–20 age class (11%, table 1). the proportion of plots containing a commercial tree without severe damage was above the defined threshold stocking level of 40–60% in all age classes (table 1, fig. 4), ranging from 67–68% in the 3–5, 6–10, and 11–15 year age classes, and increasing to 75% in the 16–20 year class. the proportion of dominant commercial trees ≥3.05 m without severe damage increased with age class with 1, 25, and 39% in the 6–10, 11–15 and 16–20 year age classes, respectively. the proportion of plots containing a commercial hardwood stem declined with age class, averaging 62, 51, 43, and 40% in the 4 age classes, respectively. conversely, the proportion of plots containing a commercial dominant softwood stem increased with age class, averaging 12, 24, 33 and 46% in the 4 age classes, respectively (fig. 5). the highest stocking rates (>80%) were restricted to softwood-dominated stands. the majority of plots with a dominant non-commercial stem also contained commercial stems (70–81% across age classes). the stocking rate of dominant commercial trees was lower (p = 0.02) in the 3–5 year age class than in the 16–20 year age class, although stocking rate was above the threshold stocking level in all age classes. the proportion of dominant commercial hardwoods was higher (p = 0.014) and the proportion of dominant commercial softwoods lower (p = 0.015) in the 3–5 year age class than in the 16–20 year age class. the proportion of plots beyond browse height (≥3.05 m) and without severe damage in the 6–10 year age class was lower than the 11–15 year (p = 0.022) and the 16–20 year age classes (p < 0.001). at least 3 commercial species accounted for ≥50% of the species composition within each age class (table 2). the majority of these species were classified with light to no damage, and the proportion of noncommercial species declined as age class increased (tables 1 and 2). the proportion of dominant commercial stems classified as hardwood declined with age class, averaging 83 ± 7.6, 69 ± 8.5, 58 ± 8.5 and 49 ± 7.2% in the 4 age classes, respectively; the opposite occurred with the proportion of dominant commercial stems classified as softwood that averaged 17 ± 7.6, 31 ± 8.5, 42 ± 8.5, and 51 ± 7.2%. red maple and yellow birch accounted for 24 and 20% of total species composition in the 3–5 year age class; no other commercial species accounted for more than 6%. in the 6–10 year class, red maple, balsam fir, and yellow birch accounted for the highest proportion of species composition (14–16% each) and in the 11–15 year age class, these 3 species accounted for 11–17% each, and red spruce 11%. red maple, balsam fir, and red spruce accounted for the greatest proportion of dominant commercial stems (21–23% each) in the 16–20 year age class; yellow birch fell to 6% (table 2). discussion overall, the impact of moose browsing on the regeneration of commercial tree alces vol. 50 andreozzi et al. – browsing and forest regeneration 71 table 1. summary values indicating the stocking of commercial tree species, stocking of commercial trees with and without severe damage, the proportion of commercial trees ≥3.05 m in height without severe damage, and the proportion of dominant commercial hardwood and softwood stems in clearcuts in northeastern vermont, 2012. rows with the same letter within columns are not statistically different (p > 0.05). age class stocking rate of dominant commercial trees (%)1 stocking rate of dominant commercial trees w/o severe damage (%)2 stocking rate of dominant commercial trees w/ severe damage (%)3 proportion of dominant commercial trees w/o severe damage and ≥3.05 m tall (%) proportion of dominant commercial hardwoods (%) proportion of dominant commercial softwoods (%) 3–5 74a 67 6 n/a 83a 17a 6–10 75ab 68 7 1a 69ab 31ab 11–15 76ab 67 9 25b 58ab 42ab 16–20 86b 75 11 39b 49b 51b 1proportion of plots containing dominant stems considered to be commercial species 2proportion of plots containing dominant commercial stems not considered to be severely damaged 3proportion of plots containing dominant commercial stems considered to be severely damaged fig. 4. stocking guide for main crown canopy of even-aged hardwood and mixed-wood stands relative to basal area, number of trees per acre, and mean stand diameter. the a-line is fully stocked, the b-line is suggested residual stocking (˜60%), and the c-line is minimum stocking (˜40%) (leak et al. 1987). the proportion (%) of commercial trees without severe damage are plotted by age class; stocking is projected to a 4" mean stand diameter. 72 browsing and forest regeneration – andreozzi et al. alces vol. 50 species in northeast vermont was considered minor. the stocking rate of commercial trees without severe damage was acceptable in all age classes based upon the minimum threshold stocking level of 40–60%, and severe damage from browsing was low in all age classes with regard to acceptable levels, ranging from 6–11% (table 1). while damage was low in all age classes, site-specific severe browsing can shift species composition (edenius et al. 2002). for example, moose drastically altered localized species composition on isle royale, michigan where browsed sites had lower overall tree density than unbrowsed sites due to decline in balsam fir and mountain ash (sorbus americana) and concurrent increase in white spruce (picea glauca) densities (snyder and janke 1976). the increasing proportion of dominant softwood stems with age indicates a possible shift to softwood-dominated stands due to selective browsing of hardwood species (fig. 5). the highest stocking rates (>80%) were restricted to softwood-dominated stands, and stands experiencing the highest levels of damage were stocked predominantly with hardwood species that had much higher damage relative to the softwood species (table 2); softwood species will likely dominate these stands as they mature. the most commercially valuable hardwood species in the study region are yellow birch and sugar maple, and they were dominant species in the youngest 2 age classes, but accounted for only 6 and 5% of dominant stems in the 16–20 year class. conversely, the commercial softwood species, balsam fir and red spruce, were minimal in the youngest age classes, but accounted for a large proportion of the dominant stems (21% each) in the 16–20 year class. red maple, a less valuable commercial species, was the most common deciduous tree species in all age classes ranging from 13–24% of dominant stems (table 2). a fig. 5. proportion (%) of plots containing either dominant commercial hardwood or softwood stems by age class in nothereastern vermont, 2012. alces vol. 50 andreozzi et al. – browsing and forest regeneration 73 table 2. species composition (%) and browse damage category of dominant stems by age class in clearcuts in northeastern vermont, 2012. age class species severe damage moderate damage light damage no damage total 3–5 american beech 0 0 2 1 3 aspen spp. 0 0 3 0 3 balsam fir 1 0 2 3 6 black cherry 1 0 0 0 1 red maple 0 0 15 8 24 red spruce 0 0 1 3 5 sugar maple 0 0 4 2 6 tamarack 0 0 0 1 2 white ash 0 0 0 0 1 white birch 1 0 1 1 3 yellow birch 1 0 13 5 20 non commercial na na na na 26 6–10 american beech 0 0 2 1 3 aspen spp. 0 0 1 0 2 balsam fir 1 1 4 9 15 red maple 1 1 12 3 16 red spruce 0 0 0 9 9 sugar maple 0 0 8 2 10 white ash 1 0 0 0 1 white birch 2 0 2 0 4 yellow birch 1 0 11 3 14 non commercial na na na na 25 11–15 american beech 1 0 1 0 3 aspen spp. 0 0 3 2 6 balsam fir 2 1 7 7 17 black spruce 0 0 0 1 1 red maple 3 2 8 0 13 red spruce 0 0 1 9 11 sugar maple 1 0 3 0 4 white birch 4 1 4 0 9 yellow birch 2 0 8 1 11 non commercial na na na na 24 16–20 american beech 1 1 3 2 6 aspen spp. 0 0 0 1 1 balsam fir 2 0 5 14 21 black spruce 0 0 0 1 1 red maple 7 1 14 1 23 table 2 continued . . . . 74 browsing and forest regeneration – andreozzi et al. alces vol. 50 similar trend occurred in new hampshire (bergeron et al. 2011) where the proportion of dominant commercial hardwood stems also declined with age class. heavy browsing pressure could potentially accelerate successional development by arresting or retarding the height development of preferred browse species in the region (mcinnes et al. 1992, davidson 1993). while previous site compositions are unknown, it is possible a shift from hardwood to softwood dominated stands may be the natural successional trend for these sites. although harvest records were unavailable for most sites, it appeared that many were originally mixed wood stands. the proportion of dominant commercial trees ≥3.05 m (beyond browse height) without severe damage increased with age class, peaking at 39% at 16–20 years (table 1). these stems are expected to recover from any moderate or light damage during future growth without browsing. in contrast, average values in adjacent northern new hampshire were 36, 60, and 71% in the 3 older age classes, suggesting that growth was more suppressed in vermont. intense browsing in areas of high moose density can arrest or retard growth of preferred browse species (bergerud and manuel 1968, angelstam et al. 2000). a study with exclosures on isle royale, michigan indicated that repeated browsing by moose retarded vertical growth of palatable species such as aspen and paper birch, and prevented stems from growing beyond browsing height resulting in a more open canopy (risenhoover and maass 1987). although heavy browsing of the same species in successive years can result in hedgy growth and lower height potential (peek et al. 1976, peek 1997), such stems can compensate if browsing declines or if removed in successive years; for example, after release of a dominant stem in forked stems (jacobs 1969) and the straightening of crooked stems with secondary growth over time (switzenberg et al. 1955, trimble 1968). a clipping study on isle royale indicated that the site-dependent survival and growth of balsam fir were related to suppression brought about by severe browsing in previous years (mclaren 1996). accurate prediction of damage is complicated by this dynamic process that is likely influenced by local site conditions, and seasonal moose density and site fidelity. in studies assessing browse damage in both southern and northern new england, time since harvest was negatively correlated with foraging intensity (faison et al. 2010, bergeron et al. 2011) which may allow compensatory growth by desirable hardwood species beyond the 16–20 year age class. however, an increasing dominance of softwood species coupled with suppressed growth of hardwood species indicates a possible shift in species composition in wmu e1. several studies have indicated change table 2 continued age class species severe damage moderate damage light damage no damage total red spruce 0 0 1 20 21 sugar maple 0 0 4 1 5 white birch 1 0 1 0 1 yellow birch 3 0 3 0 6 non commercial na na na na 14 alces vol. 50 andreozzi et al. – browsing and forest regeneration 75 in forest composition due to heavy moose browsing. in finland, heikkila et al. (2003) measured reduced height of preferred browse species resulting in the release of conifers from competition. on isle royale, moose prevented aspen, birch, and balsam fir from growing into the canopy, with little impact to spruce, resulting in a forest with fewer trees in the canopy, a well-developed understory of shrubs and herbs, and an increase in spruce biomass (mcinnes et al. 1992). similarly, selective pressure resulted in rapid occupation of spruce (picea spp.) as the dominant species in study stands in russia (abaturov and smirnov 2002). a similar trend is possibly occuring in northeast vermont where coniferous species account for >50% of total species composition in the 16–20 year age class (table 2). a reduction in moose density, as implemented in the study area, may also reduce future browsing pressure and provide for the release of preferred hardwood species. a population reduction in newfoundland in the early 1960s resulted in dramatic decline in the proportion of white birch and balsam fir stems browsed in 6–11 and 12–17 year old stands (bergerud et al. 1968). high-density moose populations have the potential to damage preferred forage species (peek 1997), but the negative impacts of over-browsing can be minimized if moose density is kept at low-moderate levels (brandner et al. 1990). in russia, a density of 0.3–0.5 moose/km2 retarded growth of preferred forage species such as aspen, whereas normal stand development occurred at 0.2–0.3 moose/km2 (abaturov and smirnov 1992). in sweden, simulated densities of 0.8–1.5 moose/km2 did not impact winter browse availability; impact was predicted at >2.0 moose/km2 (persson et al. 2005). both northern vermont and new hampshire are classified as a combination of spruce-fir, northern hardwood, and mixed forest types (degraaf and yamasaki 2001), and presumably measurable differences in forest regeneration reflect different moose density. bergeron et al. (2011) found a direct correlation between browse damage and moose density in northern new hampshire; the region with highest density had most damage. densities in northeast vermont were estimated at 1.2–1.8 moose/km2 in 1999–2009 and were probably higher than those in the highest density region of new hampshire estimated at 0.8–1.5 moose/km2 for the same time period (c. alexander, pers. comm., k. rines, nhfg wildlife biologist, pers. comm.). in both states significant differences were found in the stocking rate of dominant commercial trees, and the proportion of both dominant hardwood and softwood commercial tree species between the youngest and oldest age classes. however, the temporal comparisons among age classes indicate that sites with high initial browse pressure are often released from that pressure and recover to commercially valuable stands. in both vermont and new hampshire, stocking rate increased and damage declined over time with relative differences seemingly influenced by local moose density. compensatory growth in the region was measurable in the 16–20 year age class, but likely begins earlier when stems grow beyond browsing height. however, heavy browsing pressure on preferred tree species may result in lower stand height as measured in vermont and a possible shift in forest composition to coniferous species. further assessment is warranted to best evaluate the extent of compensatory tree growth in response to reduction in browsing due to forest aging and/or moose population density. acknowledgements we are grateful to the numerous commercial and private landowners for their cooperation and providing access to their property including, but not limited to, silvio 76 browsing and forest regeneration – andreozzi et al. alces vol. 50 o. conte national fish and wildlife refuge, plum creek, heartwood forestland fund, and devost leasing. we thank d. bergeron, c. alexander (vtfw moose biologist), and k. rines (nhfg moose biologist) for providing useful data. we are also thankful to t. millette, h. stabins, w. leak, and m. yamasaki for providing knowledge and guidance throughout the project, and to j. trudeau, n. fortin, and j. comeau for assistance in the field. references abaturov, b. d., and k. a. smirnov. 1992. formation of stands on clearings in forests with different moose population density. bulletin moskovskava obshestva isputatelij prirodi otdelenie biologii 97: 3–12. ———, and ———. 2002. effects of moose population density on development of forest stands in central european russia. alces supplement 2: 1–5. angelstam, p., p. e. wikberg, p. danilov, w. e. faber, and k. nygren. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland, and russian karelia. alces 36: 133–145. bergeron, d. h., p. j. pekins, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. the journal of wildlife management 32: 729–746. ———, ———, and h.whalen. 1968. the harvest reduction of a moose population in newfoundland. the journal of wildlife management 32: 722–728. bergström, r., and k. danell. 1986. moose winter feeding in relation to morphology and chemistry of six tree species. alces 22: 91–112. brandner, t. a., r. o. peterson, and k. l. risenhoover. 1990. balsam fir on isle royale: effects of moose herbivory and population density. ecology 71: 155–164. bubenik, a. b. 1997. behavior. pages 173222 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. carvell, k. l. 1967. the response of understory oak seedlings to release after partial cutting. west virginia university agricultural experiment station, bulletin 553. morgantown, west virginia, usa. davidson, d. w. 1993. the effects of herbivory and granivory on terrestrial plant succession. oikos: 23–35. degraaf, r. m., and m. yamasaki. 2001. new england wildlife: habitat, natural history, and distribution. university press of new england, hanover, new hampshire, usa. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forests. silva fennica 36: 57–67. faison, e. k., g. motzkin, d. r. foster, and j. e. mcdonald. 2010. moose foraging in the temperate forests of southern new england. northeastern naturalist 17: 1–18. heikkila, r., p. hokkanen, m. kooiman, n. ayguney, and c. bassoulet. 2003. the impact of moose browsing on tree species composition in finland. alces 39: 203–213. jacobs, r. d. 1969. growth and development of deer-browsed sugar maple seedlings. journal of forestry 67: 870–874. leak, w. b. 2007. accuracy of regeneration surveys in new england northern hardwoods. northern journal of applied forestry 24: 227–229. ———, d. s. solomon, and p. s. debald. 1987. silvicultural guide for northern hardwood types in the northeast (revised). research paper ne-603. u.s. alces vol. 50 andreozzi et al. – browsing and forest regeneration 77 department of agriculture, forest servie, northeastern forest experiment station, broomall, pennsylvania, usa. mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73: 2059–2075. mclaren, b. e. 1996. plant-specific response to herbivory: simulated browsing of suppressed balsam fir on isle royale. ecology 77: 228–235. millette, t. l., d. slaymaker, e. marcano, c. alexander, and l. richardson. 2011. aims-thermal a thermal and high resolution color camera system integrated with gis for aerial moose and deer census in northeastern vermont. alces 47: 27–37. north east state foresters association (nefa). 2007. the economic importance and wood flows from vermont's forests. north east state foresters association, concord, new hampshire, usa. peek, j. m. 1997. habitat relationships. pages 351–375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. ———, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3–65. persson, i. l., k. danell, and r. bergström. 2005. different moose densities and accompanied changes in tree morphology and browse production. ecological applications 15: 1296–1305. pruss, m. t., and p. j. pekins. 1992. effects of moose foraging on browse availability in new hampshire deer yards. alces 28: 123–136. renecker, l. a., and c. c. schwartz. 1997. food habits and feeding behavior. pages 403–439 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17: 357–364. scarpitti, d. 2006. seasonal home range, habitat use, and neonatal habitat characteristics of cow moose in northern new hampshire. university of new hampshire, durham, new hampshire, usa. ———, c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. snyder, j. d., and r. a. janke. 1976. impact of moose browsing on boreal-type forests of isle royale national park. american midland naturalist 95: 79–92. switzenberg, d. f., t. c. nelson, and b. c. jenkins. 1955. effect of deer browsing on quality of hardwood timber in northern michigan. forest science 1: 61–67. thompson, i. d., and r. w. stewart. 1997. management of moose habitat. pages 377-401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. thompson, m. e., j. r. gilbert, g. j. matula jr., and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in maine. alces 31: 233–245. trimble, g. r. 1968. form recovery by understory sugar maple under unevenaged management. u.s. department of agriculture, forest service research note ne-89. northeastern forest experiment station, broomall, pennsylvania, usa. vermont fish and wildlife department (vtfw). 2008. big game management 78 browsing and forest regeneration – andreozzi et al. alces vol. 50 plan 2010-2020: creating a road map for the future. pages 40-52. vermont fish and wildlife department, waterbury, vermont, usa. ———. 2011. 2011 vermont willdife harvest report moose. vermont fish and wildlife department, waterbury, vermont, usa. alces vol. 50 andreozzi et al. – browsing and forest regeneration 79 impact of moose browsing on forest regeneration in northeast vermont methods study area field measurements data analysis results discussion acknowledgements references alces29_181.pdf alces21_267.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces(25)_112.pdf opinions about moose and moose management at the southern extent of moose range in connecticut andrew m. labonte1, howard j. kilpatrick1, and john s. barclay2 1connecticut department of energy and environmental protection, wildlife division, 391 route 32, north franklin, connecticut 06254; 2wildlife conservation research center, university of connecticut, 1376 storrs road, unit 4087, storrs, connecticut, usa. 06269. abstract: increasing moose (alces alces) populations in the northeastern united states present new challenges for wildlife managers who must balance beneficial and adverse aspects of moose populations. it is important that managers understand stakeholder attitudes and values about moose and incorporate such into outreach and management programs. the objective of this research was to assess landowner and hunter perceptions about status, management, and concerns associated with a small moose population in connecticut. the majority of landowners and hunters correctly believed that <100 moose existed in connecticut, half believed that the population was increasing but had no opinion about appropriate size, and few had ever observed a moose in connecticut or been involved in a moose-vehicle accident (mva). landowner support for viewing areas was high and moose hunting low unless mvas increased; support for hunting moose was high among hunters. if human-moose conflicts increase, principally mvas, we expect reduced public support for the resident moose population. proactive education and management are suggested to reduce human-moose conflicts, mvas, and increase acceptance of hunting as a possible population management tool. alces vol. 49: 83–98 (2013) key words: alces alces, connecticut, moose, human dimensions, survey. moose (alces alces) populations have increased throughout northern new england over the past 30 years presenting management challenges to balance their beneficial and potentially adverse aspects (wattles and destefano 2011). moose provide intrinsic economic value to both consumptive and non-consumptive users (schwartz and bartley 1991), with watching and hunting as major revenue generators (wolfe 1987, timmermann and rodgers 2005), especially in northern new england. populations reaching levels sufficient for recreational opportunities may result in higher levels of adverse consequences in the form of moose-vehicle accidents (mvas) and ecological damage (mirick 1999, timmermann and rodgers 2005), although such conflicts can also occur in small populations and suburban areas (mcdonald et al. 2012). assessing attitudes of various stakeholder groups toward a wildlife species is useful to understand societal support and opposition about current and potential management decisions (bath and enck 2003), and importantly, incorporating stakeholder attitudes into outreach and management programs (teel et al. 2002). natural resource agencies increasingly emphasize stakeholder participation in decision-making (lauber and knuth 1997) and management of human-wildlife interactions (ericsson 2003) to implement plans (flanigan 1987, hartig and thomas 1988, pinkerton 1991, landre and knuth 1993), strengthen public relationships (landre and knuth 1993), and reduce conflict (erickson 1979, twight and patternson 1979, nelkin 1984, blahna and yontsshepherd 1989). 83 relative to other big game species, human dimensions (hd) research with moose was initially limited in north america (wolfe 1987). an evaluation of articles from 1974-2001 in alces indicated that the majority of early hd research pertained to hunting of moose or mvas, with less attention to public values and attitudes towards moose (ericsson 2003). in the northeastern united states, states have used hd research to evaluate public opinion about their initial moose management programs in vermont (alexander 1993), new hampshire (donnelly and vaske 1995), and new york (lauber and knuth 1997, 1999). relative to nonconsumptive recreation in new hampshire, silverberg et al. (2001) measured knowledge level, attitudes, and motivation of wildlife viewers at a moose viewing site. and recently, hd research was used to evaluate public opinions about moose to provide effective educational strategies to reduce human-moose conflicts, principally mvas, in prince george, british columbia (mcdonald et al. 2012). however, similar hd information is non-existent in connecticut, the southern extent of moose range in new england. although the potential for moose populations to continue expanding in connecticut is unclear, developing management strategies and programs that are both effective and acceptable to the public is important relative to managing human-moose conflicts. our objective was to survey landowners and hunters about the status, management, and associated concerns with the moose population in connecticut. study area and background connecticut (12,548 km2) was the fourth most densely populated area (3,500,000 people, 278 people/km2) in the united states at the time of this research (connecticut economic resource center 2006, 2010). located in southern new england, it is bounded on the south by long island sound, and by the states of rhode island to the east, massachusetts to the north, and new york to the west. connecticut is about half forested (55.6%), 20% developed or barren, 16.7% turf, grass, or agricultural field, 4.4% wetlands (non-forested, forested, and tidal), and 3.2% water (hochholzer 2010). historic accounts suggest that moose existed in connecticut prior to the 18th century (trumbull 1797, deforest 1964); however, goodwin (1935) noted that at the beginning of the 18th century there was no record of moose in connecticut. further, the lack of archaeological deposits of moose suggests that they likely existed in low numbers, if at all (n. bellantoni, connecticut state archeologist, pers. commun.). a few reports of transient moose occurred between 1916 and 1956 (connecticut wildlife 2000), and on 18 september 1956, the board of fisheries and game (currently the department of energy and environmental protection, deep) passed an emergency regulation that gave full protection to moose in connecticut. sporadic reports of moose occurred until the early 1990s (kilpatrick et al. 2003), and in 1992 the deep began documenting all credible sightings and mvas. in 1996 a question was added to the annual deer hunter questionnaire asking them to report any moose observation during the deer (odocoileus virginianus) hunting season. in 1998, the wildlife division of deep adopted a directive (deep2431-d1) that outlined procedures for responding to problem moose situations that included hazing, capture and relocation, and euthanasia. since 2000, observations of cows with calves confirmed the establishment of a small resident population (kilpatrick et al. 2003). an empirical model based on public sightings of moose reported to the deep conservatively estimated the population at ∼64 in 2004 (labonte and kilpatrick 2006) with ∼75 present at the 84 opinions about moose – labonte et al. alces vol. 49, 2013 time of this survey in 2008 (labonte 2011). despite low moose numbers, connecticut was experiencing 2–4 mvas annually (deep, unpublished data) and deep staff were exploring options to implement a moose management strategy to address increasing mvas. however, it was unknown if the general public or hunting community would support a management strategy that included moose hunting given the minimal population. understanding public and hunter opinions about moose and moose management is essential for developing an effective moose management plan in connecticut. methods based on the distribution of moose sightings by the public (kilpatrick et al. 2003), hunters (labonte et al. 2008), and reported mvas (deep, unpublished data) (fig. 1), northern connecticut was selected as the study area for the landowner survey. based on geographic features and an assessment of human population densities, towns in northern connecticut were delineated into 3 groups for the landowner survey (fig. 1) and were used for landscape level comparisons. towns were grouped as central (n = 13), eastern (n = 16), and western (n = 20) (table 1, fig. 1). landowner survey a database containing the names and addresses of landowners in the 49 study towns was obtained from municipal town offices. we set a sampling rule to include private landowners and removed all identifiable outliers (e.g., limited liability companies, corporations, companies, schools, churches, trustees, towns). we deleted duplicate landowner records (i.e., multiple ownerships) to compile a list of landowners with an equal likelihood of being randomly selected and receiving a single survey. we calculated minimum sample sizes required for each landscape based on a stratified random sampling approach (scheaffer et al. 1996). a mail survey was chosen because it can include complex questions, access geographically dispersed groups, and recipients can reply at their convenience with low potential for social desirability bias (decker et al. 2001). we used a 3-wave survey with a variation of the repeated mailing technique (dillman 1978). surveys were mailed to randomly selected landowners stratified among the 3 landscapes (eastern, central, western) in january 2008; 2 follow-up surveys were mailed to non-respondents about 4–8 weeks apart. after 3 attempts to contact landowners by mail, we contacted a sub-sample of nonrespondents by telephone to assess nonresponse bias. we used likert-scale questions in the surveys (likert-scale numbers indicated by each response were used to calculate mean response scores) to assess beliefs and experiences about wildlife (5-point scale), concerns about moose, support for hunting (5point scale), and acceptability of situations involving moose (6-point scale). there were 3 general types of questions with 3 response categories: 1) landowner beliefs and experiences (agree, neutral, disagree), 2) landowner opinions about management (support, neutral, oppose), and 3) landowner concerns (acceptable, not acceptable/no action, not acceptable/action). the study protocol and survey were reviewed and approved by the connecticut wildlife division, the northeast wildlife damage management cooperative, and the institutional review board (irb), office of research compliance at the university of connecticut; the irb chair deemed the survey exempt from irb status. surveys were conducted in accordance with federal guidelines in which minors (<18 years of age) were excluded, results were not identifiable to individuals, and surveys involved no risk to individuals. alces vol. 49, 2013 labonte et al. – opinions about moose 85 fig. 1. the study area was in the state of connecticut located in the northeastern portion of the united states. landowner surveys were conducted in 2008 in the northern portions of connecticut (shaded areas) where most moose-vehicle accidents (•) occurred, while hunter surveys were conducted at town halls ( ) located throughout the state. 86 opinions about moose – labonte et al. alces vol. 49, 2013 hunter survey we selected 31 of 169 (18%) town clerks to distribute the survey to any resident or non-resident hunter purchasing a connecticut firearms hunting license or combination hunting/fishing license; towns and sampling period were selected based on the highest volume of hunting license sales in 2004. surveys were distributed during 3 sampling periods (january, april, and october 2008) which were chosen to obtain a representative sample of each hunter group; many hunters purchase a license to pursue game during a specific season and the timing coincided with peak issuance. packets containing an instruction letter, return envelope, and specific number of surveys were mailed to the town clerks before each sample period; the number of surveys per town was based upon the respective volume of 2004 license sales. town clerks were instructed to provide a survey to every other individual that purchased a hunting or combination hunting/ fishing license; upon completion, they collected the survey and mailed all after each sampling period. we generated questions to evaluate hunting activity, participation in outdoor-related activities, and perceptions and opinions about connecticut's moose population. we used a 5-point likert-scale question to assess support for hunting and grouped responses into 3 categories: support, neutral, or oppose. the review and approval of the study protocol and survey were identical to the landowner survey. analysis we treated ordinal-level (likert scale) data as interval-level data as previous studies have validated the use of such data in survey research (nunnally and bernstein 1994, zinn and andelt 1999, daley et al. 2004). we calculated levene's test (p < 0.05) for equality of variances and the kolmogorov-smirnov test of normality; based on these results we used the kruskal-wallis test (p < 0.05) for all analysis at the landscape level, and the mann-whitney u test (p < 0.05) for comparisons between landowners and hunters. pearson chi-square tests (p < 0.05) were used to examine nominal level variables and compare responses between respondents and non-respondents. all analyses were conducted using systat 12.0 (systat 2007). results respondent demographics landowner survey — surveys were returned from 622 of 2,023 landowners (35.7% eastern, 31.3% central, 37.9% western); proportionally, 66% from the first, 20% from the second, and 14% from the final mailing. there was no difference among landscapes in gender (χ2 = 3.44, p < 0.178) and age of respondents (χ2 = 0.410, p < 0.999); 56.4% were male and the mean age of all respondents was 54.4 (sd = 14.7) years. after 3 attempts by mail, we contacted 51 non-respondents by telephone to assess non-response bias for specific questions. hunting was not commonly allowed in any landscape but was higher in the western (16%, χ2 = 13.6, p < 0.001) and eastern (14.7%, χ2 = 20.3, p < 0.001) than the central landscape (3%). table 1. human densities and landscape level (eastern, central, western) characteristics in connecticut, 2008. location eastern central western number of towns 16 13 20 population 79/km2 185/km2 71/km2 % forest 65.4 29.8 67.9 % commercial/ residential 14.2 43.2 11.7 % turf/agriculture 12.4 21.1 12.6 % wetlands 4.6 2.8 3.8 % water 2.3 1.8 3.3 % other 1.1 1.3 0.7 alces vol. 49, 2013 labonte et al. – opinions about moose 87 hunter survey — surveys were completed by 446 of 485 hunters (91.9%) and due to this high response rate, we did not assess non-response bias. gender of hunters was primarily male (97.6%) and the mean age was 48.1 (sd = 12.5) years; the majority had harvested deer (65.2%) and a few bear (7.0%) and moose (3.6%). the majority (>60%) would participate in nonconsumptive moose recreation (watching, photography) and half (50.8%) would hunt moose (table 2). landowner beliefs and experiences with wildlife most landowners believed that wildlife and management were important, and the mean response scores were similar across all landscape levels except for huntingrelated questions. in general, the majority of landowners were not unsupportive of hunting, but 30-60% were neutral/or disagreed with some aspect of hunting (table 3). knowledge about moose landowner survey — landowners were asked to identify the moose from 3 sketches depicting a deer, moose, and bear. responses were combined as no differences existed among landscapes (χ2 = 1.562, p = 0.458); most (90.3%) correctly selected the image of the moose with the remainder selecting the deer. respondent and nonrespondent opinions about the number of moose existing in connecticut were not different (χ2 = 2.316, p = 0.128) and no adjustments were made. all responses were combined because no differences (χ2 = 4.315, p = 0.634) among landscapes existed in perceptions about how many moose exist in connecticut. the majority (64%) correctly estimated that there were <100 moose in connecticut, and >90% estimated <500 moose (table 2). landowner-hunter comparisons — a similar proportion of landowners (63.9%) and hunters (67.4%) believed that <100 moose existed in connecticut (χ2 = 1.31, p = 0.253) (table 2). more hunters (27.7%) than landowners (18.5%) believed that <10 moose existed in connecticut (χ2 = 11.9, p = 0.001); although both were <10%, conversely, more landowners than hunters believed that >500 moose existed (χ2 = 8.6, p = 0.003). the primary source of information influencing opinions about the size of the moose population was from other table 2. landowner and hunter opinions about the moose population in connecticut, usa, 2008. lower case n refers to # of respondents. percent of respondents survey question landowner hunter number of moose (n) 590 408 0 3.0 6.9 <10a 18.5 27.7 <100a 63.9 67.4 100–499 28.0 29.0 >500 8.0 3.5 status of moose population (n) 606 430 increasing 51.8 67.6 decreasing 7.8 <1.0 stable 10.0 11.6 no opinion 30.4 20.0 opinion of moose population (n) 604 427 too high 3.0 3.9 too low 25.9 40.6 just right 15.7 19.2 no opinion 54.9 36.1 activities would participate in (n) 626 404 watching moose 62.1 33.8 photographing moose 50.7 27.5 hunting moose 10.7 50.8 other 2.0 1.0 none 20.0 19.0 aincludes all respondents who indicated 0 or <10. 88 opinions about moose – labonte et al. alces vol. 49, 2013 table 3. landowner beliefs and experiences about wildlife in connecticut, usa, 2008. beliefs and experiences about wildlife % responsea agree neutral disagree no opinion mean response scoresb c e w c e w c e w c e w c e w hc pc n i notice birds and wildlife around me daily 98 99 96 1 0 2 1 1 1 0 0 1 1.65 1.77 1.70 4.60 0.101 626 observing and learning about wildlife is important to me 88 92 89 10 7 7 2 1 3 0 0 1 1.34 1.47 1.40 3.12 0.210 624 hunting animals for any purpose should not be permitted 19 12 22 22 16 12 58 71 65 2 2 1 −0.54 −0.89 −0.68 7.66 0.022 623 it is important to manage some wild animal populations 84 86 86 9 5 9 6 8 4 1 1 1 1.08 1.16 1.18 3.09 0.214 622 wild animal populations should be managed for the benefit of all people 68 69 74 16 16 13 14 13 13 1 3 1 0.78 0.84 0.81 0.38 0.826 620 participation in hunting helps people appreciate wildlife and natural processes 36 53 44 23 23 15 36 22 37 4 3 4 −0.01 0.40 0.03 8.62 0.013 623 if wildlife populations are abundant, it is ok to use them as a natural renewable resource 53 65 55 22 15 24 19 17 17 6 3 5 0.45 0.71 0.49 5.13 0.077 613 regulated hunting is an acceptable use of a natural resource 65 76 69 15 9 12 16 13 17 4 2 3 0.63 0.94 0.70 9.88 0.007 621 c = central, e = eastern, w = western. alikert scale ranged from −2 (“strongly disagree”) to 2 (“strongly agree”). to evaluate percentages, responses were truncated into “agree, neutral, disagree.” bnot included in analysis are the number of respondents who choose “no opinion.” ch and p values for kruskal-wallis test statistic comparison between eastern, central, and western groups. a l c e s v o l . 4 9 , 2 0 1 3 l a b o n t e e t a l . – o p in io n s a b o u t m o o s e 8 9 sources (33.1%) for landowners and personal experience (37%) for hunters. opinions about moose landowner survey — respondent and non-respondent opinions about the status of connecticut's moose population (χ2 = 5.997, p = 0.112) and the number of moose in connecticut (χ2 = 6.374, p = 0.095) were not different and no adjustments were made. no difference among landscapes (χ2 = 0.835, p = 0.659) existed between the proportion believing that the moose population was either increasing (about half) or decreasing, or that believed it was too high (∼3%) or too low (∼25%) (χ2 = 2.71, p = 0.257); therefore, responses were combined (table 2). likewise, no landscape differences existed (χ2 = 2.68, p = 0.262) and responses were combined for the proportion of landowners (∼70%) who would support designating viewing areas for moose watching. landowner-hunter comparisons — about half (51.8%) of landowners and 2/3 of hunters believed that connecticut's moose population was increasing, but few (3 and 4%, respectively) believed the population was too high (table 2). more hunters (68%) than landowners (52%) believed that the population was increasing, and fewer that it was decreasing (χ2 = 33.1, p <0.001). there was no difference (χ2 = 0.559, p = 0.455) in the proportion of landowners and hunters who believed that the population was too low or too high; although, measurably more hunters thought the population was too low (40.6 vs. 25.9%; table 2). from a list of 3 potential moose-related activities if moose were common in connecticut, landowners favored watching and photography (62.1 and 50.7%), and hunters favored hunting and watching (50.8 and 33.8%). the proportion of landowners and hunters who would participate in watching (χ2 = 60.8, p < 0.001), photographing (χ2 = 41.9, p < 0.001), or hunting moose (χ2 = 247.6, p < 0.001) was different. the proportion of landowners and hunters who would not participate in any moose activity was similar (∼20%; χ2 = 0.057, p < 0.811) (table 2). interactions with moose landowner survey — a minority (15%) of landowners observed moose in 29 towns and differences existed among landscapes (χ2 = 14.3, p = 0.001). twice as many landowners observed moose in western (27.0%) than central (12.0%, χ2 = 13.6, p < 0.001) and eastern landscapes (12.6%, χ2 = 6.07, p = 0.014) which were not different (χ2 = 0.031, p = 0.860) (table 4). an additional 51 landowners reported moose tracks or other sign with the same landscape differences (χ2 = 13.3, p = 0.001); more moose tracks and sign were observed in western (21.8%) than in central (7.8%, χ2 = 13.2, p < 0.001) and eastern (10.0%, χ2 = 3.99, p = 0.046) landscapes which were not different (χ2 = 0.464, p = 0.496) (table 4). only landowners in western landscapes had been in a mva (n = 4) in connecticut. although the rate of mva experiences in any landscape was low over all (<5%), landscape differences existed (χ2 = 8.29, p = 0.016) (table 4). landowners in western landscapes (4.9%) were in more mvas than those in central landscapes (<1.0%, χ2 = 7.45, p = 0.006); there was no difference between western and eastern (1.0%, χ2 = 2.71, p = 0.100) or eastern and central (χ2 = 0.001, p = 0.979) landscapes (table 4). hunter survey — moose were observed by 20% of hunters (n = 91) in 36 towns, with 71 others observing tracks or scat in 14 towns where sightings occurred, as well as in 13 other towns. landowner concerns about moose landowner concerns were not different among landscapes regarding health, safety, 90 opinions about moose – labonte et al. alces vol. 49, 2013 or damage-related issues (h = 0.059–2.115, 0.742 >p >0.347), and were combined for analysis (table 5). the majority were only concerned about mva, with <20% very concerned (table 5). moose population management landowner survey — responses were combined because mean scores were not different among landscapes (h = 1.44–5.59, 0.487 > p > 0.061) for any scenario regarding moose population management (table 6). a minority (31%) of landowners supported hunting as a method to control moose populations in connecticut based on their current level of concern; their support was highest if hunting was carefully regulated and controlled by the state, or if the moose population and number of mvas were increasing (both 54%). conversely, the vast majority of hunters (83-88%) supported hunting under all scenarios (table 6). the proportion of landowners and hunters who supported hunting was different “if it was carefully regulated and controlled by the state” (u = 53,194, χ2 = 211.53, table 4. landowner interactions with moose in connecticut, usa, 2006–2007. moose-human interactions % response yes no c e w c e w χ2 pa observed moose 12.0 12.6 27.0 88.0 87.4 73.0 14.30 0.001 in yard 1.5 4.6 5.3 98.5 95.4 94.7 5.56 0.062 outside yard 3.8 3.4 13.7 96.2 96.6 86.3 13.55 0.001 crossing road 5.8 2.3 13.7 94.2 97.7 86.3 9.88 0.007 other 3.5 5.7 5.3 96.5 94.3 94.7 1.26 0.531 observed tracks/scat 7.8 10.0 21.8 92.2 90.0 78.2 13.30 0.001 moose-vehicle accident 1.0 (0.0b) 1.0 (0.0b) 4.9 (4.0b) 99.0 99.0 95.1 8.29 0.016 e = eastern (n = 87), c = central (n = 343), w = western (n = 95). aχ2 and p values for pearson chi-square comparison between eastern, central, and western groups. bmva reported just in connecticut. table 5. landowner concerns about moose interactions in connecticut, usa, 2008. concerns about moose % response mean response scoresa hb pb no concern some concern very concerned no opinion encountering a moose 67.4 24.9 4.0 3.7 1.47 1.263 0.532 the cost of residential property damage caused by moose 57.2 30.9 4.9 7.1 1.61 2.115 0.347 being injured in a motor vehicle accident that involves a moose 28.0 50.7 18.6 2.8 2.33 1.385 0.500 potential problems that moose may cause to the ecosystem 52.5 31.3 4.9 11.3 1.66 0.596 0.742 overall current level of concern related to moose 57.3 34.6 3.4 4.7 1.58 0.662 0.718 alikert scale ranged from 1 (“not concerned”) to 4 (“very concerned”). to evaluate percentages, “slightly concerned” and “somewhat concerned” responses were truncated into “some concern.” bh and p values for kruskal-wallis test statistic comparison between eastern, central, and western groups. alces vol. 49, 2013 labonte et al. – opinions about moose 91 table 6. landowner and hunter opinions about managing moose populations using hunting in connecticut, usa, 2008. concerns about moose % response mean response scoresa support neutral oppose hb pb uc pc χ2 land hunt land hunt land hunt land hunt land land based on your current level of concern? 31 na 29 na 40 na −0.23 2.05 0.35 if your level of concern increases? 47 na 25 na 29 na 0.20 3.98 0.13 if hunting were carefully regulated and controlled by the state? 54 88 22 6 24 6 0.34 1.41 2.82 0.24 53,194 0.00 211.5 if you knew that the moose population would be maintained at its current level? 41 83 30 8 29 9 0.09 1.23 2.69 0.26 49,524 0.00 206.2 if you knew that hunting is currently allowed in other new england states? 41 na 30 na 29 na 0.10 5.59 0.06 if you knew the likelihood of a human fatality was greater d? 54 85 26 8 21 7 0.44 1.37 1.44 0.48 18,731 0.00 268.0 alikert scale ranged from −2 (“strongly oppose”) to 2 (“strongly support”). to evaluate percentages, “strongly support” and “support” were truncated into “support,” and “oppose” and “strongly oppose” were truncated into “oppose.” bh and p values for kruskal-wallis test statistic comparison between eastern, central, and western groups. cu and p values for mann-whitney u test between landowners and hunters. dif you knew the likelihood of a human fatality was greater for a moose-vehicle accident than a deer-vehicle accident and that the moose population and number of moosevehicle accidents were increasing in connecticut? na = not asked on survey. 9 2 o p in io n s a b o u t m o o s e – l a b o n t e e t a l . a l c e s v o l . 4 9 , 2 0 1 3 p < 0.001), “if they knew that the moose population would be maintained at its current level” (u = 49,524, χ2 = 206.22, p < 0.001), and “if the moose population and number of mvas was increasing in connecticut” (u = 18,731, χ2 = 268.01, p < 0.001) (table 6). the range of responses was evenly distributed (13–18% per response, n = 6 responses) for those either primarily supporting or opposing hunting moose in connecticut (table 7). in general, those supporting hunting of moose wanted to either hunt moose or linked human-moose conflicts with need for hunting. conversely, those opposed to hunting moose were either unsupportive of hunting or believed that the population/conflict rate was too low (table 7). landowner opinions about roadside sightings and moose-vehicle accidents no differences existed at the landscape level in opinions about roadside sightings (h = 3.7–5.8, 0.15 > p > 0.054), mvas (h = 0.61–2.8, 0.23 > p > 0.73), or fatalities resulting from a mva (h = 2.2 – 3.0, p > 0.22). the proportion of landowners who deemed “it not acceptable and some action should be taken” increased substantially in all categories if the overall problem of mvas rose (table 8). discussion although few landowners hunted or permitted hunting on their property, observing and learning about wildlife was important to most landowners and they were supportive of designating viewing areas for moose. hunting activity, beliefs, and experiences with wildlife if hunting was involved, and direct interactions with moose and mvas were influenced by landscape. but, knowledge, opinions about moose and moose management, and concerns about moose were similar across landscapes despite landscape differences in moose experiences, albeit experiences were low (<20%) in all landscapes. we found that landowner and hunter knowledge about moose abundance was limited, as in massachusetts 20 years ago (vecellio et al. 1993). a small number (<50) of landowners and hunters combined believed no moose existed in connecticut. the main source of information about moose for landowners was from non-deep sources, table 7. landowner responses regarding reasons why they primarily supported or opposed hunting to control moose populations in connecticut, usa, 2008. primarily supported hunting n % respondents regulated hunting is a legitimate method to control moose population growth 306 18.1 moose threaten human safety 254 15.1 deep officials are well trained to handle problems associated with moose 252 14.9 moose population is too high or may become too high 244 14.5 moose cause damage to crops or property 244 14.5 want the opportunity to hunt moose 222 13.2 don't know 101 6.0 other 63 3.7 primarily opposed to hunting moose are not a threat to human safety at their current level 211 16.3 moose do not cause enough damage to warrant management 205 15.8 moose population is too low and does not warrant management 198 15.3 do not support hunters killing moose 190 14.6 disagree with hunting 181 14.0 do not support deep killing moose 176 13.6 do not know 85 6.6 other 51 3.9 alces vol. 49, 2013 labonte et al. – opinions about moose 93 table 8. landowner opinions about roadside sightings and moose-vehicle accidents in connecticut, usa, 2008. % response acceptable not acceptable/no action not acceptable/ action mean response scoresa concerns about moose c e w c e w c e w c e w hb pb a moose is on or near a busy highway occasionally 35.6 39.2 23.7 13.9 13.4 19.6 50.5 47.4 56.7 3.31 3.29 3.60 3.742 0.154 moose are frequently on or near busy highways 14.6 19.8 10.2 10.9 14.6 9.2 74.5 65.6 80.6 4.13 4.01 4.36 5.837 0.054 1 moose-vehicle collision occurs each year statewide 38.1 34.4 31.6 21.5 36.5 34.7 40.4 29.2 33.7 3.16 3.05 3.11 0.618 0.734 2-5 moose-vehicle collisions occur each year statewide 26.5 26.6 20.4 15.5 18.1 18.4 58.0 55.3 61.2 3.80 3.78 3.93 1.009 0.604 6-10 moose-vehicle collisions occur each year statewide 18.1 21.3 10.5 15.4 7.9 14.7 66.5 70.8 74.7 4.14 4.26 4.40 2.878 0.237 >10 moose-vehicle collisions occur each year statewide 13.2 17.8 9.4 12.4 7.8 8.3 74.4 74.4 82.3 4.39 4.49 4.68 2.746 0.253 a human fatality results from a motorist hitting a moose in connecticut 16.7 20.0 10.5 21.0 24.4 23.2 62.4 55.6 66.3 4.08 3.82 4.23 2.964 0.227 2-5 human fatalities result from a motorist hitting a moose in connecticut 10.8 13.3 6.3 14.2 10.0 10.4 75.0 76.7 83.3 4.52 4.56 4.69 3.069 0.216 e = eastern, c = central, w = western. alikert scale was 1 (“acceptable”), 2 (“not acceptable/no management action taken”), 3 (“not acceptable/action should be taken”). bh and p values for kruskal-wallis test statistic comparison between eastern, central, and western groups. 9 4 o p in io n s a b o u t m o o s e – l a b o n t e e t a l . a l c e s v o l . 4 9 , 2 0 1 3 whereas hunters were most influenced by personal experience and deep communications. it is not surprising that ∼25% of landowners and hunters believed <10 moose existed (table 2), because few ever observe a moose in connecticut. many landowners and hunters had no opinion about the moose population status (20–30%) or how many moose should exist in connecticut (35–55%) (table 2). the low response rate probably reflects their lack of experience, familiarity, and interest in moose. riley and decker (2000) also found a large portion of people lacked opinions about cougars in montana, presumably for the same reasons. they suggested that lack of opinion may indicate 1) a lack of general concern in the everyday lives of residents, 2) stakeholder perceptions that managers do not listen to stakeholder concerns, or 3) distrust in delegation of decision-making to managers. overall, the majority of landowners had few concerns about moose except mvas. less than half supported using hunting as a method to control moose populations in connecticut based on their current knowledge of population levels, as opposed to hunters who were strongly supportive. more than half of landowners were supportive of moose hunting if it was carefully regulated and controlled by the state. although all forms of hunting are controlled by state fish and wildlife agencies, kilpatrick et al. (2007) found that landowners often are unaware of regulations or requirements that govern wildlife resources and expressed increased support for certain regulations or requirements although they already existed. predictably, if the number of roadside sightings, mva, or the number of related human fatalities increased, the proportion finding such unacceptable also increased. given that the first reported mva in connecticut occurred in 1995 and the annual rate remains low (2.3 mva per year), it is not surprising that residents are minimally concerned about moose. overall, few landowners (<1%) had ever been involved in a mva in connecticut. if the frequency of moose sightings along roads increases substantially, support for controlling moose populations will presumably increase regardless of the number of mvas or human fatalities. about 50% believed that a moose near a busy highway was unacceptable requiring action, and 58% believed action was required at 2–5 mva per year, the current reported mva rate (table 7). a similar situation occurred with elk (cervus elaphus) in urban areas of flagstaff, arizona (lee and miller 2003), where most respondents were concerned about being in an automobile accident involving an elk or seeing one along a roadside. the collective ability for humans to accept the presence and consequences of any wildlife species will eventually define the wildlife acceptance capacity (wac) for that species (decker and purdy 1988). in anchorage, alaska where moose populations exceed habitat carrying capacity (wac is either higher or lower), only half of residents supported moose hunting (whittaker et al. 2001). in british columbia, mcdonald et al. (2012) found that most respondents suggested reducing attractants or relocating moose to alleviate moosehuman conflicts, presumably over hunting, however sample size was small (n <100). acceptance of hunting among certain stakeholders may be influenced more by basic beliefs about hunting which are based on fundamental value orientations toward use or protection of wildlife (fulton et al. 1996, zinn et al. 1998). in connecticut, because moose are of such low numbers and few residents have any direct experience with moose, an associated wac is probably not measurable or is exceedingly high. we expect a reduction in wac if moosehuman conflicts increase measurably and alces vol. 49, 2013 labonte et al. – opinions about moose 95 advocate for a proactive management strategy that would increase public education about moose, mvas, and the potential role of hunting to help protect human safety. educational efforts should improve public awareness through posted warnings about local moose on department of transportation variable message boards (vmbs), erecting moose-crossing signs in appropriate areas, and meeting with stakeholder groups. the effectiveness of vmbs and signs to reduce mvas is unknown, but they should alert drivers otherwise unaware about moose in connecticut. a multi-faceted strategy should increase public awareness and education about moose in connecticut and aid in developing a long-term moose management program beyond simply minimizing mvas. acknowledgements we thank g. chasko, p. curtis, j. enck, m. gregonis, m. ortega, d. may, and r. ricard for reviewing surveys and drafts of this manuscript, and j. brooks, t. goodie, p. lewis, t. muni, and a. ocampo for assisting with data collection. this project was supported by the university of connecticut, college of agriculture and natural resources, department of natural resources and the environment, wildlife conservation research center, the northeast wildlife damage management cooperative, and the connecticut department of energy and environmental protection, wildlife division, federal aid in wildlife restoration project 49-35. references alexander, c. e. 1993. the status and management of moose in vermont. alces 29: 187–195. bath, a. j., and j. w. enck. 2003. wildlifehuman interactions in national parks in canada and the usa. national park service research review 4: 1–32. blahna, d. j., and s. yonts-shephard. 1989. public involvement in resource planning: toward bridging the gap between policy and implementation. society and natural resources 2: 209–227. connecticut economic resource center. 2006. state-by-state analysis. rocky hill, connecticut, usa. (accessed december 2010). ———. 2010. connecticut cerc state profile. rocky hill, connecticut, usa. (accessed december 2010). connecticut wildlife. 2000. wildlife management through the century. connecticut wildlife 20(6): 14–17. daley, s. s., d. t. cobb, p. t. bromley, and c. e. sorenson. 2004. landowner attitudes regarding wildlife management on private land in north carolina. wildlife society bulletin 32: 209–219. decker, d. j., t. l. brown, and w. f. siemer. 2001. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. ———, and k. g. purdy. 1988. toward a concept of wildlife acceptance capacity in wildlife management. wildlife society bulletin 16: 523–527. deforest, j. w. 1964. history of the indians of connecticut from the earliest known period to 1850. archon books, hamden, connecticut, usa. dillman, d. a. 1978. mail and telephone surveys: the total design method. wiley –interscience, new york, new york, usa. donnelly, m. p., and j. j. vaske. 1995. predicting attitudes towards a proposed moose hunt. society and natural resources 8: 307–319. erickson, p. a. 1979. the role of the public. pages 309–321 in p. a. erickson, editor. environmental impact assessment: 96 opinions about moose – labonte et al. alces vol. 49, 2013 http://www.cerc.com/content/resources.asp http://www.cerc.com/content/resources.asp http://www.cerc.com/townprofiles/customer-images/connecticut.pdf http://www.cerc.com/townprofiles/customer-images/connecticut.pdf principles and applications. academic press. new york, new york, usa. ericsson, g. 2003. of moose and man: the past, the present, and the future of human dimensions in moose research. alces 39: 11–26. flanigan, f. h. 1987. public involvement in the chesapeake bay, usa program. pages 512–517 in s. k. majumdar, l. w. hall, and h. m. austin, editors. contaminant problems and management of living chesapeake bay resources. pennsylvania academy of science. easton, pennsylvania, usa. fulton, d., m. j. manfredo, and j. lipscomb. 1996. wildlife value orientations: a conceptual and measurement approach. human dimensions of wildlife 1: 24–47. goodwin, g. 1935. the mammals of connecticut. state geological and natural history survey, bulletin 53. hartford, connecticut, usa. hartig, j. h., and r. l. thomas. 1988. development of plans to restore degraded areas in the great lakes. environmental management 12: 327–347. hochholzer, h. 2010. connecticut's forest resource assessment and strategy; building a better tomorrow for connecticut's forests today. department of environmental protection, bureau of natural resources, forestry division, hartford, connecticut, usa. kilpatrick, h. j., r. riggs, a. m. labonte, and d. celotto. 2003. history and status of moose in connecticut 2002. bureau of natural resources, wildlife division, department of environmental protection, hartford, connecticut, usa. ———, a. m. labonte, and j. s. barclay. 2007. acceptance of deer management strategies by suburban landowners and bowhunters. journal of wildlife management 71: 2095–2101. labonte, a.m. 2011. an assessment of moose (alces alces american) and moose management in connecticut. ms thesis, university of connecticut, storrs, connecticut, usa. ———, and h. j. kilpatrick. 2006. moose (alces alces) in connecticut: establishment, expansion, and future expectations. proceedings of the 63rd annual northeast fish and wildlife conference. groton, connecticut, usa. ———, ———, and w. reid. 2008. 2007 connecticut deer summary report. department of environmental protection, bureau of natural resources, wildlife division, hartford, connecticut, usa. landre, b. k., and b. a. knuth. 1993. the role of agency goals and local context in great lakes water resources public involvement programs. environmental management 17: 153–165. lauber, t. b., and b. a. knuth. 1997. fairness in moose management decisionmaking: the citizen's perspective. wildlife society bulletin 25: 776–787. ———, and ———. 1999. measuring fairness in citizen participation: a case study of moose management. society and natural resources 11: 19–37. lee, m. e., and r. miller. 2003. managing elk in the wildland-urban interface: attitudes of flagstaff, arizona residents. wildlife society bulletin 31: 185–191. mcdonald, a. m. h., r. v. rea, and g. hesse. 2012. perceptions of moosehuman conflicts in an urban environment. alces 48: 123–130. mirick, p. g. 1999. back in the company of moose. massachusetts wildlife 49: 17–25. nelkin, d. 1984. approaches to participation in natural resource decisions. pages 175–182 in c. w. churchman, a. h. rosenthal, and s. h. smith, editors. natural resource administration. westview press, boulder, colorado, usa. nunnally, j. c., and i. h. bernstein. 1994. psyclandtric theory. third edition. mcgraw-hill, new york, new york, usa. alces vol. 49, 2013 labonte et al. – opinions about moose 97 pinkerton, e. 1991. locally based water quality planning contributions to fish habitat protection. canadian journal of fish and aquatic science 48: 1326–1333. riley, s. j., and d. j. decker. 2000. wildlife stakeholder acceptance capacity for cougars in montana. wildlife society bulletin 28: 931–939. scheaffer, r. l., w. mendenhall iii, and l. ott. 1996. elementary survey sampling. fifth edition. duxbury press, wadsworth publishing company inc, raritan, new jersey, usa. schwartz, c. c., and b. bartley. 1991. reducing incidental moose mortality: considerations for management. alces 27: 227–231. silverberg, j. k., p. j. pekins, and r. a. robertson. 2001. impacts of wildlife viewing at dixville notch wildlife viewing area. proceedings of the 2001 northeast recreation research symposium, bolton landing, new york, usa. systat software. 2007. systat 12.0 statistics iii. systat software, inc., san jose, california, usa. teel, t. l., r. s. krannich, and r. h. schmidt. 2002. utah stakeholders' attitudes toward selected cougar and black bear management practices. wildlife society bulletin 30: 2–15. timmermann, h. r., and a. r. rodgers. 2005. moose: competing and complementary values. alces: 85–120. trumbull, b. d. 1797. a complete history of connecticut, civil and ecclesiastical, from the emigration of its first planters from england, in 1630 to 1713. hudson and goodwin, hartford, connecticut, usa. twight, b. w., and j. j. patterson. 1979. conflict and public involvement: measuring consensus. journal of forestry 77: 771–776. vecellio, g. m., r. d. deblinger, and j. e. cardoza. 1993. status and management of moose in massachusetts. alces 29: 1–7. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. whittaker, d., m. j. manfredo, p. j. fix, r. sinnott, s. miller, and j. j. vaske. 2001. understanding beliefs and attitudes about an urban wildlife hunt near anchorage, alaska. wildlife society bulletin 29: 1114–1124. wolfe, m. l. 1987. an overview of the socioeconomics of moose in north america. swedish wildlife research supplement 1: 659–675. zinn, h. c., and w. andelt. 1999. attitudes of fort collins, colorado residents toward prairie dogs. wildlife society bulletin 27: 1098–1106. ———, m. j. manfredo, j. j. vaske, and k. wittmann. 1998. using normative beliefs to determine the acceptability of wildlife management actions. society and natural resources 11: 649–662. 98 opinions about moose – labonte et al. alces vol. 49, 2013 opinions about moose and moose management at the southern extent of moose range in connecticut study area and background methods landowner survey hunter survey analysis results respondent demographics landowner beliefs and experiences with wildlife knowledge about moose opinions about moose interactions with moose landowner concerns about moose moose population management landowner opinions about roadside sightings and mooseicle accidents discussion acknowledgements references epizootiology of elaphostrongylus alces in swedish moose margareta stéen1, ing-marie olsson ressner2, bodil olsson3, and erik petersson4 1department of anatomy, physiology and biochemistry, swedish university of agricultural sciences, p. o. box 7090, se-750 07 uppsala, sweden; 2swedish chemicals agency (kemi), p. o. box 2, se-172 13 sundbyberg, sweden; 3tns sifo, p.o. box 115 00, se-404 30 gothenburg, sweden; 4department of aquatic resources, swedish university of agricultural sciences, se-178 93 drottningholm, sweden abstract: a total of 961 harvested and 241 unharvested moose (alces alces) carcasses and parts from throughout sweden were examined for elaphostrongylus alces from 1985 to 1989. when available, the central nervous system and skeletal muscles were searched for adult nematodes, and lungs and feces were examined for first-stage larvae. the parasite was distributed throughout sweden with highest prevalence (56%) in the central region and lowest in the south (13%). prevalence was highest in calves and old moose (>9 years) and lowest in middle-aged animals (5–9 years), with no statistical difference between sexes, although prevalence trended higher in young males. body condition and abundance of elaphostrongylus alces were negatively correlated, and condition was poorer in unharvested than harvested moose. a short (39–73 days) prepatent period was documented, and calves as young as 1.5 months were infected. these results indicate the importance of continued surveillance of elaphostrongylus alces, particularly because a warming climate will likely increase abundance of intermediate mollusk hosts and possibly cause increased infection of moose. alces vol. 52: 13–28 (2016) key words: alces alces, climate, body condition, elaphostrongylus alces, intermediate host, gastropods, moose, prepatent period, protostrongylidae, sweden the moose (alces alces) population in scandinavia began to rise in the 1970s, peaking in the mid-1980s in sweden. with few large predators at that time, it was not unusual to find dead or sick animals (hörnberg 2001, stéen et al. 2005), and in the 1980–1990s, high mortality was noted in both swedish and norwegian moose, as well as in semidomestic reindeer (rangifer tarandus). a previously unknown disease, elaphostrongylosis (stéen and rehbinder 1986, stuve 1986), was reported in the 1980s and sick animals were characterized by locomotive abnormalities such as ataxia, incoordination, swaying of the hindquarters, broad and stamping gait, and a certain way of hypermetria that suggested paralysis of ascending proprioceptive nerve fibers (stéen and roepstorff 1990). a previously undescribed species of elaphostrongyline nematode with a dorsal-spine larva, elaphostrongylus alces (stéen et al. 1989) was invariably associated with sick and dead moose (stéen and rehbinder 1986). parasites of the genera parelaphostrongylus and elaphostrongylus belong to the subfamily elaphostrongylinae (protostrongylidae, metastrongyloidea, nematoda). species of the genus parelaphostrongylus (p. tenuis, p. odocoilei, p andersoni) affect the central nervous system (cns) and skeletal muscle corresponding author: margareta stéen, department of anatomy, physiology and biochemistry, swedish university of agricultural sciences, po. box 7068, se-750 07 uppsala, sweden, margareta.steen@slu.se 13 mailto:margareta.steen@slu.se fasciae of nearctic cervids in north america including white-tailed deer (odocoileus virginianus), black-tailed deer (o. hemonious hemonious), mule deer (o. h. columbianus), and occasionally wapiti (cervus canadensis) and moose (alces alces spp.). species of elaphostrongylus (e. alces, e. cervi, e. panticola, e. rangiferi) affect the cns, the peripheral nerve system (pns), and the skeletal muscle fasciae in eurasian cervids including moose, red deer (cervus elaphus), maral deer (c. e. sibiricus), roe deer (capreolus capreolus), and reindeer (lankester 2001). in the new world, as in the old world, central nervous disorders and mortality occur in wild cervids infected with elaphostrongyline nematodes (anderson 1964, lankester 2001, 2010). representatives of the genera elaphostrongylus and parelaphostrongylus are also harmful to domestic ruminants (lankester 2001). both elaphostrongylus spp. and parelaphostrongylus spp. develop from the first to third larval stage (l1–l3) in their gastropod intermediate host, and develop from the l3 to the adult (l5) stage in their cervid (final) host (olsson et al. 1998, olsson 2001). specific identification of adult protostrongylids and first-stage larvae (l1) in feces of swedish moose was a result of multiple studies. the morphology of e. alces was initially described by stéen et al. (1989) and (stéen and johansson 1990), and subsequent comparison of specific proteins in protostrongylid l1 indicated that l1 and adult e. alces had the same protein pattern in moose, but differed from the l1s and adult protostrongylid parasites in other wild ruminants (stéen et al. 1993). experimental infection of captive moose indicated that l1 collected from wild moose caused elaphostrongylosis, and l1 excreted from infected and sick moose and transmitted to terrestrial snails (arianta arbustorum) in which larvae develop (lankester et al. 1998), were identified as e. alces using genomic dna (gajadhar et al. 2000) and single-strand conformation polymorphism (sscp) analysis (chilton et al. 2005, huby-chilton et al. 2006). collectively, these studies indicate that protostrongylid larvae in swedish moose are e. alces. given the prevalence and deleterious effect of this disease, our objective was to determine if e. alces is related to age, sex, condition, and geographic distribution of moose in sweden. study area sweden was divided into 6 regions from the far north (69°03′36″n 20°32′55″e) to the far south (55°20′13″n 13°21′34″e) to determine the distribution and prevalence of elaphostrongylosis (fig. 1). sweden is sheltered by the scandinavian mountains and has a continental climate with large differences in temperature and precipitation between summer and winter, and a relatively small amount of precipitation (swedish meteorological and hydrological institute [smhi]). summer temperatures are similar to those in north america and asia at similar latitude, although due to the gulf stream, winter in sweden is typically milder (smhi 2015). methods hunting begins on the first monday of september in northern sweden, on the second monday of october in the south, and seasons end in december or january (swedish association for hunting and wildlife management 2015). prior to the hunting seasons (1986 and 1987), we sent hunters report cards (hunting site, sex, and approximate age) and wrapping materials to pack body parts (i.e., lungs, feces, spinal cords, and mandibles). moose carcasses and parts (n = 1137) were examined in 5 consecutive years (1985– 1989); 896 (79%) were associated with harvested moose (1986, 1987, 1989) and 241 (21%) were from non-harvested animals (i.e., euthanized or found dead; 1985–1989). 14 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 we used 1020 lungs and 1084 fecal samples to identify presence of elaphostrongylid l1, 655 spinal cords (membranes) to identify presence of adult worms, and 636 mandibles to measure fat content (table 1). age was determined by dental wear (gasaway et al. 1978) or from information provided by hunters; 151 animals were not aged due to lack of information. five age classes were established: 1) calves were ≤12 months), 2) yearlings were >12 and ≤24 months, 3) young animals were >24 months and ≤5 years, 4) middle-aged animals were >5 and ≤9 years, and 5) old animals were >9 years. sex was determined from the whole carcass or hunter information. evaluation of body condition was by visual inspection, location, and appearance of body fat (n = 948), and/or by measuring fat content (%) in mandible bone marrow (n = 636; engelsen etterlin et al. 2009). three categories of condition were established: bd ac z y x w s t u c b d eo m h f arc�c circle 66°33'39" n treriksröset 69°03'36''n, 20°32'55''e smygehuk 55 °20'13''n, 13°21'34'' region 1: n= 135 region 2: n=558 region 3: n=37 region 4: n=100 region 5: n=40 region 6: n=26 n k g fig. 1. map of sweden with county codes and 6 regions: region 1 = the laplandic counties (ac, z, bd, and y), region 2 = southern part of norrland (w, s, and x county), region 3 = northern svealand (c, u, t, and b county), region 4 = southern svealand and northern götaland (d, p, r, e, and o county), region 5 = småland (f, h, and g county), and region 6 = southern götaland (n, l, k, and m county). dots represent locations of moose infected with elaphostrongylus alces; n = harvested moose. alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 15 normal, poor (below normal), and emaciated (lack of adipose tissue). the fat content in bone marrow was measured with standard techniques under specified assay conditions and techniques (nmkl no 131, nordic committee on food analysis 1989) and also used to assign condition: normal = 75– 94%, poor = 16–<75%, and emaciated = 0.4–<16% fat content. assigning condition from visual inspection (without measuring fat content) was considered reliable because of the strong correlation between the condition category assigned from fat measurements and visual inspection of the same animals (rs = −0.801, p < 0.001, n = 592). bodies/parts were inspected for adult e. alces worms and l1 with necropsy procedures described previously (stéen and rehbinder 1986, stéen et al. 1997, 1998) and included examination of muscle fasciae, the cranial cavity, brain, and spinal cord membranes and epidural space of the spinal cord (stéen and rehbinder 1986, stéen et al. 1997). lungs from all animals were palpated and inspected for nodules, and 20 g samples of minced lungs and feces were processed to detect l1 (baermann 1917). l1s were identified as protostrongylids quantified in a counting chamber under a stereo microscope, expressed as larvae per gram of wet feces (lpg), and classified into 7 levels of relative abundance ranging from none (0) to heavy (6) (national veterinary institute, sweden). there were 4 categories of infection: 0 = uninfected, 1 = in the epidural space but not in lungs or feces, 2 = in lungs but not feces, and 3 = in feces. animals were categorized as either infected or uninfected (presence or absence of l1 and/or e. alces worms) for certain statistical comparisons (e.g., sex or age groups, prevalence in population or region), data management and statistics data were tested for normal distribution and seasonal variation, and if not normally distributed, normality was achieved with log-transformation. a peak function analysis was used to identify the best fit to the relationship between bone marrow fat and season (tablecurve software, systat 2002). a mean value was calculated for harvested animals and this value was applied together with the individual values for remaining table 1. total number of moose (harvested/unharvested) and sample location/type – epidural space of the spinal cord (epidural), lungs, feces, mandibles – used to study elaphostrongylus alces in sweden, 1985–1989. moose epidural lungs feces mandibles sex males 386/87 173/85 343/84 369/85 260/36 females 456/147 223/144 404/145 427/145 262/57 unknown 54/7 23/7 37/7 49/7 21/– age group calves 457/107 187/103 392/102 434/103 270/48 yearlings 227/36 113/36 52/12 220/36 182/11 young 40/27 23/27 39/27 38/27 36/10 middle aged 31/19 27/19 28/19 26/19 30/5 old 6/37 4/37 6/37 6/37 5/16 unknown 135/14 65/13 110/14 123/14 20/1 total 896/241 419/236 784/236 847/237 543/93 16 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 animals. the residuals for all animals were calculated (harvested moose were not combined as above), and adjusted values were calculated by adding the residual to the common mean. this produced a few values >100% that were not further corrected in subsequent analyses. bone marrow fat (adjusted for seasonal variation) was subsequently analysed using generalized linear models. body condition was also analysed with generalized linear models, modeling the probability of being in normal condition (see above) assuming a binary distribution of the response variable. the total parasite infection or parasites found in either the lungs, feces, or in the epidural space were similarly corrected, and the probability of being infected was tested with respect to 3 predictors (age, sex, region). the age when calves were infected was estimated with birth date information from each county. comprehensive data were available from 5 counties: västerbotten (ac in region 1), västra götaland (o in region 4), kalmar (h in region 5), kronoberg (g in region 5), and södermanland (d in region 4) (fig. 1). in 3 counties (h, g, and d) the mean value + sd (malmsten 2014) was used as the birth date, and in 2 counties (ac and o) the mean value + sd was estimated (broberg 2004). birth dates for the counties without data were estimated using a multiple imputation (proc mi in sas statistical software, sas 2014) with a markov chain monte carlo method in which longitude and latitude of resident cities were used with the number of imputations set to 60. other than 3 counties with a minor inconsistency (3–4 days), the approach produced an acceptable trend of earlier birth dates in southern sweden, and the dates corresponded well with the span of birth dates reported by a national hunting organization (swedish association for hunting and wildlife management 2015) (table 2). the mean category of infection (0–3) in each age group was calculated to illustrate the relationships among age (mean age of group), category of infection, and body condition. these values were used to develop a contour graph using sigmaplot software (systat 2008) where body condition, age group, and infection category were interpolated. results infection, age and sex age of moose was skewed towards young animals (table 1), and age in the two groups (harvested and unharvested) was not distributed evenly. unharvested moose were older than those harvested for combined age classes, calves, and by sex (table 3). the average age of harvested animals (n = 761) was 10.4 months (95% ci = 9.6 – 10.4; range = 0–15 years), and 22.3 months (ci = 17.4–28.4; range = 0–20 years) for unharvested animals (n = 227). females were older in the yearling, middle-aged, and combined age groups. a slight majority (57%) of the harvested sample (n = 896) was infected with l1 and/or adult e. alces worms. the prevalence was similar between sexes in each age class for l1 in lungs, l1 in feces, and adult worms in the epidural space of the spinal cord (fig. 2). there was a tendency (p = 0.074) toward higher prevalence in males than females in the young age class. worms were found in the epidural space of the spinal cord in animals 3 months to 2 years old, but not in animals 3 to 9 years old; worms were found in a single 10-year old moose. the abundance of l1 in lungs (n = 784) was high in calves and yearlings, lower at 3–4 years of age, and minimal in adults. nearly the entire sample (98%) of unharvested moose (n = 241) was infected with e. alces (fig. 3). worms were found in the epidural space of the spinal cord in 3 month to 4 year-old animals. the average age of alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 17 infected calves was 4.8 months (95% ci = 4.7–4.9). no worms were found in the epidural space of the spinal cord in 5–9 yearold moose, but worms reappeared at 10–16 years of age. the prevalence of adult worms in the epidural space of the spinal cord was 36% in the combined data (harvested and unharvested, n = 655); the prevalence of l1 in lungs (n = 1020) and feces (n = 1084) was 64 and 53%, respectively. the prevalence (worms/ l1) was 66% overall; 88% in old moose, 74% in yearlings, 67% in calves, 55% in young, and 48% in middle-aged animals. there were differences (p < 0.001) in frequency of infection among age groups; the oldest animals had the highest frequency of infection (l1) and the middle-aged the lowest. the frequency of worms in the epidural space of the spinal cord was high in calves/ yearlings, leveled out at 4 years, and then was not identified until 10–16 years at low frequency. the abundance of l1 in lungs of old animals was at the highest level (6). body condition body condition of harvested animals (n = 981) was either normal (40% overall, 24% calves) or poor (59%, 75% calves). in unharvested moose (n = 227), body condition was normal in 38% overall, with calves and old animals lower; 25% calves, 45% young, and 29% old animals were in normal condition. for all moose, body condition and category of infection were correlated (rs = 0.215, p < 0.001). in separate age classes, this table 2. prevalence of elaphostrongylus alces (adjusted for julian date and age of the sampled moose) and birth date of moose in swedish counties. birth dates marked with an asterisk are observed values; others are estimated (see data management and statistics). region county mean prevalence (%) n birth date (julian date) birth date 1 ac västerbotten 43.8 82 167* 16 june 1 bd norrbotten 51.4 13 168 17 june 1 z jämtland 53.5 8 171 20 june 1 y västernorrland 58.2 21 164 13 june 2 w dalarna 63.2 32 160 9 june 2 x gävleborg 67.7 457 157 6 june 2 s värmland 49.4 36 159 8 june 3 b stockholm 61.7 2 149 29 may 3 c uppsala 100.0 3 152 1 june 3 t örebro 67.4 17 156 5 june 3 u västmanland 79.3 15 154 3 june 4 d södermanland 100.0 5 148* 28 may 4 e östergötland 27.6 12 150 30 may 4 o västra götaland 53.9 12 153* 2 june 5 f jönköping 39.1 16 151 31 may 5 g kronobergs 28.2 10 144* 24 may 5 h kalmar 32.2 11 143* 23 may 6 k blekinge 37.4 5 140 20 may 6 m skåne 0 8 144 24 may 6 n halland 41.5 12 146 26 may 18 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 correlation was found in yearlings (rs = 0.262, p < 0.001, n = 239), young (rs = 0.463, p < 0.001, n = 67), middle-aged (rs = 0.441, p = 0.002, n = 47), and old animals (rs = 0.456, p = 0.003, n = 40), but not in calves (rs = 0.054, p = 0.239, n = 471). for all moose, body condition was correlated inversely with category of infection (rs = 0.084, p = 0.025, n = 721); separate correlations were found in yearlings (rs = 0.213, p = 0.002, n = 227) and young animals (rs = 0.398, p = 0.011, n = 40). figure 4 illustrates the probability of normal body condition relative to age and category of infection, indicating that calves have poor body condition regardless of category of infection, and that some middle-aged animals have normal body condition despite high abundance of l1 in feces. old individuals were generally in normal body condition if not infected, although few were without infection. bone marrow fat content (n = 615) varied annually (table 4, fig. 5). on average, harvested animals had higher fat content (93%, 95% ci = 91 – 96) than unharvested animals (70%, ci = 66 – 75) with values corrected for time of year, sex, and age class (table 4). in a combined sample, a negative correlation was found between bone marrow fat content and category of infection (rs = −0.212, p < 0.001, n = 635). this negative correlation was found in calves (rs = −0.131, p = 0.020, n = 319), yearlings (rs = −0.223, p = 0.002, n = 193), young (rs = table 3. age in months (mean and 95% ci) of swedish moose examined for elaphostrongylus alces, 1985–1989. the column to the far right gives level of significance between harvested and euthanized + dead moose (for the last three rows the t-tests are performed on log transformed data; the data presented in the table are back-transformed values). if all sexed animals are combined, the sexes differed in age (p < 0.05). age class sex harvested unharvested t-value calves females 4.3 (4.1–4.4) 8.5 (8.1–8.8) 22.3, p < 0.001 males 4.2 (4.0–4.4) 8.1 (7.8–8.4) 21.1, p < 0.001 all calves‡ 4.2 (3.5–5.0) 8.3 (6.6–9.9) 4.28, p < 0.001 yearlings* females 18.4 (17.7–19.0) 19.2 (17.5–20.9) 0.91, p = 0.362 males 17.7 (17.0–18.4) 17.3 (15.6–19.0) 0.36, p = 0.716 combined 18.0 (16.9–19.1) 18.3 (15.4–21.1) 0.17, p = 0.863 young females 46.5 (42.8–50.4) 47.4 (43.0–51.7) 0.28, p = 0.782 males 42.0 (36.9–47.1) 44.6 (37.4–51.8) 0.58, p = 0.562 combined‡ 44.7 (42.0–47.4) 46.2 (42.9–49.5) 0.70, p = 0.481 middle-aged* females 90.3 (85.0–95.5) 93.4 (86.7–100.2) 0.74, p = 0.461 males 84.0 (74.4–93.6) 76.8 (65.5–88.1) 0.98, p = 0.333 combined‡ 89.0 (86.0–92.1) 89.1 (85.1–93.0) 0.01, p = 0.994 old females 144.0 (106–181.7) 154.2 (141.3–167.2) 0.52, p = 0.606 males 120.0 (144.6–195.4) 144.0 (100.5–187.5) 0.56, p = 0.580 combined‡ 146.0 (139.1–152.9) 153.4 (150.6–156.2) 1.94, p = 0.053 all females§ 12.3 (10.7–14.1) 29.6 (21.0–41.6) 4.73, p < 0.001 all males§ 9.0 (8.0–10.0) 14.2 (10.7–18.7) 3.92, p = 0.004 all moose 10.4 (9.6–11.4) 22.3 (17.4–28.4) 5.74, p < 0.001 *sexes differ by age class (all causes of death included). ‡includes individuals not sexed. §sexes differ with all age classes combined. alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 19 −0.618, p < 0.001, n = 46), and old (rs = −0.736, p < 0.001, n = 21), but not middleaged moose (rs = −0.319, p = 0.062, n = 35). time of infection the earliest identification of a calf diagnosed with elaphostrongylosis was at ~1.5 months on 21 july in region 3, county of calves yearlings young mid. age old 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 epidural space fr eq u en cy 0.0 0.2 0.4 0.6 0.8 1.0 females males lungs 0.0 0.2 0.4 0.6 0.8 1.0 feces total e. alces fig. 2. the mean abundance (frequency) of elaphostrongylus alces measured in adult worms in the epidural space of the spinal cord, and larvae in lungs and feces of harvested moose, sweden, 1985– 1989. the values are least-squared means and 95% confidence limits from a generalized linear model. the interaction between age groups and sex was a categorical predictor and julian date was a continuous predictor. 20 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 uppsala (table 5). the abundance of l1 was category 6 in the lungs and 4 in feces, and the calf was in normal body condition. the earliest calf death where worms were found in the epidural in the spinal cord was on 10 october (~4 months old) in region 2, county of värmland; the abundance of l1 was category 6 in the lungs and 3 in feces. worms were first found in harvested calves on 15 september (~3 months old). fig. 4. contour plot of the probability of normal body condition in moose (n = 613) versus age and 3 categories of elaphostrongylus alces infection intensity, sweden, 1985–1989. ca lve s ye arl ing s yo un g mid dle ag ed old p ro po rt io n (% ) 0 20 40 60 80 100 uninfected e. alces worms in epidural space only e. alces l1 in lungs e. alces l1 in feces h hh h hd d d d d fig. 3. the 4 categories of elaphostrongylus alces infection within 5 age groups of swedish moose, 1985–1989. h bars represent harvested moose (n = 761) and d bars represent unharvested moose (euthanized or found dead; n = 227). alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 21 table 4. analysis of bone marrow fat (%) and body condition of harvested and unharvested moose, sweden, 1985–1989. values are mean ± se with sample size in parentheses. pair-wise comparisons (t-values) of harvested and unharvested animals are provided in each age group (raw). means denoted by the same letter in each column (percent fat and body condition separately) were not different (p < 0.05). the values for percent fat are adjusted for time of year causing certain values to be >100% (see data management and statistics). variable age class harvested euthanized or dead statistic bone marrow fat (%) calves 83.6 ± 0.8a (270) 60.5 ± 1.8a (48) t = 11.6 p < 0.001 yearlings 97.1 ± 0.9b (182) 69.9 ± 3.8b (11) t = 6.88 p < 0.001 young 98.7 ± 2.1b (36) 83.6 ± 4.0c (10) t = 3.32 p < 0.001 middle aged 99.6 ± 2.3b (30) 77.4 ± 5.7bc (5) t = 3.62 p < 0.001 old 104.6 ± 5.7b (5) 63.3 ± 3.2ab (16) t = 6.34 p < 0.001 all animals 96.7 ± 1.3 (523) 70.9 ± 1.7 (92) t =11.6 p < 0.001 probability of normal condition calves 0.17 ± 0.03a (368) 0.25 ± 0.06a (102) z = 0.81 p = 0.420 yearlings 0.48 ± 0.04b (204) 0.30 ± 0.12c (35) z = 1.19 p = 0.235 young 0.50 ± 0.11bc (40) 0.47 ± 0.15b (27) z = 0.25 p = 0.800 middle aged 0.74 ± 0.12c (30) 0.64 ± 0.21b (17) z = 0.34 p = 0.734 old 0.84 ± 0.17abc (6) 0.12 ± 0.06ab (34) z = 2.53 p = 0.012 all animals 0.33 ± 0.02 (648) 0.29 ± 0.05 (215) z = 0.56 p = 0.578 julian date 25 50 75 100 125 150 175 200 225 250 275 300 325 350 b od y co nd iti on (b on e m ar ro w fa t % ) 0 20 40 60 80 100 fig. 5. the dependence of moose body condition (expressed as percent fat) on day of the year (julian date). in this analysis harvested moose are represented by a single value (julian date = 292.80; percent fat = 66.05). the line in the figure is the estimated peak function; a lorentzian peak function; y = 45.58 + [− 34.65/(1 + (((x – 115.05)/62.57))^2)]; f3,89 = 6.23, r2 = 0.174, p < 0.001. the peak are estimate to julian date = 115.05, which is 25 april. 22 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 the abundance of l1 was category 0 in the lungs and 6 in feces. first stage larvae (infection intensity = 6) were found in lungs from 14 august (~2 months old) to 4 june the following year (~12 months old). abundance of l1 in calves ranged from categories 1–6 by 2 months old, and the lung infection remained high; 81% had an l1 abundance category of 4–6 in the first year. the excretion of larvae began at a low level (2) on 14 august, and calves continued to excrete larvae throughout the first year at all levels of abundance (1–6). the prevalence of infection in harvested moose (n = 896) differed among regions (fig. 6), ranging from 13% in southernmost region 6 to 56% in region 3 (fig. 1 and table 2). infection was most prevalent in central sweden, least prevalent in southern sweden, and similar (p < 0.05) in southern and northern sweden. discussion although parasites at low abundance are generally less harmful to their host, when the host population increases rapidly, as with swedish moose in the 1970–80s (hörnberg 2001, stéen et al. 2005), an increasing risk to the individual and host population is possible (toft 1991). the proportion of elaphostrongylosis (symptoms of nervous disorder and/or emaciation) varies among age-classes in moose, with young animals more prone to illness (stéen et al. 2005). similarly, we found that e. alces worms located in the epidural space of the spinal cord were more prevalent in calves and yearlings, and only occasionally found in adults. the high abundance measured in young animals may simply reflect that the swedish moose population is skewed towards young animals (sand et al. 2011). conversely, abundance of l1 in lungs and feces was highest in old moose, and lowest in young and middle-aged moose. both stuve (1986) and stéen et al. (2005) suggested that e. alces most frequently infects males and young animals; however, we found no difference in the abundance within the epidural space, lungs, or feces between sexes or age groups of harvested moose, only a tendency toward males in the young age group. similarly, male reindeer calves with dominant mothers had higher abundance of e. rangiferi than female calves, and it was suggested that because these calves had better access to forage, they were at greater risk of ingesting infected gastropods (halvorsen 1986a). calf weight is dependent on summer browse availability in a table 5. age (in days) of moose calves infected by elaphostrongylus alces, sweden, 1985–1989. parasite location n mean age ± sd min age max age epidural 281 140.8 ± 20.0 50 215 lung 348 140.2 ± 20.8 50 215 feces 416 141.0 ± 17.6 101 215 region 1 2 3 4 5 6 p re va le nc e 0.1 0.2 0.3 0.4 0.5 0.6 0.7 a ab a b b b fig. 6. the prevalence of elaphostrongylus alces in regions of sweden. the values are calculated with logistic regression, regions was a categorical predictor and julian date a continuous predictor. the values are mean ± se; means with the same letter are not different (p > 0.05). only harvested moose were used in the analysis. alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 23 cow’s home range, with access to and quality of forage related to its relative status (saether and heim 1993). stuve (1986) attributed the difference in infection rate between sexes in older moose to physiological changes associated with the rut, as suggested with reindeer (halvorsen 1986b). a novel finding of our study was that l1 were found in lungs and feces of calves by 21 july, and adult worms in the epidural space by 15 september, or ~50–100 days after birth (broberg 2004, malmsten 2014). this prepatent period aligns with experimental infections of e. alces in moose in which patent infection was realized 39–73 days post-infection (stéen et al. 1997). because calves sample vegetation in the first days of life to promote development of rumen microbes (syroechkovsky et al. 1989), their potential to exposure to e. alces l3 is almost immediate. not surprisingly, adult e. alces were identified in the epidural space of the caudal vertebral canal in 2 other calves harvested in september (handeland and gibbons 2001). further, calves and yearlings were most frequently infected in the epidural space of the spinal cord which seemingly corroborates that moose shed most e. alces l1 during their early years, after which a sharp drop in larval shedding and low numbers of adult worms in older animals occur (stuve 1986, stéen et al. 2005). in both harvested and unharvested moose, e. alces worms were found in the epidural space of the spinal cord of animals aged 3–4 months to 4 years, not in middleaged animals, and again at 10–16 years. conversely, high levels of larvae were found in lungs and feces irrespective of age. we believe that the low frequency of worms in older animals, despite having l1 in lungs and shed larva, is due to migration from the cns/pns into the muscle fasciae, as with some other elaphostrongylins (lankester 2001). the pattern of e. alces adults migrating out to the muscle fasciae, presumably due to an immune response in the epidural space (stéen et al. 1997, 1998), differs somewhat from that of e. rangiferi, e. cervi, and p. tenuis. the latter are believed to remain in the cns as adult worms during their entire life (in the subdural or subarachnoid space, inside the meninges), although e. rangiferi also migrates to the muscle fasciae (hemmingsen et al. 1993). e. rangiferi, e. cervi, and p. tenuis may realize an immunological harbor within the cns, as might p. andersoni that is associated with blood vessels and connective tissues where females deposit eggs (lankester 2001). we hypothesize that e. alces worms are attacked by the immune system in the epidural space, and they migrate to the muscle fasciae where, with lower immunological defense, they deliver most of their larvae. after ingestion, l3 migrate from the gastrointestinal (gi) tract to the perineal cavity along the mesenchyme nerves, and into the abdominal wall associated with the more posterior lateral nerves. it is likely that e. alces does not need to enter the cns parenchyma to develop to the 5th stage (adult), as other elaphostrongylus spp., but remains epidurally-associated with lateral nerves of the pns and finally migrates to the muscle fasciae (olsson et al. 1998). the lack of worms in the epidural space of the spinal cord in moose during their prime could be explained by this migration; however, it could also reflect an immune response to prevent reinfection as described for p. andersoni that realizes declining larval output as deer age with few adult worms in deer >1 year old. further, repeated infection in whitetailed deer resulted in sharp decline in larval numbers and a strong cellular response to adult worms (lankester 2001). worms in the epidural space of older moose could simply be a reinfection associated with a weaker immune system, or an initial infection. whether some l3s migrate directly to muscle 24 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 fasciae without being associated with neural tissue is unknown. infected animals, on average, had lower body condition than uninfected animals except for middle-aged animals in their prime. calves were in poorer condition regardless of category of infection (as expected for young, growing animals), middle-aged were likely in normal condition despite high shedding rate of l1 in feces, and old individuals were in normal condition if uninfected. thus, infection, not age per se, seemed to reflect relative body condition. however, individual variation of immunological response to the parasite presumably exists because some individuals die young, others remain in normal condition through prime, and old animals are increasingly susceptible. in contrast with e. alces, no protostrongylid l1 of e. cervi were recovered from iberian red deer fawns (cervus elaphus hispanicus) (vicente and gortázar 2001). prevalence of e. cervi l1 increased with age of deer (vicente et al. 2006) which is opposite to our findings with e. alces in moose; both had higher infection rates in young males than females. the e. cervi pattern corresponds with that in reindeer in which e. rangiferi infects the host late in the season, remaining at the same intensity for at least 3 years (halvorsen et al. 1985). it appears that e. cervi and e. rangiferi have more similar and longer evolutionary relationships to each other and their respective hosts than e. alces. moose have a long, independent evolutionary history from the alceini and the plio-pleistocene, suggesting a peculiar adaption and habitat restriction of the species (niedziałkowska et al. 2014), and presumably, a relatively short evolutionary period with e. alces that could be less adapted with its host than e. cervi and e. rangiferi. it is possible that e. alces is more pathogenic to its host because both harvested and unharvested moose of below normal or emaciated body condition were infected with e. alces. in 2-year old moose, stuve (1986) found that infected moose were lighter (carcass weight) than uninfected moose, yet conversely, stéen et al. (1997) found that moose experimentally infected with e. alces retained normal weight when fed ad libitum. it remains unclear, however, if poor body condition is an indirect or direct effect of the parasite, that emaciation is either directly caused by an inflammatory response due to an epidural localization, or that elaphostrongylosis causes locomotor disorders making it difficult to move and feed (stéen and rehbinder 1986, stéen and roepstorff 1990, stéen et al. 2005). in summary, different morphology (stéen et al. 1989, stéen and johansson 1990, gibbons et al. 1991, lankester et al. 1998), genetics (gajadhar et al. 2000, chilton et al. 2005, huby-chilton et al. 2006), location (stéen et al. 1997, 1998) (epidural for e. alces, subdural/subarachnoid for e. rangiferi), and life span and host age relationships with infection (lankester 2001) suggest different, and perhaps, ongoing evolutionary adaption in elaphostrongylus species with their hosts. of further consequence is that rising temperatures, and a warmer and wetter climate are predicted to increase habitat, distribution, and abundance of mollusk hosts (halvorsen and skorping 1982, halvorsen et al. 1985), which in turn could lead to higher infection rates in cervids (handeland and slettback 1994, halvorsen 2012). although moose are not necessarily in poor condition when infected with e. alces, condition and parasite abundance were correlated. we therefore suggest continued surveillance of this disease and its specific consideration in management of moose in sweden. acknowledgements we thank w. e. faber, department of natural resources, central lakes college, brainerd, minnesota, usa for his feedback on the manuscript. we are grateful for all alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 25 technical help and assistance from employees, and earlier employees h. mann, s. persson, and i. forssell at the department of parasitology, national veterinary institute and swedish university of agricultural sciences, uppsala, sweden. last, but not least, we thank swedish hunters for their help and cooperation in data collection. financial support for this study was provided by the swedish environmental protection agency, stockholm, sweden. references anderson, r. c. 1964. neurological disease in moose experimentally infected with pneumostrongylus tenius from white-tailed deer. veterinary pathology 1: 289–322. doi: 10.1177/030098586400100402. baermann, g. 1917. eine einfache metode zur auffindung von ancylostoma(nematoden-) larven aus erdproben. mededeel uithet geneesk lab. te weltevreden, feestbundel, batavia, pp. 41–47 (in german). broberg, m. 2004. reproduction in moose: consequences and conflicts in timing of birth. doctoral thesis, gothenburg university, gothenburg, sweden. chilton, n. b., f. huby-chilton, m. w. lankester, and a. a. gajadhar. 2005. a method for extracting genomic dna from individual elaphostrongyline (nematoda: protostrongylidae) larvae and differentiation of elaphostrongylus spp. from parelaphostrongylus spp. by pcr assay. journal of veterinary diagnostic investigation 17: 585–588. doi: 10.1177/104063870501700612. engelsen etterlin, p., a. neimanis, d. gavier-widen, and c. hård af segerstad. 2009. postmortal hullbedömning av hull hos tamdjur och vilda djur (postmoral examination of body condition in pet animals and wildlife). national veterinary institute, uppsala, sweden (in swedish). gajadhar, a., t. steeves-gurnsey, j. kendall, m. lankester, and m. stéen. 2000. differentation of protostrongylid dorasal-spined larvae by pcr amplication of its-2 rdna. journal of wildlife diseases 36: 713–722. doi: 10.7589/0090-3558-36.4.713. gasaway, w. c., d. b. harkness, and r. a. rausch. 1978. accuracy of moose age determinations from incisor cementum layers. journal of wildlife management 42: 558–563. doi: 10.2307/3800818. gibbons, l. m.,o. halvorsen, and g. stuve. 1991. revision of the genus elaphostrongylus cameron (nematoda, metastrongyloidea) with particular reference to species of the genus occurring in norwegian cervids. zoologica scripta 20(1): 15–26. doi: 10.1111/j.1463-6409.1991.tb00272.x. halvorsen, o. 1986a. on the relationship between social status of host and risk parasitic infection. oikos 47: 71–74. doi: 10.2307/3565921. –––. 1986b. epidemiology of reindeer parasites. parasitology today 12: 334–339. doi: 10.1016/0169-4758(86)90053-0. –––. 2012. reindeer parasites, weather and warming of the arctic. polar biology 35: 1209. doi: 10.1007/s00300-012-1209-0. –––, and a. skorping. 1982. the influence of temperature on growth and development of the nematode elaphostrongylus rangiferi in the gastropods arianta abustrum and euconulus fulvus. oikos 38: 285–290. doi: 10.2307/3544666. –––, a. skorping, and k. hansen. 1985. seasonal cycles in the output of first stage larvae of the nematode elaphostrongylus rangiferi from reindeer, rangifer tarandus tarandus. polar biology 5: 49–54. doi: 10.1007/bf00446045. handeland, k., and l. m. gibbons. 2001. aspects of the life cycle and pathogenesis of elaphostrongylus alces in moose (alces alces). journal of parasitology 87: 1054–1057. doi: 10.1645/0022-3395 (2001)087[1054:aotlca]2.0.co;2. –––, and t. slettbakk. 1994. outbreaks of clinical cerebrospinal elaphostrongylosis in reindeer (rangifer tarandus tarandus) 26 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 in finnmark, norway, and their relation to climatic conditions. journal of veterinary medicine b 41(1–10): 407–410. doi: 10.1111/j.1439-0450.1994.tb00244.x. hemmingsen, w., o. halvorsen, and a. skorping. 1993. migration of adult elaphostrongylus rangiferi (nematoda: protostrongylidae) from the spinal subdural space to the muscles of reindeer (rangifer tarandus). journal of parasitology 79: 728–732. doi: 10.2307/3283612. huby-chilton, f., n. b. chilton, m. w. lankester, and a. a. gajadhar. 2006. single-strand conformation polymorphism (sscp) analysis as a new diagnostic tool to distinguish dorsal-spined larvae of the elaphostrongylinae (nematoda: protostrongylidae) from cervids. veterinary parasitology 135: 153–162. doi: 10.1016/j.vetpar.2005.08.001. hörnberg, s. 2001. changes in population density of moose (alces alces) and damage to forests in sweden. forest ecology and management 149: 141–151. doi: 10.1016/s0378-1127(00)00551-x. lankester, m. w. 2001. extrapulmonary lungworms of cervids. pages 228–278 in w. m. samuel, a. a. kocan, and m. pybus, editors. parasitic diseases of wild mammals, 2nd edition. iowa state university press, ames, ia. –––. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose population. alces 46: 53–70. –––, i.-m. c. olsson, m. stéen, and a.a. gajadhar. 1998. extra-mammalian larval stages of elaphostrongylus alces (nematoda: protostrongylidae), a parasite of moose (alces alces) in fennoscandia. canadian journal of zoology 76: 33–38. doi: 10.1139/z97-168. malmsten, j. 2014. reproduction and health of moose in southern sweden. doctoral thesis, swedish university of agricultural sciences, uppsala, sweden. niedziałkowska, m., k. j. hundertmark, b.jezdrzejewska, k.niedziałkowski, v. e. sidorovich, m. gorny, r. veeroja, e. j. solberg, s. laaksonen, h.sand, v. a.solovyev, m. shkvyria, j.tiainen,i.m.okhlopkov,r.juskaitis, g. done, v. a. borodulin, e. a. tulandin, and w. jezdrzejewski. 2014. spatial structure in european moose (alces alces): genetic data reveal a complex population history. journal of biogeography 41: 2173–2184. doi: 10.1111/jbi.12362. nordisk metodik-kommitté för livsmedel (nordic committee on food analysis). 1989. sbr (schmid-bondzynski-ratslaff), no. 131. esbo, finland. http://www.evira.fi (accessed july 2015). olsson, i.-m. 2001. elaphostrongylus alces – transmission, larval morphology and tissue migration. veterinary medicine licentiate thesis, swedish university of agricultural sciences, uppsala, sweden. –––, m. w. lankester, a. a. gajadhar, and m. steén. 1998. tissue migration of elaphostrongylus spp. in guinea pigs (cavia porcellus). journal of parasitology 84: 968–975. doi: 10.2307/ 3284629. saether, b. e., and m. heim. 1993. ecological correlates of individual variation in age at maturity in female moose (alces alces); the effects of environmental variability. journal of animal ecology 62: 482–489. doi: 10.2307/5197. sand, h., n. jonzén, h. andrén, and j. månsson. 2011. strategier för beskattning av älg. (strategies for moose management). forest facts, report from the swedish university of agricultural sciences, no. 2 4, p 4 (in swedish). sas 2014. version 9.4 ts level 1m0, x64_7pro platform, sas institute inc., cary, north carolina, usa. stéen, m., c. g. m. blackmore, and a. skorping. 1997. cross-infection of moose (alces alces) and reindeer (rangifer tarandus) with elaphostrongylus alces and elaphostrongylus rangiferi (nematoda, protostrongylidae): effects on parasite morphology and prepatent period. alces vol. 52, 2016 stéen et al. – epizootiology of elaphostrongylus alces 27 http://www.evira.fi veterinary parasitology 71: 27–38. doi: 10.1016/s0304-4017(97)00013-7. –––, a. g. chabaud, and c. rehbinder. 1989. species of the genus elaphostrongylus, parasite of swedish cervidae. a description of e. alces n. sp. annales de parasitologie, humanie et comparee 64: 134–142. –––, and c. johansson. 1990. elaphostrongylus spp. from scandinavian cervidae–a scanning electron study (sem). rangifer 10: 39–46. doi: 10.7557/2.10.1.789. –––, i.-m. olsson, and e. broman. 2005. diseases in a moose population subjected to low predation. alces 41: 37–48. –––, s. persson, and l. hajdu. 1993. protostrongylidae in cervidae and ovibus moscatus; a clustering based on isoelectric focusing on nematode proteins. rangifer 13: 221–223. doi: 10.7557/ 2.13.4.1121. –––, and c. rehbinder. 1986. nervous tissue lesions caused by elaphostrongylosis in wild swedish moose. acta veterinaria scandinavia 27: 326–342. –––, and l. roepstorff. 1990. neurological disorder in two moose calves (alces alces l.) naturally infected with elaphostrongylus alces. rangifer, special issue 3: 399–406. doi: 10.7557/2.10.3.887. –––, i. y. warsame, and a. skorping. 1998. experimental infection of reindeer, sheep and goats with elaphostrongylus spp. (nematoda, protostrongylidae) from moose and reindeer. rangifer 18: 73–80. doi: 10.7557/2.18.2.1448. stuve, g. 1986. the prevalence of elaphostrongylus cervi infection in moose (alces alces) in southern norway. acta veterinaria scandia 27: 397–409. syroechkovsky, e. e., e. v. rogacheva, and l. a. renecker. 1989. moose husbandry. pages 369–386 in r. j. hudson, k. r. drew, and l. m. baskin, editors. wildlife production systems. economic utilization of wild ungulates. cambridge university press, cambridge, england. swedish association for hunting and wildlife management. 2015. http:// jagareforbundet.se (accessed july 2015). swedish meteorological and hydrological institute (smhi). 2015. http:// www.smhi.se (accessed september 2015). systat. 2002. table curve 2d, version 5.01. systat software inc., san jose, ca. –––. 2008. sigmaplot for windows version 11.0, build 11.0.0.75. systat software inc., san jose, california, usa. toft, c. a. 1991. an ecological perspective: the population and community consequences of parasitism. pages 319–343 in c. a. toft, a. aeschlimann, and l. bolis, editors. parasite-host: association coexistence or conflict? oxford university press, oxford, england. vicente, j., i. g. fernández de mera, and c. gortazar. 2006. epidemiology and risk factors analysis of elaphostrongylosis in red deer (cervus elaphus) from spain. parasitology research 98: 77–85. doi: 10.1007/s00436-005-0001-2. –––, and c. gortázar. 2001. high prevalence of large spiny-tailed protostrongylid larvae in iberian red deer. veterinary parasitology 96: 165–170. doi: 10.1016/ s0304-4017(00)00425-8. 28 epizootiology of elaphostrongylus alces – stéen et al. alces vol. 52, 2016 http://jagareforbundet.se http://jagareforbundet.se http://www.smhi.se http://www.smhi.se epizootiology of elaphostrongylus alces in swedish moose study area methods data management and statistics results infection, age and sex body condition time of infection discussion acknowledgements references alces28_101.pdf winter distribution of moose at landscape scale in northeastern vermont: a gis analysis thomas l. millette, eugenio marcano, and danelle laflower geoprocessing laboratory, mount holyoke college, south hadley, massachusetts, usa. abstract: a gis analysis of landscape scale distribution of moose (alces alces) in northern vermont during winter 2010 showed that most moose were located at elevations of 300–600 m, with little discernible elevational gradient. slope and aspect were not correlated with locations as moose were distributed in the study area with the relative amount in each descriptive class. the distribution of >85% moose based on noaa cover types was in deciduous, mixedwood, and coniferous stands relative to their availability; locations in scrub/shrub and wetlands were higher and lower than expected, respectively. higher resolution aims imagery indicated that moose used mixedwoods more and coniferous stands less than available. the most significant landscape characteristic influencing the location of moose was proximity to forest openings/timber cuts that presumably provide important seasonal browse. alces vol. 50: 17–26 (2014) key words: alces alces, distribution, gis, moose, vermont, winter. introduction most landscape analyses of moose (alces alces) have been focused upon coarse-scale location and description of home ranges, or fine-scale seasonal habitat selection and utilization through use of radio-telemetered animals, aerial markrecapture surveys, and/or fieldwork (courtois and beaumont 2002, courtois et al. 2002, potvin and courtois 2004, poole and stuartsmith 2005, scarpitti et al. 2005, dussault et al. 2006, gillingham and parker 2008, van beest et al. 2010). although increased use of gps collars has provided more accurate and plentiful locations of individuals, most studies are limited by animal sample size due to the difficulty and cost associated with monitoring the broader population itself. a recently developed airborne thermal vertical-imaging system integrated with gis has presented the opportunity to identify and map locations of hundreds of moose during mid-winter (millette et al. 2011), with subsequent exploration of landscape attributes associated with their locations. because locations are accurately geocoded with gps by the thermal imaging system, it is possible to use a wide variety of off-the-shelf gis databases to model habitat characteristics. the landscape distribution and winter habitat use examined here is thought to be unique in that no such winter “snapshot” of a regional moose population has had such high sample size of animals. methods study area the study area was 682 km2 within wildlife management unit (wmu) e1located in the northeastern corner of vermont and bordered by new hampshire and quebec (fig. 1). the area is topographically expressive, and heavily forested with expansive maple (acer saccharum, a. pensylvanicum) and american beech (fagus grandifolia), stands of balsam fir (abies balsamea), spruce (picea rubens, p, glauca, corresponding author: thomas millette, geoprocessing laboratory, mount holyoke college, south hadley, ma, usa. 17 p. mariana), hemlock (tsuga canadensis), and eastern white pine (pinus strobus); conspicuous evidence of timber harvesting existed throughout. the estimated moose density based on a rolling 3-year average of moose sightings by november deer hunters was 0.89 moose/km2 (c. alexander, vermont fish and wildlife department). similarly, the estimated density was 0.84 moose/km2 during the aerial thermal census (millette et al. 2011) that produced the data for this study. data the study area was sampled with 35 survey units (su) distributed (relatively) evenly throughout. each was laid out nonrandomly in a gis to account for the variety of topographic settings and land cover types, while avoiding major changes in elevation along flight lines to maintain constant height above ground level (agl); thus, the image swath-width was as constant as possible during flights. the total area surveyed was 131.6 km2 or 20% of wmu e1. to insure that land cover types along flight line transects were representative of wmu e1, a gis overlay analysis was used to compare the proportions along transects with those from the noaa coastal service center land cover data (noaa 2006) (fig. 2). this analysis indicated that the relative proportions of cover types were almost identical between transects and the entire wmu such that no cover type was underor oversampled. details of the sampling design can be found in millette et al. (2011). the data were developed using the aims-thermal airborne imaging system. the sensor array pairs a 16-bit radiometric thermal camera to detect warm targets on a cold background, and simultaneously acquires 8-bit high resolution color photos to identify specific heat sources. unlike most aerial thermal systems used in previous research, the aims-thermal acquires its imagery vertically like a mapping system rather than using a low-oblique viewing angle while panning across the landscape. the vertical orientation of the cameras causes minimal screening effect in coniferous stands that is more typical of systems with oblique look-angles, and preserves uniform scale and spatial resolution throughout each image allowing detailed measurements within an image. a complete description of the aims-thermal system can be found in millette et al. (2011). the aims-thermal system was deployed in january and february 2010 over a 4-week period when 6 flights were flown between 0700 and 1100 hr. in total, these flights produced 94,605 thermal images and 12,530 high-resolution color images under continuous snow cover and sky conditions ranging from heavy overcast to bright sunshine. snow cover never exceeded 45 cm and no restrictive crust layers existed. fig. 1. location of the study area in wmu e1 in northeastern vermont, usa. 18 winter distribution of vermont moose – millette et al. alces vol. 50, 2014 all imagery exposure times and associated flight data were processed into gis attribute tables that support creation of shape‐ files containing photo centers for each exposure from the thermal and natural color cameras, as well as the flight path of the aircraft. this table also provides the framework for the integration of related spatial data such as sampling transects, flightlines, topography, and vegetation and facilitates the landscape-scale analysis described here. all gis database development operations were done using software developed by the researchers; all gis analyses used arcgis software tools. gis analyses an assessment of moose locations relative to landscape attributes was conducted to explore whether relationships or patterns existed that would describe habitat selection during the winter study period. the locations of 112 observed moose were used in a series of gis overlay procedures with the usgs national elevation data (gesch et al. 2002), the noaa coastal service center land cover data (noaa 2006), and the national agriculture imagery program (naip) vermont digital color orthophotography (2009) to examine the distribution of locations relative to elevation, slope, aspect, land cover type, and land management practices. all gis data were generalized to 90 m spatial resolution to limit landscape data heterogeneity. elevation was divided into 7–100 m classes with the majority (90%) of the landscape ranging from 300–700 m. slope was divided into 4 classes of <2.5°, 2.5°–5°, 5°– 10°, and >10°. aspect was divided into fig. 2. comparison of noaa land cover types with imagery transects that indicates the similarity between availability of cover types in the study area and the actual survey area in winter 2010, northeastern vermont, usa. alces vol. 50, 2014 millette et al. – winter distribution of vermont moose 19 the 8 cardinal and inter-cardinal points of the compass. the analysis of locations relative to land cover was done using classifications from the noaa csc land cover data with 30 m spatial resolution, which was generalized to 90 m to reduce artificial heterogeneity in the land cover data derived from landsat tm data, and to allow the spatial resolution of the land cover data to better match the 1.3 ha footprint of each thermal image. a second classification created from the aims-thermal natural color data with 3.2 cm spatial resolution was visually interpreted for each 2.2 ha image containing a moose. the aims-thermal classification differs from the noaa data since it had to be generalized into deciduous, mixedwood, and coniferous cover types due to snow cover which prevented accurate delineation of wetlands, scrub/shrub, and grassland. we performed a series of chi square goodness of fit tests (snedecor and cochran 1989) to determine if the distribution of moose was random among the different classes of elevation, slope, aspect, and land cover: x2 ¼ x o � eð þ2 e ð1þ where o = the number of observed moose in each category and e = the expected number of moose if the distribution were random and determined only by the proportion of the area sampled. for this test we counted the number of pixels in each gis layer (e.g., dem, land cover) that had a moose and those that did not, as well as the total number of pixels in each category. we then estimated the number of pixels that should be expected if the distribution were random. this estimation was adjusted to sum to 112, the number of moose observed. a similar analysis was done to test the randomness of moose distribution relative to the parameter distance to forest openings/cut areas. in this analysis, there is no underlying image to count pixels, so we generated approximately 10 random ground points within each survey unit (totaling 341) using the gis. the distance from these points and the locations of moose to cut areas were then compared in a similar way. results elevation the elevational distribution of moose was not random (p = 0.012); however, locations were not clustered at one elevational range. the majority of moose (78%) were at low to mid elevation (300–599 m), as was most of the study area (71%). at higher elevations (>599 m) use (21%) was less than available (28%), and use declined sharply above 699 m (fig. 3). slope there was no relationship between any slope category and locations (p = 0.444); the proportion of locations was correlated with (similar to) the proportion available in each category. the majority of locations (∼88%) were on slopes of <10° and were evenly distributed (27–32%) among the 3 classes of lower slope. moose were found at the highest slope category (>10°) at a rate of 13%, similar to what was available (11%) (fig. 4). aspect there was no relationship between aspect and location (p = 0.932) with moose located in all aspect categories (fig. 5). proportional use was highest in the east (20%) and lowest in the north (7%). the proportion of locations in north-northeast-east directions (38%) was slightly higher than in southeast-south-southwest directions (35%) with more solar exposure (fig. 5). land cover the land cover analysis indicated that the proportional use of cover types was 20 winter distribution of vermont moose – millette et al. alces vol. 50, 2014 not entirely proportional with availability (p = 0.001), although the proportional use (locations) and availability in the 3 most common cover types (deciduous, mixedwood, coniferous) were similar (88 and 87%, fig. 6), indicating no preference for any forest type. use of scrub/shrub was ∼3 x higher than available (4%), and conversely, use of wetlands was negligible with 4% availability (fig. 6); the scrub/shrub cover type represented young forest openings. there were certain differences between the analyses with the noaa and aims land cover data. availability of the 3 major forest cover types and use of the deciduous cover type was similar in both analyses; however, use was measurably lower (30 vs. 48%) in mixedwood and higher in coniferous (17 vs. 6%) in the noaa analysis than the aims analysis (fig. 7). in part, the difference was due to the reallocation of 13 moose (12% of total moose identified) fig. 3. the distribution of moose observations by elevation class (usgs national elevation data) indicating that most moose (78%) were located at 300–600 m, yet moose were observed at higher elevations similar to availability (χ2 =16.34, p = 0.0121) in winter 2010, northeastern vermont, usa. 36 32 30 14 32 29 27 13 28 28 34 11 0 5 10 15 20 25 30 35 40 <2.5 2.5-5 5.1-10 >10 m oo se o bs er va � on s slope classes in degrees moos e observa�ons by slope clas s percent moos e observa�ons percent study area fig. 4. the even distribution of moose observations by slope class (usgs national elevation data) indicating no relationship (χ2 =2.68, p = 0.444) between slope and location, including steep slopes in winter 2010, northeastern vermont, usa. alces vol. 50, 2014 millette et al. – winter distribution of vermont moose 21 located in the wetland and scrub/shrub categories in the noaa classification into either deciduous, mixedwood, or coniferous classes in the aims classification. further, the more accurate mapping of mixedwood and coniferous stands supported by the 3.2 cm aims imagery reclassified certain moose from coniferous to the mixedwood cover type. distance to timber cuts based upon visual analysis of >100,000 thermal and natural color images asso‐ ciated with the 2010 census, we had strong 8 13 22 13 13 12 15 16 7 12 20 12 12 11 13 14 10 12 17 11 10 13 15 12 0 5 10 15 20 25 north northeast east southeast south southwest west northwes t m oo se o bs er va � on s aspect moos e observa�ons percent moos e observa�ons percent study area fig. 5. the even distribution of moose observations by aspect class (usgs national elevation data) indicating no relationship (χ2 = 3.04, p = 0.9319) between aspect and location in winter 2010, northeastern vermont, usa. fig. 6. the distribution of moose observations by noaa land cover type indicating the uneven use (χ2 =22.39, p = 0.001) of shrub/scrub (higher) and wetlands (lower); major forest cover types were used relative to availability and accounted for the majority of observations (88%) in winter 2010, northeastern vermont, usa. 22 winter distribution of vermont moose – millette et al. alces vol. 50, 2014 anecdotal evidence that moose locations were influenced by the relative distance to forest openings associated with timber harvest. an analysis using vermont orthophotography at 1.0 m spatial resolution was done to measure the radial distance from each location to the nearest forest opening. the distribution of locations relative to the pro‐ ximity of forest openings was non-random (p < 0.0001). the strong, direct relationship was evident as 65% of all locations were within 100 m, 85% within 300 m, and 99% within 700 m of a forest opening (fig. 8). fig. 7. the distribution of moose observations with aims land cover types in winter 2010, northeastern vermont, usa. locations increased in mixedwood and declined in conifer relative to the proportional distributions based on noaa cover types. fig. 8. the distribution of moose observations relative to distance to forest opening/timber cut (vermont 2009 orthophoto data) in winter 2010, northeastern vermont, usa. a strong correlation (χ2 =133.09, p <0.0001) existed between distance and the proportion of locations with the majority of locations at <100 m. alces vol. 50, 2014 millette et al. – winter distribution of vermont moose 23 discussion this gis analysis of the winter distribution of moose at the landscape level indicated that, for the most part, moose were located throughout the study area in proportion with available cover types, and were little influenced by elevation to 600 m, slope, or aspect. we further investigated whether moose distribution between 300–500 m was influenced by availability of cover type and forest openings but found no pattern. the only obvious deviation in use and availability of cover types was in wetlands (lower) and scrub/shrub (higher); forage use is presumably minimal in wetlands during winter whereas scrub/shrub areas were likely regenerating forest providing preferred winter browse in the region (thompson et al. 1995, scarpitti et al. 2005, bergeron et al. 2011, andreozzi et al. 2014). both analyses with the noaa and aims land cover data were reasonably consistent with most locations either in deciduous or mixedwood forest areas (as expected) and fewer in coniferous forest. however, a lower observation rate in the coniferous cover type could possibly reflect reduced sightability due to higher canopy cover. due to the inaccessibility of the survey transects and resource limitations, no independent attempt was made to estimate the sightability or error rate of moose not captured in imagery. therefore, the 93 thermal images containing moose were analyzed to test if the camera lens parallax produced different probabilities of detection inside and outside the image nadir due to screening effects of trees. images with moose were divided into 5 zones, each representing 20% of the image area from the westedge to the east-edge, and the numbers of observations were totaled for each zone. the distribution of moose across these zones indicated that there was no apparent screening effect due to lens parallax since more observations were at the edges of images where parallax distortions are highest (fig. 9; millette et al. 2011). with regard to the noaa land cover assignments (fig. 6), the number of observations in coniferous stands was similar to the available coniferous forest, as it was in the deciduous and mixedwood stands, suggesting that the aims-thermal sensor with its vertical view angle did not suffer from the screening effects of coniferous canopy; overall, coniferous stands (with fig. 9. the distribution of aims imagery parallax observations (93 images with 112 moose) indicating that minimal screening effects probably occurred in the coniferous cover type (see millette et al. 2011). 24 winter distribution of vermont moose – millette et al. alces vol. 50, 2014 and without moose) as seen in the aims color imagery were not considered densely stocked with tight canopy closure. interestingly, in adjacent northern new hampshire, moose were also observed in all cover types with less powerful infrared technology (adams et al. 1997). a substantial number of observations moved from coniferous to mixedwood stands when using the aims versus noaa classification. because the aims land cover classification is the product of fine (3.2 cm) spatial resolution color imagery processed by an experienced photo-interpreter, it is considered to be more accurate than the noaa land cover classification derived from machine-processed landsat tm data that is less sensitive to distinctions between coniferous and mixedwood forest. analysis of most observations that were reassigned from the noaa coniferous cover type to the aims mixedwood cover type indicated that, although conifers were present, they represented <20% of the total forest cover in each image. therefore, we believe that the aims analysis provided more accurate use of cover types. high use of mixedwood forest during winter was also measured in northwestern quebec (courtois et al. 2002) and northern new hampshire (scarpitti et al. 2005). mixedwood forest can be ideal winter habitat if it contains openings that provide preferred winter forage and coniferous canopy that provides thermoregulatory cover if needed. about 20% of moose were observed bedded in the study and sheltered by a conifer in either coniferous or mixedwood stands. the most striking landscape metric identified in this study was the strong relationship between moose locations and proximity to forest openings/timber cuts (fig. 8). although this relationship is widely recognized (see peek 1997), this study is based on a very large sample size of moose that can be analyzed at both the landscape and local scale. because locations were in proportion to the availability of cover types, it is apparent that timber harvesting activity both influenced winter habitat use and was extensive throughout the study area. given that use of regenerating forest is generally temporal (about 10–15 years), regular use of these surveys could identify shifting habitat use and/or sites with high winter fidelity that are often of concern relative to adequate forest regeneration. for example, andreozzi et al. (2014) identified certain 10–20 year old clear-cuts in the study area that had poor regeneration. although this study was conducted principally to provide a population estimate, it also provided gis imagery databases that can be explored for current and temporal analyses of habitat use. specifically, high resolution aerial imagery can be used to produce detailed forest metrics of tree species, stocking density, dbh measurements, and shrub condition in areas of dense winter moose populations. further, given the concentration of wintering moose, it is possible to cost-effectively task additional survey flights at lower altitudes to produce color imagery with sufficient spatial resolution (1.0 cm) for detailed assessments of size, age, and sex of individuals, twinning rates, and potentially health status based on weight and coat condition. future studies that simultaneously measure habitat use and population characteristics with this technology will have the distinct advantages of large sample size and accurate temporal information regarding changes in cover type, land use, and moose population size and distribution. references adams, k. p., p. j. pekins, k. a. gustafson, and k. bontaites. 1997. evaluation of infrared technology for aerial moose surveys in new hampshire. alces 33: 129–139. alces vol. 50, 2014 millette et al. – winter distribution of vermont moose 25 andreozzi, h. a., p. j. pekins, and m. l. langlais. 2014. impact of moose browsing on forest regeneration in northeast vermont. alces 50: 67–79. bergeron, d. h., p. j. pekins, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. courtois, r., and a. beaumont. 2002. a preliminary assessment on the influence of habitat composition and structure on moose density in clear-cuts of northwestern quebec. alces 38: 167–176. ———,c.dussault,f. potvin, and g.daigle. 2002. habitat selection by moose (alces alces) in clear-cut landscapes. alces 38: 177–192. dussault, c., r. courtois, and j. p. ouellet. 2006. a habitat suitability index model to assess moose habitat selection at multiple spatial scales. canadian journal of forest research 36: 1097–1107. gesch, d., m. oimoen, s. greenlee, c. nelson, m. steuck, and d. tyler. 2002. the national elevation dataset. photogrammetric engineering and remote sensing 68: 5–11. gillingham, m., and k. parker. 2008. the importance of individual variation in defining habitat selection by moose in northern british columbia. alces 44: 7–20. millette, t. l., d. slaymaker, e. marcano, c. alexander, and l. richardson. 2011. aims-thermal a thermal and highresolution color camera system integrated with gis for aerial moose and deer census in northeastern vermont. alces 47: 27–37. naip – 1mclrnaip digital ortho photography (vermont). 2009. vermont center for geographic information, waterbury, vermont, usa. noaa coastal services center (noaa). 2006. c-cap zone 65 2006-era land cover classification of landsat scenes. noaa ocean service, coastal services center, charleston, south carolina, usa. peek, j. m. 1997. habitat relationships. pages 351–375 in a.w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. poole, k., and k. stuart-smith. 2005. finescale winter habitat selection by moose in interior montane forests. alces 41: 1–8. potvin, f., and r. courtois. 2004. winter presence of moose in clear-cut black spruce landscapes: related to spatial pattern or to vegetation? alces 40: 61–70. scarpitti, d., c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. snedecor, g., and w. cochran. 1989. statistical methods, 8th edition. iowa state university press, ames, iowa, usa. thompson, m. e., j. r. gilbert, g. j. matula jr., and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31: 233–245. van beest, f., m. atle, l. loe, and j. milner. 2010. forage quantity, quality and depletion as scale-dependent mechanisms driving habitat selection of a large browsing herbivore. journal of animal ecology 80: 771–785. 26 winter distribution of vermont moose – millette et al. alces vol. 50, 2014 winter distribution of moose at landscape scale in northeastern vermont: a gis analysis introduction methods study area data gis analyses results elevation slope aspect land cover distance to timber cuts discussion references alces(23)_107.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 moose antler morphology and asymmetry on isle royale national park kenneth j. mills1,3 and rolf o. peterson2 1department of biological sciences, michigan technological university, houghton, michigan, usa 49931; 2school of forestry and wood products, michigan technological university, houghton, michigan, usa 49931. abstract: isle royale national park, an island archipelago in lake superior, supports moose at higher density (1–4/km2) relative to most other north american sites. we compared antler size and asymmetry measurements from isle royale moose that died of natural causes to measurements available for other regional moose populations in published literature. we used these comparisons to test predictions that antlers of isle royale moose would be smaller and more asymmetric that other regional populations due to the high population density and the resulting ecological conditions on isle royale. moose on isle royale follow the same patterns of antler development as elsewhere, reaching maximum size at 7–8 years of age with slight declines after age 10–12. however, these moose develop antlers that are much smaller than all measured north american subpopulations. antler size was most comparable to moose from scandinavia where moose exist at comparably high population density. boone and crockett score, which is commonly used to compare antler size, performed poorly at ranking individuals with large antlers suggesting that more biologically relevant measures such as antler volume should be considered for comparisons of antler size. pedicle constriction was found to be a reliable indicator of senescence among old bulls. antler asymmetry was negatively related to antler size and was more extreme than asymmetry measured in alaskan moose. moose age had no detectable effect on the degree of antler asymmetry. in general, bull moose on isle royale develop smaller, more asymmetric antlers than other north american subpopulations which exist at lower density, consistent with the hypothesis that these qualities are related to nutrient limitation caused by high population density. results, however, may also reflect genetic differences and artifacts of sampling. alces vol. 49: 17–28 (2013) key words: alces alces, antler, asymmetry, development, isle royale national park, moose. moose (alces alces) develop large ant‐ lers during a relatively short growing period, requiring an intake of nutrients and expenditure of energy above that required for maintenance of basal functions (stewart et al. 2000). the ability to acquire and allocate resources necessary for antler development is influenced by factors such as age, body size, nutrition, genetics, and population and environmental conditions (sæther and haagenrud 1985, clutton-brock and albon 1989, markusson and folstad 1997, stewart et al. 2000, strickland and demarais 2000, bowyer et al. 2001, schmidt et al. 2001). as secondary structures in sexually dimorphic cervids, antlers have significance in sexual selection and are correlated with social dominance and mating success (cluttonbrock and albon 1989, bartoš 1990, solberg and sæther 1994, pélabon and joly 2000, stewart et al. 2000). these developmental, morphological, and sociobehavioral attributes allow antlers to be useful parameters in ecological research. 3present address: wyoming game and fish department, po box 850, pinedale, wy, 82941. kenneth.mills@wyo.gov 17 antler size typically increases until bulls reach maximum body and antler size between the ages of 5 and 10 years (stewart et al. 2000, bowyer et al. 2001). after age 10, antler size tends to decline (sæther and haagenrud 1985, bubenik 1990, bubenik 1998, stewart et al. 2000, bowyer et al. 2001), and simultaneously there is increasing evidence of physical senescence (hindelang and peterson 1994). age and body mass, then, both influence energetic investment in antler development (scribner and smith 1990). antler development patterns of isle royale moose that die of wolf predation and other natural causes will reflect overall nutritional condition as well as the culling influence of mortality factors. also, the large number of relatively old moose in the population (peterson 1977) should illuminate the poorly understood influence of senescence on antler development (bubenik 1998). asymmetry, defined as random deviations from perfect bilateral symmetry, is present to varying degrees in all bilateral morphological traits (palmer and strobeck 1986, bubenik 1990, bowyer et al. 2001). antlers are bilateral secondary structures and, therefore, portray differential degrees of asymmetry which depend on developmental stability, environmental quality, and individual fitness (e.g., nutritional status, inbreeding, injury, parasite load, age) and thus may be useful for comparisons between individuals and populations (palmer and strobeck 1986, clutton-brock and albon 1989, solberg and sæther 1994, alados et al. 1995, folstad et al. 1996, møller et al. 1996, markusson and folstad 1997, pélabon and van breukelen 1998, pélabon and joly 2000, bowyer et al. 2001, schmidt et al. 2001). antler asymmetry has an inverse relationship with antler size for many cervid species, which may be indicative of relative individual fitness regardless of age (markusson and folstad 1997, pélabon and van breukelen 1998, bowyer et al. 2001, ditchkoff et al. 2001). population wide stressors, such as reduced nutrition, may also manifest themselves through patterns in antler asymmetry and thus measures of antler asymmetry at broader scales may also be useful for comparisons between populations. reduced predator species diversity has allowed moose population density to reach uncommonly high levels on isle royale national park compared to most other north american subpopulations (peterson 1995, karns 1998, peterson et al. 2003), where a relative shortage of nutrition could reduce individual fitness and limit the ability of bull moose to allocate excess energy toward antler development (brown 1990). nutritional restriction due to high density may also manifest itself in the degree of antler asymmetry at the scale of the individual and the population (pélabon and van breukelen 1998, pélabon and joly 2000, bowyer et al. 2001). likewise, wolf predation and starvation are the only significant sources of mortality for moose on isle royale (peterson 1977, peterson 1999), so age structure and thus antler characteristics likely differ from other populations where antler morphology has been studied (gasaway et al. 1987, nygrén 2000, stewart et al. 2000, bowyer et al. 2001). therefore, antler characteristics may provide a basis for comparing condition and nutritional status of moose at isle royale and other geographic sites (bowyer et al. 2001). herein we assess antler size relative to age and antler asymmetry relative to age and antler size for bull moose collected on isle royale national park. we predict that patterns of antler development and asymmetry will follow similar general patterns measured for other north american populations. however, we also expect that antlers for moose on isle royale will be smaller and more asymmetric than other north american populations due to the nutritional 18 moose antler morphology – mills and peterson alces vol. 49, 2013 restriction caused by high population density (see also peterson et al. 2011). study area moose have existed on isle royale (544 km2) for the past century and in the last half-century they have been cropped by an unmanipulated population of gray wolves (canis lupus). both species have been protected since the establishment of isle royale national park in 1940 (mech 1966). wolf and moose populations have been counted each year since 1959. both predator and prey exist at relatively high density, with moose fluctuating from about 500 (1/km2) to over 2,000 (4/km2) animals during 1959– 2002, with a mean of 2.03 ± 0.11/km2 (se; range = 0.92–4.45/km2) during that period (peterson 1999, r. peterson, unpublished data). population densities for moose in other regions of north america are generally below 1/km2 (karns 1998). likewise, moose populations located on the nearest mainland in southwest ontario and northeast minnesota, the likely source for moose on isle royale, generally range from 0.20–0.40/km2 (mech 1966, karns 1998, ontario ministry of natural resources, unpublished data). methods skulls of male moose with polished antlers were collected during field studies at isle royale during 1970–2001. ages of moose were estimated from counts of annular cementum lines. antler size was measured in accordance with the boone and crockett club (b&c) scoring system (boone and crockett club 2011, gasaway et al. 1987). a net dry score for each set of antlers, tallied in inches, was calculated as follows: [spread + (2 � smallest palm length) + (2 � smallest palm width) + (2 � smallest beam circumference) + (2 � least number of points)] (see boone and crockett club 2011 for details on scoring methods). the remaining measurements were recorded in centimeters (gasaway et al. 1987). the largest diameter of both left and right pedicles on each skull was measured to study how this skull character varies with age. some pedicles showed an apparent constriction at the point where the antler joins the pedicle, which has not been described previously in the scientific literature (fig. 1). therefore, both constricted and unconstricted pedicle measurements were taken for these individuals in order to quantify this morphological trait. the constricted measurement was taken at the area of greatest constriction just before the antler base, while the unconstricted measurement was taken directly medial to the constricted area. scoring systems such as b&c may have limitations that affect the results of comparative studies (gasaway et al. fig. 1. constriction of the pedicle (outlined in white) just medial to the base of the antler was evident for many antlered bulls collected from isle royale national park. alces vol. 49, 2013 mills and peterson – moose antler morphology 19 1987, bubenik 1998). therefore, we also determined antler volume to directly measure antler size using water displacement. prior to measurement, each antler was saturated in water until all air pockets were filled prior to measurement. in order to measure the accuracy of this technique, we determined volume for 10 antlers, 3 times each. each individual measurement for each antler was compared to the mean of the 3 measurements for that antler to determine the error of each measurement. finally, the total mean error of the 30 measurements was calculated to confirm that the error was within acceptable limits (i.e., < 5%). we then compared two of the most used measures of antler size, b&c score and spread (boone and crockett club 2011), to the respective total volume measurement for each individual to determine the degree to which these scores accurately estimate antler size using exponential regression. second-order polynomial equations were fitted to data relating antler character size to moose age to evaluate variation in antler size with age and age-related growth of antlers compared to that of alaskan moose as measured by bowyer et al. (2001). a dunnett's test (zar 1999) was used to determine if the mean maximum sizes for the 20 largest isle royale moose for both b&c score and spread were smaller than the same measurements from multiple subpopulations of north american moose, as determined by gasaway et al. (1987), and moose from finland as determined by nygrén (2000). we also plotted comparative growth curves for isle royale moose, selected north american subpopulations, and a swedish subpopulation of moose as adapted from gasaway et al. (1987). growth curves were determined by using 3-year running averages except for the oldest and youngest age classes, which are presented as actual means. we pooled individuals in the 14 year age class and older for the isle royale subpopulation. relative antler asymmetry was determined by taking the difference between the large and small side of each measured antler parameter for each individual (i.e., palm width, palm length, beam circumference, number of points, pedicle diameter, and volume) divided by the respective large side for each measured antler parameter for that individual (e.g., [large palm width – small palm width] ÷ large palm width = relative asymmetry of the palm width for that individual moose). we then assessed the relationship between relative asymmetry and moose age using linear regression. we also used linear regression to measure the relationship between relative asymmetry and the mean size of the respective antler parameter. we used a one-sample t-test to compare the mean relative asymmetry for palm width, palm length, beam circumference, and number of points for isle royale moose to the mean relative asymmetry of the respective measures for alaskan moose as determined by bowyer et al. (2001). we tested whether asymmetry was fluctuating or directional for each lateral antler character using a wilcoxon signedrank test (see palmer and strobeck 1986, zar 1999, pélabon and joly 2000, bowyer et al. 2001). results the total number of skulls in the sample was 106, but not all parameters could be measured for some specimens because of weathering prior to collection. antlers for isle royale moose were smaller than alaskan subpopulations in palm width, palm length, beam circumference, number of points and spread (fig. 2, 3). for b&c score and spread, isle royale moose were smaller than all other north american subpopulations measured (all p <0.05; table 1). antler spread from isle royale 20 moose antler morphology – mills and peterson alces vol. 49, 2013 isle royale y = –0.2219× 2 + 4.6907× –3.5515 r2 = 0.2962 p<0.001 y = –0.6256×2 + 13.112× –12.102 r2 = 0.3778 p<0.001 y = –1.0131×2 + 20.658× +9.3598 r2 = 0.3638 p<0.001 y = –24.816×2 + 513.7× –419.44 r2 = 0.278 p<0.001 y = –0.0876×2 + 1.8086× +7.3762 r2 = 0.3642 p<0.001 y = –0.043×2 + 0.897× + 1.1549 r2 = 0.194 p<0.001 isle royale alaska p a lm w id th ( cm ) 45 40 35 30 25 20 15 10 5 0 25 20 15 10 5 0 4500 4000 3500 3000 2500 2000 1500 1000 500 0 p a lm le n g th ( cm ) b & c s co re ( in ) v o lu m e ( m l ) b e a m c ir cu m fe re n ce ( cm ) 110 100 90 80 70 60 50 40 30 20 10 0 160 140 120 100 80 60 40 20 0 age (years) 0 2 4 6 8 10 12 14 age (years) 0 2 4 6 8 10 12 14 n u m b e r o f p o in ts 12 11 10 9 8 7 6 5 4 3 2 1 0 fig. 2. regression analyses of antler characteristics in relation to age of bull moose collected from isle royale national park. raw data was used to generate a second order polynomial regression equation for isle royale moose. regression lines for alaskan moose were obtained from bowyer et al. (2001). sample sizes for the isle royale sample are as follows: palm width, n = 68; number of points, n = 74; palm length, n = 67; beam circumference, n = 91; b&c score, n = 64; volume, n = 68. alces vol. 49, 2013 mills and peterson – moose antler morphology 21 moose was also smaller than the palmate antler category from finland (∣q∣ = 3.1696, p <0.05), and was marginally different from the non-palmate antler category (∣q∣ = 1.9245, p ≈ 0.05; table 1). isle royale moose also appear to have maximum antler spread similar to that of moose from sweden, although raw data were not available for the swedish subpopulation (fig. 3). for moose at isle royale, maximum antler size is reached between the ages of 7 and 8 years for all measured parameters, except for b&c score, which reached its maximum at 6 years (fig. 2, 3). generally, a slight decrease in size occurred after 10–12 years of age, with incipient physi‐ cal senescence (fig. 2). this was evident by the malformed or misshapen antlers of several senescent individuals (see bubenik 1998). the volume measurement technique was determined to be accurate to within a mean of 1.9 ± 0.3% (range = 0.2–5.5%). age-related change in antler volume was similar to other size measurements, reaching a maximum at age 7, then decreasing more slightly after age 10 (fig. 2). the relationship between b&c score and total volume (left + right) was exponential and variable for individuals with high b&c scores (fig. 4a). antler spread also was exponentially related to total volume and was more variable as spread increased (fig. 4b). pedicle diameter portrayed the same antler development pattern as other parameters, reaching maximum size at 8 years (fig. 5a). however, it did not appear to decline as an indication of senescence as other parameters did. pedicle constriction was present in some moose as early as 7 years and increased with age to a maximum at 16–18 years (fig. 5b). the degree of relative asymmetry was not related to moose age for any bilateral antler parameter (all p > 0.458), but was negatively related to antler size for most bilateral antler categories including volume (f = 0.27, p = 0.002; fig. 6), palm width (f = 1.61, p = 0.000), beam circumference (f = 10.82, p = 0.001), and number of points (f = 0.74, p = 0.000). relative asymmetry had no relationship with antler size for palm length (f = 0.07, p = 0.799) or pedicle diameter (f = 0.15, p = 0.697) the degree of relative asymmetry for isle royale moose was much larger than in alaskan moose for palm length, palm width, and beam circumference but was not different for number of points (table 2). wilcoxon signed-rank tests showed that left and right antler sides were not different for palm length, palm width, beam circumference, number of points, volume, or pedicle diameter (z = 0.061, p = 0.952; z = 1.056, p = 0.291; z = 0.002, p = 0.998; z = −0.836, p = 0.403; z = 0.679, p = 0.497; z = 0.808, p = 0.419, respectively). fig. 3. comparative growth curves for selected north american subpopulations and a swedish subpopulation of moose as adapted from gasaway et al. (1987). curves are plotted by using 3-year running averages except for the oldest and youngest age classes, which are actual means. for the isle royale national park subpopulation (n = 76), individuals in the 14 year age class and older are pooled. 22 moose antler morphology – mills and peterson alces vol. 49, 2013 discussion population density for moose on isle royale, where there is predation only by gray wolves, is an order of magnitude higher than most other areas of north america (peterson 1999), but comparable to many moose ranges in scandinavia (0.8–1.8/km2; cederlund and markgren 1987, hörnberg 2001). isle royale moose, to a greater extent than other moose populations, are also subjected to strong selection by wolf predation, and are thereby more naturally regulated than other hunted populations. these two ecological characteristics make interpopulation comparisons involving moose at isle royale particularly compelling. however, it is necessary to address this difference in terms of sample selection when comparing datasets collected from individuals subjected to natural mortality and those collected from hunter-killed individuals. neither sample is randomly selected; in the case of isle royale, individuals were collected after death from natural causes, and so probably include proportionately higher numbers of individuals in poor condition and/or older age classes. with other datasets, individuals were measured table 1. antler spread and boone and crockett score for the 20 largest moose from selected regions of north america and finland. (data adapted from gasaway et al. 1987 and nygrén 2000). spread (cm) boone and crockett score subspecies/region mean se max. mean:max. mean se max. n gigas alaska1 182.6 2.64 207 0.88 247.1 0.71 255 20 gigas x andersoni yukon and northwest territories1 170.2 2.18 191.8 0.89 232.9 1.57 247.3 20 gigas x andersoni northern british columbia2 154.7 2.64 172.7 0.90 215.7 0.91 229.1 20 andersoni western canada (except north british columbia) and minnesota2 154.7 2.41 178 0.87 217.3 1.27 226.9 20 andersoni x americana ontario2 151.6 2.79 181.6 0.83 201.3 1.35 211.6 20 americana eastern canada and maine2 154.4 2.49 181.9 0.85 202.9 2.73 238.6 19 shirasi western usa3 133.9 2.69 151.9 0.88 188.2 1.73 205.5 20 andersoni isle royale2 107.0 3.18 129.4 0.83 133.4 2.04 151.7 20 alces finland palmate 114.8 0.46 149 0.77 511 nonpalmate 111.9 0.86 139 0.81 1considered alaska-yukon moose by boone and crockett club. 2considered canadian moose by boone and crockett club. 3considered shiras moose by boone and crockett club. alces vol. 49, 2013 mills and peterson – moose antler morphology 23 following hunter harvest, which would introduce biases based on hunter selection (e.g., hunter selection for larger than average bulls, antler size restrictions imposed by wildlife management agencies). the mean:maximum ratios presented in table 1 suggest that the regional datasets are likely similar, and therefore comparable. it is likely that the true maximum antler size realized by isle royale moose is larger than that presented herein, but cast antlers that are significantly larger than the largest represented in this dataset are rarely found during fieldwork on the island (r. peterson, michigan technological university, unpublished data). this evidence suggests that these datasets have at least acceptable levels of comparability, but comparisons should still be considered with a y = 369.46e0.0212x r 2 = 0.7808 p < 0.001 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140 160 boone & crockett score (in) t o ta l v o lu m e ( m l ) b y = 0.0019x3.196 r 2= 0.758 p < 0.001 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 spread (cm) t o ta l v o lu m e ( m l ) fig. 4. regression analyses of the relationship between total antler volume and (a) boone & crockett score (n = 65) or (b) spread (n = 65) for antlered bull moose collected from isle royale national park. a y = 0.0173x2 + 0.0085x – 0.0498 r 2 = 0.2987 p < 0.001 0 2 4 6 8 10 age (years) m e a n p e d ic le c o n st ri ct io n ( m m ) b y = –0.2584x2 + 6.6782x + 25.188 r 2 = 0.6205 p < 0.001 0 20 40 60 80 100 120 0 2 4 6 8 10 12 14 16 18 0 2 4 6 8 10 12 14 age (years) l a rg e st p e d ic le d ia m e te r (m m ) fig. 5. regression analysis for (a) pedicle constriction (n = 91) and (b) pedicle diameter (n = 95) in relation to age of antlered bull moose collected from isle royale national park. y = –5e-05x + 0.2754 r 2= 0.1403 p = 0.002 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 1000 2000 3000 4000 5000 6000 7000 8000 mean volume (ml) r e la tiv e a sy m m e tr y fig. 6. linear regression between relative asymmetry of left and right antler sides against mean antler volume for bull moose collected from isle royale national park (n = 68). 24 moose antler morphology – mills and peterson alces vol. 49, 2013 caution due to the potential biases caused by differences in sampling methodologies (e.g., sample sizes, sampling duration, sample collection protocols). moose present on isle royale develop smaller antlers than all other reported sub‐ populations in north america, and their antlers are similar to or smaller than two subpopulations reported for scandinavia. however, antler development through age follows much the same patterns as other populations, reaching a maximum size after 7 to 8 years, which is maintained until senescence at around age 12 (gauthier and larsen 1985, bowyer et al. 2001). the fact that isle royale moose appear to have a restricted ability to produce larger antlers should be a function of ecological conditions on the island, with nutrient limitation induced by high population density being the most fundamental difference between this island population and those in mainland areas. this is demonstrated when comparing antler size of isle royale moose to antler measurements collected from moose in southwest ontario. the maximum antler spread and b&c score for isle royale was 22.2 cm or 49.6 in smaller than the mean of the 19 largest moose measured from the mainland ontario population (table 1). this analysis suggests a significant reduction in antler size in the century following moose colonization on the island, with the primary difference between these groups being population density (karns 1998, peterson 1999, r. o. peterson, unpublished data, ontario ministry of natural resources, unpublished data). most comparative antler studies use composite scores of linear measures such as the b&c scoring system or simply antler spread. these scores are easy to calculate, but may have significant limitations (gasaway et al. 1987, bubenik 1998). we determined that b&c score and spread do not accurately rank large antlered individuals, in many cases ranking larger individuals below smaller individuals (figs. 4a and b). the b&c score, therefore, may have limited usefulness when comparing antler size between similar populations, especially when comparing primarily large antlered bulls. volume, on the other hand, should be a more accurate measure of antler size because it is directly related to energetic investment during antler development. this suggests that researchers should consider biologically relevant morphological metrics such as volume when conducting comparative studies on antlers. moose numbers on isle royale are naturally regulated with no human interference, which allows individual moose the opportunity to reach ages when signs of senescence would be expected. in most cases, the second order polynomials used in regression estimated reductions in size for older individuals. despite this, most measured antler parameters had only slight reductions in antler size for post-prime age individuals, table 2. comparison of mean relative asymmetry (ra; large small/large) for antler characters from 1,501 harvested alaskan moose and antlered bull moose collected from isle royale national park. data for antler characters from alaska were obtained from bowyer et al. (2001) and gasaway et al. (1987). alaska isle royale antler character ra se ra se n t p palm width 0.10 0.002 0.20 0.033 67 3.04 0.003 palm length 0.07 0.002 0.16 0.034 66 2.60 0.011 beam circumference 0.03 0.001 0.06 0.011 100 2.46 0.016 # points 0.19 0.005 0.20 0.026 74 0.44 0.660 alces vol. 49, 2013 mills and peterson – moose antler morphology 25 which was consistent for moose measured in alaska (bowyer et al. 2001). however, there was a small proportion of old and senescent individuals that developed small and drastically asymmetric antlers (see bubenik 1998). pedicle constriction may be a better indicator of declining reproductive vigor in older individuals. pedicle constriction was observed in both large and small antlered individuals as well as individuals with normal and abnormal antler morphology. constriction was first apparent in some bulls that were 7 years of age, the same age that antlers begin to reach their maximum, mature size, and it increased with age, though not all older individuals had measurable restrictions. a. b. bubenik (pers. commun.) suggested that pedicle constriction resulted from testosterone insufficiency, which may begin well after sexual maturity and increase with reproductive senescence. antler asymmetry for moose on isle royale was fluctuating and was most pronounced among moose with small antlers at the extremes of age and development. although some older, senescent individuals developed very small and asymmetric antlers (see bubenik 1998), overall there was little evidence to suggest that age has any governing effect on antler asymmetry. therefore, antler asymmetry should be a valid indicator of individual fitness and condition regardless of age, with the individuals in the best condition developing the largest and most symmetric antlers. likewise, asymmetry may also provide a basis for comparisons of fitness and condition between populations. in this case, isle royale moose portrayed greater degrees of relative asymmetry than alaskan subpopulations, the only subpopulation for which asymmetry measurements were available (bowyer et al. 2001). high levels of antler asymmetry population-wide, as measured for moose from isle royale, may reflect more nutrient limitation and developmental instability. in general, bull moose on isle royale develop smaller, more asymmetric antlers than other north american subpopulations, even those within the same geographic region, suggesting that these qualities are the result of nutrient limitation caused by high population density (peterson et al. 2011). these findings are consistent with the evidence of slight dwarfism associated with high population density and lack of selection by wolf predation during the first half of the 20th century (peterson et al. 2011). however, this correlative study did not quantify or eliminate other potential contributing factors, such as genetic founder effects and effects of sampling methodology. this study also supports the contention that antlers are useful indicators for both individual and population condition (e.g., markusson and folstad 1997, pélabon and van breukelen 1998, strickland and demarais 2000, schmidt et al. 2001), although future research should attempt to specifically evaluate fitness in relation to measures of antler size and asymmetry for moose. acknowledgements we thank p. dewitt, e. parker, c. peterson, and d. suhonen for assistance in antler measurements, and financial assistance from the u.s. national science foundation (deb-0918247), the u.s. national park service (co-op agreement no. j631005n0040003), and rop, the robbins chair at michigan technological university in sustainable management of the environment. references alados, c. l., j. escós, and j. m. emlen. 1995. fluctuating asymmetry and fractal dimension of the sagittal suture as indicators of inbreeding depression in dama and dorcas gazelles. canadian journal of zoology 73: 1967–1974. bartoš, l. 1990. social status and antler development. pages 442–459 in 26 moose antler morphology – mills and peterson alces vol. 49, 2013 g. a. bubenik and a. b. bubenik, editors. horns, pronghorns, and antlers: evolution, morphology, physiology and social significance. springer-verlag, new york, new york, usa. boone and crockett club. 2011. http:// www.boone-crockett.org/pdf/sc_moose. pdf . bowyer, r. t., k. m. stewart, j. g. kie, and w. c. gasaway. 2001. fluctuating asymmetry in antlers of alaskan moose: size matters. journal of mammalogy 82: 814–824. brown, r. d. 1990. nutrition and antler development. pages 426–441 in g. a. bubenik and a. b. bubenik, editors. horns, pronghorns, and antlers: evolution, morphology, physiology and social significance. springer-verlag, new york, new york, usa. bubenik, a. b. 1990. epigenetical, morphological, physiological, and behavioral aspects of evolution of horns, pronghorns, and antlers. pages 3–113 in g. a. bubenik and a. b. bubenik, editors. horns, pronghorns, and antlers: evolution, morphology, physiology and social significance. springer-verlag, new york, new york, usa. ———. 1998. evolution, taxonomy and morphophysiology. pages 77–124 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 1st edition. smithsonian institution press, washington, d. c., usa. cederlund, g., and g. markgren. 1987. the development of the swedish moose population, 1970–1983. swedish wildlife research supplement. 1: 55–61. clutton-brock, t. h., and s. d. albon. 1989. red deer in the highlands. bsp professional books, oxford, england. ditchkoff, s. s., r. l. lochmiller, r. e. masters, w. r. starry, and d. m. leslie jr. 2001. does fluctuating asymmetry of antlers in white-tailed deer (odocoileus virginianus) follow patterns predicted for sexually selected traits. proceedings of the royal society of london, b 268: 891–898. folstad, i., p. arneberg, and a. j. karter. 1996. antlers and parasites. oecologia 105: 556–558. gasaway, w. c., d. j. preston, d. j. reed, and d. d. roby. 1987. comparative antler morphology and size of north american moose. swedish wildlife research supplement 1: 311–325. gauthier, d. a., and d. g. larsen. 1985. geographical variation in moose antler characteristics, yukon. alces 21: 91–101. hindelang, m., and r. o. peterson. 2000. skeletal integrity in moose at isle royale national park: bone mineral density and osteopathology related to senescence. alces 36: 61–68. hörnberg, s. 2001. changes in population density of moose (alces alces) and damage to forests in sweden. forest ecology and management 149: 141–151. karns, p. d. 1998. population distribution, density and trends. pages 125–140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 1st edition. smithsonian institution press, washington, d. c., usa. markusson, e., and i. folstad. 1997. reindeer antlers: visual indicators of individual quality? oecologia 110: 501–507. mech, l. d. 1966. the wolves of isle royale. united states national park service fauna series number 7. united states government printing office, washington, d. c., usa. møller, a. p., j. j. cuervo, j. j. soler, and c. zamora-muñoz. 1996. horn asymmetry and fitness in gemsbok, oryx g. gazella. behavioral ecology 7: 247–253. nygrén, k. 2000. directional asymmetry in moose. alces 36: 147–154. palmer, a. r., and c. strobeck. 1986. fluctuating asymmetry: measurement, analysis, pattern. annual review of ecology and systematics 17: 391–421. alces vol. 49, 2013 mills and peterson – moose antler morphology 27 http://www.boone-crockett.org/pdf/sc_moose.pdf http://www.boone-crockett.org/pdf/sc_moose.pdf http://www.boone-crockett.org/pdf/sc_moose.pdf pélabon, c., and l. van breukelen. 1998. asymmetry in antler size in roe deer (capreolus capreolus): an index of individual and population fitness. oecologia 116: 1–8. ———, and p. joly. 2000. what, if anything, does visual asymmetry in fallow deer antlers reveal? animal behaviour 59: 193–199. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. national park service scientific monograph series no. 11. united states government printing office, washington, d. c., usa. ———. 1995. the wolves of isle royale: a broken balance. willow creek press, minocqua, wisconsin, usa. ———. 1999. wolf-moose interaction on isle royale: the end of natural regulation? ecological applications 9: 10–16. ———, d. beyer, m. schrage, and j. räikkönen. 2011. phenotypic variation in moose: the island rule and the moose of isle royale. alces 47: 125–133. ———, j. a. vucetich, r. e. page, and a. chouinard. 2003. temporal and spatial aspects of predator-prey dynamics. alces 39: 215–232. sæther, b.-e., and h. haagenrud. 1985. geographical variation in the antlers of norwegian moose in relation to age and size. journal of wildlife management 49: 983–986. schmidt, k. t., a. stien, s. d. albon, and f. e. guinness. 2001. antler length of yearling red deer is determined by population density, weather and early lifehistory. oecologia 127: 191–197. scribner, k. t., and m. h. smith. 1990. genetic variability and antler development. pages 460–473 in g. a. bubenik and a. b. bubenik, editors. horns, pronghorns, and antlers: evolution, morphology, physiology and social significance. springer-verlag, new york, new york, usa. solberg, e. j., and b.-e. sæther. 1994. male traits as life-history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammology 75: 1069–1079. stewart, k. m., r. t. bowyer, j. g. kie, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36: 77–83. strickland, b. k., and s. demarais. 2000. age and regional differences in antlers and mass of white-tailed deer. journal of wildlife management 64: 903–911. zar, j. h. 1999. biostatistical analysis. 4th edition. prentice hall, upper saddle river, new jersey, usa. 28 moose antler morphology – mills and peterson alces vol. 49, 2013 moose antler morphology and asymmetry on isle royale national park study area methods results discussion acknowledgements references alces26_14.pdf alces29_263.pdf alces21_191.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 moose habitat use throughout gros morne national park krystal kerckhoff1, brian e. mclaren1, shane p. mahoney2 and tom w. knight3 1lakehead university, faculty of natural resources management, 955 oliver road, thunder bay, on, canada p7b 5e1; 2government of newfoundland and labrador, sustainable development and strategic science division, 2 canada drive, st. john’s, nl, canada a1b 4j6; 3gros morne national park, p.o. box 130, rocky harbour, nl, canada a0k 4n0. abstract: previous research indicated high variability in availability and habitat use by female moose in the lowlands of gros morne national park (gmnp), newfoundland and labrador, an area dominated by bogs and forest. here, we extend the earlier analysis with an additional 7 female moose (alces alces americana) occupying the park highlands, a region dominated by heath and shrub vegetation with forest limited to sheltered valleys, typical of interior and highland parts of the province. resource selection function (rsf) models with differences in habitat use between moose resident in the 2 regions and 2 moose that migrated from the lowlands in winter to the highlands in summer were rejected. in summer, more use of closed-canopy forest types occurred on the lowlands, while more use of non-forest habitat types occurred on the highlands. as before, we found that selection of disturbed forest is a winter phenomenon on the lowlands of gmnp; the same series of habitat types associated with disturbance were avoided in summer. summer migration by about 20% of gmnp moose to the highlands suggests that foraging opportunities are better during that season than in winter, a motivation for migration perhaps augmented by an overabundance of moose on the lowlands and unfavourable temperatures in disturbed areas that might otherwise serve as lowland foraging areas. an observation of more clustered relocations of moose on the highlands than on the lowlands of gmnp is consistent with our conclusion that moose use habitats within the highlands and lowlands of newfoundland and labrador very differently. we recommend 2 approaches to moose management for these different landscapes, both within gmnp and elsewhere in newfoundland and labrador. alces vol. 49: 113–125 (2013) key words: gros morne national park, habitat selection, moose, newfoundland and labrador, resource selection functions. according to habitat selection theory (fretwell and lucas 1970), individuals distribute themselves in a manner proportional to the quantity or quality of limiting resources available in each of several foraging patches, larger habitat units, and still larger landscapes. for ungulates, habitat selection should be driven by an individual’s ability to sense and select higher-quality food items or foraging areas (mcnaughton 1985, fryxell 1991). habitat selection that involves migration between 2 different landscapes can arise in a seasonal climate where different fitness opportunities (or forage availability) are offered by each landscape, but where the difference is less during the growing season (holt and fryxell 2011). moose (alces alces americana) in newfoundland and labrador, canada presumably distribute themselves optimally according to habitat selection theory in each of 2 typical landscapes in this province, the “highlands” and the “lowlands.” we explore this idea with analysis of summer and winter location data from gps-collared moose, using the assumption that more forested, lowland landscapes are, on average (i.e., throughout the year), superior to the 113 highlands where some moose migrate during summer. this paper is motivated by previous study of gros morne national park (gmnp), an area of 1,805 km2 in western newfoundland, where approximately 20% of the female moose population migrates within the park from forested, coastal lowlands (< 400 m above sea level) in winter to relatively open highlands (between 400 and 800 m) during summer (mclaren et al. 2000). our interpretation is that spending summer in a highland landscape offers an advantage to this fraction of the moose population. we compared seasons of activity of the resident moose in the highlands and lowlands of gmnp, and described their finer-scale activity in terms of frequency of smaller movements and the densities in which these smaller movement clusters occur throughout, by comparing the 2 landscapes. during winter, snow limits accessibility to forage more on the highlands than in the coastal lowlands (martin 2004), a motivation for migration that is consistent with empirical evidence from other studies of moose (reviewed by ball et al. 2001). as an additional explanation for the moose migration within gmnp, as suggested in mclaren et al. (2000), summer migration from the lowlands may be a means to avoid black bear (ursus americana) predation on calves, because the highlands may offer easier escape from this predator given the longer sightlines in open habitats. this second idea would be similar to the explanation for why woodland caribou (rangifer tarandus caribou) often migrate up mountain slopes (bergerud et al. 1984), and for why elk (cervus canadensis) migrate between high and low elevations in alberta (hebblewhite and merrill 2007, hebblewhite et al. 2008). however, because moose densities are about tenfold higher in newfoundland than in other parts of their range in north america (mclaren et al. 2004), creating obvious effects on hampering regeneration in the forests of the coastal plain of gmnp (connor et al. 2000, mclaren et al. 2004, gosse et al. 2011, humber and hermanutz 2011), and because only a fraction of the population migrates, we favour limited forage availability in the lowlands as the primary factor for summer moose migration. to explore this hypothesis, which is consistent with the holt and fryxell (2011) model for migration, we compare the frequency of the fine-scale summer movement on the lowlands and the highlands, and compare resource selection functions (rsfs) for highland and lowland moose in gmnp. study area gmnp is located on the gulf of the st. lawrence on the northern peninsula of newfoundland. its lowlands, which encompass parts of the western newfoundland forest and the coastal plain sub-region of the northern peninsula forest (damman 1983), are characterized by weather influences from the gulf, producing moderate levels of annual precipitation (900–1000 mm) and cold and snowy winters (300–350 mm is in the form of snow; hare 1952). its highlands, which are situated in the long range barrens (damman 1983), are similarly influenced by the gulf, but with an orographic effect that creates a harsher climate, having annual precipitation and snowfall on average double that of the lowlands (watson 1974). the mean annual temperature on the highlands is 4.5 °c colder than that of the lowlands (banfield 1983). in 1878, one female and one male moose were introduced to newfoundland from nova scotia, and in 1904, 2 male and 2 female moose were introduced from new brunswick (pimlott 1953). moose first inhabited the northern peninsula of newfoundland by the 1940s (caines and deichmann 1989). while moose are currently found in all ecoregions of newfoundland, 114 habitat use throughout gros morne park – kerckhoff et al. alces vol. 49, 2013 their density varies considerably. in the late 1970s when gmnp was being established, moose increased first on the highlands and by the 1980s moose had increased throughout the park (connor et al. 2000). at the time of the gps collaring, surveys using stratified random blocks estimated the moose population at 7,377 ± 1,249 (4.1 ± 0.7 moose/km2; mclaren et al. 2000; gmnp, unpublished data). in 2007, population size was estimated separately for the two landscapes, at 3,975 ± 1,287 in the lowlands (4.2 ± 1.4 moose/km2) and 788 ± 223 sd in the highlands (0.9 ± 0.3 moose/km2); densities in partial surveys of the park were estimated in 2009 at 5.9 moose/km2 on the lowlands and 1.1 moose/km2 on the highlands (gmnp, unpublished data). methods habitat classification taylor and sharma (2010) classified habitat types on the lowlands and highlands of gmnp from a single-image subset of 2, 10-m multispectral spot-5 satellite images (recorded 20 june 2006) with a k-means unsupervised classification. classes were reorganized and described using information from aerial photographs and forest inventories, and local expert knowledge and field visits. ten habitat types resulted for the lowlands (table 1), and 6 for the highlands (table 2). collectively, the lowlands comprise 938 km2 or 52% of the park, of which 417 km2 or 44% is moose habitat in forest or disturbed forest types; the highlands comprise 867 km2 or 48% of the park, 641 km2 or 74% of which is moose habitat, but only a fraction of which is forest (table 3). the classifications in table 1 and 2 are the reference for our description of habitat use by moose. moose locations in june 1997, 12 adult female moose (11 with at least one calf) were immobilized and fitted with gps collars (lotek engineering, inc.; mclaren et al. 2000; table 4). the collars were set to attempt a fix at 3-h intervals. remote downloading occurred in september 1997, november 1997, and march 1998. the collars were removed in november 1998, and the remaining data records were collected at that time. location accuracy was found to be dependent on collar position in relation to topography and canopy, but 95% of all differentially corrected data from test collars had ± 25 m accuracy (moen et al. 1997, mclaren et al. 2000). all 2-dimensional fixes were removed from the dataset, and only differentially corrected locations were used in the current analysis. depending on collar functioning, locations were recorded over a 4–15.5 month period (table 4). five of the collared moose were year-round residents in the lowlands, 5 were year-round residents in the highlands, and the remaining 2 migrated seasonally between the 2 landscapes. data analysis the dataset was divided into summer and winter seasons following vander wal and rodgers (2009). six moose were used for calculation of seasonal transition dates; 3 in the lowlands and 3 in the highlands with sufficient data records to span most of a calendar year. for these moose, cumulative distance travelled was calculated in arcview version 9 (esri, redlands, california) and plotted against time beginning with 1 january. winter was defined as the period when rate of travel was less than the mean rate, estimated from the points of inflection of the best-fit logistic curves to the plots, where the estimated changes from winter to summer and from summer to winter are symmetric around the inflection points. curvefitting used the logistic regression program in the statistical package for the social sciences (spss), version 18 (also used for all subsequent analysis). the median dates alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 115 for the start and end of winter were estimated from the 3 curves for each of the 2 landscapes and used to define the seasons for all subsequent analysis. summer and winter home ranges and core-use areas were calculated using the fixed-kernel method in home range tools (rodgers et al. 2007) with gaussian (bivariate normal) distributions, reporting the 95% and 50% isopleths for ranges and cores, respectively. the bandwidth size was determined by finding the smallest proportion of the reference bandwidth that allowed one continuous outer line to encompass the table 1. habitat descriptions from a lowlands classification of gros morne national park, newfoundland, canada. habitat type description category mature softwood forest softwood dominated, especially balsam fir (abies balsamea); some mixed stands with white birch (betula papyrifera). closed-canopy forest closed spruce forest softwood dominated (balsam fir and black spruce, picea mariana); other species include tamarack (larix laricina), trembling aspen (populus tremuloides) and alder (alnus spp.); site condition can be wet. some stands of scrub forest. closed-canopy forest closed mixed forest balsam fir dominated with some mixed stands (balsam fir, white birch). stem density can be very high. younger mixed stands (∼30 years since disturbance) are included. closed-canopy forest young softwood forest softwood dominated with high content of hardwoods; canopy >50% and 6–9 m in height. closed-canopy forest open softwood forest balsam fir dominated with 25–50% open canopy; white birch can be significant; some tree regeneration (heights of 1–4 m). open-canopy forest open mixed forest softwood dominated with 25–50% open canopy. sometimes wet. trees shorter than in closed mixed forest; some tree regeneration. open-canopy forest open hardwood forest hardwood dominated with 25–50% open canopy. often originally a mixed forest where regeneration of balsam fir does not occur. open-canopy forest sparse softwood forest softwood dominated (balsam fir, black spruce) with <25% canopy; limited regeneration; ferns and grass very prominent (<50% of ground cover); forest canopy is very broken consisting of mostly remnant forest from past disturbance; low density young black spruce <6 m height; pockets of conifer regeneration <4 m height can be present. disturbed forest: sparse canopy with herb/grass ground cover herb-hardwood forest dominant plants include ferns, grass and raspberry (rubus spp.) >50% of ground cover; very sparse forest canopy; some remnant white birch with alder or elderberry (sambucus racemosa). very little balsam fir. scattered spruce <4 m height. includes forested areas that have not regenerated after severe disturbance. disturbed forest: sparse canopy with herb/grass ground cover herb forest dominant plants include ferns and grass (>50% of ground cover); exposed soil is common; large amounts of dead material (standing or fallen) and scattered remnant trees. little regeneration >30 cm height. mostly forested areas that have not regenerated after severe disturbance. disturbed forest: sparse canopy with herb/grass ground cover 116 habitat use throughout gros morne park – kerckhoff et al. alces vol. 49, 2013 polygons (worton 1989). areas of open water, wetlands, and rock barrens were excluded from each of the resulting polygons and the remaining area was divided into the habitat types appropriate to the landscape. fine-scale habitat use examined areas where a minimum of 3 consecutive gps locations < 24 h apart occurred, with distances between them of < 50 m. this definition of an important habitat patch was arbitrary, but based on an inference that foraging and other activities such as bedding take place with shorter travel distances. mean weekly travel distances, as well as distances between the habitat patches, were calculated for each moose, for summer and winter separately, and then compared across seasons using repeatedmeasures analysis of variance (anova). minimum travel distances were calculated in all cases as straight lines between successive location points. rsfs (manly et al. 2002) were modelled 6 times each using logistic regression from pooled locations of all individuals: 1) based on number of locations in each habitat type within the home range, compared to its area on the surrounding landscape, for describing summer habitat use by residents and migrants using the highlands in a marginal model; 2) in a similar marginal model for describing winter habitat use by residents and migrants using the lowlands; 3) in a conditional model based on number of locations for each moose in each habitat type within its table 2. habitat descriptions from a highlands classification of gros morne national park, newfoundland, canada. habitat type description category open softwood forest balsam fir and some black spruce in a closed canopy ranging to <75% open; dense pockets of krummholz (locally known as tuckamore). open heath and fen and bog interspersed. closedto opencanopy forest scrub forest trees <4 m height. open heaths, fens, and bogs throughout (>50% of area). open-canopy forest shrub predominantly low shrubs (<1 m height), interspersed with fens, bogs, and small pockets of scrub forest. associated with transition from fen and tundra to scrub forest. can be wet. non-forest tundra low heath vegetation comprised of sedges (carex spp.), caribou moss (cladonia spp.) and crowberries (empetrum spp.); <20% rock, but few shrubs or trees. fairly dry. non-forest fen sedge meadows with fens throughout. non-forest rock barren boulder fields and exposed rock. very little vegetation. non-forest table 3. habitat availability in gros morne national park, newfoundland, canada. habitat type availability on landscape area (km2) percent lowlands mature softwood forest 69.6 7 closed spruce forest 27.6 3 closed mixed forest 65.5 7 young softwood forest 56.7 6 open softwood forest 62.5 7 open mixed forest 43.6 5 open hardwood forest 39.0 4 sparse softwood forest 19.3 2 herb-hardwood forest 20.0 2 herb forest 13.5 1 highlands open softwood forest 184.7 21 scrub forest 133.5 15 shrub 130.0 15 tundra 130.3 15 fen 62.6 7 alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 117 table 4. first and last dates of collaring, record length, and home range area in summer and winter from fixed-kernel estimates using a 95% isopleth for 12 gpscollared moose. this table also shows median seasonal transition dates, and lengths of summer and winter for moose using the two landscapes year-round in gros morne national park, newfoundland, canada. migrating moose are identified by an asterisk. id landscape first day collared last day recording record length (days) home range size (km2) median seasonal transition dates median season length (days) summer winter winter to summer summer to winter summer winter 15 lowlands 25-jun-97 13-oct-98 468 11.9 12.1 18-apr-98 11-oct-98 173 181 16 lowlands 25-jun-97 13-oct-98 468 13.2 13.1 18-apr-98 11-oct-98 173 181 19 lowlands 25-jun-97 05-nov-97 130 2.9 1.8 18-apr-98 11-oct-98 173 181 21* lowlands 25-jun-97 16-jan-98 201 — 5.5 18-apr-98 11-oct-98 173 181 22* lowlands 26-jun-97 18-jun-98 352 — 12.3 18-apr-98 11-oct-98 173 181 25 lowlands 26-jun-97 21-jun-98 355 8.3 12.1 18-apr-98 11-oct-98 173 181 26 lowlands 26-jun-97 15-nov-97 139 4.2 2.4 18-apr-98 11-oct-98 173 181 17 highlands 25-jun-97 13-oct-98 468 11.3 10.8 30-apr-98 24-oct-98 174 180 18 highlands 25-jun-97 27-feb-98 242 6.8 8.7 30-apr-98 24-oct-98 174 180 20 highlands 25-jun-97 17-mar-98 262 6.6 8.2 30-apr-98 24-oct-98 174 180 21* highlands 25-jun-97 16-jan-98 201 5.7 — 30-apr-98 24-oct-98 174 180 22* highlands 26-jun-97 18-jun-98 352 8.0 — 30-apr-98 24-oct-98 174 180 23 highlands 26-jun-97 13-oct-98 467 9.2 7.2 30-apr-98 24-oct-98 174 180 24 highlands 26-jun-97 01-jun-98 335 7.0 4.6 30-apr-98 24-oct-98 174 180 1 1 8 h a b it a t u s e t h r o u g h o u t g r o s m o r n e pa r k – k e r c k h o f f e t a l . a l c e s v o l . 4 9 , 2 0 1 3 home range, compared to its area on the surrounding landscape, for describing summer and winter habitat use by residents of the lowlands; 4) in a similar conditional model for describing summer and winter habitat use by residents of the highlands; 5) in a conditional model based on number of locations for each moose in each habitat type within its core-use area, compared to its area in the home range, for describing finer-scale summer and winter habitat use by residents of the lowlands; and 6) in a similar conditional model for describing summer and winter habitat use within the core-use areas of residents of the highlands. in the first 2 (marginal) models, one moose resident on the highlands was removed because of too few locations (id 18, table 4). habitat use by residents in the 2 landscapes, habitat use by the 2 migrant moose, and differences in habitat use between the winter and summer seasons were statistically compared in a mixed-effects model with random intercepts and coefficients (gillies et al. 2006). to determine the most parsimonious regression models, corrected akaike’s information criteria (aicc) and model deviance were compared to a model with random variables for each individual moose. a compound symmetric structure was assumed, meaning that covariance among all responses of an animal was assumed constant (skrondal and rabe-hesketh 2004) and habitat availability was also assumed constant over time (manly et al. 2002). these assumptions limit the applicability of the rsfs to the time period studied. random intercepts and coefficients for all habitat types experiencing some use were estimated, and coefficients significantly > 1 were defined as selection of a habitat type. calculations were all relative to open softwood forest as a reference habitat type, which was defined similarly for both landscapes. results home-range size varied considerably among individual moose, and there was no consistent size difference by landscape for either winter (f1,11 = 0.57, p = 0.58) or summer (f1,11 = 1.53, p = 0.06; table 4). there was also no difference in winter and summer home-range sizes on either the lowlands (f1,11 = 0.26, p = 0.88) or the highlands (f1,11 = 0.33, p = 0.67). the mean distances travelled during a one-year period were 309 km for residents on the lowlands and 267 km for residents on the highlands. moose travelled less in winter than in summer. weekly travel distances varied according to season (f1,11 = 106.35, p < 0.001; fig. 1). there was no difference in weekly travel distances by landscape (f1,11 = 0.75, p = 0.47). the summer season differed in length between the 2 landscapes, but only by a day (table 4). summer, defined by moose travel rates, started and ended close to 2 weeks earlier on the lowlands. the best-fit marginal rsf models describing habitat use by residents and migrants showed consistent selection of fig. 1. weekly distance travelled (km) for moose in gros morne national park, newfoundland, canada. alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 119 habitats in summer when they occupied the highlands together, and variable selection of habitats in winter when they occupied the lowlands together (table 5); however, for both seasons, models with differences in habitat use between residents and migrants were rejected. for the remaining 4 rsfs, conditional models with residents and migrants pooled, the best-fit were those including seasonal differences. in summer, habitat types used less than expected based on availability at both the home-range and core-use scales were herb-hardwood forest and herb forest (table 6). closed spruce forest, closed mixed forest, and young softwood forest were among the top habitat types selected in summer relative to open softwood forest, but were not selected more than expected. the pattern was generally reversed in winter on the lowlands; herbhardwood forest and herb forest, along with open hardwood forest and sparse softwood forest, were all selected by resident moose at both the home-range and core-use scales. at the home-range scale, young softwood forest and both closed forest types were also selected in winter. in the rsfs calculated for moose resident on the highlands, the fen, tundra, and shrub habitat types were selected in summer at the home-range scale, while only the fen and tundra types were selected at the core-use scale. the pattern was similar in winter, but scrub forest was also selected at the home-range scale and shrub, not tundra, was selected at the core-use scale. the overall trend in summer was more use of closed-canopy forest types on the lowlands and more use of non-forest habitat types on the highlands. selection of disturbed forest is a winter phenomenon on the lowlands of gmnp; the same category of habitat types is avoided in summer. defined by repeated occupation of an area with travel distances < 50 m apart, most fine-scale habitat patches on the lowlands were categorized as disturbed forest (50/127) or as open-canopy forest (47/127). there were an additional 22 fine-scale habitat patches identified in young softwood forest, while only 8 of the 127 fine-scale habitat selections on the lowlands were in closedcanopy forest. there were 13.5 habitat patches per 100 km2 on the lowlands, but more on the highlands (18.3/100 km2) where the majority were in open softwood forest (70/159). straight-line distances travelled between fine-scale habitat patches were table 5. ranking of habitat types, from most to least selected, for highland residents (n = 5; n = 3,252) and migrants (n = 2; n = 2,919) during summer on the highlands, and lowland residents (n = 5; n = 2,013) and migrants (n = 2; n = 1,018) during winter on the lowlands, where lower-case n refers to total number of locations in home ranges used to calculate resource selection functions (rsfs); gros morne national park, newfoundland, canada. habitat types significantly selected (p < 0.05) by at least 4 of 5 residents or both of the migrants are shown in boldface. the open softwood forest is a reference habitat (shown in italics). residents migrants summer on the highlands (1) fen fen (2) tundra tundra (3) shrub shrub (4) open softwood forest open softwood forest (5) scrub forest scrub forest winter on the lowlands (1) closed spruce forest herb-hardwood forest (2) herb forest closed spruce forest (3) closed mixed forest herb forest (4) mature softwood forest mature softwood forest (5) young softwood forest closed mixed forest (6) herb-hardwood forest sparse softwood forest (7) sparse softwood forest open hardwood forest (8) open hardwood forest young softwood forest (9) open mixed forest open mixed forest (10) open softwood forest open softwood forest 120 habitat use throughout gros morne park – kerckhoff et al. alces vol. 49, 2013 greater in summer than in winter (f1,120 = 36.28, p = 0.01), a consistent pattern for moose in both landscapes (f1,120 = 0.08, p = 0.93); there was no difference in travel distances between patches by landscape (f1,120 = 0.01, p = 0.99). discussion despite variation among individual moose in habitat selection in the park’s more diverse and forested lowlands, as reported earlier (mclaren et al. 2009), we are able to show with rsfs that more use of closed-canopy forest occurs in summer, likely as a means of heat avoidance. conversely, selection of disturbed and open-canopy forest is a winter phenomenon on the lowlands; the same category of habitat types is avoided during summer. in the cooler highlands, selection of non-forest habitat types may reflect less need to escape heat in the summer relative to the lowlands, and perhaps a means to escape insects. in winter, where moose populations are locally at higher densities according to both aerial surveys (gmnp, unpublished data) and the frequency of our identified winter habitat patches, selecting disturbed and opencanopy forest on the lowlands may be matched to optimal foraging, while selecting fen and shrub on the highlands may be matched to travel through areas where snow is packed along trails that reduces the energy cost of locomotion (telfer and kelsall 1979). table 6. ranking of habitat types, from most to least selected, within home ranges and core-use areas for resident moose in gros morne national park, newfoundland, canada: 5 moose in lowlands and 4 moose in highlands; lower-case n refers to total number of locations in home ranges or in core-use areas used to calculate rsfs. habitat types selected or avoided (p < 0.05) in proportion to their available area are shown in boldface. the open softwood forest is a reference habitat (shown in italics). summer habitat ranking winter habitat ranking lowlands home range (n = 3,765) core-use area (n = 1,485) home range (n = 2,013) core-use area (n = 1,679) (1) closed spruce forest closed mixed forest (1) herb forest herb-hardwood forest (2) young softwood forest closed spruce forest (2) herb-hardwood forest herb forest (3) closed mixed forest young softwood forest (3) young softwood forest sparse softwood forest (4) open softwood forest mature softwood forest (4) open hardwood forest open hardwood forest (5) mature softwood forest open softwood forest (5) closed mixed forest open mixed forest (6) sparse softwood forest open hardwood forest (6) closed spruce forest closed spruce forest (7) open hardwood forest open mixed forest (7) sparse softwood forest young softwood forest (8) open mixed forest sparse softwood forest (8) open mixed forest closed mixed forest (9) herb-hardwood forest herb-hardwood forest (9) mature softwood forest open softwood forest (10) herb forest herb forest (10) open softwood forest mature softwood forest highlands home range (n = 2,954) core-use area (n = 1,609) home range (n = 1,914) core-use area (n = 1,619) (1) fen fen (1) fen fen (2) tundra tundra (2) shrub shrub (3) shrub shrub (3) tundra tundra (4) open softwood forest open softwood forest (4) scrub forest scrub forest (5) scrub forest scrub forest (5) open softwood forest open softwood forest alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 121 in areas where snow is deep, reducing the energy cost of travel may be more important than avoiding competition for food. the most straightforward way of describing habitat use is in terms of density (holt and fryxell 2011). to approximate local moose densities, the total area in habitat types selected by moose could be substituted for an average density over the entire landscape areas. if habitat types selected during winter, based on the core-use areas of gps-collared moose occupying the lowlands in this season, are used to represent the best habitat types (herb-hardwood, herb, sparse softwood, and open hardwood forests), winter density would be 20.6 moose/km2, almost 5 x larger than the density estimate across the lowlands in the march 2007 survey (4.2 moose/km2). if moose remaining on the highlands in winter similarly used only those habitat types selected in core-use areas by the gps-collared subset (fen and shrub), their density would be 1.7 moose/km2, about twice the landscape density estimate (0.9 moose/km2) for the highlands. the lowland winter habitat types are essentially abandoned during summer in favour of habitat types providing thermal cover (closedcanopy forest); this change, combined with 20% of the population migrating to the highlands (mclaren et al. 2000), reduces effective summer density on the lowlands. meanwhile, summer migrants, according to the 2007 winter lowland population estimate, should double the corresponding winter estimate for the highlands, where tundra, roughly equal in area to shrub, is simply substituted as a preferred habitat in summer. presumably, seasonal abundance of forage is one benefit to spending the summer on the highlands. thus, 2 landscapes in gmnp provide insight into habitat selection by moose in newfoundland and labrador. we find that moose adapt seasonally to the park’s lowlands and highlands. moose adopting either of 2 strategies, year-round residence in one landscape or migration between landscapes, do not appear to select habitat differently when they occupy the same landscape. this point parallels the consensus for migration in a review and study in sweden (ball et al. 2001) that concluded that snow depth is the likely driver for moose migration. what differs in our study is insight into the advantages in summer for the fraction of moose opting to return to an otherwise less hospitable landscape, that being the snowy highlands. if we accept a conclusion from a québec study that movement rates for moose are better indicators of forage availability than home range size (dussault et al. 2005), and that the habitat patches in open softwood forest on the highlands offer more forage in summer than that provided on average in the disturbed or open-canopy forest on the lowlands, we are describing a situation similar to what has been described for predatorfree svaldbard reindeer (rangifer tarandus platyrhynchus) (bremset hansen et al. 2009). in this case, populations in overgrazed range move to areas of higher forage biomass, not higher forage quality. further, although plant phenology from spring through early summer is generally associated with increasing forage quality (klein 1990), the nitrogen content in forage declines initially after snowmelt (van der wal et al. 2000). thus, migrant moose may travel upland in gmnp to maximize biomass consumption while tracking delayed plant phenology in the cooler highlands climate. it is recommended that moose management in gmnp consider 2 landscapes (the lowlands and the highlands) as separate management units due to differences both in habitat types they offer and densities of moose they support. park management plans should ensure landscape connectivity for moose migrating between the highlands and lowlands. on this note, management across newfoundland and labrador that is both 122 habitat use throughout gros morne park – kerckhoff et al. alces vol. 49, 2013 effective and adaptable need not be dependent on defining discrete populations of moose, but should be in the context of the 2 very different landscapes the province offers to moose. acknowledgements dr. a. rodgers, centre for northern forest ecosystem research, ontario ministry of natural resources, and dr. d. eastman, university of victoria, provided advice and constructive criticism on portions of this paper submitted by its lead author as a msc thesis in forestry at lakehead university. two anonymous reviewers and alces editor dr. e. addison provided helpful comments to improve this manuscript. s. taylor, gmnp, provided invaluable assistance in interpreting habitat types and in analyzing gps data. this project was funded and coordinated by the institute for biodiversity, ecosystem science and sustainability of the government of newfoundland and labrador, with additional funding and in-kind support provided by parks canada and lakehead university. the original fieldwork and preliminary data analysis occurred under the guidance of the inland fish and wildlife division of the government of newfoundland and labrador, with assistance by d. anions and c. mccarthy (formerly of gmnp) and technical advice from dr. a. rodgers. references ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7: 39–47. banfield, c. e. 1983. climate. pages 37– 106 in g. r. south, editor. biogeography and ecology of the island of newfoundland. w. junk publishers, boston, massachusetts, usa. bergerud, a. t., h. e. butler, and d. r. miller. 1984. anti-predator tactics of calving caribou: dispersion in mountains. canadian journal of zoology 62: 1566– 1575. bremset hansen, b., i. herfindal, r. aanes, b.-e. sæther, and s. henriksen. 2009. functional response in habitat selection and the trade-offs between foraging niche components in a large herbivore. oikos 118: 859–872. caines, p., and h. deichmann. 1989. resource description and analysis: gros morne national park. gros morne national park, parks canada, rocky harbour, newfoundland, canada. connor, k. j., w. b. ballard, t. dilworth, s. mahoney, and d. anions. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 36: 111–132. damman, a. w. h. 1983. an ecological subdivision of the island of newfoundland. pages 163–205 in g. r. south, editor. biogeography and ecology of the island of newfoundland. w. junk publishers, boston, massachusetts, usa. dussault, c., r. courtois, j.-p. ouellet, and i. girard. 2005. space use of moose in relation to food availability. canadian journal of zoology 83: 1431–1437. fretwell, s. d., and h. l. lucas. 1970. on territorial behaviour and other factors influencing habitat distribution in birds. acta biotheoretica 19: 16–36. fryxell, j. m. 1991. forage quality and aggregation by large herbivores. american naturalist 138: 478–498. gillies, c. s., m. hebblewhite, s. e. nielsen, m. a. krawchuk, c. l. aldridge, j. l. frair, d. j. saher, c. e. stevens, and c. l. jerde. 2006. application of random effects to the study of resource selection by animals. journal of animal ecology 75: 887–898. gosse, j., l hermanutz, b. mclaren, p. deering, and t. knight. 2011. degradation of boreal forests by non-native herbivores in newfoundland’s national parks: recommendations for ecosystem alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 123 restoration. natural areas journal 31: 331–339. hare, f. k. 1952. the climate of the island of newfoundland: a geographical analysis. geographic bulletin 2: 36–88. hebblewhite, m., and e. h. merrill. 2007. multiscale wolf predation risk for elk: does migration reduce risk? oecologia 152: 377–387. ———, ———, and g. mcdermid. 2008. a multi-scale test of the forage maturation hypothesis in a partially migratory ungulate population. ecological monographs 78: 141–166. holt, r., and j. fryxell. 2011. theoretical reflections on the evolution of migration. pages 17–31 in e. milner-gulland, j. fryxell, and a. sinclair, editors. animal migration: a synthesis. oxford university press, oxford, united kingdom. humber, j. m., and l. hermanutz. 2011. impacts of non-native plant and animal invaders on gap regeneration in a protected boreal forest. biological invasions 13: 2361–2377. klein, d. r. 1990. variation in quality of caribou and reindeer forage plants associated with season, plant part, and phenology. rangifer special issue 3: 123–130. manly, b. f. j., l. l. mcdonald, d. l. thomas, t. l. mcdonald, and w. p. erickson. 2002. resource selection by animals, second edition. kluwer academic publishers, dordrecht, netherlands. martin, c. 2004. climatology and historical snowcover of the big level plateau, gros morne national park, newfoundland. msc thesis, memorial university of newfoundland, st. john’s, newfoundland, canada. mclaren, b., c. mccarthy, and s. mahoney. 2000. extreme moose migrations in gros morne national park, newfoundland. alces 36: 217–232. ———, b. a. roberts, n. djan-chékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 45–59. ———, s. taylor, and s. h. luke. 2009. how moose select forested habitat in gros morne national park, newfoundland. alces 45: 125–135. mcnaughton, s. j. 1985. ecology of a grazing ecosystem: the serengeti. ecological monographs 55: 259–294. moen, r., j. pastor, and y. cohen. 1997. accuracy of gps telemetry collar locations with differential correction. journal of wildlife management 61: 530–539. pimlott, d. h. 1953. newfoundland moose. transactions of the north american wildlife conference 18: 563–581. rodgers, a. r., a. p. carr, l. b. hawthorne, l. smith, and j. g. kiem. 2007. home range tools for arcgis. version 1.1. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. skrondal, a., and s. rabe-hesketh. 2004. generalized latent variable modeling: multilevel, longitudinal, and structural equation models. chapman and hall, new york, new york, usa. taylor, s., and r. sharma. 2010. assessing the status of forest recovery in disturbance affected areas using spot 5 imagery in gros morne national park. gros morne national park, parks canada, rocky harbour, newfoundland, canada. telfer, e. s., and j. p. kelsall. 1979. studies of morphological parameters affecting ungulate locomotion in snow. canadian journal of zoology 57: 2153–2159. vander wal, e., and a. r. rodgers. 2009. designating seasonality using rate of movement. journal of wildlife management 73: 1189–1196. van der wal, r., n. madan, s. v. lieshout, c. dormann, r. langvatn, and s. d. albon. 2000. trading forage quality for quantity? plant phenology and patch 124 habitat use throughout gros morne park – kerckhoff et al. alces vol. 49, 2013 choice by svalbard reindeer. oecologia 123: 108–115. watson, w. b. 1974. the climate of gros morne national park, newfoundland. atmospheric environmental service, department of environment, st. john’s, newfoundland, canada. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home range studies. ecology 70: 164–168. alces vol. 49, 2013 kerckhoff et al. – habitat use throughout gros morne park 125 moose habitat use throughout gros morne national park study area methods habitat classification moose locations data analysis results discussion acknowledgements references diversity and abundance of terrestrial gastropods in voyageurs national park, mn: implications for the risk of moose becoming infected with parelaphostrongylus tenuis tim cyr1, steve k. windels2, ron moen3, and jerry w. warmbold2,4 1integrated biosciences graduate program and natural resources research institute, university of minnesota duluth, duluth, minnesota 55811; 2voyageurs national park, 360 highway 11 east, international falls mn, 56649; 3natural resources research institute, 5013 miller trunk highway, duluth, minnesota 55811; 4present address: university of south dakota, 414 east clark st, vermillion, south dakota 57069. abstract: voyageurs national park (vnp) has a stable population of about 40–50 moose (alces alces). recent declines in moose abundance in adjacent areas in northern minnesota raise concerns about the long-term viability of moose in vnp. the parasitic nematode parelaphostrongylus tenuis has been documented in moose in vnp and has been implicated in moose declines in other populations. terrestrial gastropods are the intermediate hosts for p. tenuis, and describing spatial and temporal differences in their abundance should increase understanding about the risk of p. tenuis infection for vnp moose at the individual and population levels. we used cardboard sheets to estimate species composition and abundance of terrestrial gastropods in representative vegetation communities in vnp. we collected a total of 6,595 gastropods representing 25 species, 22 terrestrial snails and 3 slugs; 8 are known vectors of p. tenuis, including the slug deroceras laeve, the most common species found. gastropods were more abundant in september than july, and in upland forests (maximum = 555 gastropods/m2) more than in wetter lowlands (20 gastropods/m2). we used location data from gpscollared moose in vnp to estimate the relative exposure of moose to gastropods that could be infected with p. tenuis larvae. the boreal hardwood forest and northern spruce-fir forest ecotypes had the highest use by moose and high abundance of p. tenuis vectors in summer, and may pose the greatest risk for infection. habitat use and the related risk of ingesting gastropod vectors varied by individual moose. our method can be extended in moose range to estimate the relative risk of p. tenuis infection. alces vol. 50: 121–132 (2014) key words: alces, meningeal worm, minnesota, moose, p. tenuis, parasite introduction the parasitic nematode parelaphostrongylus tenuis can be fatal to moose (alces alces) (anderson 1964), and was the probable cause of 5% of mortality of radio-collared moose in northwestern minnesota and >20% of incidentally-recovered moose in northern minnesota (murray et al. 2006, wünschmann et al. 2014). the infection causes weakness in the hindquarters, circling, tilting of the head, and increased fearlessness of humans (anderson and prestwood 1981). infections can be lethal and cause mortality indirectly through increased risk of predation or accidents (lankester et al. 2007, butler et al. 2009, wünschmann et al. 2014). voyageurs national park (vnp) in northern minnesota maintains a stable, low-density population of about 40–50 moose, and p. tenuis infection and associated mortality has been documented in and surrounding vnp (windels 2014). though the effect on moose at the population level in vnp is unknown, previous studies suggest it is unlikely to be a major 121 mortality source at the currently low whitetailed deer (odocoileus virginianus) density (3–6 deer/km2) (whitlaw and lankester 1994b). the normal lifecycle of p. tenuis includes white-tailed deer as the definitive host and terrestrial gastropods as intermediate hosts (lankester and anderson 1968). white-tailed deer ingest infected gastropods while foraging and gastropods become infected with p. tenuis by crawling over or near infected deer feces (lankester 2001). however, only 0.1–4.2% of gastropods collected in minnesota and ontario were infected with p. tenuis larvae (lankester and anderson 1968, lankester and peterson 1996). at those infection rates, a white-tailed deer would need to consume up to 1000 gastropods to encounter a single larva (lenarz 2009). however, lankester and peterson (1996) reasoned that even at such low rates of infection in gastropods, the high rates of infection measured in white-tailed deer in the region (≤91%, slomke et al. 1995) is explained by the large volume of vegetation eaten on and near the ground over a few months in the autumn. infection rates in white-tailed deer derived from winter fecal samples have ranged from 67–90% from the 1970s to the present in vnp (gogan et al. 1997, vanderwaal et al. 2014). white-tailed deer are the definitive host of p. tenuis but moose, an aberrant host, also ingest infected gastropods during foraging and become infected. initial signs of p. tenuis infection can appear in moose as early as 20 days after experimental infection (lankester 2002). gastropods are necessary for p. tenuis to complete its life cycle. therefore, knowledge of gastropod populations in vnp may help managers better understand the role of p. tenuis in local moose population dynamics. the distribution and habitat preferences of terrestrial gastropods in vnp have not been studied previously. extrapolation from studies of gastropod communities in different regions of minnesota and the surrounding areas is possible (e.g., from northwestern minnesota [nekola et al. 1999] or rock outcrops in northeastern minnesota [nekola 2002]). gastropods exhibit habitat preferences that result in variation in presence or density across vegetation communities or other habitat features, and few studies have examined their abundance and diversity at fine spatial scales (moss and hermanutz 2010). the risk of p. tenuis infection is presumably influenced by vector density and could vary within a population because individual moose demonstrate differential habitat use (gillingham and parker 2008). fine-scale habitat use derived from gps collars can help clarify the risk of p. tenuis infection to individuals and populations of moose. combined, individual differences in habitat use and variability among habitat types in gastropod diversity and abundance may result in differential risk of moose and other cervids to p. tenuis infection (vanderwaal et al. 2014). in this study we surveyed terrestrial gastropod species on the kabetogama peninsula in vnp. our objectives were to 1) estimate the abundance and diversity of terrestrial gastropods in different ecotypes, with particular focus on known vectors of p. tenuis, 2) document changes in gastropod abundance over the growing season, and 3) compare the use of cover types by gps-collared moose to density of p. tenuis vectors to estimate the encounter risk of individual moose. study area voyageurs national park (48.50° n, 92.88° w) is an 882 km2 protected area comprised of a mixture of forested land (61%) and large lakes (39%) along the u.s.-canada border. moose are primarily restricted to the kabetogama peninsula (windels 2014), a 300 km2 roadless area in the center of vnp, and have remained relatively stable 122 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 since the 1990s with density ranging from 0.14–0.19 moose/km2 (windels 2014). white-tailed deer density in winter during the study ranged between 3–6/km2 (gogan et al. 1997, unpublished data of vnp). vegetation is a mix of southern boreal and laurentian mixed conifer-hardwood forests comprised primarily of a mosaic of quaking aspen (populus tremuloides), paper birch (betula papyrifera), balsam fir (abies balsamea), white spruce (picea alba), white pine (pinus strobus), red pine (p. resinosa), jack pine (p. banksiana), and black spruce (picea mariana) (faber-landgendoen et al. 2007). soils range from thin, sandy loams over bedrock to poorly draining clays at lower elevations (kurmis et al. 1986). beaver-created wetlands and associated seral stages are abundant (johnston and naiman 1990). temperatures vary from −40 to 36 °c, with an average annual temperature of 1.4 °c. mean annual precipitation is 62 cm, with most precipitation falling between may and september (kallemeyn et al. 2003). methods we used the “ecotype”-level vegetation classification derived from the usgs-nps vegetation map (hop et al. 2001) to select the 10 most common terrestrial ecotypes on the kabetogama peninsula to sample for gastropods. we excluded 4 of these because they were too wet to sample with our methodology: poor conifer swamps, rich hardwood swamps, wet meadows, and shrub bogs. the remaining 6 ecotypes comprised 80% of the non-aquatic vegetation communities (table 1); 4 were dry uplands (rock barrens with trees, northern spruce-fir forests, boreal hardwood forests, and northern pine forests) and 2 wet lowland ecotypes (northern shrub swamp and rich conifer swamp). we randomly selected 5 patches (polygons) within each of the 6 ecotypes within a restricted area to facilitate access to sampling sites (fig. 1) assuming that these sites were representative of those across the entire peninsula. at each site we sampled during a single over-night period at approximately 1-month intervals in each of 4 periods: 6– 20 june, 29 july–3 august, 18–25 august, and 9–14 september. we used 0.25 m2 cardboard sampling squares (50 � 50 cm) placed on ground vegetation to collect gastropods (lankester and peterson 1996, hawkins et al. 1998, nankervis et al. 2000, maskey 2008). we randomly selected a starting sample point and direction within each polygon such that a 100-m sampling transect would fit entirely within the polygon. we placed 10 corrugated cardboard squares on the 100-m transect and verified that all were in the same ecotype. the cardboard was placed directly on the soil or duff layer after rocks and branches were cleared from the sampling site. the cardboard was saturated with water and covered with a 0.36 m2 sheet of 3-mm thick clear plastic. sheets were set in the morning and retrieved ∼24 h later. the wetness of each sheet was estimated as the percentage of the bottom that was visibly damp. all slugs table 1. area (km2) and % total area covered by each of 6 terrestrial vegetation ecotypes sampled on the kabetogama peninsula, voyageurs national park (vnp), minnesota, usa, juneseptember 2011. area calculations exclude lakes and ponds. ecotype classifications are according to the us-national vegetation classification system applied to vnp (hop et al. 2001). ecotype area (km2) % northern spruce-fir forest 66 23 boreal hardwood forest 62 21 northern pine forest 52 18 treed rock barrens 39 13 northern shrub swamp 8 3 rich conifer swamp 5 2 total 232 80 alces vol. 50, 2014 cyr et al. – gastropod vectors of p. tenuis in vnp 123 and snails on the underside of the cardboard were collected and stored in plastic jars with damp paper towels. subsequent identification was to the lowest taxonomic level possible using available keys (burch 1962, nekola 2007, j. nekola, minnesota department of natural resources, pers. comm.). in 3 cases, we lumped 2 closely related species together that could not be reliably differentiated by morphological characteristics: zonitoides nitidus and z. arboreus, nesovitrea electrina and n. binneyana, and euconulus alderi and e. fulvous. we identified potential gastropod vectors of p. tenuis based on a literature review (lankester and anderson 1968, gleich et al. 1977, upshall et al. 1986, rowley et al. 1987, platt 1989, lankester and peterson 1996, whitlaw et al. 1996, nankervis et al. 2000, lankester 2001). we considered the 100-m sample transect the sample unit and tested for the effects of ecotype and sample period on abundance of gastropod groups (total gastropods, snails only, slugs only) using factorial anova. we also tested for an interaction between ecotype and sampling period. we used bonferroni corrections when making post-hoc comparisons between main effects (ecotype and sample period) and set statistical significance at p = 0.05. we obtained gps locations at 15-min intervals from 11 adult moose (9f:2m) wearing gps collars to measure habitat use during june-september 2010. spatial data were displayed using arcgis 10.1 with arcgis spatial analyst (esri, redlands, ca, usa 2012), and home ranges were calculated in the geospatial modeling environment fig. 1. primary moose range (dashed line) and terrestrial gastropod sampling area (crosshatched area) in voyageurs national park, minnesota, usa, 6 july-14 september 2010. 124 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 (2012 spatial ecology llc) running via arcgis 10.1 and r 3.0.1. we calculated the proportion of locations in each ecotype for individual moose. we calculated a relative measure of p. tenuis transmission risk to moose in different ecotypes by comparing the abundance of gastropods in each ecotype to habitat use in each ecotype. mean monthly habitat use (i.e., proportion of all locations within an ecotype) varied little from june to september; all differences were <5% between months for any ecotype. we therefore used the mean proportion of use for the entire june-september period to estimate an overall risk of p. tenuis infection by ecotype during summer. we also evaluated variation in relative risk of p. tenuis infection to individual moose. risk value was calculated by multiplying the proportion of each ecotype used by a moose by the mean density of potential p. tenuis gastropod vectors measured in each ecotype. we scaled the risk value for each moose to the highest individual risk value to compare relative risk of infection among individual moose. our indices of risk assume that 1) gastropod infection rates (i.e., proportion of gastropods infected with p. tenuis larvae) did not vary among gastropod species, among habitat types, or over the sampling time, and 2) the likelihood of a moose ingesting a potentially infected vector gastropod in a given ecotype is proportional to the density of known vectors of p. tenuis in that ecotype. our index of risk does not consider morbidity or mortality for infected moose, because the severity and duration of the infection can be highly variable (lankester 2002, 2010). results we collected 6,595 gastropods representing 9 families and 25 species (3 slug species and 22 snail species; table 2), and successfully classified 62% of slugs and 50% of snails. we could not identify 3,116 (47%) of the gastropods because they were damaged beyond recognition during collection and storage, or were juveniles that can be difficult to identify accurately even to the family level (j. nekola, pers. comm.). the total number of snails/m2 (including unidentified) increased from july to september in all ecotypes combined (anova, f3,29 = 8.7, p < 0.001). the treed rock barren cover type had the lowest snail density (7.1/m2) for all sampling periods combined. the northern pine forest and northern spruce-fir forest ecotypes had the most snails for all periods combined, increasing from 7.3 and 10.2 snails/m2 in july to 23.7 and 22.8 snails/m2 in september, respectively (fig. 2). overall, slug density was relatively constant over time within each ecotype, and at lower density than snails. slug density (including unidentified) was more variable over time than snail density (fig. 3). slug density in all 4 sampling periods combined was lowest (1.3/m2) in the rich conifer swamp ecotype and highest in the northern pine (6.2/m2) and northern spruce-fir forests (6.9/m2). northern shrub swamp (3.3/m2) and rich conifer swamp (1.3/m2) had lower slug densities than the other 4 ecotypes (anova, f5,29 = 20.88, p < 0.001). cardboard wetness increased as the survey progressed (anova, f3,29 = 165.8, p < 0.001); for example, mean wetness was 47% in survey 1 and 90% in survey 4. within ecotypes, cardboard wetness in the treed rock barren ecotype was lower (51%) than in the other 5 ecotypes (range = 75– 82%; anova, f5,29 = 44.3, p < 0.001). eight of the collected species are known vectors of p. tenuis and comprised 32% of the sample. the slug deroceras laeve was the most common vector collected (26% of total captures), was present in every ecotype, and most common in the northern spruce-fir forest ecotype. two other slug vectors were pallifera hemphili and a deroceras specimen that we could not identify to species, but alces vol. 50, 2014 cyr et al. – gastropod vectors of p. tenuis in vnp 125 assumed was a p. tenuis vector like its congener d. leave. the snails discus cronkhitei, zonitoides nitidus+arboreas, strobilops spp., and cochlicopa sp., known vectors of p. tenuis, were ∼11% of the sample and found across all surveys and sample sites (table 2). risk of p. tenuis infection was highest in northern spruce-fir forests (fig. 4). the northern spruce-fir ecotype had the highest use by moose (35% of total locations) and also had the second highest estimated density of p. tenuis vectors. treed rock barrens had the fourth highest use by moose (8%) table 2. composition of terrestrial gastropods collected in voyageurs national park, minnesota, usa, june-september 2011. gastropod species were identified to the lowest taxonomic level possible; 62% of slugs and 50% of snails were classified. group family species count % total captures p. tenuis vectors slug limacidae deroceras laeve 906 26.0 slug limacidae deroceras sp. (but not d. leave) 13 0.4 slug philomycidae pallifera hemphili 6 0.2 snail endodontidae discus cronkhitei 55 2.0 snail strobilopsidae strobilops spp. 145 4.0 snail valloniidae cochlicopa sp. 6 0.2 snail zonitidae zonitoides (nitidus+arboreas) 159 4.6 total 1290 37.4 non-vectors snail endodontidae helicodiscus parallelus 7 0.2 snail endodontidae punctum californicum 7 0.2 snail endodontidae punctum minutissimum 2 <0.1 snail endodontidae punctum spp. 4 0.1 snail oxychilidae nesovitrea (electrina+binneyana) 105 3.0 snail pupillidae columella simplex 6 0.2 snail pupillidae gastrocopta pentodon 6 0.2 snail pupillidae gastrocopta sp. 11 0.3 snail pupillidae vertigo spp. 319 9.0 snail pupillidae unknown 143 4.0 snail succineidae oxyloma retusa 19 0.5 snail valloniidae cochlicopa lubricella 11 0.3 snail valloniidae zoogenetes harpa 62 2.0 snail zonitidae euconulus (alderi + fulvous) 638 18.0 snail zonitidae guppya sterkii 6 0.2 snail zonitidae striatura milium 29 0.8 snail zonitidae striatura exigua 7 0.2 snail zonitidae striatura ferrea 6 0.2 snail zonitidae vitrina limpida 326 9.0 snail zonitidae unknown 461 13.0 total 2175 61.4 126 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 and the third highest p. tenuis vector density, suggesting moderate risk. boreal hardwood forests were also a moderate risk ecotype based on their relatively high use and low vector density. rich conifer swamps and northern shrub swamps were low risk fig. 2. mean (+se) number of snails/m2 (including unidentified) measured in each of 6 ecotypes for a single over-night period in each of 4 sampling periods in juneseptember, 2011 in voyageurs national park, minnesota, usa. sample periods were: survey 1 = 6–20 june, survey 2 = 29 july – 3 august, survey 3 = 18–25 august, survey 4 = 9–14 september. fig. 3. mean (+se) number of slugs/m2 (including unidentified) measured in each of 6 ecotypes for a single over-night period in each of 4 sampling periods in juneseptember, 2011 in voyageurs national park, minnesota, usa. sample periods were: survey 1 = 6–20 june, survey 2 = 29 july – 3 august, survey 3 = 18–25 august, survey 4 = 9–14 september. alces vol. 50, 2014 cyr et al. – gastropod vectors of p. tenuis in vnp 127 ecotypes because of their relatively low use (5% and 7%) and low density of p. tenuis vectors (fig. 4). moose displayed variability in individual risk of infection as a result of differential habitat use. ten of 11 moose had relative risk scores of 0.68–1.0, and relative risk differed by ≤32% for the majority of moose. moose v09 was an exception as it spent little time in gastropod rich habitats and had a much lower risk of infection (0.21) relative to the other moose (table 3). discussion gastropod density, and more specifically density of known vectors of p. tenuis, differed among the ecotypes and sample periods. similar to previous studies, ecotypes of mixed conifer-deciduous forest types had the highest gastropod densities (gleich et al. 1977, kearney and gilbert 1978, nankervis et al. 2000). the increasing density of gastropods and potential p. tenuis vectors from summer to fall is also consistent with previous studies in northern minnesota (lankester and peterson 1996). d. laeve was the most abundant gastropod found in our study area, and is likely the most important vector of p. tenuis. most larvae in infected gastropods are presumably in the infective stage (i.e., third stage) by early july (lankester and peterson 1996) corresponding to our sampling period between mid-june and september. the cardboard sampler method is meant to provide a relative measure of gastropod diversity and abundance, and it is critical that they be as uniform as possible in shape, thickness, and wetness. all were saturated with water at the time of deployment but dried at different rates depending on habitat features (e.g., soil moisture, rockiness, exposure) and weather conditions (e.g., dry and windy vs. calm and humid). cardboard wetness varied from 0–100% at collection and this wide variation could skew the estimates of gastropod abundance because they are less likely to be found on dry cardboard (unpublished data, vnp). variation in cardboard wetness could be minimized by distributing the cardboard after the warmest part of the day and checking them before the warmest part of the next day, which would be especially important in the longer and warmer days in july and early august. past studies indicate lower boreal hardwood forest northern pine forest northern shrub swamp northern sprucefir forest rich conifer swamp treed rock barren 5 7 9 11 13 15 17 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 g as tr o p o d v ec to rs /m 2 o f tr ap propor�on of habitat use fig. 4. relative risk of moose encountering p. tenuis gastropod vectors in 6 ecotypes in voyageurs national park, minnesota, usa, july-september 2010. 128 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 gastropod abundances in early summer (lankester and anderson 1968, kearney and gilbert 1978, lankester and peterson 1996), and although these studies did not report the relative wetness of cardboard sheets, they may be biased low if sheets were drier in early summer. cardboard samplers may underestimate the total density of gastropods in an area, as the number of gastropods in the soil underneath cardboard samplers has been reported higher than those attached to the cardboard samplers (hawkins et al. 1998). by combining information about gastropod density and relative habitat use, we assessed the relative risk of p. tenuis infection for moose in different habitat types (fig. 4). we likewise calculated risk values for individual moose (table 3). these methods can also be used to compare risk of infection between different geographic areas or populations. however, we caution that the assumptions associated with our methods need to be considered carefully because seasonal variation of infection rates in gastropod hosts is not well understood (lankester and anderson 1968, kearney and gilbert 1978, lankester and peterson 1996). high whitetailed deer density has been correlated with increased infection rates of gastropods (lankester and peterson 1968) and moose (whitlaw and lankester 1994a) at larger spatial scales. a recent study found no correlation between white-tailed deer abundance and p. tenuis infection at smaller spatial scales within vnp (vanderwaal et al. 2014), although the range of deer abundance was limited across sites. risk of p. tenuis infection varies among individual moose because of differences in habitat use within respective home ranges. it will also be influenced by landscape composition and the availability of different habitats within an area. for example, the western half of the kabetogama peninsula has more area covered by the higher risk table 3. proportional habitat use and individual risk of moose encountering p. tenuis infected gastropods in the kabetogama peninsula, voyageurs national park, minnesota, usa, june-september 2010. risk value is calculated by multiplying the proportion of each ecotype used by a moose by the mean density of p. tenuis gastropod vectors measured in each ecotype. the relative index of risk is the risk value scaled to the highest risk value found for an individual moose in 2010 (i.e., moose v05). proportion habitat use moose # northern pine forest northern sprucefir forest treed rock barren boreal hardwood forest northern shrub swamp rich conifer swamp risk value relative index of risk v05 0.43 0.18 0.02 0.24 0.00 0.03 9.50 1.00 v06 0.42 0.10 0.02 0.17 0.03 0.05 8.96 0.94 v07 0.09 0.34 0.11 0.18 0.07 0.04 8.91 0.94 v14 0.05 0.35 0.14 0.19 0.10 0.04 8.76 0.92 v07 0.04 0.37 0.10 0.18 0.06 0.03 7.80 0.82 v10 0.05 0.33 0.10 0.19 0.06 0.03 7.66 0.81 v18 0.07 0.19 0.26 0.19 0.02 0.00 7.64 0.80 v17 0.13 0.25 0.08 0.18 0.03 0.02 7.44 0.78 v12 0.10 0.32 0.03 0.21 0.06 0.05 7.23 0.76 v08 0.01 0.37 0.00 0.27 0.02 0.01 6.49 0.68 v09 0.00 0.17 0.00 0.13 0.09 0.11 2.04 0.21 alces vol. 50, 2014 cyr et al. – gastropod vectors of p. tenuis in vnp 129 boreal hardwood and northern spruce-fir ecotypes, and conversely, the eastern half of the park contains more of the drier, low risk treed rock barrens and northern pine ecotypes. vanderwaal et al. (2014) found that p. tenuis infection rates in white-tailed deer increased as the proportion of vector-rich habitats such as mixed conifer-hardwood forest increased within a local area. while our methods only considered coarse habitat use in our risk index, moose behavior within individual ecotypes is presumably also important. moose may prefer to bed in certain ecotypes (e.g., in lowland habitats in hot weather) and feed in others (peek 1997), and even if gastropods are abundant in certain ecotypes, the risk of p. tenuis infection should be less in areas less preferred for foraging. risk of infection may also be affected by individual preferences for forage choice, previous exposure to p. tenuis, health status, genetics, body mass/longevity (ezenwa et al. 2006), and other factors not considered here. acknowledgments we thank m. lankester and two anonymous reviewers for comments that improved our manuscript. we thank n. walker, b. olson, and w. chen for project assistance. funding for this work was provided by voyageurs national park, usgsnps national resource preservation program, the environment and natural resources trust fund, the integrated biosciences graduate program at the university of minnesota duluth, and the natural resources research institute. the north american moose conference and workshop newcomers travel grant provided a travel grant to the senior author. references anderson, r. c. 1964. neurologic disease in moose infected experimentally with pneumostrongylus tenuis from whitetailed deer. pathologia veterinaria 1: 289–322. ———, and a. k. prestwood. 1981. lungworms. pages 266–317 in w. r. davidson, f. a. hayes, v. f. nettles, and f. e. kellogg, editors. diseases and parasites of the white-tailed deer. miscellaneous publications number 7, tall timbers research station, tallahassee, florida, usa. burch, j. b. 1962. how to know the eastern land snails. w. c. brown company, dubuque, iowa, usa. butler, e., m. carstensen, m. schrage, d. pauly, m. lenarz, and l. cornicelli. 2009. preliminary results from the 2007–2008 moose herd health assessment project. 2009 summaries of wildlife research findings. minnesota department of natural resources, st. paul, minnesota, usa. ezenwa, v., s. a. price, s. altizer, n. d. vitone, and c. cook. 2006. host traits and parasite species richness in even and odd toed hoofed mammals, artiodactyla and perissodactyla. oikos 115: 526–537. faber-landgendoen, d., n. aaseng, k. hop, and m. lew-smith. 2007. field guide to the plant community types of voyageurs national park. u.s. geological survey techniques and methods 2–a4. gillingham, m. p., and k. l parker. 2008. the importance of individual variation in defining habitat selection by moose in northern british columbia. alces 44: 7–20. gleich, j. g., f. f. gilbert, and n. p. kutscha. 1977. nematodes in terrestrial gastropods from central maine. journal of wildlife diseases 13: 43–46. gogan, p. j. p., k. d. kozie, and e. m. olexa. 1997. ecological status of moose and white-tailed deer at voyageurs national park, minnesota. alces 33: 187–201. hawkins, j. w., m. w. lankester, and r. r. a. nelson. 1998. sampling terrestrial 130 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 gastropods using cardboard sheets. malacologia 39: 1–9. hop, k., d. faber-landgendoen, m. lewsmith, n. aaseng, and s. lubinski. 2001. voyageurs national park, minnesota: usgs-nps vegetation mapping program. u.s. department of interior, u.s. geological survey, upper midwest environmental sciences center, la crosse, wisconsin, usa. johnston, c. a., and r. j. naiman. 1990. aquatic patch creation in relation to beaver population trends. ecology 71: 1617–1621. kallemeyn, l. w., k. l. holberg, j. a. perry, and b. y. odde. 2003. aquatic synthesis for voyageurs national park. usgs information and technology report 2003-0001. kearney, s. r., and f. f. gilbert. 1978. terrestrial gastropods from the himsworth game preserve, ontario, and their significance in parelaphostrongylus tenuis transmission. canadian journal of zoology 56: 688–694. kurmis, v., s. l. webb, and l. c. merriam. 1986. plant communities of voyageurs national park, minnesota, u.s.a. canadian journal of botany 64: 531–540. lankester, m. w. 2001. extrapulmonary lungworms of cervids. pages 228–278 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals, 2nd edition. iowa state university press, ames, iowa, usa. ———, 2002. low-dose meningeal worm (parelaphostrongylus tenuis) infections in moose (alces alces). journal of wildlife diseases 38: 789–795. ———, 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53–70. ———, and r. c. anderson. 1968. gastropods as intermediate host of meningeal worm, pneumostrongylus tenuis, dougherty. canadian journal of zoology 46: 373–383. ———, and w. j. peterson. 1996. the possible importance of wintering yards in the transmission of parelaphostongylus tenuis to white tailed deer and moose. journal of wildlife diseases 32: 31–38. ———, ———, and o. ogunremi. 2007. diagnosing parelaphostrongylosis in moose (alces alces). alces 43: 49–59. lenarz, m. s. 2009. a review of the ecology of parelaphostrongylus tenuis in relation to deer and moose in north america. pages 70–75 in m. w. doncarlos, r. o. kimmel, j. s. lawrence, and m. s. lenarz, editors. 2009 summaries of wildlife research findings. minnesota department of natural resources, st. paul, minnesota, usa. maskey, j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph. d. thesis, university of north dakota, grand forks, north dakota, usa. moss, m., and l. hermanutz. 2010. monitoring the small and slimy – protected areas should be monitoring native and nonnative slugs (mollusca: gastropoda). natural areas journal 30: 322–327. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. nankervis, p. j., w. m. samuel, s. m. schmitt, and j. g. sikarskie. 2000. ecology of meningeal worm, p. tenuis (nematode), in white-tailed deer and terrestrial gastropods of michigan's upper peninsula with implications for moose. alces 36: 163–181. nekola, j. c. 2002. distribution and ecology of terrestrial gastropods in northwestern minnesota. final report to 2001– 2001 nongame and natural heritage research program, minnesota alces vol. 50, 2014 cyr et al. – gastropod vectors of p. tenuis in vnp 131 department of natural resources, st. paul, minnesota, usa. ———. 2007. key to the terrestrial gastropod genera of wisconsin and nearby states. (accessed may 2012). ———, m. barthel, p. massart, and e. north. 1999. terrestrial gastropod survey of igneous outcrops in northeastern minnesota. final report to 1998 nongame and natural heritage research program, minnesota department of natural resources st. paul, minnesota, usa. peek, j. m. 1997. habitat relationships. pages 351–376 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington d. c., usa. platt, t. r. 1989. gastropod intermediate hosts of parelaphostrongylus tenuis (nematoda: metastrongyloidea) from northwestern indiana. journal of parasitology 75: 519–523. rowley, m. a., e. s. loker, and j. f. pagels. 1987. terrestrial gastropod hosts of parelaphostrongylus tenuis at the national zoological park's conservation and research center, virginia. journal of parasitology 73: 1084–1089. slomke, a. m., m. w. lankester, and w. j. peterson. 1995. infrapopulation dynamics of parelaphostrongylus tenuis in white-tailed deer. journal of wildlife diseases 31: 125–135. upshall, s. m., m. d. b. burt, and t. g. dilworth. 1986. parelaphostrongylus tenuis in new brunswick: the parasite in terrestrial gastropods. journal of wildlife diseases 22: 582–585. vanderwaal, k., s. k. windels, b. t. olson, t. vannatta, and r. moen. 2014. spatial epidemiology of liver fluke and meningeal worm in white-tailed deer in northern minnesota, usa. international journal of parasitology. in press. wünshmann, a., a. g. armien, e. butler, m. schrage, b. stromberg, j. b. bender, a. f. firshman, and m. carstensen. 2014. necropsy findings in 62 opportunistically collecting freeranging moose (alces alces) from minnesota, usa (2003–2013). journal of wildlife diseases. doi: 10.7589/201402-037. windels, s. k. 2014. 2014 voyageurs national park moose population survey report. natural resource data series nps/voya/nrds-2014/645. national park service, fort collins, colorado, usa. whitlaw, h. a., and m. w. lankester. 1994a. a retrospective evaluation of the effects of parelaphostongylosis on moose populations. canadian journal of zoology 72: 1–7. ———, and ———. 1994b. the co-occurrence of moose, white-tailed deer, and parelaphostrongylus tenuis in ontario. canadian journal of zoology 72: 819–825. ———, ———, and w. b. ballard. 1996. parelaphostrongylus tenuis in terrestrial gastropods from white-tailed deer winter and summer range in northern new brunswick. alces 32: 75–83. 132 gastropod vectors of p. tenuis in vnp – cyr et al. alces vol. 50, 2014 http://www.uwlax.edu/biology/faculty/perez/perez/perezlab/research/snailkey%20-%20april%2027%20updates.pdf http://www.uwlax.edu/biology/faculty/perez/perez/perezlab/research/snailkey%20-%20april%2027%20updates.pdf http://www.uwlax.edu/biology/faculty/perez/perez/perezlab/research/snailkey%20-%20april%2027%20updates.pdf http://www.uwlax.edu/biology/faculty/perez/perez/perezlab/research/snailkey%20-%20april%2027%20updates.pdf diversity and abundance of terrestrial gastropods in voyageurs national park, mn: implications for the risk of moose becoming infected with parelaphostrongylus tenuis introduction study area methods results discussion acknowledgments references cadmium geochemistry of soils and willow in a metamorphic bedrock terrain and its possible relation to moose health, seward peninsula, alaska larry p. gough1, paul j. lamothe2, richard f. sanzolone2, larry j. drew1, julie a. k. maier3, and john h. schuenemeyer4 1u.s. geological survey, national center, ms 954, reston, virginia 20192; 2u.s. geological survey, box 25046, denver federal center, denver, colorado 80225; 3university of alaska, p.o. box 756720, fairbanks, alaska 99775; 4southwest statistical consulting, llc, 960 sligo st, cortez, colorado 81321, usa. abstract: the regional geochemistry of soil and willow over paleozoic metamorphic rocks in the seward peninsula, alaska is potentially high in cadmium (cd), and willow, a preferred browse of moose, bioaccumulates cd. local moose show clinical signs of tooth wear and breakage and have been declining in population for unknown reasons. willow leaves (all variants of salix pulchra) and a-, b-, and c-horizon soils were sampled near 2 mining prospects suspected to be high in cd. although al, cd, co, cu, fe, mo, ni, pb, and zn were examined, our focus in this exploratory study was on the level of cd in the 3 soil horizons and willow between and within the 2 prospects and their vicinity. we used an unbalanced, one-way, hierarchical analysis of variance (anova) to investigate the geochemistry of soils and willow at various distance scales across the 2 prospect areas that were separated by ∼80 km; sites within a location were approximately 0.5 km apart and replicate samples were separated by ∼0.05 km. cd concentration was significantly different in willow between and within sites, and within sites for all soil horizons. specifically, this exploratory study identified highly elevated levels of cd in willow growing over paleozoic bedrock in the seward peninsula at both prospects and over the paleozoic geologic unit in general. potential negative effects for moose are discussed. alces vol. 49: 99–111 (2013) key words: alaska, alces alces, cadmium, health, mineralized soil, moose, plasma-mass spectrometry, willow. introduction in 2002 the united states geological survey (usgs) initially studied the relationship between regional geology and the geochemistry of soils and vegetation that occur in specific geologic terrains. specifically, how soil geochemistry and the uptake and bioaccumulation of toxic trace elements by native vegetation might ultimately affect the health of grazing herbivores (eisler 1985, brazil and ferguson 1989, gough et al. 2009); this relationship is increasingly important if animal health is threatened (glooschenko et al. 1988). moose (alces alces) are an essential cultural and economic resource in northern regions, thus their health and numbers are a primary management focus of resource agencies (maier et al. 2005, schmidt et al. 2008). local accounts of excessive tooth breakage (all moose ≥7 years old had broken incisiform teeth) and enamel defects in a declining moose population on the seward peninsula, alaska raise special concern (smith 1992, rozell 2003, stimmelmayr et al. 2006), yet the etiology of enamel defects are unclear. we propose that a possible explanation for this local moose issue is elevated 99 concentrations of cd in their preferred willow (salix spp.) browse, because high willow consumption can expose moose to elevated concentrations of cd (gough et al. 2002). in excess, cd has numerous adverse physiological effects on mammals (arnold et al. 2006, kabata-pendias 2011) including tooth and bone construction, uterus and mammary gland development, general growth inhibition, and renal tubular dysfunction (eisler 1985, larison et al. 2000). excess cd also competes with cu, zn, and ca for active sites on enzymes, phytochelatins, and cysteine-rich metal-binding proteins (metallothioneins). in general, there is a direct linear relationship between cd concentration in plant material and soils (kabata-pendias 2011). uptake in plants is affected by soil ph, carbonate and clay content, and cd in plants is associated with its affinity for sulfhydryl groups and other side chains of proteins (kabata-pendias 2011). uptake by plants is also affected by a number of physical and chemical soil features; as soil ph decreases, uptake increases (hough et al. 2003), and uptake generally increases as the total amount in soil increases. low microbial soil activity in soils, as in the study area, enhances oxic soil conditions which enhances uptake; conversely, permafrost and low soil temperatures reduce uptake. however, we reported previously that cd is bio-accumulated in willow at levels several times higher than that in other native vegetation, up to 10–100 × greater at the same location (larison et al. 2000, gough et al. 2002, 2006). in areas of alaska that lack diversity of winter forage species like the seward peninsula, moose consume willow almost exclusively and are known to remove >55% of the current annual twig growth (bowyer and neville 2003). we hypothesize that bioaccumulation of cd by willow in areas of alaska naturally high in cd may be detrimental to the health of moose (gough et al. 2002) either by being directly toxic (nephropathy or poor bone construction) and/or by inducing cu deficiency (frank et al. 2000). the purpose of this pilot study was to describe the biogeochemistry of cd in soil and willow in an area with documented physical abnormalities in moose and regionally elevated graphite and cd concentrations in bedrock (j. slack, usgs, pers. comm.). study area the study occurred on the seward peninsula, alaska at 2 locations, the quarry prospect and big hurrah transects (fig. 1); both locations have a long history of placer gold mining (collier et al. 1908, kaufman 1986, read and meinert 1986). the 2 locations were separated by ∼80 km, collection sites within each location were approximately 0.5 km apart, and within a site near replicate soil samples were collected ∼0.05 km apart. the a-, b-, and c-horizon soil and willow leaf samples were collected from 21 sites combined. location 1 (quarry prospect transect) had 10 sampling sites located northeast of the teller road between the sinuk and cripple rivers at approximately 64° 42′ n latitude and 165° 45′ w longitude (fig. 1). the area of arctic tundra/shrub tundra was at ∼230 m elevation and extended from the quarry prospect (an excavated pit with abundant sulfide mineralization) northeast for 3 km. bedrock geology of the area is composed of paleozoic metamorphic rocks (till et al. 1986, till et al. 2011), and based on the map of bundtzen et al. (1994), is within both the massive marble and the graphitic schist and quartzite members, the latter described as either carbonaceous, finegrained mudstones or mylonites. these units are known to be potentially high in cd (werdon et al. 2005b). location 2 (big hurrah transect) with 11 sites was east of the council road in an area 100 cadmium and moose health – gough et al. alces vol. 49, 2013 of arctic tundra/shrub tundra at elevation of ∼120–150 m. the sampling transect circumnavigated a low hill (identified on usgs c-5 quadrangle map as hill 596) and was in the southern half of section 33 at approximately 64° 40′ n latitude and 164° 15′ w longitude. bedrock geology of the area is defined as ordovician to precambrian graphitic schist and quartzite on the north, west, and south sides of the hill, and ordovician to precambrian schist on the east; both units are part of the mixed unit as identified by till et al. (1986, 2011) and werdon et al. (2005a, b). like location 1, these geologic units are known to be potentially high in cd (werdon et al. 2005b). methods soil samples in general, soils of the seward peninsula ecoregion (nowacki et al. 2002), sometimes referred to as the norton sound highlands, are classified as pergelic cryaquepts to pergelic cryorthents (rieger et al. 1979). these soils belong to the soil orders inceptisol and entisol, respectively, are both poorlyand well-drained, underlain by permafrost, and commonly form in gravely colluvium. depth to permafrost varies depending on aspect (slope orientation) and elevation and was between 15–90 cm. soil sample pits were dug to a depth that included the c-horizon. each sample was a mixture of soil that originated most commonly from the weathering of colluvium, bedrock, and loess. a-, b-, and c-horizon materials were collected, rocks were removed, and approximately 0.5 kg of the material was put into paper soil bags. soil samples were dried under forced air at ambient temperature. the air-dried samples were disaggregated in a mechanical mortar and pestle and sieved at 2 mm (10 mesh), and the minus-2-mm fraction was saved for fig. 1. simplified geology of the southern seward peninsula, alaska, the study area and transect locations within (geology and base map after till et al. 2011). alces vol. 49, 2013 gough et al. – cadmium and moose health 101 further analysis. a split of the minus-2-mm material was ground to pass through a 0.15mm sieve, using an agate shatter box. this material was subjected to chemical analysis using inductively coupled plasma-mass spectrometry (icp-ms) following a 4-acid digestion protocol (crock et al. 1999, briggs and meier 2002). a subset of a-, b-, and c-horizon soils was examined by quantitative x-ray diffraction (xrd) for their bulk mineralogical composition (gough et al. 2008). plant samples plant sampling was limited to the leaf material of the ubiquitous willow of the region, salix pulchra (tealeaf willow; fig. 2). although many willows are not considered preferred browse species because of the presence of tannins and alkaloids (hansjoachim et al. 1979), s. pulchra contains relatively low amounts of these 2 substances and is actually preferred by moose. this species is quite common in areas throughout alaska and canada occurring within forests, at and above tree line, and in arctic tundra with adaptability and propensity to form hybrids (hultén 1968). it is easy to identify in the field, even without flowers or seeds, because of its broad, diamond-shaped to elliptical leaves and its tendency to retain the previous year’s leaves and stipules; the latter trait makes the shrubs quite easy to identify at a distance. it is obvious from field observations that moose browsed on both leaves and twigs. the leaf material from at least 3 adjacent shrubs (in a radius ∼5 m from a soil pit) fig. 2. this female moose is foraging in a stand of salix pulchra (tealeaf willow) near the shore of norton sound; s. pulchra is common throughout the southern seward peninsula, alaska. 102 cadmium and moose health – gough et al. alces vol. 49, 2013 was composited, placed in cloth sample bags, and allowed to air dry. in the laboratory leaf material was placed in teflon® beakers, submerged and rinsed in deionized water, and drained; this process was repeated 3 times. the material was then rinsed briefly with deionized water and allowed to drip drain, and forced air was used to dry the material at ambient room temperature. samples were ground in a wiley® mill to pass a 2-mm sieve, ashed in an oven at 450–500 °c for 18 h, digested using the same 4-acid protocol as the soil samples, and analyzed using icp-ms (briggs and meier 2002). statistics the study design was constructed to investigate differences in levels of cd in willow and soil geochemistry between and within locations. an unbalanced, one-way, hierarchical anova was performed (systat 11, systat® software, inc.) to assess possible significance where cd was the response variable. the analyses were performed on the log base 10-transformed data because of the right-skewed nature of the data (miesch 1976). because the prospect locations and samples within locations were purposefully selected, this is considered a ‘fixed effects’ model procedure. this statistical design allows the partitioning of the total measured natural variation into its component parts, level 1 and level 2. level 1 is the comparison of means between locations (the quarry prospect and big hurrah areas) and level 2 compares the means within locations; the nearby samples are used to estimate the error term. all samples were analyzed in a random sequence to help negate any systematic errors that might occur in either sampling or analysis. factor analysis is a multivariate statistical procedure designed to describe variability by partitioning it into some smaller number of common factors and a component unique to each variable (schuenemeyer and drew 2011). it was used as an exploratory tool to examine possible correlations among the element concentrations. the goals were to 1) determine if the factors can be interpreted according to some geochemical association, and 2) determine if factors vary within and between willow leaves and soil horizons. results soil analysis although the soils sampled in the seward peninsula (mostly pergelic cryaquepts to pergelic cryorthents, rieger et al. 1979) contain transported loess material, they are predominantly residual, organic in nature, and composed of weathered metamorphic bedrock and loess. the samples were analyzed for numerous elements (gough et al. 2008); however, here we focus on the biogeochemistry of cd and 8 other metals. the quarry prospect and big hurrah transects had 10 cd samples in 7 levels (sample locations) and 11 cd samples in 8 levels, respectively, for the a-, b-, and c-horizon soils. there were 13 cd willow samples in 9 levels in quarry prospect and 8 samples in 6 levels in big hurrah. the bulk mineralogical composition of selected soil samples determined by quantitative xrd is presented in table 1. the hierarchical anova using log base 10 values (table 2) indicated that cd concentrations in all 3 soil horizons were similar at the level 1 effect (between locations); conversely, the level 2 (within locations) effect was significant (p < 0.05; table 2). we caution that sample sizes were small. summary statistics for the concentration of elements in willow and the a-, b-, and chorizon soils at the quarry prospect and big hurrah transects are presented in table 3. the main purpose of this table and units (log base 10) is to provide descriptive analysis (mean, standard deviation) and comparison among the elements and soil horizons. alces vol. 49, 2013 gough et al. – cadmium and moose health 103 this comparison is best made when data are in log units since 1or 2 observations can skew the mean and/or standard deviation (sd). the following is a descriptive analysis based upon inspection of means and sd and is not based on the results of statistical tests. these data (table 3) are useful as preliminary geochemical baseline values for the 2 locations. sample means for cd concentration among the soil horizons were higher in the big hurrah than quarry prospect, but not significantly different (p > 0.05); the same pattern occurred for cu, fe, mo, and ni concentrations. conversely, sample means for al, co, pb, and zn were higher at quarry prospect in all soil horizons. willow analysis the anova for cd concentrations in willow leaves and soils is presented in table 2; both the level 1 and the level 2 (level 1) effects were highly significant (p < 0.001). the mean concentrations of cd, fe, ni, and zn from the horizon samples were consistent between the 2 transect locations (table 3). factor analysis for the a-, b-, and c-horizon soils, the variables were logarithmically-transformed element concentrations expressed in parts per million (ppm). the choice of 3 common factors was made after examining the data, and a varimax (orthogonal) rotation was used. since all concentrations are in log base 10 of ppm, factor analysis was performed on the covariance matrix. the factor analysis with the willow data was performed similarly except that mo and pb were omittedta b le 1 . b u lk m in er al o g y (q u an ti ti v e x r d ; g o u g h et al ., 2 0 0 8 ) fo r re p re se n ta ti v e sa m p le s o f a -, b -, an d c -h o ri zo n tu n d ra so il s d ev el o p ed fr o m b ed ro ck an d lo es s, s ew ar d p en in su la . [o p ϵt , p re ca m b ri an m ix ed u n it o f th e n o m e g ro u p (c h lo ri te -r ic h sc h is t an d m ar b le ); o p ϵs q , o rd o v ic ia n to p re ca m b ri an m ix ed u n it o f th e n o m e g ro u p (g ra p h it ic sc h is t an d q u ar tz it e) ; — , m in er al w as n o t o b se rv ed ]. w ei g h t p er ce n t s am p le id en ti fi er s o il h o ri zo n r o ck u n it l o ss o n ig n it io n q u ar tz p o ta ss iu m fe ld sp ar p la g io cl as e c al ci te d o lo m it e a m p h ib o le p y ri te g o et h it e a p at it e r u ti le p ea t 0 5 a k 0 11 a a o p ϵt 8 .8 3 9 0 .4 2 .3 — — — — 1 .3 0 .1 0 .6 9 .8 0 5 a k 0 11 b b 3 .5 3 6 0 .8 1 .5 — 0 .1 — — 2 .3 0 .6 0 .5 3 .5 0 5 a k 0 11 c c 4 .2 3 5 0 .4 1 .3 — — — — 3 .7 — 0 .7 5 .1 0 5 a k 0 2 1 a a o p ϵt 4 1 1 9 1 .1 1 .6 — 0 .2 — 0 .1 2 .3 0 .2 0 .1 3 9 0 5 a k 0 2 1 b b 3 .5 4 4 2 .5 2 .1 0 .2 — — — 3 .2 0 .3 0 .1 6 .2 0 5 a k 0 2 1 c c 3 .5 4 6 1 .8 2 .1 0 .2 — — — 3 .7 0 .3 0 .1 6 .5 0 5 a k 1 3 1 a a o p ϵs q 1 0 4 5 1 .6 2 .1 — — — — 2 .4 — 0 .2 1 7 0 5 a k 1 3 1 b b 8 .8 4 7 2 2 .2 — — — 0 .1 2 .3 0 .1 — 1 3 0 5 a k 1 3 1 c c 8 .8 4 8 1 .8 2 .2 — — — — 2 .5 0 .2 — 1 2 104 cadmium and moose health – gough et al. alces vol. 49, 2013 because of the presence of censored data (less than the detection limit). the factor analysis is presented in table 4 with the largest absolute values highlighted in each row. the numbers under the factor column headings are loadings (weights) of a chemical element on a factor. element loadings may be considered to be the correlation between an element and a factor. for example, in the a-horizon, cd loads heavily (0.778) on factor 1 (i.e., cd and factor 1 are strongly associated), and lightly on factors 2 (0.126) and 3 (0.198), and has a unique component of 0.339; the unique component usually contains error which is difficult to isolate. in total, the factor loading patterns were consistent across the 3 soil horizons. this is illustrated by factor 1 in the a-horizon and factor 2 in the band c-horizons loading heavily on cd, pb, and zn, factor 2 in the a-horizon and factor 3 in the band c-horizons loading heavily on co and fe, and factor 3 in the a-horizon and factor 1 in the band c-horizons loading heavily on cu, mo, and ni (table 4). note that the variability accounted for by factors 1 and 2 is approximately the same, so the fact that the pattern appeared in factor 1 in the a-horizon and factor 2 in the band c-horizons is not important. unfortunately, there is no clear factor pattern in willow and the data set was too small to justify a specification of more than 3 factors. table 2. results of a hierarchical anova of cadmium (cd) concentration (n = 21; data are in log base 10) measured in 3 soil horizons and willow leaves in the quarry prospect and big hurrah regions, seward peninsula, alaska, usa. source sum of squares degrees of freedom mean squares f-ratio p-value a-horizon soil level 1 0.029 1 0.029 0.517 0.499 level 2(level 1) 3.266 13 4.487 4.487 0.038* error 0.336 6 0.015 total sum of squares 3.631 b-horizon soil level 1 0.098 1 0.098 2.285 0.181 level 2(level 1) 4.849 13 0.373 8.700 0.007* error 0.257 6 0.043 total sum of squares 5.204 c-horizon soil level 1 0.128 1 0.128 1.978 0.209 level 2(level 1) 4.604 13 0.354 5.489 0.023* error 0.387 6 0.065 total sum of squares 5.119 willow level 1 1.622 1 1.622 150.362 0.0001* level 2(level 1) 4.333 13 0.333 21.648 0.001* error 0.092 6 0.015 total sum of squares 6.047 *, significant at the 0.05 probability level. alces vol. 49, 2013 gough et al. – cadmium and moose health 105 table 3. summary statistics for the concentration of selected elements measured in willow leaves and a-, b-, and c-horizon soils in the quarry prospect and big hurrah regions, seward peninsula, alaska, usa. cadmium (cd) results are highlighted; "—" = not determined due to the presence of values below the detection limit. the detection ratio expresses the number of values above the detection limit to the total number of analyses. willow a-horizon soil b-horizon soil c-horizon soil mean std dev detection mean std dev detection mean std dev detection mean std dev detection element log base 10 log base 10 ratio log base 10 log base 10 ratio log base 10 log base 10 ratio log base 10 log base 10 ratio quarry prospect transect al, ppm 1.818 0.185 10:10 4.729 0.172 10:10 4.822 0.153 10:10 4.819 0.153 10:10 cd, ppm 0.478 0.553 10:10 0.080 0.546 10:10 –0.099 0.653 10:10 –0.050 0.637 10:10 co, ppm –0.578 0.390 10:10 1.040 0.135 10:10 1.092 0.146 10:10 1.163 0.145 10:10 cu, ppm 0.935 0.167 10:10 1.341 0.099 10:10 1.341 0.135 10:10 1.356 0.105 10:10 fe, ppm 1.992 0.130 10:10 4.550 0.083 10:10 4.628 0.091 10:10 4.674 0.114 10:10 mo, ppm — — 6:10 –0.207 0.172 10:10 –0.258 0.204 10:10 –0.272 0.146 10:10 ni, ppm 0.072 0.228 10:10 1.346 0.168 10:10 1.399 0.185 10:10 1.467 0.146 10:10 pb, ppm — — 2:10 1.608 0.676 10:10 1.641 0.676 10:10 1.653 0.661 10:10 zn, ppm 2.390 0.243 10:10 2.445 0.537 10:10 2.433 0.602 10:10 2.448 0.601 10:10 big hurrah transect al, ppm 1.868 0.208 10:11 4.665 0.159 11:11 4.749 0.138 11:11 4.754 0.117 11:11 cd, ppm 1.187 0.231 11:11 0.133 0.303 11:11 0.005 0.357 11:11 0.073 0.365 11:11 co, ppm 0.358 0.335 11:11 0.908 0.323 11:11 1.019 0.353 11:11 1.101 0.295 11:11 cu, ppm 0.826 0.100 11:11 1.702 0.203 11:11 1.813 0.171 11:11 1.878 0.190 11:11 fe, ppm 2.022 0.159 11:11 4.560 0.253 11:11 4.673 0.238 11:11 4.697 0.222 11:11 mo, ppm — — 7:11 1.083 0.219 11:11 1.134 0.186 11:11 1.155 0.209 11:11 ni, ppm 0.810 0.200 11:11 1.683 0.239 11:11 1.782 0.229 11:11 1.832 0.193 11:11 pb, ppm — — 0:11 1.166 0.214 11:11 1.241 0.169 11:11 1.266 0.150 11:11 zn, ppm 2.211 0.137 11:11 2.173 0.172 11:11 2.268 0.202 11:11 2.310 0.195 11:11 1 0 6 c a d m iu m a n d m o o s e h e a l t h – g o u g h e t a l . a l c e s v o l . 4 9 , 2 0 1 3 table 4. factor analysis of the concentration values (n = 21; data are in log base 10) for al, cd, co, cu, fe, mo, ni, pb, and zn measured in 3 soil horizons and willow leaves in the quarry prospect and big hurrah regions, seward peninsula, alaska, usa. the symbol “—” indicates not calculated because of the presence of censored data; also not used in the cumulative variance calculation (see text). element factor 1 factor 2 factor 3 unique component factor loadings for the a-horizon al, ppm 0.352 0.568 0.545 cd, ppm 0.778 0.126 0.198 0.339 co, ppm 0.200 0.975 0.005 cu, ppm 0.263 0.902 0.118 fe, ppm 0.120 0.831 0.246 0.234 mo, ppm −0.112 −0.257 0.921 0.073 ni, ppm 0.510 0.791 0.110 pb, ppm 0.871 0.199 −0.210 0.158 zn. ppm 0.968 0.217 −0.108 0.005 cumulative variance 0.277 0.551 0.824 factor loadings for the b-horizon al, ppm −0.121 0.296 0.474 0.673 cd, ppm 0.318 0.800 0.238 0.202 co, ppm 0.171 0.982 0.005 cu, ppm 0.967 0.173 0.026 fe, ppm 0.300 0.128 0.858 0.157 mo, ppm 0.901 −0.177 −0.267 0.085 ni, ppm 0.789 0.457 0.160 pb, ppm −0.215 0.890 0.157 0.138 zn. ppm 0.984 0.164 0.005 cumulative variance 0.291 0.578 0.839 factor loadings for the c-horizon al, ppm 0.333 0.599 0.523 cd, ppm 0.329 0.792 0.276 0.188 co, ppm 0.203 0.976 0.005 cu, ppm 0.958 0.069 fe, ppm 0.212 0.159 0.875 0.164 mo, ppm 0.929 −0.110 −0.218 0.076 ni, ppm 0.847 0.359 0.143 pb, ppm −0.225 0.866 0.223 0.150 zn. ppm 0.979 0.189 0.005 cumulative variance 0.302 0.584 0.853 alces vol. 49, 2013 gough et al. – cadmium and moose health 107 discussion in order to assess the scale of spatial variability in the concentration of cd and other elements in soils and willow across the landscape, sampling occurred at 2 mineralized prospects separated by 80 km. the greatest difference in cd concentration in soils occurred within locations across all soil horizons and not between the locations, indicating general uniformity in landscape geochemistry. for willow, an important proportion of the total biogeochemical variability of cd occurred between and within locations. when one examines the distribution of cd, these trends may be due to variation in soil mineralogy, especially in the amount of amorphous graphite present because it has been associated with cd. unfortunately, because the graphite in soils is amorphous, it is not detectable in the quantitative xrd procedure. differences in the transition metals cd, co, ni, and zn may be explained by variability in the amount of graphite in the bedrock because in this terrain high graphite content correlates with high levels of transition metals (j. slack, usgs, pers. commun.). for these elements, the geochemistry of the bedrock appears to affect the biogeochemistry of the willow. together, these trace element data show consistency among the soil horizons whereas, because of the too small data set, the pattern for willow could not be characterized. this exploratory study identified elevated levels of bioavailable cd in soils developed over paleozoic metamorphic bedrock and local willow leaves on the seward peninsula, alaska. typical cd content across a variety of plant foodstuffs (grasses, grains, vegetables, fruits) ranges from 0.005–1.3 ppm dry weight (kabata-pendias 2011), whereas in this study we found much higher levels of 0.65–42.0 ppm cd in willow; the location means were 3.0 and 15.0 ppm. this corresponds to previous reports of high cd concentrations in willow in colorado (larison et al. 2000) and alaska (gough et al. 2002). however, the levels from the seward peninsula are higher than those reported for willow in the colorado ore belt (larison et al. 2000). because willow can bioaccumulate cd, its role in the health of the local moose population is of concern given the endemic tooth breakage and negative physiological effects associated with elevated cd in mammals. a direct moose tissue analysis was not performed, but would be warranted in the area. acknowledgements the authors thank personnel of the bering straits native corporation, and table 4. continued element factor 1 factor 2 factor 3 unique component factor loadings for willow al, ppm 0.103 0.992 0.005 cd, ppm 0.993 0.005 co, ppm 0.656 0.688 0.096 cu, ppm −0.551 0.693 fe, ppm 0.608 0.623 mo, ppm — ni, ppm 0.659 0.711 0.235 0.005 pb, ppm — zn. ppm 0.172 −0.726 0.126 0.428 cumulative variance 0.314 0.530 0.735 108 cadmium and moose health – gough et al. alces vol. 49, 2013 especially i. anderson of the land and resources department for anecdotal historical moose information and allowing us access to native corporation lands. the bulk mineralogy for soils was provided by d. eberl, usgs, boulder. we also thank a. till, geologist with the usgs in anchorage, for providing the geologic base used in our figures and for her guidance in the field. references arnold, s. m., r. l. zarnke, v. l. tracey, m-a. chimonas, and a. frank. 2006. public health evaluation of cadmium concentrations in liver and kidney of moose (alces alces) from four areas of alaska. science of the total environment 357: 103–111. bowyer, r. t., and j. t. neville. 2003. effects of browsing history by alaskan moose on regrowth and quality of feltleaf willow. alces 39: 193–202. brazil, j., and s. ferguson. 1989. cadmium concentrations in newfoundland moose. alces 25: 52–57. briggs, p. h., and a. l. meier. 2002. the determination of forty-two elements in geological materials by inductively coupled plasma-mass spectrometry. chapter i in j. e. taggart, editor. analytical methods for chemical analysis of geologic and other materials. u. s. geological survey open-file report 02-223. u. s. geological survey, denver, colorado, usa. bundtzen, t. k., r. d. reger, g. m. laird, d. s. pinney, k. h. clautice, s. a. liss, and g. r. cruse. 1994. preliminary geologic map of the nome mining district. state of alaska division of geology and geophysical surveys, public-data file 94-39. alaska division of geological & geophysical surveys, fairbanks, alaska, usa. collier, a. j., f. l. hess, p. s. smith, and a. h. brooks. 1908. the gold placers of parts of seward peninsula, alaska, including the nome, council, kougarok, port clarence, and goodhope precincts. u. s. geological survey bulletin 328. crock, j. g., b. f. arbogast, and p. j. lamothe. 1999. laboratory methods for the analysis of environmental samples. pages 265–287 in g. s. plumlee and m. j. logsdon, editors. reviews in economic geology volume 6a, the environmental geochemistry of mineral deposits, part a, processes, techniques, and health issues. society of economic geologists, inc., littleton, colorado, usa. eisler, r. 1985. cadmium hazards to fish, wildlife, and invertebrates a synoptic review. u. s. fish and wildlife service biological report 85 (1.2). u. s. fish and wildlife service, laurel, maryland, usa. frank, a., r. danielsson, and b. jones. 2000. the ‘mysterious’ disease in swedish moose. concentrations of trace elements in liver and kidneys and clinical chemistry. comparison with experimental molybdenosis and copper deficiency in the goat. the science of the total environment 249: 107–122. glooschenko, v., c. downes, r. frank, h. e. braun, e. m. addison, and j. hickie. 1988. cadmium levels in ontario moose and deer in relation to soil sensitivity to acid precipitation. science of the total environment 71: 173–186. gough, l. p., j. g. crock, and w. c. day. 2002. cadmium accumulation in browse vegetation, alaska implication for animal health. pages 77–78 in h. c. w. skinner and a. berger, editors. geology and health closing the gap. oxford university press, new york, new york, usa. ———, ———, b. wang, w. c. day, d. d. eberl, r. f. sanzolone, and p. j. lamothe. 2008. substrate geochemistry and soil development in boreal forest and tundra ecosystems in the yukon-tanana upland and seward peninsula, alaska. u. s. geological survey scientific alces vol. 49, 2013 gough et al. – cadmium and moose health 109 investigations report 2008-5010. u. s. geological survey, reston, virginia, usa. ———, r. eppinger, p. h. briggs, and s. giles. 2006. biogeochemical characterization of an undisturbed highly acidic, metal-rich bryophyte habitat, east-central alaska. arctic, antarctic, and alpine research 38: 522–529. ———, p. j. lamothe, r. f. sanzolone, l. j. drew, and j. a. k. maier. 2009. the regional geochemistry of soils and willow in a metamorphic bedrock terrain, seward peninsula, alaska, 2005, and its possible relation to moose. u.s. geological survey open-file report 2009–1124. u.s. geological survey, reston, virginia, usa. hans-joachim, g. j., g. o. batzli, and d. s. seigler. 1979. patterns in the phytochemistry of arctic plants. biochemical systematics and ecology 7: 203–209. hough, r. l., s. d. young, and n. m. crout. 2003. modeling of cd, cu, ni, pb and zn uptake, by winter wheat and forage maize, from a sewage disposal farm. soil use and management 19: 19–27. hultén, e. 1968. flora of alaska and neighboring territories. stanford university press, stanford, california, usa. kabata-pendias, a. 2011. trace elements in soils and plants, 4th edition. crc press, boca raton, florida, usa. kaufman, d. s. 1986. surficial geologic map of the solomon, bendeleben, and southern part of the kotzebue quadrangles, western alaska. u.s. geological survey miscellaneous field studies map mf01838-a. larison, j. r., g. e. likens, j. w. fitzpatrick, and j. g. crock. 2000. cadmium toxicity among wildlife in the colorado rocky mountains. nature 406: 181–183. maier, j. a. k., j. m. ver hoef, a. d. mcguire, r. t. bowyer, l. saperstein, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics effects of scale. canadian journal of forestry research 35: 2233–2243. miesch, a. t. 1976. geochemical survey of missouri methods of sampling, laboratory analysis, and statistical reduction of data. u.s. geological survey professional paper 954-a. u. s. government printing office, washington, d. c., usa. nowacki, g., p. spencer, m. fleming, t. brock, and m. t. jorgenson. 2002. unified ecoregions of alaska 2001. u. s. geological survey open-file report 02-297. (accessed october 2012). read, j. j., and l. d. meinert. 1986. goldbearing quartz vein mineralization at the big hurrah mine, seward peninsula, alaska. economic geology 81: 1760– 1774. rieger, s., d. b. schoephorster, and c. e. furbush. 1979. exploratory soil survey of alaska. u.s. department of agriculture, soil conservation service, washington, d. c., usa. rozell, n. 2003. the mystery of the broken moose teeth. alaska science forum article 1669. university of alaskafairbanks, fairbanks, alaska, usa. schmidt, j. i., k. j. hundertmark, r. t. bowyer, and k. g. mccracken. 2008. population structure and genetic diversity of moose in alaska. journal of heredity advance access. (accessed october 2012). schuenemeyer, j. h., and l. j. drew. 2011. statistics for earth and environmental scientists. john wiley & sons, hoboken, new jersey, usa. smith, t. e. 1992. incidence of incisiform tooth breakage among moose from the seward peninsula, alaska, usa. alces supplement 1: 207–212. stimmelmayr, r., j. a. k. maier, k. person, and j. battig. 2006. incisor tooth breakage, enamel defects, and periodontitis in a declining alaskan moose population. alces 42: 65–74. 110 cadmium and moose health – gough et al. alces vol. 49, 2013 http://agdc.usgs.gov/data/projects/fhm http://agdc.usgs.gov/data/projects/fhm http://jhered.oxfordjournals.org/cgi/content/full/esn076v1 http://jhered.oxfordjournals.org/cgi/content/full/esn076v1 http://jhered.oxfordjournals.org/cgi/content/full/esn076v1 till, a. b., j. a. dumoulin, b. m. gamble, d. s. kaufman, and p. i. carroll. 1986. preliminary geologic map and fossil data, solomon, bendeleben, and southern kotzebue quadrangles, seward peninsula, alaska. u.s. geological survey open-file report 86-276. (accessed october 2012). ———, ———, m. b. werdon, and h. a. bleick. 2011. bedrock geologic map of the seward peninsula, alaska: u.s. geological survey scientific investigations map 3131, scale 1:500,000. (accessed october 2012). werdon, m. b., r. j. newberry, d. j. szumigala, j. e. athey, and s. a. hicks. 2005a. bedrock geologic map of the big hurrah area, northern half of the solomon c-5 quadrangle, seward peninsula, alaska. report of investigations 20051b. state of alaska division of geological and geophysical surveys, fairbanks, alaska, usa. ———, d. s. p. stevens, r. j. newberry, d. j. szumigala, j. e. athey, and s. a. hicks. 2005b. explanatory booklet to accompany geologic, bedrock, and surficial maps of the big hurrah and council areas, seward peninsula, alaska. report of investigations 20051a. state of alaska division of geological and geophysical surveys, fairbanks, alaska, usa. alces vol. 49, 2013 gough et al. – cadmium and moose health 111 http://pubs.er.usgs.gov/ http://pubs.er.usgs.gov/ http://pubs.usgs.gov/sim/3131/ http://pubs.usgs.gov/sim/3131/ cadmium geochemistry of soils and willow in a metamorphic bedrock terrain and its possible relation to moose health, seward peninsula, alaska introduction study area methods soil samples plant samples statistics results soil analysis willow analysis factor analysis discussion acknowledgements references alces29_213.pdf alces(25)_48.pdf alces24_56.pdf alces24_22.pdf rodgersar text box alces(23)_227.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces26_37.pdf alces21_507editorialreviewcom.pdf alces vol. 21, 1985 alces26_73.pdf alces22_253.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces24_97.pdf alces27_74.pdf alces22_323.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces27_226erratumcropped.pdf alces28_165.pdf alces22_139.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces(23)_49.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces21_447.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces21_359.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 echinococcus granulosus genotype g8 in maine moose (alces alces) anne lichtenwalner1, nirajan adhikari1, lee kantar2, emily jenkins3 and janna schurer3 1university of maine animal health lab, 5735 hitchner hall, orono, me 04469; 2maine dept. of inland fisheries and wildlife, 650 state st., bangor, me 04401; 3dept. of veterinary microbiology, western college of veterinary medicine, university of saskatchewan, saskatoon, sk, canada, s7n 5b4. abstract: during a 2012 survey of harvested moose (alces alces) in maine (usa), an incidental finding of hydatid cysts was found in 39% (21 of 54) of lung sets examined. cytology of cyst contents was consistent with echinococcus granulosus. the g8 genotype was identified based on pcr and dna sequencing of a 470 base pair region of the nadh dehydrogenase subunit 1 (nad1) mitochondrial gene. the hydatid cysts were the northern, or cervid genotype and this is the first confirmed report of e. granulosus in maine moose. the atlantic regions of the northern usa and canada were not previously thought to be endemic regions for e. granulosus. it is presumed that either domestic dogs or eastern coyotes (canis latrans) are the definitive host. alces vol. 50: 27–33 (2014) key words: alces alces, echinococcus granulosus, hydatid cyst, moose, zoonosis. moose (alces alces) populations are in general decline in much of the northeastern united states, but recent population estimates in maine are reasonably stable based on aerial surveys across core moose range where density ranges from 0.4–4 moose/ km2 (unpublished data, maine department of inland fisheries and wildlife [ifw]; kantar and cumberland 2013). outside of alaska, this regional population is the largest in the united states occupying remote commercial, boreal forestland in northern maine to predominantly hardwood forests in massachusetts. given the substantial ecological, economic, and cultural importance of moose in the region, periodic studies of moose health are warranted to predict potential threats to the moose population, and to monitor for zoonotic disease or other diseases transferrable between wildlife and livestock. echinococcus granulosus is a parasitic cestode (“tapeworm”) with 2 hosts; a carnivore (usually canidae) definitive host, and an intermediate ungulate host (eckert et al. 2001). in the definitive host, the tapeworm resides in the small intestine and causes little pathology. the adult tapeworm releases eggs into the feces of the definitive host and these eggs may persist for long periods in the environment (eckert et al. 2001). after ingestion of these eggs by the intermediate host, immature forms of the tapeworm migrate into host tissues, developing into cystic structures containing protoscolices (immature heads) of the tapeworm. the tapeworm life cycle is completed when a suitable definitive host consumes cyst-containing tissue of an intermediate host. the definitive host then becomes infected with the tapeworm, which develops to the adult stage within the host’s small intestine. at least 10 genotypes of e. granulosus have been identified using molecular methods, and circulate in unique host assemblages around the world. the g8 and g10 27 strains are associated with the “northern” or “sylvatic” biotype of e. granulosus, whose definitive host is a wild canid such as a wolf (canis lupis) or coyote (c. latrans) and whose intermediate host is usually a wild cervid such as elk (cervus canadensis), deer (odocoileus spp.), caribou (rangifer tarandus) or moose. this is the first published report of echinococcus granulosus in maine moose and also the first report of the g8 genotype in the northeastern united states. study area & methods during the 2012 maine moose hunting season, a partial survey of harvested moose was conducted in collaboration with ifw to determine the prevalence of a novel dictyocaulus spp. lungworm observed in preceding seasons (a. b. lichtenwalner, unpublished data). in order to most efficiently collect lung samples from harvested moose, one hunter check-in site was staffed during the fall hunt. hunters returning from wildlife management districts (wmd) 2, 3, 4, 5, and 6 were expected to use this check point for confirmation of their permit and data collection by the ifw (fig. 1). during the previous year’s hunt (2011), these wmds yielded 309 ± 85 moose each with an overall success rate of 76% (# killed/# permits). during the 2012 hunt, these wmds yielded 384 ± 144 moose with a success rate of 86% (fig. 1, 3; ). in total, lung sets (except a single case with one lung) from 54 moose were examined for cysts; 41 were collected at the hunter check-in site and 13 were collected by ifw staff and delivered to the university of maine animal health lab (umahl), including 3 from wmd 1. hunters provided the permit number, seal number, general location where the animal was shot, gender, and estimated age of the moose, along with a bag containing the trachea and lungs. the tissues were collected from the hunters within approximately one day of death. tissues were transported on ice to the umahl where they were visually inspected and dissected using routine biosafety precautions. the lungs were flushed with a saline solution to recover any lungworms present in the airways, followed by dissection along the airways to recover lungworms. fluid was aspirated from selected pulmonary cysts using a 20 gauge needle and syringe. the cyst contents were spun down at 1500 rpm for 5 min. aliquots of the sediment were examined as wet mounts, and placed into either 10% formalin for histology, or 95% ethanol. the ethanol-fixed cyst sediment was shipped to western college of veterinary medicine at the university of saskatchewan, saskatchewan, canada for dna extraction and genotyping. the sediment was re-suspended in approximately 1 ml of 70% ethanol, and dna was extracted as previously described (schurer et al. 2013). briefly, primers were used to amplify a 470 base pair region of the nadh dehydrogenase subunit 1 (nad1) mitochondrial gene (bowles and mcmanus 1993). electrophoresis (110v, 30 min) using 1.5% agarose gel and redsafe nucleic acid staining solution (chembio ltd, hertfordshire, united kingdom) was conducted to resolve pcr products, and bands were visualized under uv light. pcr products were purified using the qiaquick pcr purification kit (qiagen inc, valencia, california, usa) and sent for sequencing (macrogen inc., seoul, korea). a staden software package (pregap 4, gap 4) was utilized to align dna sequences which were ultimately submitted to genbank™ (national center for biotechnology information), and identified to the genotype level by comparison to reference sequences. results pulmonary cysts, consistent with hydatid cysts, were found in 39% (21 of 54) of 28 e. granulosus g8 in maine – lichtenwalner et al. alces vol. 50, 2014 http://www.maine.gov/ifw/hunting_trapping/hunting/harvest.htm http://www.maine.gov/ifw/hunting_trapping/hunting/harvest.htm moose lung samples. moose with cysts were from 16 townships representing all 6 wmds; effectively, the geographic spread included the entire study area stretching east to west across northern maine (fig. 1); only 1 township had >1 case. although exact numbers of cysts were not measured, the relative amount was assessed as none, few (∼10 or fewer), moderate (11–100), or many (>100) cysts. of the 21 positive moose, 15 had a few cysts (14 adult females, 1 male calf) and 6 had many cysts (4 adult females, 2 unknown). in some lungs, 100s of cysts were present. the subpleural and interstitial cysts were pale white to yellow, firm to the touch, and ruptured easily during handling. when observed as a fresh mount, the cytology of the cyst contents was consistent with echinococcus granulosus. fresh and fixed cytology samples showed clearly defined protoscolices, numerous calcareous granules, a thin cyst wall, and granular cyst fluid debris (fig. 2, 3). pcr amplification and dna sequencing of the nad1 locus identified cyst fluid samples as the g8 strain fig. 1. map of wildlife management districts (wmd) and townships in the northern maine study area. areas from which unaffected moose lungs were collected are shaded grey; areas from which moose lungs were positive for echinococcus-type cysts are shaded black. the number of lung sets sampled and % positive samples are listed under each wmd (n; %). alces vol. 50, 2014 lichtenwalner et al. – e. granulosus g8 in maine 29 of e. granulosus and were most closely related to the g8 genotype identified in a moose from minnesota (genbank™ accession number ab235848.1, nakao et al. 2007). sequences from moose in the current study were submitted to the ncbi genbank™ database (accession nos: kc839819, kc839820 and kc839821). discussion this is the first report of echinococcus granulosus in maine, and is unusual since this parasite was not thought as endemic to the northeastern united states due to the current absence of wolves (http:// www.fws.gov/midwest/wolf/aboutwolves/ wolfpopus.htm). in a 1941 evaluation of 20 ill moose in maine, e. granulosus was not found (sweatman 1952). the sylvatic form of e. granulosus is present across most of canada, where it has been reported at prevalences as high as 67% in moose of northern ontario (addison et al. 1979). in a summary report concerning e. granulosus in north america, reported sylvatic cases (of wolves, moose, caribou, mule deer and coyotes) were reported in alaska, canada, northern minnesota, and northern california, but not in maine (eckert et al. 2001). further, e. granulosus was not reported in maine moose by state wildlife biologists working during the last 20 years (l. kantar and k. morris (retired), ifw; h. gibbs, university of maine; pers. comm.). in the current study, only a small percentage (1–5% of the total moose killed during 2012 by wmd) of moose was sampled, but rates of 11–67% were detected, suggesting that e. granulosus may be widely distributed in the maine moose population. the pathology of e. granulosus in the moose host is unclear. in most lungs only a few cysts were found, and cysts appeared to be surrounded by well-aerated lung tissue. despite 100s of cysts in specific lungs, the lung tissues appeared relatively normal. in general, these lung cysts did not appear to be highly pathogenic in the intermediate host. certainly large numbers of cysts might be expected to reduce lung capacity and possibly impede mobility and escape from predators. as in this study, a larger and more quantitative study of e. granulosus in canadian moose suggested that abnormally high cyst numbers occur in a small percentage of fig. 2. fresh mount of aspirate from representative lung cyst. the unstained, unfixed material was imaged at 400x total magnification. protoscolices with calcareous granules can be seen within the cyst fluid. fig. 3. cyst aspirate material after formalin fixation, embedding and staining (hematoxylin and eosin). the protoscolices (a) are contained within a brood capsule (b). photo captured at 100x total magnification. 30 e. granulosus g8 in maine – lichtenwalner et al. alces vol. 50, 2014 http://www.fws.gov/midwest/wolf/aboutwolves/wolfpopus.htm http://www.fws.gov/midwest/wolf/aboutwolves/wolfpopus.htm http://www.fws.gov/midwest/wolf/aboutwolves/wolfpopus.htm affected moose (messier et al. 1989). although cysts occur in other internal organs (addison et al. 1979, eckert et al. 2001), we only investigated lungs in this study. both g8 and g10 are found in north american wildlife (thompson et al. 2006, schurer et al. 2013), and g8 was identified in a road-killed moose in minnesota (bowles et al. 1994). recently, wild elk and mule deer in idaho and montana have been identified with e. granulosus cysts. examination of grey wolves in these regions revealed the adult worm in their intestinal tracts (foreyt et al. 2009); however, the genotype of the ungulate cysts was not reported. it is not established that this report represents a true range expansion of e. granulosus, or if it has been simply undetected previoulsy. identifying e. granulosus g8 in maine moose suggests that either wild coyotes or possibly domestic dogs served as the definitive host for this parasite, as wolves are not present in maine. coyotes could serve as sylvatic definitive hosts in regions of the northeastern usa and canada where wolves have previously been extirpated or were historically absent (sweatman 1952), thus enabling the local life cycle. the parasite could have been introduced by wildlife translocations or anthropogenic movement of infected domestic dogs, similar to how echinococcus species were introduced in regions of north america and elsewhere (davidson et al. 1992, hoberg et al. 1994, lind et al. 2011, jenkins et al. 2012). surveys of coyotes and other canid populations may be warranted in the wmds of origin to establish the source of infection. the g8 strain of e. granulosus identified in the current study is genetically and biologically distinct from the pastoral biotype associated with genotypes g1-g3 in sheep and buffalo elsewhere in the world. the sylvatic biotype (g8 and g10 strains) has been associated with human disease in north america, but relatively rarely and with mild pathology compared to the pastoral biotype (lamy et al. 1993, himsworth et al. 2010, nakao et al. 2010). for example, the g8 genotype was identified in a 1999 report of cystic hydatid disease in an alaskan woman (castrodale et al. 2002, mcmanus et al. 2002), while the g10 genotype was the most likely identification for a neural hydatid cyst identified in a child in saskatchewan, canada in 2008 (himsworth et al. 2010). human infection occurs due to ingestion of the tapeworm eggs acquired from the feces of the definitive canid host; most people become infected with e. granulosus through cohabitation with, or sharing contaminated environments with infected domestic dogs. moose pose no direct risk of e. granulosus infection to hunters or people in contact with moose carcasses. however, in regions endemic for sylvatic e. granulosus, offal from moose carcasses should be buried or burned to prevent scavenging by wild and domestic canids, and cooked or frozen before feeding to dogs. people should avoid direct contact with carnivore fecal material, and veterinarians should advise regular tapeworm treatment for domestic dogs at risk of exposure to infective e. granulosus cysts through scavenging or diet. we advise further study to assess the distribution, ecology, and overall effect of e. granulosus on maine moose. acknowledgements this work is based upon research supported in part by hatch grant #me0-l-500514-13 from the usda national institute of food and agriculture. it is published and distributed in furtherance of cooperative extension work, acts of congress of may 8 and june 30, 1914, by the university of maine and the u.s. department of agriculture cooperating. cooperative extension and other agencies of the usda provide equal opportunities in programs and alces vol. 50, 2014 lichtenwalner et al. – e. granulosus g8 in maine 31 employment. this is maine agricultural and forestry experiment station publication #3318. references addison, e. m., a. fyvie, and f. j. johnson. 1979. metacestodes of moose, alces alces, of the chapleau crown preserve, ontario. canadian journal of zoology 57: 1619–1623. bowles, j., d. blair, and d. p. mcmanus. 1994. molecular genetic characterisation of the cervid strain (“northern form”) of echinococcus granulosus. parasitology 109: 215–221. ———, and d. p. mcmanus 1993. nadh dehydrogenase 1 gene sequences compared for species and strains of the genus echinococcus. international journal for parasitology 23: 969–972. castrodale, l. j., m. beller, j. f. wilson, p. m. schantz, d. p. mcmanus, l. h. zhang, f. g. fallico, and f. d. sacco. 2002. two atypical cases of cystic echinococcosis (echinococcus granulosus) in alaska, 1999. the american journal of tropical medicine and hygiene 66: 325–327. davidson, w. r., m. j. appel, g. l. doster, o. e. baker, and j. f. brown. 1992. diseases and parasites of red foxes, gray foxes, and coyotes from commercial sources selling to fox-chasing enclosures. journal of wildlife diseases 28: 581–589. eckert, j., b. gottstein, d. heath, and f. j. liu. 2001. prevention of echinococcosis in humans and safety precautions. pages 239–247 in j. eckert, m. a. gemmell, f. x. meslin, and z. s. palowski, editors. who/oie manual on echinococcosis in humans and animals: a public health problem of global concern. world organisation for animal health, paris, france. foreyt, w. j., m. l. drew, m. atkinson, and d. mccauley. 2009. echinococcus granulosus in gray wolves and ungulates in idaho and montana, usa. journal of wildlife diseases 45: 1208–1212. himsworth, c. g., e. jenkins, j. e. hill, m. nsungu, m. ndao, r. c. andrew thompson, c. covacin, a. ash, b. a. wagner, a. mcconnell, f. a. leighton, and s. skinner. 2010. emergence of sylvatic echinococcus granulosus as a parasitic zoonosis of public health concern in an indigenous community in canada. the american journal of tropical medicine and hygiene 82: 643–645. hoberg, e. p., s. miller, and m. a. brown. 1994. echinococcus granulosus (taeniidae) and autochthonous echinococcosis in a north american horse. the journal of parasitology 80: 141–144. jenkins, e. j., a. s. peregrine, j. e. hill, c. m. somers, k. m. gesy, b. barnes, b. gottstein, and l. polley. 2012. detection of a european strain of echinococcus multilocularis in north america. emerging infectious diseases 18: 1010–1012. kantar, l. e., and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose abundance in maine. alces 49: 29–37. lamy, a. l., b. h. cameron, j. g. leblanc, j. a. culham, g. k. blair, and g. p. taylor. 1993. giant hydatid lung cysts in the canadian northwest: outcome of conservative treatment in three children. journal of pediatric surgery 28: 1140–1143. lind, , e. o., m. juremalm, d. christensson, s. widgren, g. hallgren, e. o. agren, h. uhlhorn, a. lindberg, m. cedersmyg, and h. wahlstrom. 2011. first detection of echinococcus multilocularis in sweden, february to march 2011. eurosurveillance 16(14): 1, id 9836. mcmanus, d. p., l. zhang, l. j. castrodale, t. h. le, m. pearson, and d. blair. 2002. short report: molecular genetic characterization of an unusually severe case of hydatid disease in alaska caused by the cervid strain of echinococcus granulosus. the american journal of 32 e. granulosus g8 in maine – lichtenwalner et al. alces vol. 50, 2014 tropical medicine and hygiene 67: 296–298. messier, f., m. rau, and m. mcneill. 1989. echinococcus granulosus (cestoda: taeniidae) infections and moose-wolf population dynamics in southwestern quebec. canadian journal of zoology 67: 216–219. nakao, m., d. p. mcmanus, p. m. schantz, p. s. craig, and a. ito. 2007. a molecular phylogeny of the genus echinococcus inferred from complete mitochondrial genomes. parasitology 134: 713–722. ———, t.yanagida, m. okamoto, j. knapp, a. nkouawa, y. sako, and a. ito. 2010. state-of-the-art echinococcus and taenia: phylogenetic taxonomy of humanpathogenic tapeworms and its application to molecular diagnosis. journal of molecular epidemiology and evolutionary genetics in infectious diseases 10: 444–452. schurer, j., t. shury, f. leighton, and e. jenkins. 2013. surveillance for echinococcus canadensis genotypes in canadian ungulates. international journal for parasitology: parasites and wildlife 2: 97–101. sweatman, g. k. 1952. distribution and incidence of echinococcus granulosus in man and other animals with special reference to canada. canadian journal of public health 43: 480–486. thompson, r., a. c. boxell, b. j. ralston, c. c. constantine, r. p. hobbs, t. shury, and m. e. olson. 2006. molecular and morphological characterization of echinococcus in cervids from north america. parasitology 132: 439–447. alces vol. 50, 2014 lichtenwalner et al. – e. granulosus g8 in maine 33 echinococcus granulosus genotype g8 in maine moose (alces alces) study area & methods results discussion acknowledgements references alces28_1.pdf alces27_31.pdf alces28_41.pdf alces(23)_311bobcornellhonoured.pdf alces vol. 23, 1987 alces24_1.pdf alces27_85.pdf alces21_55.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces(25)_11.pdf alces29_187.pdf alces27_127.pdf alces28_111.pdf alces21_127.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces28_215.pdf alces21_279.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces26_24.pdf alces24_126.pdf mass, morphology, and growth rates of moose in north dakota william f. jensen1, jason r. smith2, james j. maskey jr.3, james v. mckenzie (deceased)1, and roger e. johnson (retired)4 1north dakota game and fish department, 100 north bismarck expressway, bismarck, north dakota 58501 usa; 2north dakota game and fish department, 3320 east lakeside road, jamestown, north dakota 58401 usa; 3university of mary, 7500 university dr., bismarck, north dakota 58504 usa; 4north dakota game and fish department, 7928 45th street ne, devils lake, north dakota 58301 usa. abstract: this paper provides predictive formulas to estimate live weights of moose (alces alces andersoni) from hunter-harvested animals and evaluate growth rates of moose in north dakota, and reviews weight-related measurements among moose populations. from 1978–1990, morphometric data were collected on 224 hunter-killed moose harvested after the rut (10 november–12 december) in north dakota. body mass increased rapidly for both sexes from 0.5 years to 1.5 years-of-age. whole weight and total body length reached an asymptote for both sexes by 5.5 years; mean whole weight appeared to decline among older individuals. although field dressed weight was the best predictor of whole weight (r2 = 0.93; n = 154), total body length provided reasonably good estimates of whole weight (r2 = 0.76; n = 153). whole weight estimates based upon shoulder height (r2 = 0.33; n = 158) and hind-foot length (r2 = 0.46; n = 163) were less reliable. we also used morphometric variables to predict field dressed weight, carcass weight, and visceral weight. field dressed weight was the best predictor of antler width (r2 = 0.72; n = 108) and antler width was a good predictor of male age (r2 = 0.70; n = 119). when compared to other north american populations, average weights of moose harvested in north dakota tended to be higher in all age classes. additionally, the calf-to-yearling growth rate of female moose in north dakota was as high, or higher than in other populations. morphometric comparisons of free-ranging moose from various north american populations had much size overlap, with southern and eastern moose populations tending to have largest average adult body mass. sexual dimorphism of mature north dakota moose (> 4.5 years) was comparable to that in other populations. alces vol. 49: 1–15 (2013) key words: alces alces, body mass, morphometrics, sex, age, growth rates, north dakota. body weights and measurements of cervids provide insight on condition and health of local populations (clutton-brock et al. 1982, sauer 1984, verme and ullrey 1984, loudon 1987). they allow for the analysis of energetic requirements, energetic capability, and other metabolic parameters (schwartz et al. 1987), and how subspecies vary in size, shape, and rate of growth (bubenik 1998, geist 1998). comparative morphometric data on north american moose (alces alces) are limited (blood et al. 1967, timmerman 1972, schladweiler and stevens 1973, peterson 1974, franzmann et al. 1978, crichton 1980, adams and pekins 1995, lynch et al. 1995, broadfoot et al. 1996). additionally, measurement and definition of morphometric variables vary among studies making comparisons among populations problematic. difficulties in accessing hunter-killed moose, and the physical labor involved in handling these 1 animals have undoubtedly limited data collection. historically, moose in north dakota were restricted to the heavily forested areas of the turtle mountains, pembina hills, and the major tributaries of the red river. accounts of early traders in the area indicated that they were not as abundant as elk (cervus elaphus) or other big game species, and apparently disappeared from the state during the early 20th century (knue 1991). by the 1960s, moose had returned to north dakota and small numbers were occupying portions of their historic range. in 1977, the first modern moose hunting season allowed the harvest of 10 moose in cavalier, pembina, and walsh counties. the expansion of moose into the relatively accessible farmland of north dakota, coupled with the willingness of local farmers to assist hunters with loading and transporting animals with farm equipment, made it feasible for the north dakota game and fish department (ndgfd) to collect morphometric data. we provide analyses of 1) age and sex-specific weights and measurements, 2) measurement-weight relationships, 3) age and morphometric relationships, and 4) growth rates of hunter-killed moose from north dakota. our goal is to provide predictive formulas for estimating whole, field-dressed, carcass, and viscera weights of hunter-killed moose and to make comparisons with other north american and european populations. methods morphometric data were collected on 224 hunter-killed moose examined between 1978 and 1990. hunters were asked whenever possible to bring their moose in whole to a check station (i.e., prior to removal of viscera, hide, head, or legs). all moose were harvested between 10 november and 12 december after the rutting season; date of kill, sex, and legal descriptions (section, township, range) for all kill sites were recorded. the distribution of the animals examined was between 47.10 and 48.99° n and 97.14 and 100.41° w (fig. 1). moose hunting units m1c and m4 are comprised primarily of aspen (populus spp.) forests with intermingled cropland; the remainder of the harvest area was drift prairie with heavy conversion to cropland (fig. 1). for a more complete description of habitat see maskey (2008). weight was measured with local grain elevator scales with an accuracy of ±4.5 kg. recorded weights were whole weight (ww) which comprised a completely intact carcass except for loss of blood and tissue resulting from gunshot wounds; field dressed weight (fdw) which included carcass weight minus all thoracic and abdominal viscera; viscera weight (vw) which included all thoracic and abdominal organs and their contents including blood and the contents of the digestive system; and carcass weight (cw) which comprised the dressed carcass minus the head, hide, and legs below the hock joint. length was measured (nearest cm) before the moose was dressed-out following the methods of peterson (1974). with the carcass laid flat on its side, with head and spinal column supported on the same plane, total body length (tbl) was measured from tip of nose to tip of tail (point of last coccyx bone, excluding hair) by following the dorsal (spinous) processes of the vertebra. with the carcass laid flat on its side and the front leg positioned so that it was straight and perpendicular to the longitudinal axis of the body, hind-foot length (hfl) was measured from the calcaneum (heel bone of hock) to the tip of hoof, and shoulder height (sh) from the superior angle of the scapula (cartilaginous top of shoulder) to the distal tip of the front hoof. antler width (aw) was measured at the greatest spread between the tines and at a right angle to the longitudinal axis of the skull. prior to additional analyses, we 2 morphology of north dakota moose – jensen et al. alces vol. 49, 2013 used t-tests to determine whether any morphometric measurements differed between male and female moose of all ages. whole weight, fdw, cw, vw, and aw were compared to other morphometric measurements using simple linear regressions. all statistical analyses were conducted using r, version 2.11.1 (r development core team 2010). incisor eruption (peterson 1955) was used to identify young-of-the-year or calves (≤6 month-of-age) and front incisors were collected from moose ≥6 months of age for cementum annuli analysis (gasaway et al. 1978, haagenrud 1978). ageing was performed by matson's laboratory, milltown, mt, and by co-author r. johnson (ndgfd) after 1980. the relationship between each morphometric variable and age was examined with linear regression using square-roottransformed age as the dependent variable; to facilitate utility, we present backtransformed equations in results. growth rates for moose have been described as falling into 3 phases: 1) a selfaccelerated phase (schwartz 1998) of near exponential growth from birth to weaning (4–6 months old) allowing the calf to follow its mother over rough terrain and obstacles (geist 1998), 2) a second phase of rapid growth from calf to yearling (16–18 months old) allowing young moose to reach a body size (250–280 kg ww) that allows yearlings to confront predators (geist 1998), and 3) a self-inhibiting growth phase where seasonal peaks and troughs in body mass occur at different times for males and females (schwartz et al. 1987). results and discussion focus on these 3 growth phases. healthy north american moose calves have a mean birth weight ranging from fig. 1. locations of 224 moose harvested in north dakota, usa (1978–1990). each dot represents the site, to the nearest section, where ≥1 moose were harvested. alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 3 12.6–18 kg for single calves (kellum 1941 in peterson 1955, franzmann et al. 1980, schwartz 1998) and 13.6 kg for twins (franzmann 1978). the mean weight for 1–3 day-old alaskan calves was 18.0 kg (n = 109; franzmann et al. 1980). average weights of captured calves < 2 weeks old in ontario averaged 15.7 kg (n = 8) for females and 17.3 kg (n = 10) for males (addison et al. 1994). the average weight of 43 captured neonate calves in alberta was 19.6 kg; however, some were captured as late as august (welch et al. 1985). lacking information on birth weights for north dakota moose, we used the range of 13–18 kg as the basis for calculating phase 1 growth rates. we determined percent change of growth rate during phase 1 by dividing calf weight at 0.5 years by the neonate weights of 13 and 18 kg, and we estimated the percent change during phase 2 by dividing the average yearling (1.5 years) weight by the average calf weight. we measured the level of sexual dimorphism for moose ≥4.5 years old. only these moose were included because they would be at or near their maximum size and this age range would best permit comparison to other studies. we calculated dimorphic ratios (male:female) for ww, fdw, cw, vw, tbl, hfl, and sh. we also used t-tests to examine whether these measures differed significantly between males and females ≥4.5 years old. results morphometric measurements sample sizes for ww, fdw, vw, and cw were obtained from 160, 166, 146, and 40 moose, respectively (tables 1 and 2). sample sizes for tbl, hfl, sh, and aw were obtained from 206, 196, 200, and 121 moose, respectively (tables 3 and 4). antler width appeared to plateau at 6.5 years and then decline in older males (table 4). morphometric relationships morphometric measurements were not significantly different for males and females of all ages. therefore, we combined sexes when conducting regression analyses for these variables. whole weight was best predicted by fdw (r2 = 0.93, n = 154), followed by tbl (r2 = 0.76, n = 153); fdw was best predicted by tbl (r2 = 0.70, n = 181) and cw (r2 = 0.65, n = 39). antler width was most highly correlated with fdw (r2 = 0.72, n = 108). all regression equations for predicting ww, fdw, cw, vw, and aw are provided in table 5. age was best predicted by ww (r2 = 0.71, n = 114), followed by fdw (r2 = 0.70, n = 166; table 6). antler width was also a reasonable estimator of male age (r2 = 0.70, n = 119; table 6). growth rates and patterns during phase 1, both post-rut male and female calves averaged 196 kg at about 7 months of age, representing a 989–1455% increase in body mass. the average growth rate would be 1% per day assuming a mean age of about 180 days, given a 1 december harvest date. the fdws of female and male calves were 61 and 72% of ww; fdw of yearling females and males were 73 and 70%, respectively. during growth phase 2, female and male calves (averaging 196 kg at 6–7 months) increased their ww over the next year by 69 and 65%, respectively. the fdw during phase 2 increased by 102 and 59.3% for females (n = 10) and males (n = 32), respectively (table 7). the fdws of female and male calves were 61 and 72% of ww; fdw of yearling females and males were 73 and 70%, respectively. whole weights, fdw, and tbl plateaued at 5.5 years for both males and females during the self-inhibiting growth phase (tables 1–4). 4 morphology of north dakota moose – jensen et al. alces vol. 49, 2013 table 1. age-weight relationship for female moose harvested in north dakota, usa (1978–1990). whole weight (kg) field dressed weight (kg) viscera weight (kg) carcass weight (kg) age n mean (± sd) range n mean (± sd) range n mean (± sd) range n mean (± sd) range 0.5 years 5 196.0 ± 21.6 176.9–226.8 4 119.1 ± 30.2 77.1–145.1 4 73.3 ± 21.1 54.4–99.8 2 127.7 ± 54.5 86.2–163.3 1.5 years 12 331.1 ± 44.1 272.2–444.5 10 241.1 ± 26.9 204.1–290.3 10 94.6 ± 25.1 68.0–154.2 2 195.0 ± 19.2 181.4–208.7 2.5 years 5 366.5 ± 20.9 331.1–385.6 6 255.8 ± 12.5 240.4–272.2 5 109.5 ± 17.0 84.8–127.0 4 197.3 ± 14.7 179.2–211.4 3.5 years 11 410.3 ± 50.9 349.3–526.2 12 285.4 ± 36.5 229.1–358.3 10 158.1 ± 27.6 95.3–172.4 1 281.2 281.2 4.5 years 2 437.7 ± 22.5 421.8–453.6 3 323.6 ± 42.1 294.8–371.9 2 138.3 ± 16.0 127.0–149.7 1 231.3 231.3 5.5 years 5 467.2 ± 39.9 417.3–512.6 5 330.7 ± 29.8 292.6–371.9 5 136.5 ± 30.5 90.7–167.8 1 220.0 220 6.5 years 2 437.7 ± 60.9 394.6–480.8 3 310.0 ± 15.9 294.8–326.6 2 127.0 ± 38.5 99.8–154.2 1 226.8 226.8 7.5 years 2 435.4 ± 25.7 417.3–453.6 2 310.7 ± 22.5 294.8–326.6 2 124.7 ± 3.2 122.5–127.0 1 239.5 239.5 8.5 years 3 444.5 ± 64.0 385.6–512.6 3 319.0 ± 41.2 281.2–362.9 3 125.5 ± 22.8 104.3–149.7 0 10.5 years 1 489.9 489.9 1 335.7 335.7 1 154.2 154.2 0 ≥ 1.5 years 43 397.7 ± 64.9 272.2–526.2 45 285.3 ± 43.2 204.1–371.9 40 127.5 ± 59.2 68.0–172.4 11 216.2 ± 29.1 179.2–281.2 table 2. weight categories (see methods) by age for male moose harvested in north dakota, usa (1978–1990). whole weight (kg) field dressed weight (kg) viscera weight (kg) carcass weight (kg) age n mean (± sd) range n mean (± sd) range n mean (± sd) range n mean (± sd) range 0.5 years 5 196.0 ± 34.3 140.6–226.8 4 141.7 ± 5.7 136.1–149.7 4 68.0 ± 12.3 49.9–77.1 0 1.5 years 27 323.6 ± 31.4 249.5–403.7 32 225.7 ± 24.5 181.4–290.3 27 96.3 ± 18.5 63.5–163.3 6 177.9 ± 24.4 147.4–207.3 2.5 years 35 402.3 ± 43.3 290.3–485.3 35 292.2 ± 26.7 217.7–349.3 30 119.1 ± 18.9 68.0–158.8 8 219.5 ± 37.9 158.8–281.2 3.5 years 22 446.3 ± 40.4 335.7–503.5 22 322.5 ± 30.1 254.0–367.4 20 121.8 ± 17.9 77.1–154.2 5 236.9 ± 32.2 195.0–270.8 4.5 years 8 444.1 ± 30.1 408.2–499.0 9 320.0 ± 36.6 272.2–381.0 8 123.6 ± 201 90.7–154.2 2 246.5 ± 4.8 243.1–249.9 5.5 years 7 496.6 ± 60.1 408.2–589.7 7 365.5 ± 22.0 340.2–408.2 6 141.7 ± 34.6 95.3–181.4 4 257.3 ± 26.3 235.0–295.3 6.5 years 7 479.5 ± 51.4 403.7–535.2 6 352.3 ± 33.5 303.9–394.6 6 121.0 ± 27.8 99.8–172.4 2 318.2 ± 88.8 255.4–381.0 7.5 years 0 1 349.3 349.3 0 0 9.5 years 1 453.6 453.6 1 335.7 335.7 1 117.9 117.9 0 ≥ 1.5 years 107 430.6 ± 69.2 249.5–589.7 113 311.7 ± 53.3 181.4–408.2 98 121.0 ± 24.1 63.5–181.4 27 245.5 ± 62.4 158.8–381.0 a l c e s v o l . 4 9 , 2 0 1 3 je n s e n e t a l . – m o r p h o l o g y o f n o r t h d a k o t a m o o s e 5 table 3. morphological measurements of female moose harvested in north dakota, usa (1978–1990). total body length (cm) hind foot length (cm) shoulder height (cm) age n mean (± sd) range n mean (± sd) range n mean (± sd) range 0.5 years 5 203.8 ± 12.2 190.0–219.7 5 71.0 ± 2.9 68.0–76.2 4 151.8 ± 9.7 143.0–164.0 1.5 years 13 248.5 ± 10.3 236.5–265.0 13 77.0 ± 2.6 71.1–81.9 13 175.5 ± 6.8 157.5–185.2 2.5 years 8 248.2 ± 10.4 232.0–264.7 6 77.9 ± 2.2 74.8–81.0 8 181.2 ± 4.7 177.3–184.9 3.5 years 15 257.5 ± 17.5 214.0–279.2 14 79.2 ± 3.4 74.1–86.1 15 188.2 ± 5.2 179.8–197.3 4.5 years 3 271.2 ± 7.7 266.1-280.0 3 79.0 ± 0.6 78.3–79.4 3 190.6 ± 6.7 182.9-194.7 5.5 years 5 273.2 ± 10.6 259.7–286.1 5 79.4 ± 2.4 77.4–82.9 5 193.4 ± 9.9 185.0–206.8 6.5 years 3 267.3 ± 5.0 263.6–272.9 3 80.7 ± 1.5 79.2–82.1 3 192.8 ± 0.4 192.4–193.1 7.5 years 2 263.1 ± 4.0 260.2–265.9 2 81.7 ± 1.2 80.8–82.5 2 190.9 ± 9.7 184.0–197.7 8.5 years 3 267.4 ± 12.1 255.2 279.4 3 77.9 ± 2.1 75.9–80.0 3 192.3 ± 7.2 188.4–200.6 10.5 years 1 261.1 261.1 1 80.3 80.3 1 189.1 189.1 ≥ 1.5 years 53 257.5 ± 14.8 214.0–286.1 50 78.6 ± 2.8 71.1–86.1 53 157.5 ± 206.8 157.5–206.8 table 4. morphological measurements of male moose harvested in north dakota, usa (1978–1990). total body length (cm) hind foot length (cm) shoulder height (cm) antler width (cm) age n mean (± sd) range n mean (± sd) range n mean (± sd) range n mean (± sd) range 0.5 years 6 201.4 ± 14.0 176.0–215.0 6 70.6 ± 2.5 66.0–73.5 6 153.0 ± 8.0 141.0–161.5 1 23.5 23.5 1.5 years 40 244.5 ± 11.9 218.0–272.3 38 77.7 ± 2.6 72.2–82.1 38 179.7 ± 7.1 167.5–194.5 28 69.9 ± 10.7 53.3–92.3 2.5 years 48 258.0 ± 12.9 227.0–288.3 46 80.0 ± 3.2 68.9–85.5 46 187.9 ± 9.4 162.0–203.5 40 89.1 ± 11.5 66.4–125.1 3.5 years 26 267.7 ± 13.9 240.0–295.3 26 81.0 ± 2.9 75.0–86.6 26 191.5 ± 8.4 176.0–206.3 26 100.1 ± 9.5 84.4–129.9 4.5 years 9 269.9 ± 9.4 257.5–282.8 7 80.1 ± 2.8 76.2–84.5 9 188.8 ± 6.3 181.0–200.5 9 107.1 ± 14.9 81.3–128.9 5.5 years 9 277.7 ± 12.9 258.0–295.9 9 82.7 ± 1.5 79.1–84.5 9 198.8 ± 6.4 189.0–205.8 9 122.8 ± 8.5 104.8–134.6 6.5 years 7 272.6 ± 12.8 255.9–289.4 7 79.2 ± 2.9 76.4–84.9 7 193.6 ± 13.1 178.5–219.7 7 128.6 ± 23.7 99.8–168.3 7.5 years 1 273.0 273.0 0 0 1 125.7 125.7 8.5 years 1 270.9 270.9 1 82.3 82.3 1 206.4 206.4 1 109.8 109.8 9.5 years 1 274.0 274.0 1 80.0 80.0 1 207.8 207.8 0 ≥ 1.5 years 142 259.0 ± 16.3 218.0–295.9 135 79.7 ± 3.1 68.9–86.6 137 187.7 ± 10.2 162.0–219.7 121 93.6 ± 21.0 53.3–168.3 6 m o r p h o l o g y o f n o r t h d a k o t a m o o s e – je n s e n e t a l . a l c e s v o l . 4 9 , 2 0 1 3 sexual dimorphism moose ≥4.5 years old had a sexual dimorphism ratio of 1.04 for ww (x = 485.8 kg for males [n = 15] and 452.1 kg for females [n = 15]), and 1.07 for fdw (x = 357.2 kg for males [n = 15] and 321.7 kg for females [n = 17]; table 8). these ratios were mid-range in comparison with other studies where weight dimorphism for ww and fdw ranged from 0.90–1.19 and 0.70–1.36, respectively (table 9). although measurements for male moose were larger for all but table 5. simple regression equations for weight-measurement relationships among moose from north dakota, usa (1978–1990). comparison n equation r2 whole wt. (ww) (kg) vs. field dressed wt.(fdw) (kg) 154 ww = (fdw×1.28) + 35.5 0.93 whole wt. (ww) (kg) vs. carcass wt. (cw) (kg) 39 ww = (cw×1.47) + 67.3 0.75 whole wt. (ww) (kg) vs. viscera wt. (vw) (kg) 156 ww = (vw×2.41) + 119.8 0.65 whole wt. (ww) (kg) vs. total body length (tbl) (cm) 153 ww = (tbl×0.36) − 520.2 0.76 whole wt. (ww) (kg) vs. shoulder height (sh) (cm) 158 ww = (sh×0.24) − 49.1 0.33 whole wt. (ww) (kg) vs. hind-foot length (hfl) (cm) 163 ww = (hfl×1.51) − 802.8 0.46 field-dressed wt. (fdw) (kg) vs. carcass wt. (cw) (kg) 39 fdw = (cw×0.84) + 104.4 0.65 field-dressed wt. (fdw) (kg) vs. viscera wt. (vw) (kg) 146 fdw = (vw×1.41) + 119.7 0.40 field-dressed wt. (fdw) (kg) vs. total body length (tbl) (cm) 181 fdw = (tbl×0.26) − 387.8 0.70 field-dressed wt. (fdw) (kg) vs. shoulder height (sh) (cm) 165 fdw = (sh×0.17) − 41.0 0.31 field-dressed wt. (fdw) (kg) vs. hind-foot length (hfl) (cm) 174 fdw = (hfl×0.11) + 196.4 0.03 carcass wt. (cw) (kg) vs. viscera wt. (vw) (kg) 32 cw = (vw×1.14) + 93.6 0.56 carcass wt. (cw) (kg) vs. total body length (tbl) (cm) 47 cw = (tbl×0.18) − 236.3 0.33 carcass wt. (cw) (kg) vs. shoulder height (sh) (cm) 40 cw = (sh×0.24) + 231.0 0.37 carcass wt. (cw) (kg) vs. hind-foot length (hfl) (cm) 46 cw = (hfl×0.013) + 207.8 0.001 viscera wt. (vw) (kg) vs. total body length (tbl) (cm) 156 vw = (tbl×0.09) + 0.011 0.40 viscera wt. (vw) (kg) vs. shoulder height (sh) (cm) 147 vw = (sh×0.05) + 22.7 0.14 viscera wt. (vw) (kg) vs. hind-foot length (hfl) (cm) 156 vw = (hfl×0.36) − 173.3 0.22 antler width (aw) (cm) vs. whole wt. (ww) (kg) 97 aw = (ww×2.57) − 137.2 0.65 antler width (aw) (cm) vs. field-dressed wt. (ww) (kg) 108 aw = (fdw×3.46) − 102.0 0.72 antler width (aw) (cm) vs. total body length (tbl) (cm) 129 aw = (tbl×0.81) − 1183.0 0.37 antler width (aw) (cm) vs. shoulder height (sh) (cm) 116 aw = (sh×0.39) + 198.8 0.13 antler width (aw) (cm) vs. hind-foot length (hfl) (cm) 125 aw = (hfl×.024) + 901.0 < 0.001 table 6. regression equations for age-measurement relationships among moose from north dakota, usa (1978–1990). comparison n equations r2 age vs. whole weight (ww) (kg) 114 age = ((ww×0.0043) – .066)2 0.71 age vs. field dressed wt. (fdw) (kg) 166 age = ((fdw×0.06) + 0.022)2 0.70 age vs. carcass wt. (cw) (kg) 39 age = ((cw×0.007) + 0.098)2 0.53 age vs. viscera wt. (vw) (kg) 145 age = ((vw×0.01) + 0.51)2 0.39 age vs. total body length (tbl) (cm) 203 age = ((tbl×0.0014) – 0.21)2 0.53 age vs. shoulder height (sh) (cm) 197 age = ((sh×0.0011) – 0.47)2 0.26 age vs. hind-foot length (hfl) (cm) 153 age = ((hfl×0.00035) + 1.31)2 0.014 age vs. antler width (aw) (cm) 119 age = ((aw×0.0016) + 0.26)2 0.70 alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 7 table 7. comparisons of mean field dressed weight (fdw) of calves and percent change from calf (0.5 years) to yearling (1.5 years) age classes from 7 north american and 4 northern european moose populations. note: broadfoot et al. (1996) used moose of 11 months of age. females males location subspecies calf fdw (kg) [n] yearling fdw (kg) [n] change calf fdw (kg) [n] yearling fdw (kg) [n] change source north dakota a. a. andersoni 119.1 [4] 241.1 [10] 102.4% 141.7 [4] 225.7 [32] 59.3% this study vermont, new hampshire, and maine a. a. andersoni 108 [23] 216 [65] 100 % 112 [23] 199 [139] 77.7% adams and pekins (1995) quebec a. a. americana 108.6 [26] 192.6 [11] 77.3% 119.6 [19] 199.3 [24] 66.6% heyland (1964) in peterson (1974) quebec 108.6 [26] 203.4 [34] 87.3% 119.6 [19] 204.3 [51] 70.8% heyland (1966) in peterson (1974) ontario a. a. andersoni 156.5 [3] 230.9 [7] 47.5% 140.2 [7] 254.5 [19] 81.2% timmerman (1972) ontario a. a. andersoni 75.3 [1] 220.9 [2] 193.4% 115.2 [1] 240.4 [1] 108.7% simkin (1962) ontario a. a. andersoni 136.3 [3] 228.9 [5] broadfoot et al. (1996) manitoba a. a. andersoni 163.3 [1] 205.8 [4] crichton (1980), and pers. comm. alberta a. a. andersoni 93.9 [27] 161.9 [28] 72.4% 95.3 [21] 152.9 [34] 60.4% blood et al. (1967) montana a. a. shirasi 84.4 [14] 163.3 [15] 93.5% 99.3 [14] 170.6 [28] 71.8% schladweiler and stevens (1972) norway, southern a. a. alces 69.8 [74] 140.2 [115] 100.9% saether et al. (1996) norway, interior a. a. alces 68.5 [298] 139.8 [210] 104.1% saether et al. (1996) norway, alpine a. a. alces 63.2 [625] 125.1 [370] 97.9% saether et al. (1996) norway, northern a. a. alces 72.9 [7] 146.0 [122] 100.3% saether et al. (1996) 8 m o r p h o l o g y o f n o r t h d a k o t a m o o s e – je n s e n e t a l . a l c e s v o l . 4 9 , 2 0 1 3 vw and hfl, only fdw and cw differed significantly between sexes (table 8). discussion general observations field dressed weights were the best estimator of ww, although reliable estimates of ww were also obtained using tbl and sh (table 5). for all sexes the best estimate of age was ww followed by fdw, and aw was also a good estimator of male age (table 6). therefore, we suggest that collection of baseline data in local populations focus on fdw, tbl, aw, and incisor collection. although age can be reasonably predicted by aw for males, and ww and fdw for moose of both sexes, cementum annuli aging remains the most accurate method for determining age. further, cw and tbl appear to be fairly good predictors of fdw, while other measurements have limited use in estimating fdw (table 5). carcass weight and vw were poorly predicted by morphometric measurements (table 5). our results also indicate that ww of both sexes can range about ±20% of the mean within an age class. schwartz et al. (1987) reported overwinter weight loss can range from as little as 7% to as high as 23% of pre-rut body mass; thus, weights of individual moose in fallearly winter that are >20% below the local population average of an age class may indicate nutritional stress or other health concerns such as parasite infection. several authors have estimated the relationship between fdw and ww with varying results. in our study the slope of the regression between fdw and ww was 1.28 for both sexes combined (table 5) which was similar to that in studies with a combined relationship for both sexes. for example, peterson (1974) and crichton (1980) calculated slopes of 1.28 and 1.31 for both sexes. other authors have reported varying results for the relationship between ww and fdw. blood et al. (1967) estimated an overall carcass yield of 50% (1.50) for all ages, and similarly, schladweiler and stevens (1973) estimated ww:fdw ratios of 1.43 for adult females and 1.33 for adult males, and broadfoot et al. (1996) estimated 1.51 for females and 1.48 for males for 11-month old captive moose. other researchers have developed estimators of ww based on a variety of morphometric measurements. franzmann et al. (1978) developed equations for estimating ww from tbl, chest girth, hfl, and sh, and wallin et al. (1996) focused on chest circumference to estimate carcass body mass of moose in sweden after removal of head, skin, lower legs, kidneys, and viscera. because the methods and outcome of analyses have varied substantially among studies, we suggest that when comparing fdw and ww between populations, eliminate as many biases as possible (e.g., sex, age, and table 8. sexual dimorphism for 7 morphometric measurements and results of t-tests comparing these measurements between sexes for moose ≥4.5 years old from north dakota, usa (1978–1990). measurement sexual dimorphism t degrees of freedom p whole weight (ww) 1.04 1.3 34 0.21 field-dressed weight (fdw) 1.07 2.2 38 0.03 carcass weight (cw) 1.27 2.4 8 0.04 viscera weight (vw) 0.93 1.1 28 0.27 total body length (tbl) 1.02 1.5 40 0.13 hind-foot length (hfl) 1.0 0.11 29 0.91 shoulder height (sh) 1.16 1.5 28 0.14 alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 9 table 9. comparisons of sexual dimorphism based on mean field dressed weight (fdw) and mean whole weight (ww) for adult females and males (≥4.5 years) from various moose populations in north america. * note: some references did not permit calculating weight using these age criteria. therefore, weights for quinn and aho (1989) were for mature animals aged as wear class vi (>6.5 years), schladweiler and stevens (1973) were for mature animals aged as wear class v (>5.5 years), blood et al. (1967) and franzmann et al. (1978) includes animals >3.5 years, geist (1998) included animals >4.5 years, and franzmann et al. (1987) were 5 year-old captive animals. females males females males location subspecies fdw (kg) [n] fdw (kg) [n] fdw sexual dimorphism ww (kg) [n] ww (kg) [n] ww sexual dimorphism source north dakota a. a. andersoni 321.7 [17] 343.2 [24] 1.07 452.1 [15] 471.3 [23] 1.04 this study quebec a. a. americana 267.3 [156] 348.0 [224] 1.30 heyland (1964, 1966) in peterson (1974) vermont, new hampshire, and maine a. a. americana 261.9 [76] 342.8 [251] 1.31 adams and pekins (1995) ontario a. a. andersoni 301 [5] 408 [10] 1.36 timmerman (1972) ontario a. a. andersoni 285.0 [3] 360.0 [2] 1.26 simkin (1962) ontario a. a. andersoni 461.0 [17] 496.0 [9] 1.07* quinn and aho (1989) manitoba a. a. andersoni 281.2 [7] 308.4 [13] 1.10 400.7 [3] 461.7 [8] 1.15 crichton (1980) alberta a. a. andersoni 205.0 [55] 220.0 [39] 1.07* 417.0 [6] 413.2 [3] 0.99* blood et al. (1967) alberta a. a. andersoni 413.0 [32] 456.0 [30] 1.10* geist (1998) alberta a. a. andersoni 412.4 [32] 461.1 [23] 1.12 lynch et al. (1995) montana a. a. shirasi 214.3 [29] 269.0 [35] 1.25* schaldweiler and stevens (1973) alaska a. a. gigas 400.5 [66] 454.6 [5] 1.14* franzmann et al. (1978) alaska a. a. gigas 499.0 [3] 594.0 [3] 1.19* franzmann et al. (1978) 1 0 m o r p h o l o g y o f n o r t h d a k o t a m o o s e – je n s e n e t a l . a l c e s v o l . 4 9 , 2 0 1 3 conversions to ww from fdw and morphometric measurements), and make only direct comparisons such as fdw and ww. growth rates and weights of north dakota moose our estimates of phase 1 growth rates were similar to that of schwartz (1998) who estimated a phase 1 accelerated growth of 1.3–1.6% per day for calves <165 days old (pre-rut). our rates were higher than those reported by addison et al. (1994) for captive calves with neonatal weights of 15.7 kg (n = 8) and 17.3 kg (n = 10) for females and males, respectively. at 187 days, or approximately 6 months old, their females averaged 149.2 kg (n = 6) and males 165.8 kg (n = 9), representing a 950 and 958% increase in mass, respectively. in short, during phase 1, north american moose display an impressive rate of growth of *1000% within a 6 month period. during growth phase 2, our female calves appeared to grow faster than male calves (tables 1 and 2), particularly when comparing fdw. this was also true for 5 of the 6 north american populations where similar comparisons could be made (table 7), suggesting that a larger proportion of body mass is being devoted to skeletal and muscle development at an earlier age in males. this additional skeletal and muscle mass may provide a selective advantage for male calves as they are often driven from their mother by courting males during the rut and must briefly survive alone. during the third (self-inhibiting) growth phase when body mass goes through seasonal peaks and troughs, the maximum weights for females and males occur midwinter and prerut, respectively (schwartz et al. 1987). body weight of north american female moose appears to plateau at 3–4 years (geist 1998), whereas weight of male moose plateaus at 5 years (peterson 1974). we found that both male and female weights peaked at 5.5 years (tables 1 and 2). while sample sizes were small, tbl also appeared to plateau at 5.5 years in north dakota (tables 3 and 4), which was similar as reported for alaskan moose (franzmann et al. 1978). comparative data are lacking, but these results suggest that north dakota moose with access to highly nutritious agricultural crops may maximize body weight earlier and continue to grow structurally longer than other moose populations. additional observations about moose growth rates in 6 of the 7 populations in which comparisons could be made, female calves weighed ≥5% less than male calves (table 7). the exception (timmermann 1972) where female calves weighted >10% more than males may be an artifact of small sample size (n = 3 female calves). overall, fdw of north american female and male calves averaged 101.4 kg (n = 69), and 112.0 kg (n = 61), respectively (table 7). estimates of ww from fdw using various correction factors are relatively common in the literature. however, actual ww of north american female and male moose calves in the literature are limited to blood et al. (1967), lynch et al. (1995), and this study. the average ww of female and male calves was 174 kg (n = 4) and 197 kg (n = 5) in alberta (blood et al. 1967), and 171 kg (n = 12) and 197 kg (n = 13) in ontario (lynch et al. 1995). we found that ww for female and male calves were identical, but male fdw appears to be higher. in short, the available data suggests that either maternal moose invest more resources in their male calves, or there is a selective advantage for male calves to develop greater muscle mass at an early age. during growth phase 2, 5 of 8 populations previously described (table 7) had heavier mean fdw for yearling males, but north dakota female yearlings grew at a alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 11 faster rate than most. during this phase, female yearlings are more likely to remain in loose association with their mother during subsequent calving, while male yearlings disperse and are likely at relative disadvantage. overall, fdw of north american female and male yearlings averaged 202.2 kg (n = 155), and 199.1 kg (n = 313), respectively. based upon the fdw versus ww proportions for north dakota moose (tables 1 and 2), the mean ww of north american female and male moose yearlings would average 278 and 285 kg, respectively. during this growth phase, geist (1998) suggested that a solitary moose must be at least 250 kg to confront predators, basing his assumption on the size of adult ussuri or “dwarf” moose of manchuria (a. a. cameloides) (heptner and nasimovitch 1967, p. 72 in geist 1998). whole weights of north dakota moose and data on fdw we provide from other studies (table 9) are partially supportive of his assumption. sexual dimorphism it should be noted that ww and fdw in north dakota are subject to biases that may or may not be obvious. for example, weights of adult males from other studies sampled during pre-rut would likely result in a higher sexual dimorphic ratio because these animals should be in optimal physical condition. the low dimorphic ratio for ww and fdw in north dakota may relate directly to our data collection period during post-rut when males weigh less. higher sexual dimorphism of fdw of 1.3 was reported in vermont, new hampshire, and maine (adams and pekins 1995), quebec (peterson 1974), and ontario (timmermann 1972; table 9). the earliest starting dates for moose hunting seasons in new hampshire, maine, ontario, quebec, and vermont range from august–october (timmermann and buss 1995), whereas, moose in this study and in alberta (26 november-6 january; blood et al. 1967) were harvested post-rut. further, the sample size of ww in alberta was small (n = 6 females and 3 males) and included animals 3.5 years of age (blood et al. 1967). we assume that the value of 1.3 probably represents the pre-rut maximum sexual dimorphism, and ratios <1.1 probably reflect post-rut leaner male weights. other dimorphic ratios we report (tbl, hfl, and sh; table 8) were low and should not vary seasonally. subspecies comparisons sample sizes reported by simkin (1962) are too small (i.e., 1-2 individuals per category) for comparative purposes, but are reported here for completeness. north dakota calf and yearling moose were larger than in all other populations except in ontario (timmermann 1972; table 7). ontario moose were sampled during the first 2 weeks of an october hunting season from forestland long managed for pulpwood with an abundance of preferred and productive habitat and forage (timmermann 1972). comparisons were also made between mean fdw and ww of mature adult moose from north dakota with 9 other north american populations (table 9); however, age criteria used for determining adult or mature moose was not uniform for all populations. for example, moose from alberta (blood et al. 1967) and alaska (franzmann et al. 1978) included animals as young as 3.5 years, and females from north dakota were comparatively shorter in tbl than those from alaska (franzmann et al. 1978). however, our comparisons suggest that female moose from north dakota are markedly heavier than all other populations, with the exception of 5 year-old captive animals in alaska (franzmann et al. 1978) and animals sampled prior to the rutting season in ontario (quinn and aho 1989). 12 morphology of north dakota moose – jensen et al. alces vol. 49, 2013 these data also indicate that moose in north dakota grow as fast or faster during their initial years when compared to other north american and european moose populations (table 7). based upon tbl, moose from north dakota reached their maximum mass by 5.5 years. range expansion of moose from northern forested areas into the agriculturally dominated landscape of north dakota provides a unique opportunity to obtain and evaluate subspecies and regional morphometric variations. during the fall, moose in north dakota are frequently observed foraging on sunflowers and other agricultural crops. although food habits information in north dakota is limited to a single study (maskey 2008), seasonal use of agricultural foods was high, with row crops (primarily corn [zea mays]) composing 11 and 22% of the fall and winter diets (sheridan co., moose hunting unit m9; fig. 1). in the turtle mountains (bottineau co., moose hunting unit m4; fig. 1) where row crops are limited, alfalfa (medicago sativa) composed 13% of the summer and fall diet representing *90% of consumed forbs. this agricultural forage may help maximize growth rate and body mass of moose in north dakota faster than would occur in traditional forest habitat. management implications environmental conditions affect body mass (sand 1996, hjeljord and histol 1999, ericsson et al. 2002), and in turn, body mass influences reproductive potential (saether and haagenrud 1983, adams and pekins 1995). monitoring moose populations is difficult, particularly in forested environments; however, measuring body weight of yearling and adult moose has potential value in predicting the nutritional and reproductive status of moose populations (adams and pekins 1995). growth rates also have predictive value in estimating reproductive potential and habitat conditions. because measuring body weight of moose is often a difficult task, having reliable equations to predict ww from other morphological parameters provides important alternatives for measuring and monitoring moose populations. the predictive equations provided here will be useful to wildlife managers for estimating various weight parameters and age related to productivity, and may aid in law enforcement and immobilization protocols when body weight and age of moose are often necessary. acknowledgements we want to express our sincere thanks to the hunters and landowners for their cooperation during this study. additionally, we would like to thank the following department personnel for their assistance on check stations: s. allen, a. anderson, a. aufforth, b. bitterman, s. brashears, e. dawson, g. enyeart, t. ferderer, t. frank, j. gulke, a. harmoning, l. johnson, m. johnson, m. johnson, m. kanzelman, s. kohn, r. knapp, j. kobriger, a. kreil, g. link, b. lynott, m. mckenna, r. parsons, r. patterson, c. penner, h. pochant, c. pulver, g. rankin, b. renhowe, s. richards, r. rollings, j. samuelson, d. schmidt, j. schulz, b. stotts, r. sohn, l. tripp, and l. vetter. finally, we thank e. addison and 2 anonymous reviewers for providing helpful editorial comments. financial support was provided by the north dakota game and fish department federal aid project w-67-r. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53–59. addison, e.m., r. f. mclaughlin, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infected and uninfected with winter ticks alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 13 (dermacentor albipictus). canadian journal of zoology 72: 1469–1476. blood, d. a., j. r. mcgillis, and a. lovass. 1967. weights and measurements of moose in elk island national park, alberta. canadian field-naturalist 81: 263–269. broadfoot, j. d., d. g. joachim, e. m. addison, and k. s. macdonald. 1996. weights and measurements of selected body parts, organs and long bones of 11-month-old moose. alces 32: 173–184. bubenik, a. b. 1998. evolution, taxonomy and morphology. pages 77–123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., usa. clutton-brock, t. h., f. e. guinness, and s. d. albon. 1982. red deer: behavior and ecology of two sexes. university of chicago press, chicago, illinois, usa. crichton, v. f. 1980. manitoba's second experimental moose hunt on hecla island. proceedings of the north american moose conference and workshop 16: 489–526. ericsson, g., j. b. ball, and k. danell. 2002. body mass of moose calves along an altitudinal gradient. journal of wildlife management 66: 91–97. franzmann, a. w. 1978. moose. pages 67-88 in j. l. schmidt and d. l. gilbert, editors. big game of north america: ecology and management. stackpole books, harrisburg, pennsylvania, usa. ———, w. b. ballard, c. c. schwartz, and t. h. spraker. 1980. physiologic and morphometric measurements in neonate moose and their cows in alaska. proceedings of the north american moose conference and workshop 16: 106–123. ———, r. e. leresche, r. a. rausch, and j. l. oldemeyer. 1978. alaskan moose measurements and weights and measurement-weight relationships. canadian journal of zoology 56: 298–306. gasaway, w. c., d. b. harkness, and r. a. rausch. 1978. accuracy of moose age determinations from incisor cementum layers. journal of wildlife management 42: 558–563. geist, v. 1998. deer of the world: their evolution, behavior, and ecology. stackpole books, mechanicsburg, pennsylvania, usa. haagenrud, h. 1978. layers in secondary dentine of incisors as age criteria in moose (alces alces). journal of mammalogy 59: 857–858. hjeljord, o., and t. histol. 1999. rangebody mass interactions of a northern ungulate a test of hypothesis. oecologia 119: 326–339. knue, j. 1991. big game in north dakota: a short history. north dakota game and fish department, bismarck, north dakota, usa. loudon, a. s. i. 1987. the influence of forest habitat structure on growth, body size, and reproduction in roe deer (capreolus capreolus l.). pages 559– 567 in c. m. wemmer, editor. biology and management of the cervidae. smithsonian institution press, washington d. c., usa. lynch, g. m., b. lajeunesse, j. willman, and e. s. telfer. 1995. moose weights and measurements from elk island national park, canada. alces 31: 199–207. maskey, jr., j. j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph. d. dissertation, university of north dakota, grand forks, north dakota, usa. peterson, r. l. 1955. north american moose. university of toronto press, toronto, canada. ———. 1974. a review of the general life history of moose. le naturaliste canadien 101: 9–21. 14 morphology of north dakota moose – jensen et al. alces vol. 49, 2013 quinn, n. w. s., and r. w. aho. 1989. whole weights of moose from algonquin park, ontario, canada. alces 25: 48–51. saether, b. e., r. anderson, o. hjeljord, and m. heim. 1996. ecological correlates of regional variation in life history of the moose alces alces. ecology 77: 1493–1500. ———, and h. haagenrud. 1983. life history of moose (alces alces): fecundity rates in relation to age and carcass weight. journal of mammalogy 64: 226–232. sand, h. 1996. life history patterns in female moose (alces alces): the relationship between age, body size, fecundity and environmental conditions. oecologia 106: 212–220. sauer, p. r. 1984. physical characteristics. pages 73-91 in l. k. halls, editor. white-tailed deer: ecology and management. stackpole books, harrisburg, pennsylvania, usa. schladweiler, p., and d. r. stevens. 1973. weights of moose in montana. journal of mammalogy 54: 772–775. schwartz, c. c. 1998. reproduction, natality and growth. pages 141–171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington d.c., usa. ———, w. l. regelin, and a. w. franzmann. 1987. seasonal weight dynamics of moose. swedish wildlife research supplement 1: 301–310. simkin, d. w. 1962. weights of ontario moose. ontario fish and wildlife review 1: 10–12. timmermann, h. r. 1972. some observations of the moose hunt in the black sturgeon area of northwestern ontario. proceedings of the north american moose conference and workshop 8: 223–239. verme, l. j., and m. e. buss. 1995. the status and management of moose in north america early 1990s. alces 31: 1–14. ———, and d. e. ullrey. 1984. physiology and nutrition. pages 91–118 in l. k. halls, editor. white-tailed deer: ecology and management. stackpole books, harrisburg, pennsylvania, usa. wallin, k., g. cederlund, and a. pehrson. 1996. predicting body mass from chest circumference in moose alces alces. wildlife biology 2: 53–58. welch, d. a., m. l. drew, and w. m. samuel. 1985. techniques for rearing moose calves with resulting weight gains and survival. alces 21: 475–491. alces vol. 49, 2013 jensen et al. – morphology of north dakota moose 15 mass, morphology, and growth rates of moose in north dakota methods results morphometric measurements morphometric relationships growth rates and patterns sexual dimorphism discussion general observations growth rates and weights of north dakota moose additional observations about moose growth rates sexual dimorphism subspecies comparisons management implications acknowledgements references wild animal research – new legal requirements in the european union margareta stéen1, katarina cvek2, and petter kjellander3 1swedish center for animal welfare, swedish university of agricultural sciences, uppsala, sweden; 2department of clinical sciences, swedish university of agricultural sciences, uppsala, sweden; 3department of ecology, grimsö wildlife research station, swedish university of agricultural sciences, riddarhyttan, sweden. abstract: the european union agreed on a directive (dir) for the protection of animals used for scientific purposes in 2010 which was implemented by member states at the onset of 2013. the dir applies to animals used for science or education that are subjected to pain, suffering, distress or lasting harm equivalent to, or higher than that caused by a needle. the dir changes the legal framework for wild animal research and requires educational and training standards of staff involved in capturing, planning, or performing research. both wild animals studied in or taken from the wild into captivity are covered by the dir. an animal welfare body must be established that includes a scientific member and at least one person responsible for animal welfare, and they must receive input from a designated veterinarian. the dir will aid and improve wild animal research because standards of animal welfare and research ethics must be met. although similar standards for moose research were employed previously in scandinavia, future moose research and conservation will likewise benefit. alces vol. 49: 127–131 (2013) key words: alces alces, animal welfare, capture, european union, legislation, marking, moose, research. introduction the european union (eu) developed a new directive (dir) for the protection of animals used for scientific purposes in 2010 (directive 2010/63/eu), that was implemented by member states (ms) at the onset of 2013. the principle reasons for the dir were to standardize legislation between ms and to improve the welfare of animals used in scientific research and procedures. this action was actualized by the increasing scientific knowledge about factors that influence animal welfare (aw), as well as the capacity of animals to sense and express pain and suffering. ethical concerns of the general public were also influential, and the desire to replace the use of animals for scientific purposes with non-animal alternatives. it was recognized that animals have an intrinsic value which must be respected, they should always be treated as sentient creatures, and their use should be restricted to benefitting human or animal health, or the environment. thus, ms must ensure that live animals are not used when a scientifically satisfactory method or testing strategy not entailing the use of live animals is available. wildlife research with wild ruminants had been regulated previously by national legislation in scandinavian countries; however, no comprehensive european regulations existed. moose (alces alces) in the wild have been used for scientific purposes in northern europe for decades and been used in theoretical and applied ecological research covering a wide range of topics: predator-prey interactions, herbivore-plant interactions, ecology, behavior, migration, veterinary medicine, disease, and moose-human 127 interactions using a variety of scientific methods and procedures, and technical and analytical approaches. moose provide an excellent case study for educating and explaining the dir to current wildlife researchers. in this paper we aim to inform researchers about the content and intent of the dir, and identify the consequences of its implementation for future moose research. description the dir states that animals taken from the wild (e.g., ruminants) shall not be used for scientific studies unless a competent authority has granted an exemption. the exemption shall only be granted if the purpose of the procedure cannot be achieved by the use of an animal which has been bred for such use. captures must be performed with care by competent persons such that an animal is not caused any avoidable pain, suffering, distress, or lasting harm. if a wild animal is found to be injured or in poor health at or after capture, it must be examined by a veterinarian or other competent person; action shall be taken to minimize the suffering of the animal. competent authorities may grant exemptions from the requirement to minimize suffering of the animal if there is scientific justification. scientists using research animals need to understand the dir legislation to identify which parts apply to and affect their research. this will facilitate experimental design and planning, minimize the number of animals used, and ensure that their research is within the legal framework. the dir applies to all animals that are used for scientific or educational purposes that are subjected to pain, suffering, distress, or lasting harm equivalent to, or higher than that caused by the insertion of a needle, in accordance with good veterinary practice. this includes moose studied in or taken from the wild and kept in captivity if subjected to procedures that cause “pain” or the equivalent to pain; e.g., an immobilized moose fitted with ear tags or radio-collar. all users of moose for scientific purposes must be authorized and registered by a competent authority. research projects must pass an ethical evaluation to receive authorization for use of animals in research or teaching; the use is evaluated and justified relative to the societal, scientific, and/or educational purpose of the project. the project should be designed such that procedures are in compliance with the requirement of the 3r’s (i.e., replacement of the use of animals for scientific studies, reduction of the number of animals used, refinement of procedures; russell and burch 1959). the evaluation of the project shall be performed in an impartial manner, the evaluation process must be transparent, and it shall weigh the predicted societal and scientific benefits, or educational value of the project against the harm and suffering of the research animals. furthermore, the evaluation shall verify that the project is designed such that procedures are performed in the most humane and environmentally sensitive manner possible. the ethical evaluation shall also determine whether the project should be evaluated retrospectively, concerning the aw and outcome of the study. for all projects, a non-technical summary written by the project leader/researcher shall be published by the ms to facilitate communication with the general public. before initiation of research, any staff involved in capturing, handling, planning, or performing research must be formally educated and trained until proven competent to perform their tasks. all users and breeders of animals for research must form an aw body which shall include the person responsible for aw and care, and in the case of a user, a scientific member; the body shall also receive input from the designated veterinarian. the 128 wild animal research – stéen et al. alces vol. 49, 2013 primary task of this body is providing advice on aw issues and the outcome of aw in projects. the body should also foster a climate of care and provide tools for the practical application and timely implementation of recent technical and scientific developments in relation to the principles of the 3 r’s to enhance the life-time experience of research animals. the advice of the aw body should be properly documented and open to scrutiny during inspections. the ms shall ensure that an animal may only be reused in a new procedure provided that the actual severity of any previous procedure was ‘mild’ or ‘moderate’ and that the general state of health and well-being has been fully restored; veterinary advice shall be taken into account regarding the lifetime experience of the animal. the ms may allow animals used in procedures to be returned to a suitable habitat appropriate to the species, provided that the state of health of the animal allows such and there is no danger to public health, animal health, or the environment. appropriate measures shall be taken to safeguard the wellbeing of the animal. application to moose a typical radio-collaring project with moose that is managed by a university or research institute in europe needs to comply with the dir by adopting the following protocol chronologically: 1) the department has to be granted a general permit to use animals in research for a limited period (years); 2) a local group (e.g., aw organization) should be formally appointed at the department level to help oversee aw; 3) the project leader should write an application describing the purpose (what, why, when, how) of the study relative to aw and how the project complies with the 3 r’s; 4) the application is signed by both the project leader and head of the local aw group; and 5) the application passes a review by a national ethics board. a permit to radio-collar a defined number of moose would then be approved for a maximum of 5 years. research studies involving capture and restraint can be stressful and cause measurable harm to moose, and can also influence experimental assumptions and data. the dir permits competent authorities to exempt the requirement to minimize the suffering of wild-caught animals found to be in poor health or injured, given scientific justification. the dir states that the assessment of health and welfare of the animals must be performed by a competent person. assessment of competence is based upon an appropriate level of understanding animal behavior, biology, and ecology of the species, and the ability to recognize poor health, injury, discomfort, pain, and distress. it is important to minimize the disturbance of a study population and understand the potential pathologies related to the capture activity, and how to prevent sickness and take appropriate actions in the case of poor health or welfare of a captured animal. proven competence is required for capture, handling, and restraint techniques including the operation and maintenance of any trapping devices. the idea of wild animals suffering is of concern for the legislating bodies, as well as scientists and the general public. concern regarding aw is related not only to marking methods, but also capture and handling procedures prior to, during, and after marking and release. combined long-term effects associated with these procedures and activity can occur at both the individual and population level. to our knowledge, the first chemical immobilization of a wild moose for research purposes in europe took place in 1975 at grimsö wildlife research area in south-central sweden (sandegren et al. 1987); > 2800 moose have been immobilized throughout scandinavia since (arnemo et al. 2006). moose are usually darted from a helicopter alces vol. 49, 2013 stéen et al. – wild animal research 129 and marking is performed under general anesthesia with typical surveillance including pulseand respiratory rate, body temperature, and sometimes arterial oxyhemoglobin saturation (spo2). chemical capture and anesthesia of free-ranging mammals involves some risk of mortality even in healthy animals. a scandinavian study on immobilization of moose estimated such mortality as 0.7% (n = 2,816) with 0.2% related directly to the immobilization procedure (arnemo et al. 2006). even if mortality is the most apparent negative side-effect of marking wild animals, other subtle, stress-induced biases are critical to identify because of their potential influence. for example, it is recommended to omit data from the initial 5 days post-capture when measuring moose movement and distribution (neumann et al. 2011). to help achieve zero mortality in scandinavian moose research, 3 factors have been identified: 1) use an experienced, trained professional capture team, 2) develop and follow a species-specific capture protocol, and 3) require a mortality assessment after any capture-related death to provide a knowledge-based evaluation of the capture protocol. this approach complies with the intent of the dir and should be followed by all moose researchers. when recapturing and re-marking the same animal, the three r′s should be employed to reduce the risk of impaired aw in moose research, conservation, and management. specific considerations are: 1) replacement strategies by which the required information may be obtained by other means than marking live wild animals, 2) educational strategies to use the fewest animals possible for providing valid information and statistical significance, and 3) refining strategies to use the most humane, least invasive marking techniques with the goal of minimizing pain and distress. these requirements are already in the spirit of most, if not all contemporary moose researchers. in addition to ecological studies with radio-collared moose, both physiological and pathological studies have been performed on captive moose housed indoors. for example, several successive swedish studies of infectious diseases in moose including brainworm (elaphostrongylus alces) and moose wasting syndrome involved calves obtained from the wild, and subsequently penned and raised in stalls (stéen et al. 1997, broman et al. 2002). the dir classifies facilities where wild animals are housed for experimental studies as ‘establishments’ and defines them as any installation, building, group of buildings, or other premises, and may include a place that is not wholly enclosed or covered, or a mobile facility. the ms must ensure that an establishment has installations and equipment suited to the species of animal housed and related research procedures. the dir states that animals removed from the wild shall not be used for scientific studies unless a competent authority has granted an exemption; of concern is the future of research requiring the capture, mark, and indoor housing of wild animals. to grant exemption for future studies of this type, competent authorities will require adequate knowledge of the behavior, biology, and veterinary studies of multiple wild species. if granted, the requirements of the dir are beneficial for all study participants including researchers, veterinarians, and field assistants since they will participate in project planning. this planning would include educational requirements of the various participants and their responsibilities, experimental design, ethical concerns, and the aw of the study animal. the practical implications of the dir should be advantageous for research with wild animals overall 130 wild animal research – stéen et al. alces vol. 49, 2013 and specifically benefit moose research and conservation globally. references arnemo, j. m., p. ahlqvist, r. andersen, f. bernsten, g. ericsson, j. odden, s. brunberg, p. segerström, and f. swenson. 2006. risk of capture related mortality in large free-ranging mammals experiences from scandinavia. wildlife biology 12: 109–113. broman, e., k. wallin, m. stéen, and g. cederlund. 2002. a wasting syndrome in swedish moose (alces alces): the background and current hypotheses. ambio 31: 409–416. directive 2010/63/eu. 2010. directive of the european parliament and the council on the protection of animals used for scientific purposes, 22 september 2010. (accessed march 2013). neumann, w., g. ericsson, h. dettki, and j. m. arnemo. 2011. effect of immobilizations on the activity and space use of female moose (alces alces). canadian journal of zoology 89: 1013–1018. russell, w. m. s., and r. l. burch, 1959. the principles of humane experimental technique. methuen, london, united kingdom. sandegren, f., l. pettersson, p. ahlqvist, and b. o. röken. 1987. immobilization of moose in sweden. swedish wildlife research supplement 1: 785–791. stéen, m., c. g. m. blackmore, and a. skorping. 1997. cross-infection of moose (alces alces) and reindeer (rangifer tarandus) with elaphostrongylus alces and elaphostrongylus rangiferi (nematoda, protostrongylidae): effects on parasite morphology and prepatent period. veterinary parasitology 71: 27– 38. alces vol. 49, 2013 stéen et al. – wild animal research 131 http://eur-lex.europa.eu wild animal research new legal requirements in the european union introduction description application to moose references shivering by captive moose infested with winter ticks edward m. addison1,2 and robert f. mclaughlin1,3 1wildlife research and development section, ontario ministry of natural resources, 300 water street, 3rd floor north, peterborough, ontario, canada k9j 8m5. abstract: occurence and rate of shivering were measured to assess thermoregulatory responses of captive moose (alces alces) infested with winter ticks (dermacentor albipictus). shivering was observed on 47 occasions in 5 of 8 infested moose calves from october to april; in contrast, 4 moose calves not infested with winter ticks did not shiver under identical weather conditions. only 5 shivering bouts occurred from october to march, all on a single day. the other 42 shivering bouts occurred in april with bouts lasting 1–103 min. during the april bouts, ambient temperature was 1– 4 °c (42 of 42), maximum wind speed was ≤12 km/h (38 of 42), and it was raining (30 of 42). shivering was associated with 23–44% hair loss in april, but not during cold weather in mid-winter despite 5–10% hair loss in march. maintaining stable core body temperature during late winter-early spring could compromise the energetic balance of wild free-ranging moose with extensive hair loss and abundant ticks, in conditions equivalent to or worse than measured in this study. alces vol. 50: 87–92 (2014) key words: alces alces, dermacentor albipictus, behavior, moose, winter ticks, shivering, thermoregulation. there are numerous studies and much conjecture as to how winter ticks (dermacentor albipictus) adversely affect moose (alces alces). effects on growth, food habits, blood composition, and bioenergetics have been studied in experimentally infested captive moose (mclaughlin and addison 1986, glines and samuel 1989, welch et al. 1990, addison and mclaughlin 1993, addison et al. 1994, addison et al. 1998). samuel (2004) and musante et al. (2007) modeled the potential blood loss from winter ticks and concluded that the elevated energy expenditure and protein imbalance associated with blood loss is deleterious to survival of moose calves with heavy infestations. mclaughlin and addison (1986) speculated that heat loss through winter tick-induced disruption of the hair coat in late winter-early spring may cause accelerated loss of body fat as measured in well fed, captive moose. in contrast, welch et al. (1990) suggested that hair loss may impose only nominal thermoregulatory costs on free-ranging moose and may possibly facilitate dissipation of excess heat in warm, early spring weather. shivering in moose is of interest because it is an involuntary form of thermogenesis that usually occurs when low ambient temperatures (ta) require an increase in metabolic rate to maintain core body temperature (tb) (i.e., the lower critical temperature, tlc) (hohtola 2004). shivering at or near the tlc has been observed in other cervids (parker and robbins 1984). our objective was to determine under what conditions shivering occurs in captive moose calves infested with winter ticks. this information will increase understanding of the metabolic impacts during late winter and early spring weather when moribund and dead moose with severe winter tick 2present address: ecolink science, 107 kennedy street west, aurora, ontario, canada l4g 2l8 3present address: r.r. #3, penetanguishene, ontario, canada l0k1p0 87 infestations are most frequently observed in the wild. methods the experiments were conducted in algonquin provincial park, ontario (45° 30′ n, 78° 35′ w) where 13 of 18 calves were captured at <2 weeks of age in may 1982; 5 calves were from other areas in central and northeastern ontario (addison and mclaughlin 1993). male and female calves were paired in each of 6 adjacent pens (29.6 × 16.5 m) located within a mixed forest stand with little undergrowth and a partial canopy (50% cover in summer) of white pine (pinus strobus), white birch (betula papyrifera), trembling (populus tremuloides) and big tooth aspen (p. grandidentata). calves were weaned as described by addison et al. (1983) and from late october to the end of the experiment were fed ad libitum a ruminant ration containing 16% crude protein, 2.5% crude fat, and 6% crude fiber (united cooperative of ontario, mississauga, ontario, canada). three treatment groups of moose were established: 4 controls without winter ticks, 4 infested with 21,000 winter tick larvae (moderate), and 4 infested with 42,000 larvae (heavy). where possible, siblings were assigned to different treatment groups. control moose were sprayed with acaricide (dursban m., dow chemical of canada ltd., sarnia, ontario, canada) twice in november and powdered with rotenone in december, january, and february to control accidental infestation by larval ticks. all moose were euthanized at the end of the experiment when their hair was dissolved and hides were checked for remaining ticks (addison et al. 1979). moose were weighed on a calibrated platform scale (canadian scale company model 7045) approximately every 7 days from 16–325 days of age (addison et al. 1994). these same 12 moose were used in concurrent studies of hair loss that was measured every 4–10 days from 23 january – 15 april 1983 by measuring the area of hair loss and depth of hair removed (mclaughlin and addison 1986). shivering and other behaviors were recorded simultaneously from 3 observation booths positioned above the back of each of a pair of 6 adjacent pens. three observers recorded observations of 2 moose in a single pen during a 2.5 h period, after which a second group of 3 observers replaced them. these 2 groups then alternated in successive 2.5 h periods throughout the day; all observations were in daylight hours. the 6 moose observed simultaneously included 2 from each treatment group; subsequently, the remaining 6 moose (2 per treatment) were observed the following day and this process continued until 27–30 h of observations were recorded monthly (october – april) for each moose. behaviour was recorded in 60, 1-min intervals during 1 h observation periods; 2403 h of observation occurred from mid-october 1982 to mid-april 1983. the ta, wind speed, and precipitation were recorded at the beginning of each observation period, and those data were attributed to each observation min in that period. results female and male calves weighed 128 ± 6 and 144 ± 15 kg when 4½ months old in early october and 200 ± 17 and 218 ± 20 kg, respectively, at the end of the experiment at 11 months of age (addison et al. 1994). the 4 control moose harbored 0, 4, 21, and 85 winter ticks at the end of the experiment, in contrast to 1179–8290 ticks recovered from the infested moose. fortytwo of 47 shivering bouts were observed in 2 individual moose in the moderate treatment group; 29 bouts were by a single moose (m4) (mclaughlin and addison 1986). the ta was relatively mild on most observation days particularly in october, 88 shivering in tick infested moose – addison and mclaughlin alces vol. 50, 2014 november, march, and april. in december – february (the coldest months) there were 59, 18, and 5 h of observations when ta was −10 to −19, −20 to −29, and −30 to −32 °c, respectively (table 1). no wind was measured in 123 h of observation, winds 20– 40 km/h occurred most frequently in october – december, and there were 14 h of observations in february when wind was >20 km/h (table 1). precipitation during observation periods was lowest in february (table 1). hours of precipitation were generally similar in september – january and april, with most april precipitation as rain. no shivering was observed from october – january and in march. one infested moose had 5 short shivering bouts within 35 min on a single day in february; ta was −10 to −11 °c with wind speed of 12 km/h during 1 bout, and 28 km/h during the other 4 bouts. these bouts averaged 5 min, ranging 1–10 min in length. the other 42 bouts occurred in april and averaged 17.8 min, ranging from 1–103 min in length. the ta was 1–4 °c, maximum wind speed was ≤12 km/h (38 of 42), and it rained during 30 of the 42 april bouts. shivering occurred when moose were standing (n = 25) or recumbent (n = 20), and when both standing and recumbent (n = 2). other activities during shivering bouts included recumbent with head rested on the ground (n = 9), walking (n = 9), ruminating (n = 8), feeding (n = 4), defecating (n = 4), drinking or eating snow (n = 2), urinating (n = 2), grooming (n = 2), rubbing against other objects (n = 1), trotting (n = 1), and galloping (n = 1). discussion shivering in moose during relatively warm and often wet april weather is of special interest because it seems inconsistent with results of prior studies. the insulative properties of moose hair are as high or higher table 1. weather characteristics during monthly observations of captive calf moose infested with winter ticks, algonquin provincial park, ontario, 1982–1983. values represent hours of observation during specific conditions; total monthly hours per category may differ due to rounding. the proportional monthly hours of rain are in parentheses. oct nov dec jan feb mar apr ambient temperature (°c) >0 64 12 10 0 0 33 50 0 to −9 1 48 26 42 19 26 10 −10 to −19 0 0 19 15 25 0 0 −20 to −29 0 0 5 3 10 0 0 −30 to −32 0 0 0 0 5 0 0 wind speed (km/h) 0 7 13 7 28 22 21 25 1–9 5 8 5 5 2 16 3 10–19 24 22 28 25 21 16 21 20–29 24 19 20 2 14 1 9 30–40 1 3 0 0 0 0 0 precipitation rain 15 (56) 10 (27) 10 (28) 7 (19) 0 13 (57) 26 (84) snow 17 27 26 29 7 10 5 alces vol. 50, 2014 addison and mclaughlin – shivering in tick infested moose 89 than that of any arctic mammal (scholander et al. 1950), hence cold weather under most conditions would not cause thermoregulatory stress to moose. increased metabolic rate indicative of moose reaching their winter tlc was not observed from −25 to −30 °c (renecker and hudson 1986), and captive, well fed moose calves were almost as cold tolerant as adults (renecker et al. 1978). absence of shivering in december, january, and most of february, despite very cold weather, was consistent with metabolic measurements and predictions (renecker and hudson 1986). in contrast, certain of our moose had extensive hair loss in april and shivering occurred in some but not all tick-infested moose during relatively mild temperatures in april. insulative properties of hair are positively correlated with the depth of hair for many arctic mammals and cattle (scholander et al. 1950, bennett 1964). however, wet pelts and skin substantially increase heat loss (scholander et al. 1950, holmes 1981) and wind reduces the insulative quality of hair coats (scholander et al. 1950, parker and robbins 1985). hair loss of our moose was generally ∼5% at the end of february, but for most infested moose increased to 23–44% by mid-april. most shivering was by a single moose with 31% hair loss during rainy and moderately windy conditions in april (mclaughlin and addison 1986). distinct individual variation in shivering was evident as 62 and 17% of shivering bouts were by 2 moose (moderate). it may seem anomalous that only 2 moose in the moderate group shivered the most, whereas only limited shivering was observed in the heavy treatment group. it is possible that the quantity of larval ticks applied in the 2 treatments was not different biologically. for example, hair loss was not substantially different between the moderate (23–31%) and heavy (28–44%) treatment groups. further, hair loss in the moderate group was highly variable (2–24%) at the conclusion of the experiment (mclaughlin and addison 1986). moose also displayed variation in grooming and rubbing behaviors within each of the treatment groups. little behavioral response was evident in our tick-infested moose in alberta during a relatively warm winter and spring (welch et al. 1990). only 1 trial occurred below −10 °c, and none below −25 °c, temperatures above the tlc of moose (renecker and hudson 1986). it has been suggested that hair loss might have a thermoregulatory advantage by dissipating excess heat during warm spring weather (welch et al. 1990). unfortunately, variables affecting thermo‐ regulation such as precipitation, wind, and radiant energy are seldom documented (parker and robbins 1985), and lack of standardization in techniques and variable diets are cause for caution when comparing bio‐ energetic studies of cervids (renecker and hudson 1986). in our study, the duration and intensity of wind and precipitation were generally similar in all months except for less precipitation in february (table 1). the only obvious weather differences between april and the prior 4 months included more precipitation as rain in april and warmer ta in march and april (table 1). there was also an exponential increase in hair loss (mclaughlin and addison 1986) and rapid growth by adult winter ticks in april (addison and mclaughlin 1988), hence higher blood loss as estimated in previous studies (see samuel 2004, musante et al. 2007). of note is that our experimental moose had >20,000 fewer ticks and much less hair loss at the end of the experiment than measured, on average, on free-ranging moose succumbing to heavy winter tick infestation in alberta (samuel and barker 1979, samuel 2004). the greater attrition of fat reservoirs in infested than non-infested moose, despite high quality food, presumably occurred over a period of 1–2 months (mclaughlin 90 shivering in tick infested moose – addison and mclaughlin alces vol. 50, 2014 and addison 1986). it is unlikely that the loss of body condition in infested moose could be attributed only to energy demands of maintaining core tb associated with hair loss. models developed by samuel (2004) and musante et al. (2007) indicate that the energy cost associated with replacing direct blood loss has a substantial metabolic impact during the 2–3 week peak period of female adult engorgement. these energy costs and higher thermoregulatory costs in march and april likely contributed to the reduced body condition of the infested calves as described by mclaughlin and addison (1986). it is also possible that due to depleted and less oxygenated red blood cells, there is less general motor activity and reduced movement, foraging, and production of body heat in heavily infested free-ranging moose. most moose experimentally infested with up to 42,000 larval winter ticks had limited hair loss and did not shiver during cold weather in mid-winter or even with substantial hair loss in spring. when shivering did occur it was for extended periods (hours) in spring, at or above freezing temperatures when moose were wet. the combination of rainy weather and severe hair loss from heavy tick infestation likely has more thermoregulatory impact than recognized previously. calves with severe hair loss and reduced body fat due to heavy winter tick infestation likely have an elevated tlc in late winter-early spring that could further compromise their survival. acknowledgements we appreciate d. j. h. fraser for his coordination of many early aspects of this study. special thanks go to a. rynard, a. macmillan, m. jefferson, v. ewing, and d. bouchard for their steadfast assistance in collection of data and for moose husbandry under adverse conditions. additional assistance was provided by c. pirie, m. a. mclaughlin, d. carlson, d. joachim and p. methner. l. smith, k. paterson, k. long, a. jones, s. gadawski, s. fraser, d. fraser, and l. berejikian assisted in the earlier care of calves. we appreciate the assistance of c. d. macinnes and g. smith and staff for their administrative support. field work was conducted at the wildlife research station in algonquin park where r. keatley, p. c. smith, and staff were of great help. thank you to r. moen, b. samuel, and anonymous reviewers for providing excellent suggestions for improvements to earlier drafts of the manuscript. references addison, e. m., f. j. johnson, and a. fyvie. 1979. dermacentor albipictus on moose (alces alces) in ontario. journal of wildlife diseases 15: 281–284. ———, and r. f. mclaughlin. 1988. growth and development of winter tick, dermacentor albipictus, on moose (alces alces). the journal of parasitology 74: 670–678. ———, and ———. 1993. seasonal variation and effects of winter ticks (dermacentor albipictus) on consumption of food by captive moose (alces alces) calves. alces 29: 219–224. ———, ———, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and uninfested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469–1476. ———, ———, and ———. 1998. effects of winter tick (dermacentor albipictus) on blood characteristics of captive moose (alces alces). alces 34: 189–199. ———, ———, and d. j. h. fraser. 1983. raising moose calves in ontario. alces 18: 246–270. bennett, j. w. 1964. thermal insulation of cattle coats. proceedings of the australian society of animal production 5: 160–166. alces vol. 50, 2014 addison and mclaughlin – shivering in tick infested moose 91 glines, m. v., and w. m. samuel. 1989. effect of dermacentor albipictus (acari: ixodidae) on blood composition, weight gain, and hair coat of moose, alces alces. experimental and applied acarology 6: 197–213. hohtola, e. 2004. shivering thermogenesis in birds and mammals. pages 241–252 in b. n. barnes and h. v. carey, editors. life in the cold: evolution, mechanisms, adaptation, application. twelfth international hibernation symposium. biological papers of the university of alaska, number 27. fairbanks, alaska, usa. holmes, c. w. 1981. a note on the protection provided by the hair coat or fleece of the animal against the thermal effects of simulated rain. animal production 32: 225–226. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)-induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. parker, k. l., and c. t. robbins. 1984. thermoregulation in mule deer and elk. canadian journal of zoology 62: 1409–1422. ———, and ———. 1985. thermoregulation in ungulates. pages 161–182 in r. j. hudson and r. g. white, editors. bioenergetics of wild herbivores. crc press, boca raton, florida, usa. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. ———, ———, k. christophersen, and c. lewis. 1978. effect of posture, feeding, low temperature, and wind on energy expenditures of moose calves. proceedings of the north american moose conference and workshop 14: 126–140. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1, federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m., and m. j. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869) on moose alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. scholander, p. f., v. walters, r. hock, and l. irving. 1950. body insulation of some arctic and tropical mammals and birds. the biological bulletin 99: 225–236. welch, d. a., w. m. samuel, and r. j. hudson. 1990. bioenergetic consequences of alopecia induced by dermacentor albipictus (acari: ixodidae) on moose. journal of medical entomology 27: 656–660. 92 shivering in tick infested moose – addison and mclaughlin alces vol. 50, 2014 shivering by captive moose infested with winter ticks methods results discussion acknowledgements references pre-parturition movement patterns and birth site characteristics of moose in northeast minnesota amanda m. mcgraw1, juliann terry2, and ron moen1 1natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811-1442; 2university of minnesota-duluth, swenson science building 207, 1035 kirby drive, duluth, mn 55812. abstract: habitat used immediately after parturition is important to survival of moose calves, though different habitat types may be functionally similar and thus contribute to the variability in habitat use reported in the literature. neonates are relatively immobile, which restricts movement of the cow-calf pair and makes both vulnerable to predation. the cow also requires adequate access to forage during the period when calf mobility is limited. we used fine-scale movement data to determine linear distance traveled to the birth site as well as habitat use by cow-calf pairs in northeast minnesota. all cows made long distance movements (x = 6 km) to the birth site where they localized in 1.72 ± 0.48 ha (95% kernel polygon) for approximately 7 ± 0.7 days. a mosaic of cover types that reflected availability across the landscape were used by the cow prior to localization at the birth site. birth site areas consisted of one cover type rather than the mosaic used before birth, and varied among cows, though bogs were used most often (40%). the small birth site area and use of bog habitat were likely a consequence of low calf mobility post-parturition. upon exiting the birth site, cow-calf pairs shifted toward use of mixed and young/regenerating forest which likely reflects the need and use for highly nutritious browse to meet the high energetic cost of lactation. alces vol. 50: 93–103 (2014) key words: alces alces, calving sites, minnesota, moose, parturition habitat. the time around parturition is critical to survival of offspring. the mother should select habitats that increase the survival of offspring and express behavior that reduces exposure of her and offspring to higher mortality risk. for species such as moose (alces alces), choices may be further restricted because the calf has limited mobility during the first weeks of life (altmann 1958, 1963). most moose give birth during a 19-day period in the month of may (sigouin et al. 1997). searches for calving sites typically take place after peak calving. opportunistic ground searches (addison et al. 1990, wilton and garner 1991), ground searches using vhf telemetry (bowyer et al. 1999, langley and pletscher 1994, leptich and gilbert 1986, scarpitti et al. 2007), and searches from aircraft for vhf collared cow-calf pairs (bailey and bangs 1980, mcgraw et al. 2011) have all been used to locate calving sites. cows will make a longer distance move followed by localization at calving, indicating that monitoring of daily locations can help identify when calving occurs (testa et al. 2000). for logistical reasons, most descriptions of pre-parturition movement patterns and birth site characterization have relied on relatively few locations in each parturition event. single daily locations are typically obtained in vhf telemetry studies to determine if a cow has localized or given birth (testa et al. 2000). ground searches for maternal beds accurately describe birth locations, but do not fully delineate the entire area used around the birth location during the postparturition period when cow-calf movements 93 are limited. aerial and ground vhf searches more accurately describe portions of the post-parturition area used by calf-cow pairs than the actual birth site. the limitations of single or few observations of cow-calf pairs during parturition may contribute to the high variation in vegetative cover and vegetation density, visibility, and proximity to water that has made describing generalized calving site characteristics difficult (addison et al. 1990, poole et al. 2007). variability of calving habitats across regions is probably also a function of available habitat types within a study area. moose are reported to birth on hill tops in quebec and ontario (addison et al. 1990, wilton and garner 1991, chekchak et al. 1998), and some swim to islands where available (addison et al. 1993). undisturbed lowland areas dominated by cedar and near water were important for calving in maine (leptich and gilbert 1986), and in new hampshire moose used mature, mixed, and coniferous forests, perhaps because open water and islands were rare (scarpitti et al. 2007). post-parturition areas had a higher than expected bog component in minnesota, though results were variable among cows (mcgraw et al. 2011). some cow moose select birth sites that provide hiding cover but do not necessarily have the highest quality or quantity of forage available (leptich and gilbert 1986, langley and pletscher 1994, bowyer et al. 1999). this is often interpreted as a trade-off between avoiding predators and meeting nutritional requirements (bowyer et al. 1999), and may be important to consider as a factor influencing calving site selection in minnesota where black bear (ursus americanus) and wolves (canis lupus) occur. some cows in british columbia calved in areas with lower forage availability that also had lower predation risk, while others calved in areas with higher forage availability and presumably higher predation risk (poole et al. 2007). understanding habitat and space use behavior during parturition may be especially important in northeast minnesota because the moose population is declining and recruitment rates in recent years are the lowest on record (delgiudice 2013). characteristics of the birth site and post-parturition areas have not been studied in detail in minnesota, in large part because post-parturition locations have been obtained from vhf telemetry flights that occur after peak calv‐ ing, and provide only one location within 2–4 weeks of calving (mcgraw et al. 2011). gps-collared moose can be used to locate calving sites by identifying a longer movement followed by localization (poole et al. 2007). our objective was to identify movement patterns indicative of calving by using location data recorded at 20 min intervals from gps-collared moose. we then evaluated fine-scale movements and habitat use of the cow while the calf had limited mobility. additional objectives were to characterize size, cover type composition, and length of time spent at the birth site by cow-calf pairs. the fine-scale gps locations allowed us to re-examine our previous data from vhf-collared moose (mcgraw et al. 2011) and compare past results with a more precise and robust dataset that can more accurately describe habitat use and movement patterns of moose during parturition. study area both the vhfand gps-collared moose studies occurred in approximately the same 3,700 km2 area of northeast minnesota (47°30′n, 91°20′w; fig. 1). land ownership is mostly public (∼82%) and includes portions of the superior national forest as well as state, county, and tribal lands. a significant portion of private ownership exists as blocks of industrial forest land. 94 moose parturition patterns – mcgraw et al. alces vol. 50, 2014 a boreal forest mix is the matrix from which moose in northeast minnesota choose a calving location. the region is part of the northern superior uplands (minnesota department of natural resources [mndnr] 2010) and is transitional from northern hardwoods in the south to canadian boreal forests in the north (pastor and mladenoff 1992). important habitat types in the home ranges of moose are young mixed conifer and deciduous forests, including aspen (populus tremuloides), paper birch (betula papyrifera), and balsam fir (abies balsamea). early successional forests (11–30 years post-disturbance) are used because forage is within reach of moose (kelsall et al. 1977). summer ranges consist largely of black spruce (picea mariana) lowlands as well as uplands and cut over areas dominated by paper birch, aspen, and balsam fir (peek et al. 1976). in early summer, moose generally use upland, lowland, and plantation areas in proportion to their occurrence (peek et al. 1976). northeast minnesota has a continental climate with severe winters and warm summers. precipitation usually occurs as snow from december–march. methods adult cow moose were darted from helicopters and fitted with gps collars (lotek fig. 1. study area in northeast minnesota where black dots indicate birth sites (n = 20) of gps collared moose, 2012. alces vol. 50, 2014 mcgraw et al. – moose parturition patterns 95 wireless, inc., newmarket, ontario, canada) in january and february 2011 (mccann et al. 2014). blood samples were taken and blood progesterone levels were used as an indication of pregnancy at capture. gps collars recorded locations every 20 min for 2 years. locations, movement, and habitat use of collared females were analyzed for patterns indicative of parturition beginning 1 may as previously defined in a vhf study (lenarz et al. 2011, mcgraw et al. 2011). the area occupied by the cow following its initial localization in may was considered the birth site because cows birth shortly after the initial localization (testa et al. 2000, poole et al. 2007). some cows remained in the immediate vicinity of the birth site, and others moved a short distance from the birth site and localized again (i.e., a secondary localization event). the area and time spent at both sites, as well as the total postparturition area were calculated. pre-parturition movement patterns we monitored pre-parturition movements using 20-min gps location data for each cow (n = 52) during the month of may. the short time scale between gps locations allowed us to calculate the distance moved each day throughout parturition. the length and duration of the movement was recorded from the last location in a cluster of foraging paths and bed sites to the initial localization at the birth site following the linear path. straight-line distances from the last feeding bout to the birth site were also calculated for comparison with past vhf studies. initial localizations at birth sites were identified by cows occupying smaller areas for longer durations and with less variation in location than foraging or bedding locations. in 2012 births were verified using helicopter searches to visually observe cow-calf pairs at the birth site shortly after parturition. vhf flight simulations we used calving dates and locations from gps-collared moose to better describe the vhf data set used previously to des‐ cribe post-parturition habitat in minnesota (mcgraw et al. 2011). we simulated postparturition locations using gps data to estimate the distance that cow-calf pairs in the vhf data set were from the birth site based on calving dates from gps-collared moose in 2012. this was necessary because calves located in the vhf data set could have been up to 4 weeks old. we also compared habitat characteristics at birth sites of gpscollared moose to the habitat characteristics of the simulated vhf data set, and to habitat characteristics of the actual vhf data set. all 12 flight dates from the 2004–2008 vhf study (mcgraw et al. 2011) were randomly assigned to each cow in the current study to simulate the location of the cow at 1200 hr using excel 2010 (microsoft corporation, redmond, washington, usa). the random assignment of flight dates was repeated to estimate the distance from the actual birth site to the simulated vhf post-parturition location. we then calculated the straight line distances from the gps birth site to the simulated vhf post-parturition site. birth site and post-parturition area birth sites (n = 20) were identified by viewing location data in googleearth (google inc. 2013 (version 7.1.1.1888, mountain view, california, usa) to locate clusters of points following long movements in may. dates and times of entry into and exit from the birth sites were identified visually. entrance into the birth site was defined as the first point in a cluster. the cow was still considered to be in the birth site when making short movements and re-localizing. the cow left the birth site when she moved and did not localize again or made a large movement from the last cluster of points at the 96 moose parturition patterns – mcgraw et al. alces vol. 50, 2014 birth site. birth site areas were calculated as 50% and 95% kernel polygons using all locations occurring between birth site entry and exit dates (fig. 2). we calculated 50% and 95% kernel polygons in the geospatial modelling environment (beyer 2012) using the plug-in estimator and a 10 m output cell size. we used the land use land cover (lulc) habitat classification system to determine cover type composition of birth sites (mcgraw et al. 2011). the lulc raster data set was derived from landsat thematic mapper (tm) images at a 30 m resolution (mndnr 2007). source imagery dates ranged from june 1995 to june 1996. the lulc classification system defined 16 cover types in northeast minnesota with >95% accuracy. more than 90% of the study area consisted of 6 terrestrial cover types: mixed, coniferous and deciduous forests, wet bog, marshes and fens, and regenerating forests (moen et al. 2011). we calculated cover type composition within 50% and 95% kernel birth site polygons using arcmap 10.1 (esri, redlands, california). cover type composition in these polygons was compared to cover type outputs of post-parturition areas identified during the vhf study (mcgraw et al. 2011) with anova. we used statistix (version 9.0; analytical software, boca raton, florida) and excel 2010. significance level for all tests was set at p = 0.05. unless otherwise noted, means are presented throughout as x � se. results all gps-collared cows with high progesterone levels (4.98 ± 0.3 ng/ml) at capture showed localization behavior indicative of calving in may. most cows that localized (46 of 52) made a long distance movement of 6 km ± 0.8 km (range = 1–33 km; fig. 2) over 17 ± 1.7 h (range = 2–57 h) before stopping at the birth site (fig. 2). linear path distances calculated using 20 min gps locations were twice as long as straight line distances from the beginning of the long distance movement to the birth site. the straight-line distance from the start of the long-distance movement to the calving site was 3 ± 0.5 km (range = 0.3–23 km). cows with low progesterone levels (0.33 ± 0.2 ng/ml) were not pregnant. all cows with low progesterone levels did not make a long distance movement and did not localize. cows remained localized at the birth site for 4 ± 0.4 days (range = 1–15 days). of the 52 cows that initially localized, 32 moved 133 ± 17.9 m (mode = 80 m; range = 30– 460 m) before localizing again for an additional 3 ± 0.6 days (range = 0–16 days). the remaining 20 cows did not move away from the site where they first localized. in total, cow-calf pairs spent 7 ± 0.7 days (range = 1–18 days) in the post-parturition area. the 50% kernel areas for 20 cows were 0.42 ± 0.06 ha and the 95% kernels were 1.72 ± 0.48 ha in the post-parturition period (fig. 3). the cover type trends indicating selection by cows were consistent, though not statistically different from post-parturition areas calculated in the previous vhf study (fig. 4). the proportion of bog cover type continued to increase as polygon size around birth sites fig. 2. distance moved by cows (n = 52) before localization at parturition. distances were measured as a linear path from locations collected every 20 min by gpscollared moose. alces vol. 50, 2014 mcgraw et al. – moose parturition patterns 97 decreased (34 ± 11% of 50% kernel polygons), while the amounts of mixed forest (24 ± 9% of 50% kernel polygons) and regenerating young forests (6 ± 5% of 50% kernel polygons) declined. primary and secondary localization sites for each cow, defined using 50% kernel polygons, were within a single cover type (table 1). bog habitats (35%) were used more than other available cover types by gps-collared cows, followed by mixed forest (25%) and conifer (20%). composition of cover types used 5 days prior to the long distance movement to birth site was variable among cows, but included a higher diversity of cover types than after localizing at the birth site (fig. 5). the trend in cover type use indicated a general movement away from variable use of mixed forests and young and regenerating stands to use of one cover type, more often bogs. 95% kernels50% kernelsgps locations cow 1 cow 0 100 200 400 meters 2 fig. 3. example of birth site areas as defined by 50% (black lines) and 95% (gray lines) kernel polygons for 2 cows. each point indicates the location of each cow at the birth site at 20 min intervals. fig. 4. change in cover type composition as the area surrounding the known cow/calf locations (ppa) and random locations were incrementally reduced from 100 to 5 ha (mcgraw et al. 2011), compared to composition of kernel birth site areas as defined using gps collar location data collected at 20 min intervals throughout the calving period. 98 moose parturition patterns – mcgraw et al. alces vol. 50, 2014 the average calving date during the gps study was 14 may (mode = 17 may; range = 3–27 may), with 70% of births occurring between 9 and 20 may. post-parturition locations from the vhf study were obtained by observing cow-calf pairs from helicopters after peak calving, between 21 may and 5 june each year (mcgraw et al. 2011). simulation based on calving dates of gpscollared cows indicated that mean calf age was 12 days (95% ci: 9.9–13.4 days) when cow-calf pairs were located during the vhf study. cow-calf locations for gps-collared cows corresponding to the simulated locations from the vhf study were 2 km (95% ci: 0.2–3.8 km) from actual birth sites. discussion in the current study in which actual birth sites were identified (n = 20), the bog cover type was used in higher proportion than its availability. however, there was still considerable variability in cover type composition among birth sites, with mixed forest used less than expected if birth site selection was random, whereas coniferous and deciduous forests were used about in proportion to availability. in the 5 days before long distance movement to birth sites, cows used a wide variety of cover types. upon movement to the birth site, cows tended to move to a specific cover type in which they localized to give birth and remained for 7 days. these data are consistent with past studies (addison et al. 1990, langley and pletscher 1994, chekchak et al. 1998, bowyer et al. 1999, scarpitti et al. 2007) demonstrating considerable variation in habitat types used table 1. proportion of birth sites (n = 20) in each cover type, based on the area inside 50% kernel birth site polygons. cover type birth site (%) home range (95% fixed kernel) northeast minnesota (%) deciduous forest 10 5 9 mixed forest 25 39 40 bogs 40 21 13 coniferous forest 20 17 23 regenerating 5 16 7 other 0 2 8 fig. 5. proportion of cover types used during periods beginning 5 days prior to the long distance movement (pre-ldm), during the long distance movement (ldm), localization at the birth site (birth), and 5 days post localization at the birth site (post-birth). alces vol. 50, 2014 mcgraw et al. – moose parturition patterns 99 as calving sites. this implies that many habitat types are functionally similar in terms of what aspects are required to successfully rear calves. as a result, specific protective measures of calving sites would be difficult to implement given the wide variety of suit‐ able habitat available. while calves are relatively immobile after birth, bog cover types were used by nearly half of cows for birth sites in minnesota. bogs likely provide hiding cover for calves as well as some foraging opportunities and access to water for cows while their movements are restricted by the calf. as the cow-calf pairs moved away from the birth site, variability of cover type composition increased, with more time spent in mixed forests as well as young and regenerating forests. movement to foraging habitat shortly after spring green up, when browse species have the highest nutritional content, likely allows cows to meet the energetic demands of lactation. despite the lag time between parturition and when cow-calf pairs were located during vhf telemetry flights, cover type composition was consistent with those observed with gps data. while the cow-calf pair moves away from the birth site after about one week, some cover types may remain important and sought out 3–4 weeks postparturition. as the post-parturition area surrounding vhf telemetry locations of cows with 12 day old calves was decreased, the proportion of bog cover type increased and mixed forest types decreased (mcgraw et al. 2011). straight-line distances for cows were half as long (3 km) as the movement lengths measured by following the linear path (6 km). using 20 min gps location data allowed us to refine movement calculations, and as a result, we identified that a greater proportion of moose in this study (88%) made long distance pre-parturition movements than reported in the literature (20%; bowyer et al. 1999). previous studies have measured straight-line distances from 2 discrete observations. while long distance movements observed in northeast minnesota are similar in length to those reported in central alaska (7.3 ± 2.3 km; bowyer et al. 1999), they are twice the length of those reported in south central alaska (4 km; testa et al. 2000). these differences are likely due to the method of measurement. in this study, duration and length of the pre-parturition movement were calculated by measuring the linear path of the cow with 20 min location data, whereas past vhf studies located cows once (testa et al. 2000) or twice daily (bowyer et al. 1999) during the calving period. more frequent locations will influence the measured distance moved per day during the long distance pre-parturition movement. calves are most vulnerable during the first weeks of life and affect cow movement until the calf is more mobile. cows remain in visual or vocal distance during the initial postparturition days when calves are less mobile (cederlund et al. 1987, van ballenberghe and ballard 2007). the duration of time spent in the birth area was ∼7 days which is consistent with the amount of time female moose were protective of birth sites at a research facility in russia (bogomolova and kurochkin 2002); however, it is shorter than the 3–4 weeks reported elsewhere (altmann 1963, bowyer 1999). this could be a result of the limitations of vhf technologies or reflect different definitions of the birth site area, ranging from the point of parturition to the larger area of restricted cow-calf movement in the weeks following birth. the size of the birth site area used during the first post-partum week tended to be small and located in a single cover type; however, this is partly a function of the coverage resolution (30 × 30 m cell size). variability among cows in terms of cover type selection for birth sites could also be an anti-predator 100 moose parturition patterns – mcgraw et al. alces vol. 50, 2014 strategy (bowyer et al. 1999). black bears (ursus americanus) and wolves (canis lupus) occur throughout moose habitat in northeast minnesota and their possible effect on birth site selection strategy should be considered, as restriction of the birth site area could be a function of predator avoidance. a cow remaining localized in a small area while its calf is relatively immobile reduces the area in which a predator may encounter the pair (bowyer et al. 1999). eventual movement from the birth site after ~7 days may be a function of depleted forage and continued predator avoidance. though the birth site area may be small, it stands to reason that the longer the pair remains at the birth site, the more likely they are to encounter a predator. cow-calf pairs that moved short distances within a few days post-partum remained in the same cover type when they localized again. these short movements could also reflect low availability of forage or disturbance (bowyer et al. 1999). the temporal scale at which we were able to monitor gps-collared cows is much finer than was previously possible, enabling us to more accurately define the spatial extent of birth site areas. it is possible that past observations using ground searches and/or telemetry overlooked small movements to secondary birth sites, resulting in calculation of smaller birth site and post-parturition areas. aerial birth site searches should be interpreted with caution if they are flown only once or twice a year after peak calving. simulated locations of cows that represented the 2004–2008 vhf flights occurred 12 days after localization of cows at the birth site, and after most cow-calf pairs would have left the area (mcgraw et al. 2011). the identification of birth site characteristics from the simulated vhf flights would have been 2 km from the actual post-parturition habitat when the calf is more mobile. the use of gps collars to collect locations every 20 minutes vastly improved our ability to identify pre-parturition movement patterns and allowed us to more accurately define the spatial extent of calving areas. future research using gps technology to record fine scale movement patterns is needed to determine when cows with calves resume normal activity levels after parturition. defining and identifying this will lead to more accurate description of post-parturition habitat use and cow-calf movements. acknowledgements funding for this work was provided by the environment and natural resources trust fund of minnesota, the university of minnesota duluth, and the natural resources research institute. partial summer support for a. mcgraw and j. terry was provided by the integrated biosciences graduate program, university of minnesota duluth. this is contribution number 566 from the center for water and the environment at the natural resources research institute, university of minnesota duluth. references addison, e. m., r. f. mclaughlin, d. j. h. fraser, and m. e. buss. 1993. observations of preand post-partum behaviour of moose in central ontario. alces 29: 27–33. ———, j. d. smith, r. f. mclaughlin, d. j. h. fraser, and d. g. joachim. 1990. calving sites of moose in central ontario. alces 26: 142–153. altmann, m. 1958. social integrations of the moose calf. animal behavior 6: 155– 159. ———. 1963. naturalistic studies of ma‐ ternal care in moose and elk. pages 233–253 in h. l. rheingold, editor. maternal behavior in mammals. wiley & sons, new york, new york, usa. bailey, t. n., and e. e. bangs. 1980. moose calving areas and use on the kenai alces vol. 50, 2014 mcgraw et al. – moose parturition patterns 101 national moose range, alaska. proceedings of the north american moose conference and workshop 16: 289–313. beyer, h. l. 2012. geospatial modelling environment (version 0.7.2.1). (software). url: . bogomolova, e. m., and y. a. kurochkin. 2002. parturition activity of moose. alces supplement 2: 27–31. bowyer, r. t., v. van ballenberghe, j. g. kie, and j. a. k. maier. 1999. birth-site selection by alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80: 1070–1083. cederlund, g., f. sandegren, and k. larsson. 1987. summer movements of female moose and dispersal of their offspring. journal of wildlife management 51: 342–352. chekchak, t., r. courtois, j-p. ouellet, l. breton, and s. st-onge. 1998. characteristics of moose (alces alces) calving sites. canadian journal of zoology 76: 1663–1670. delgiudice, g. d. 2013. 2013 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. (accessed march 2013). kelsall, j. p., e. s. telfer, and t. d. wright. 1977. the effects of fire on the ecology of the boreal forest, with particular reference to the canadian north: a review and selected bibliography. occasional paper #323. canadian wildlife service, ottawa, canada. langley, m. a., and d. h. pletscher. 1994. calving areas of moose in northwest‐ ern montana and southeastern british columbia. alces 30: 127–135. leptich, d. j., and j. r. gilbert. 1986. characteristics of moose calving sites in northern maine as determined by multivariate analysis: a preliminary investigation. alces 22: 69–81. lenarz, m. s. 2011. 2011 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. (accessed march 2011). mccann, n. p., r. a. moen, and s. k. windels. 2014. identifying thermal refugia for a cold-adapted mammal facing climate change. oecologia (in review). mcgraw, a. m., r. a. moen, and m. schrage. 2011. characteristics of postparturition areas of moose in northeast minnesota. alces 47: 113–124. minnesota department of natural resources (mndnr). 2007. landsat based land use-land cover. (accessed april 2013). ———. 2010. ecological classification system: laurentian mixed forest prov‐ ince. (accessed february 2011). moen, r. a., m. e. nelson, and a. edwards. 2011. radiotelemetry locations, home ranges, and aerial surveys in minnesota. alces 47: 101–112. pastor, j., and d. j. mladenoff. 1992. the southern boreal-northern hardwood forest border. pages 216–240 in h. h. shugart, r. leemans, and g. b. bonan, editors. a systems analysis of the global boreal forest. cambridge univer‐ sity press, cambridge, united kingdom. peek, j. m., d. l. urich, and r. t. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3–65. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammology 88: 139–150. scarpitti, d. l., p. j. pekins, and a. r. musante. 2007. characteristics of neonatal moose habitat in northern new hampshire. alces 43: 29–38. 102 moose parturition patterns – mcgraw et al. alces vol. 50, 2014 http://www.spatialecology.com/gme http://www.spatialecology.com/gme http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2011.pdf http://deli.dnr.state.mn.us/metadata.html?id=l250000120604 http://deli.dnr.state.mn.us/metadata.html?id=l250000120604 http://deli.dnr.state.mn.us/metadata.html?id=l250000120604 http://www.dnr.state.mn.us/ecs/212/index.html http://www.dnr.state.mn.us/ecs/212/index.html sigouin, d., j-p. ouellet, and r. courtois. 1997. geographical variation in the mating and calving periods of moose. alces 33: 85–95. testa, j. w., e. f. becker, and g. r. lee. 2000. movements of female moose in relation to birth and death of calves. alces. 36: 155–162. van ballenberghe, v., and w. b. ballard. 2007. population dynamics. pages 223–246 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. wilton, m. l., and d. l. garner. 1991. preliminary findings regarding elevation as a major factor in moose calving site selection in south central ontario, canada. alces 27: 111–117. alces vol. 50, 2014 mcgraw et al. – moose parturition patterns 103 pre-arturition movement patterns and birth site characteristics of moose in northeast minnesota study area methods pre-rturition movement patterns vhf flight simulations birth site and post-arturition area results discussion acknowledgements references a novel method of performing moose browse surveys rachel l.w. portinga1,2,3 and ron a. moen1,2 1biology department, university of minnesota duluth, 1049 university drive, duluth, mn 55812; 2university of minnesota duluth natural resources research institute, 5013 miller trunk hwy, duluth, mn, 55811 abstract: we measured browse availability and use along foraging paths of gps radio-collared moose (alces alces) in northeastern minnesota to estimate diet composition and browse species preference. on foraging paths during summer and winter we counted twigs via traditional methods for comparison with a novel method that attempted to better simulate moose foraging behavior. twigs were collected and used to develop diameter at point of browsing – biomass regressions for each browse species. these regressions, different under open and closed canopy, were used to estimate biomass consumption on foraging paths and to compare 4 approaches. the average diets were similar to previously measured regional diets, and importantly, our data identified variance among individual seasonal diets. our field method allowed us to better quantify and compare diet composition and browse selection of individual free-ranging moose directly on foraging paths. alces vol. 51: 107–122 (2015) key words: bite size, browse availability method, browse selection, diameter-biomass regressions, diet composition, minnesota large herbivores like moose (alces alces) view their food resources at the landscape, patch, and feeding station levels (senft et al. 1987). at the landscape level moose choose which patches to visit based on the spatial distribution of browse density and forage availability within each patch. at the patch level, moose must choose where to forage based on the available browse species, and tree and shrub heights at different feeding stations. younger patches can provide large quantities of high quality browse while older patches that have grown out of reach provide less browse (schwartz 1992). within a feeding station, bite size is based on the tradeoff between cropping and processing (spalinger and hobbs 1992). moose need to consume about 130 g dry mass/kg body weight0.75 daily in summer and about 40 g dry mass/kg body weight0.75 daily in winter (renecker and hudson 1985). using an average bite size of 1.02 g/bite (renecker and hudson 1986), this equates to at least 13,000 bites in summer and 4,000 bites in winter for a 454 kg (1000 lb) moose. winter consumption may be up to 50% higher depending on browse availability and species composition (hjeljord et al. 1994). this large demand for forage forces moose to move between patches and feeding stations in order to consume enough biomass. browse availability and bite size have been measured by following moose or moose tracks in snow and counting the number of available twigs per species, the number of bites per species, and measuring diameter at point of browsing, dry mass, and twig length (risenhoover 1987, shipley et al. 1998). locations of moose were found via radio telemetry (risenhoover 1987, hjeljord et al. 1990) or finding a track crossing a road (shipley et al. 1998). these methods were largely opportunistic and data collection was either clumped (location 3present address: hibbing community college, 1515 25th st e, hibbing, mn 55746 107 every hour for 2 days) or spread temporally (1–2 tracks weekly). another typical method is to measure browse availability in plots along randomly placed straight transects instead of following moose foraging paths. this provides an estimate of absolute browse density in a patch, rather than an estimate of browse availability encountered by moose. we measured intensively used feeding patches with 3 different protocols and a randomly placed straight transect in northeastern minnesota. our new protocol (the large feeding station method) attempted to simulate how a moose browses, which we contrasted with measurements along the foraging path and with absolute browse density. study area this study was conducted in northeastern minnesota where moose were previously collared for a vhf telemetry study (fig. 1) (lenarz et al. 2011). these forests transition between the canadian boreal and northern hardwood forests and experience a continental climate with short warm summers and severe winters (heinselman 1996). most was part of the superior national forest with the remaining either state, county, tribal, or industrial forest land fig. 1. the study area was within the superior national forest in northeastern minnesota. each black dot represents one measured foraging path in winter and a dark gray dot represents a summer foraging path. 108 novel browse surveys – portinga and moen alces vol. 51, 2015 (lenarz et al. 2010, moen et al. 2011). more specific details are provided in the minnesota moose research and management plan (mndnr 2011). methods regressions and estimating bite mass summer leaves were collected between july and september 2012, and winter twigs between january and april 2012 and 2013; twigs from both years were combined in the regression analyses. we clipped (standard garden clippers) browsed (∼3 cm below the browse point) and unbrowsed twigs of all browse species (table 1). samples were bagged and labeled with the location, date, and species. all browsed and unbrowsed twigs and leaves were stored at 2–3 °c prior to measurements. these twigs were used to develop diameter-biomass regressions for each season (telfer 1969). in summer we collected stripped twigs of each species which we clipped directly above the first unbrowsed petiole. a winter bite was equal to the twig biomass and a summer bite the leaf biomass from one twig, both with current annual growth >5 cm. on browsed twigs we measured (nearest 0.01 mm) the diameter at point of browsing and on unbrowsed twigs the simulated diameter at point of browsing. in summer, the simulated point of browsing was the diameter underneath the last stripped petiole. the wet weight of winter twigs and stripped summer leaves was weighed to the nearest 0.01 g. after weighing the wet mass of leaves, they were placed in the same bag with the corresponding twig. all unbrowsed table 1. the common and scientific names for each potential browse species in northeastern minnesota and seasons in which each species is consumed. “rare” species make up <1% of the diet at large feeding station paths. “not browsed” species were not consumed along the foraging paths. common name scientific name winter summer balsam fir abies balsamea common not browsed red maple acer rubrum common common mountain maple acer spicatum common common alder alnus rugosa rare rare juneberry amelanchier spp. common common paper birch betula papyrifera common common bog birch betula pumila not browsed rare red-osier dogwood cornus stolonifera common rare hazel corylus cornuta common rare black ash fraxinus niger not browsed rare white pine pinus strobus rare rare balsam poplar populus balsamifera rare rare quaking aspen populus tremuloides common common pin cherry prunus pennsylvanicus common common choke cherry prunus virginianus common common oak quercus spp. not browsed rare willow salix spp. common common elderberry sambucus pubens not browsed rare mountain ash sorbus decora rare common alces vol. 51, 2015 portinga and moen – novel browse surveys 109 twigs in both seasons were stored in labeled bags. all unbrowsed summer and winter twigs were dried at 60 °c for 48 h in a drying oven. dried twigs in winter and dried leaves in summer were stored at room temperature before being weighed to the nearest 0.01 g. most winter twigs (74%) and summer leaves (90%) were measured within 5 days of removal from the drying oven; the remainder was measured 6–9 days later. gps collars we captured adult moose in february and early march 2011 by darting them from helicopters. gps radio-collars (sirtrack ltd. and lotek wireless) fitted to each moose were programmed to transmit a location every 20 min. animal capture and handling protocols met the guidelines recommended by the american society of mammalogists (sikes et al. 2011) and were approved by university of minnesota and national park service animal care and use committees (#0912a75532). measuring browse availability summer browse availability was measured between 25 july and 14 september 2012, and winter browse availability between 3 january and 22 march 2013. browse availability was measured at the patch scale which we identified from the gps locations – patches had a concentrated number of moose locations. we used a handheld garmin gps to reach our pre-identified patches and then searched for a feeding station to identify a foraging path. a feeding station was defined as a plant or clump of plants with browsed twigs that were accessible when the forefeet of a moose are stationary (goddard 1968, novellie 1978, senft et al. 1987). a foraging path was defined as a trail of feeding stations within a patch. summer foraging paths were measured 1 to 15 days after the moose departed, and winter foraging paths were measured 3 to 17 days after departure. patches were considered accessible if they were on public land and we could access them by walking <2 km on a trail and/or <550 m from a trail. we measured winter patches containing 29 foraging paths from 8 moose (6f, 2m), and summer patches containing 31 foraging paths from 7 moose (5f, 2m). we defined a large feeding station as a location that appeared to have ≥10 bites. at all sites we measured browse under 4 different protocols to produce 4 foraging path types: 1) large feeding stations along the foraging path, 2) random plots along the foraging path, 3) random feeding stations along the foraging path, and 4) plots along a straight transect through the area containing the foraging path. each path type consisted of 10 measurement plots. large feeding station plots — the first large feeding station encountered was the first plot of the site and marked as a waypoint on the handheld gps. the plot or feeding station to be measured was a half circle with radius of 99.1 cm (39 in), with the center of the back side (straight line diameter) held at the approximate place where the moose stood. tracks in winter, other sign in either season, or placement of bites relative to open space were also used to determine where the moose stood and the direction it faced. at each large feeding station we counted the unbrowsed and browsed twigs of each browse species between 0.5 and 3 m above the ground (table 1; shipley et al. 1998). each cut-off twig was considered a bite. although an occasional large feeding station had <10 bites, we included it as a large feeding station because the observer estimated it had at least 10 bites. this only occurred at 10 of 290 (3%) large feeding stations in winter and 36 of 297 (12%) in summer. 110 novel browse surveys – portinga and moen alces vol. 51, 2015 we established the foraging path type from the first large feeding station by following tracks and browsing sign to the next large feeding station, marked it as the second waypoint on the gps, and repeated the measurements (fig. 2). plots could not overlap and this process continued until 10 large feeding stations had been measured on the foraging path. random plots on the foraging path — we created the random plot path type by stopping along the foraging path and repeating our browse measurements in random plots. a list of random distances between 5 and 14 m was generated using microsoft excel, and in the field we established the random plots using these distances in the gps “find” feature (fig. 2). random feeding stations — if a random plot had been browsed (evident bites), then that random plot was also defined as a random feeding station. if no browsed bites were in the random plot, we followed the foraging path to the next browsed twig (even if only one bite) and this became the location of the next random feeding station (fig. 2), eventually creating the random feeding path type. straight transect plots — after completing the large feeding station, random plot, and random feeding station measurements, we established a straight line transect that returned to the first plot. along this transect we stopped at random distances between 5 and 14 m until 10 plots were measured. if we reached the first large feeding station plot before completing 10 plots, we lengthened the transect. if, however, the cover type changed past the first plot and <10 plots were measured, we established a new transect in a random direction within the same cover type; 10 of 29 straight transects were angled in winter and 15 of 31 in summer. 3 1 2 8 9 54 7 6 = large feeding station (≥ 10 bites) = random plot = random feeding station (≥ 1 bite) fig. 2. a diagram of how we measured a foraging path. plot 1 is a large feeding station plot with ≥10 bites. plot 2 is a random plot. because plot 2 did not have any bites taken we stop at the next bite which becomes plot 3, a random feeding station plot. plot 4 is the second large feeding station plot. plot 5 is the second random plot with 1–9 bites, so it is also the second random feeding station plot. plot 6 is the third large feeding station plot. plot 7 is the third random plot that had ≥10 bites, so it is also the third random feeding station plot and the fourth large feeding station plot. plot 8 is the fourth random plot. plot 9 is the fourth random feeding station that had ≥10 bites, so it is also the fifth large feeding station plot. we continued until there were 10 plots of each type. alces vol. 51, 2015 portinga and moen – novel browse surveys 111 some cover types had little available browse making the foraging path difficult to follow in summer when 10 of 30 foraging paths had <10 plots in all path types. if no bites were found within 20 m of the previous feeding station when moving forward, we assumed the moose stopped foraging. effectively this meant that there were <10 large feeding stations, random feeding stations, and/or random plots in that foraging path. snow tracking in winter allowed us to more easily identify the foraging path; thus, 10 plots in all path types were measured in 28 of 30 foraging paths. canopy cover was measured 3 times with a densiometer (every 8th plot) to produce an average value in each patch. twigs collected from sites with 0–50% canopy closure were considered grown in open canopy, and twigs from sites with 70–100% canopy closure were considered grown in closed canopy. twigs from sites with 51–69% canopy cover were not used in the regressions or bite size summary statistics. statistical analyses biomass-diameter at point of browsing regressions, anovas on browse density, fig. 3. the percent of random feeding stations measured in each size category (line) and the percent of bites consumed at all feeding stations of a given size category (bar) in winter and summer. the dashed line separates the small feeding stations (≤9 bites) from the large feeding stations. in winter, 57% of the random feeding stations were considered large but they accounted for 86% of the consumed bites. in summer, 49% of the random feeding stations were considered large but they accounted for 82% of the consumed bites. 112 novel browse surveys – portinga and moen alces vol. 51, 2015 kruskal-wallis comparisons of diet, pearson χ2 goodness of fit tests, and bonferroni z-tests were all performed in jmp 10.0. significance level was set at p = 0.05. regressions — simulated diameters at point of browsing and dry masses of twigs from the unbrowsed winter twigs were log10 transformed and used to make 2 separate diameter-biomass regressions for each of the main browse species. the first regression used twigs grown in open canopy (0–50% shaded) and the second twigs from closed canopy (70–100% shaded). similarly, 2 summer regressions were made using leaf dry mass of each browse species. the raw data are found in ward (2014) and only results are presented here. statistics on bite size diameter and bite mass were calculated for each species. a t-test was used to test for statistical differences between the average diameter at point of browsing in open and closed canopy in both seasons for each species. available browse density — browse density was estimated as twig counts and as biomass. to obtain the total number of available twigs per path, we added the number of available twigs and the number of browsed bites. we estimated the total biomass originally available (browsed or unbrowsed) along a foraging path by multiplying the number of twigs of a given species by the average biomass of one bite of that species. for foraging paths in 0–50% shade, we used the average biomass values from open canopy regressions. likewise, we used the average biomass values from closed canopy regressions for foraging paths in 51–100% shade. although the closed canopy regressions were developed with twigs grown in 70–100% shaded areas, we felt the foraging paths in 51–69% shade were better classified as closed canopy than open canopy. balsam fir was not included in summer browse density estimates because it is not typically part of the summer diet. available and consumed browse density along each of the 4 path types were estimated using twig counts and biomass in both seasons. the length of each path was calculated by measuring the length of a line passing through all of the plots of each path type. the area of the foraging path was considered twice this distance to represent the ability of moose to browse either side of the foraging path. to calculate browse density we divided the twig count (available or consumed) by the area of the foraging path. these same calculations were made using biomass and twig counts. the browse density on large feeding station paths was compared with those on the random feeding, random, and straight transect paths using an anova of the log transformed data. diet composition — diet composition was calculated for each moose on the 4 path types in both seasons. we made a weighted average of those diet compositions to estimate diet composition for all moose on each path type in winter and summer. species were considered rare when they made up <1% of the average diet (shipley et al. 1998) at large feeding station paths. the percentage of the diet consisting of rare species is reported in the tables (but not text) to illustrate how a few individual moose consumed many bites of rare species. each individual diet had at least one browse species not identified on the foraging paths. because these data were not normally distributed and no transformation could correct this skewedness, we used a kruskalwallis test to test for significant differences between diet composition on the 4 path types. a kruskal-wallis test was also used to test for differences between each individual diet. browse species selection — we also determined the selection for each browse species from a combined average of all moose and for each individual using the alces vol. 51, 2015 portinga and moen – novel browse surveys 113 data from large feeding station paths. a pearson χ2 goodness of fit test and a bonferroni z-test were performed on the availability and use of all browse species for all moose combined and each individual moose (neu et al. 1974). a species was considered “positively selected”, “negatively selected”, or neither if there was a significantly larger, smaller, or equal proportion of browsed versus available twigs. results regressions all of the twig diameter – biomass regressions had slopes significantly different from zero. the slopes ranged from 0.58– 2.80 in winter and 0.45–2.07 in summer. in winter, 75% of the regressions had an r2 >0.60, and in summer 43% had an r2 >0.60. there was no consistent pattern between the open canopy or closed canopy regression slopes being larger or smaller (ward 2014). bite size across all species in winter, the mean diameter at point of browsing was 3.0 ± 0.02 mm in open canopy (range = 0.5–9.0 mm) and 3.1 ± 0.1 mm in closed canopy (range = 0.2–8.4 mm) (table 2). in summer, the mean across species was 2.3 ± 0.02 mm in open canopy (range = 0.02–11.1 mm) and 2.4 ± 0.04 mm in closed canopy (range = 0.2–6.1 mm) (table 3). using the regressions found in ward (2014), we calculated the average biomass consumed per bite for each browse species (tables 2 and 3). in winter, pin cherry had the largest bite size (2.3 ± 1.4 g) under closed canopy and the smallest bite size under open canopy (0.4 ± 0.1 g). mountain maple had the smallest bite size under closed canopy (0.4 ± 0.2 g). mountain ash had the largest (1.7 ± 1.4 g) and quaking aspen the smallest bite size (0.3 ± 0.2 g) under closed canopy in summer. bite density at feeding stations one purpose of establishing the random feeding station plots was to estimate the frequency of feeding stations of different sizes occurring along foraging paths. in winter 57% of random feeding station plots (n = 281) had ≥10 or more bites, and in summer 49% (n = 267). in both seasons at least 80% of twig consumption on the foraging path was from feeding stations with ≥10 bites, although moose occasionally consumed <10 bites at a station. browse density total available browse density was measured at 29 patches in winter and 30 patches in summer. it was significantly different among the 4 path types in both seasons using either method (winter twigs: f3, 112= 62.7, summer twigs: f3, 118 = 32.5, winter biomass: f3, 112 = 84.3, summer biomass: f3, 120 = 16.8, pall < 0.0001). likewise, density of consumed browse was also significantly different in winter and summer among the 4 path types (winter twigs: f3, 112 = 63.4, summer twigs: f3, 120 = 31.2, winter biomass: f3, 112 = 70.9, summer biomass: f3, 119 = 5.0, pall < 0.0025). as expected, both available and consumed browse densities were highest at large feeding station paths, followed by random feeding station, random plot, and straight transect paths (table 4). the average available browse density estimated by biomass at large feeding stations was 53% higher in summer (15.2 ± 1.7 g/m2) than winter (9.9 ± 1.0 g/m2). conversely, density estimated by twig counts was ∼2.5x larger in winter (15.2 ± 1.6 twigs/m2) than in summer (5.9 ± 0.6 twigs/m2). large feeding station paths had ∼60% more available twigs (727 ± 3) 114 novel browse surveys – portinga and moen alces vol. 51, 2015 in winter than in summer (460 ± 37), whereas the available biomass was ∼2.5x larger in summer (1166 ± 88 g) than winter (471 ± 26 g). the same seasonal differences existed for consumed twigs and biomass. the distance walked in winter to complete the large feeding station paths (27.6 ± 2.0 m, n = 29) was about half that in summer (50.5 ± 4.9 m, n = 31). the available and consumed browse density for each browse species was largest at large feeding station paths followed by random feeding station, random plot, and straight transect paths. the one exception (based on twig count) was that the highest browse density of hazel was found on the straight transect path in summer (when hazel is rarely consumed). table 2. summary statistics on browsed twigs in winter for all browse species. open canopy indicates twigs grown in locations shaded 0–50% and closed canopy indicates twigs grown in locations shaded 70–100%. p-values indicate t-test results between the diameter at point of browsing (dpb) of each species in open and closed canopy. we did not find enough individual twigs of juneberry, paper birch, pin cherry, or willow in closed canopy to calculate reliable averages for those categories. diameter at point of browsing (mm) species canopy average ± se minimum maximum average bite ± se (g) n p balsam fir** open 2.7 ± 0.1 0.9 6.5 1.6 ± 0.3 82 0.002 closed 2.2 ± 0.1 1.0 4.0 1.2 ± 0.2 50 red maple** open 3.5 ± 0.1 1.3 7.4 0.7 ± 0.3 125 0.009 closed 4.1 ± 0.1 2.7 6.9 1.4 ± 0.5 27 mountain maple* open 2.8 ± 0.3 1.5 4.6 0.6 ± 0.3 47 0.019 closed 2.4 ± 0.3 0.4 4.9 0.4 ± 0.2 56 juneberry open 2.4 ± 0.1 0.9 4.5 0.5 ± 0.1 161 0.583 closed na na na na 8 paper birch open 2.7 ± 0.1 0.6 4.8 0.8 ± 0.1 188 na closed na na na na 7 hazel open 2.7 ± 0.1 1.1 5.3 0.6 ± 0.1 301 0.104 closed 2.8 ± 0.1 1.1 4.5 0.6 ± 0.1 132 red-osier dogwood*** open 3.5 ± 0.1 1.5 6.1 1.1 ± 0.1 332 <0.0001 closed 4.3 ± 0.2 2.0 6.6 1.4 ± 0.4 40 quaking aspen open 3.5 ± 0.1 0.9 6.8 0.9 ± 0.1 209 0.155 closed 3.2 ± 0.1 1.0 5.7 0.7 ± 0.4 32 pin cherry open 2.4 ± 0.1 0.6 4.9 0.4 ± 0.1 216 na closed na na na na 6 choke cherry open 3.0 ± 0.3 1.5 4.8 0.7 ± 0.1 53 0.120 closed 2.6 ± 0.4 0.2 4.1 0.4 ± 0.1 20 willow open 3.1 ± 0.1 0.5 6.4 0.9 ± 0.1 501 na closed1 na na na na 0 mountain ash* open 4.3 ± 0.1 1.6 6.8 1.3 ± 0.3 43 0.045 closed 3.7 ± 0.1 1.2 8.4 0.7 ± 0.5 53 combined open 3.0 ± 0.02 0.5 9.0 na 2388 closed 3.1 ± 0.1 0.2 8.4 na 454 alces vol. 51, 2015 portinga and moen – novel browse surveys 115 consumption rate the pattern of consumption rate was similar to that of consumed browse density. the proportion of consumed twigs was highest on the large feeding station paths and declined progressively to the random feeding station, random plot, and straight transect paths. consumption was 45% in summer and 35% in winter on the large feeding station paths. overall, it was 23–45% on all paths except the straight transects where rates were 13% in winter and 9% in summer. diet composition season — at least 70% of all bites (all moose) consumed in winter along the 4 path types consisted of hazel, paper birch, willow, and quaking aspen. the remaining 30% consisted of balsam fir, juneberry, mountain maple, red maple, red-osier dogwood, pin table 3. summary statistics on browsed twigs of all species in summer. open canopy indicates twigs grown in locations shaded 0–50% and closed canopy indicates twigs grown in locations shaded 70–100%. pvalues indicate t-test results between the diameter at point of browsing (dpb) of each species in open and closed canopy. we did not find enough individual twigs of red maple in open canopy or pin cherry, willow, or mountain ash in closed canopy to calculate reliable averages for those categories. diameter at point of browsing (mm) species canopy mean ± se minimum maximum mean bite ± se (g) n p red maple open na na na na 14 0.349 closed 2.8 ± 0.2 1.3 6.0 1.4 ± 0.3 27 mountain maple*** open 2.3 ± 0.03 0.5 4.7 0.7 ± 0.1 675 <0.0001 closed 3.0 ± 0.1 0.5 4.9 1.0 ± 0.1 264 juneberry open 1.6 ± 0.04 0.1 3.2 0.5 ± 0.04 149 0.145 closed 2.1 ± 0.3 0.2 4.2 1.0 ± 0.4 20 paper birch** open 2.3 ± 0.1 0.02 5.1 0.8 ± 0.1 316 0.003 closed 2.0 ± 0.1 0.6 3.8 0.5 ± 0.1 84 hazel open 1.6 ± 0.1 0.5 3.5 0.7 ± 0.04 105 0.739 closed 1.6 ± 0.1 0.6 2.5 0.6 ± 0.1 48 red-osier dogwood*** open 2.9 ± 0.1 1.5 5.7 1.3 ± 0.1 41 0.001 closed 2.1 ± 0.2 0.5 4.4 0.7 ± 0.1 26 quaking aspen*** open 3.1 ± 0.2 0.5 11.1 1.4 ± 0.2 169 <0.0001 closed 1.6 ± 0.1 0.3 4.3 0.3 ± 0.2 53 pin cherry open 2.2 ± 0.1 0.6 4.2 0.8 ± 0.1 53 na closed na na na na 0 choke cherry open 2.2 ± 0.1 1.0 4.1 0.8 ± 0.1 44 0.085 closed 2.0 ± 0.1 0.8 3.9 0.8 ± 0.1 80 willow*** open 2.3 ± 0.1 0.5 5.5 0.9 ± 0.1 242 <0.0001 closed na na na na 14 mountain ash open 4.0 ± 0.1 2.0 7.0 1.1 ± 0.1 72 0.802 closed na na na na 7 all species open 2.3 ± 0.02 0.02 11.1 na 2071 na closed 2.4 ± 0.04 0.2 6.1 na 627 116 novel browse surveys – portinga and moen alces vol. 51, 2015 cherry, and choke cherry. rare species were alder, mountain ash, balsam poplar, and white pine (table 5). in summer 70% of bites consisted of mountain maple, willow, and paper birch on large feeding station, random feeding station, and random plot paths. the remaining 30% was juneberry, red maple, pin cherry, choke cherry, quaking aspen, and mountain ash. rare species were hazel, balsam poplar, red-osier dogwood, balsam fir, alder, bog birch, black ash, oak, elderberry, and white pine. on straight transects at least 70% of consumed twigs were mountain maple, willow, quaking aspen, and species considered rare (table 5). path type — despite the general similarities in diet diversity, all browse species comprised different portions of the winter diet on the 4 path types (kruskal-wallis, h3 > 12.3, p < 0.007) except paper birch and hazel (kruskal-wallis, h3 < 1.2, p > 0.60; table 6). in summer juneberry, quaking aspen, and mountain ash comprised different portions of the diet on all 4 path types in summer (kruskal-wallis, h3 > 8.1, p < 0.045; table 5). no difference existed among the 4 path types for red maple, mountain maple, paper birch, cherry, and willow (kruskal-wallis, h3 < 5.7, p > 0.13). individuals — diets based on twigs consumed on large feeding station paths varied individually and from the pooled average (tables 6 and 7). one winter example of this individual difference was female moose 31180 that consumed 26% red maple and 50% hazel (4 paths combined) compared to the group average of 5% red maple and 26% hazel (table 6); red maple was more available in her foraging patches. an example in summer was male moose 31190 that consumed 10% mountain maple and 61% willow (4 paths combined) compared to the group average of 41% mountain maple and 21% willow (table 7). browse species selection the average diet in winter (all moose combined) was different from that available (v29 ¼ 3122, p < 0.0001). a bonferroni z-test on the combined data indicated that juneberry, red maple, mountain maple, paper birch, red-osier dogwood, and quaking aspen were eaten more than available in summer. hazel was eaten less than available, and cherry and willow were used in proportion to availability (table 8). individual diets were also different from browse availability on their respective foraging paths (all moose: v2�9 � 74:6, p < 0.0001 for all moose). the average summer diet (all moose combined) was also different from available (v28 ¼ 840, p < 0.0001), as were individual diets (all moose: v2�8 � 43:9, p < 0.0001). table 4. available browse density and consumed browse density along four path types in summer and winter measured by twigs/m2 ± se and biomass (g)/m2 ± se. w = winter, s = summer. method season large feeding station random feeding station random plot straight transect available # twigs w 15.4 ± 1.6 2.3 ± 0.2 2.0 ± 0.2 1.4 ± 0.2 s 5.9 ± 0.6 2.0 ± 0.2 1.8 ± 0.3 1.1 ± 0.1 biomass w 9.9 ± 1.0 1.7 ± 0.1 1.5 ± 0.1 1.0 ± 0.1 s 15.2 ± 1.7 6.8 ± 1.9 4.5 ± 0.8 2.9 ± 0.4 consumed # twigs w 5.3 ± 0.6 2.1 ± 0.1 0.5 ± 0.1 0.2 ± 0.03 s 2.7 ± 0.3 1.0 ± 0.3 0.4 ± 0.1 0.1 ± 0.03 biomass w 4.0 ± 0.4 0.5 ± 0.04 0.4 ± 0.04 0.2 ± 0.02 s 6.7 ± 0.7 2.4 ± 0.4 1.0 ± 0.2 0.3 ± 0.04 alces vol. 51, 2015 portinga and moen – novel browse surveys 117 a bonferroni z-test on the combined summer data indicated that red maple, mountain maple, cherry, and mountain ash were eaten more than available in summer, willow less than available, and juneberry, paper birch, and quaking aspen proportional to availability (table 8). discussion we initially chose to measure large feeding stations (≥10 bites) because field observations indicated that these sites were common and theory (senft et al. 1987) supports the strategy of such foraging behavior. by contrasting browse density along a foraging path at large feeding stations with alternate routes, we demonstrated how moose increased effective browse density by selecting a specific foraging path. for example, moose took at least 80% of their bites at large feeding stations with ≥10 bites. the identification of large feeding stations provided a fast and efficient manner to measure browse availability and consumption along presumed foraging paths, and this method can also be used to evaluate the relative quality of browsed and unbrowsed patches (ward 2014, ward and moen 2014) this method avoids 2 potential complications associated with the straight transect table 5. diet composition (average percent of diet ± se) measured on four path types. averages and se were weighted by moose. rare includes species that made up <1% of the diet at large feeding station paths. 29 foraging paths were measured in winter 2013 and 31 were measured in summer 2012. winter species large feeding station random feeding station random plot straight transect hazel 27 ± 7 26 ± 8 27 ± 9 28 ± 8 paper birch 26 ± 7 26 ± 6 25 ± 6 18 ± 6 willow 11 ± 5 14 ± 6 13 ± 6 11 ± 5 quaking aspen 7 ± 3 8 ± 4 10 ± 5 13 ± 6 juneberry 6 ± 2 5 ± 2 4 ± 1 4 ± 2 red maple 5 ± 3 4 ± 2 5 ± 3 4 ± 4 red-osier dogwood 5 ± 4 3 ± 3 3 ± 3 10 ± 11 balsam fir 4 ± 2 6 ± 2 6 ± 3 2 ± 2 mountain maple 4 ± 3 3 ± 1 2 ± 1 2 ± 1 cherry 3 ± 1 2 ± 1 2 ± 1 2 ± 1 rare 2 ± 2 2 ± 1 2 ± 1 5 ± 6 summer mountain maple 42 ± 11 45 ± 10 43 ± 11 25 ± 11 willow 21 ± 8 21 ± 9 28 ± 11 23 ± 11 paper birch 11 ± 3 9 ± 4 6 ± 4 6 ± 5 cherry 9 ± 4 7 ± 4 6 ± 4 3 ± 5 quaking aspen 8 ± 4 10 ± 3 8 ± 3 14 ± 7 mountain ash 4 ± 2 3 ± 2 4 ± 4 0 juneberry 2 ± 1 3 ± 2 2 ± 1 8 ± 5 red maple 1 ± 1 0 0 7 ± 4 rare 1 ± 0.3 1 ± 0.4 0.2 ± 0.1 10 ± 7 118 novel browse surveys – portinga and moen alces vol. 51, 2015 method: 1) measuring random locations, and 2) empty plots. the foraging path approach eliminates these concerns by ensuring plentiful data at actual foraging locations. arguably, it also reflects the browse availability a moose would actually perceive. randomly placed plots in straight transects are often empty, which would mean that many more plots would be required to accurately estimate the availability of patchy browse. our method avoids empty plots, incorporates distance moved between feeding stations, and table 6. diet composition of individual moose in winter 2013 measured by twigs consumed at large feeding station paths. there are diets for eight collared moose. 31189 and 31190 are male, the rest are females. n is the number of foraging paths measured. rare species made up <1% of the combined moose diet at large feeding stations. moose number species all moose 31166 31174 31175 31178 31180 31182 31189 31190 hazel 27 21 38 29 13 50 33 9 68 paper birch 26 14 41 15 57 9 3 56 3 willow 11 5 9 6 5 3 quaking aspen 7 28 16 12 <1 8 2 juneberry 6 18 1 8 9 1 red maple 5 26 9 red-osier dogwood 5 15 <1 38 4 balsam fir 4 2 4 15 1 14 mountain maple 4 16 1 1 25 cherry 3 11 1 5 6 <1 3 6 rare 1 2 5 1 n 29 2 2 3 3 4 2 5 3 table 7. diet composition of individual moose in summer 2012 measured by twigs consumed at large feeding stations only. there are diets for seven collared moose. 31189 and 31190 are male, the rest are females. n is the number of sites measured. rare species made up <1% of the combined moose diet at large feeding stations. moose number species all moose 31166 31168 31175 31178 31180 31189 31190 mountain maple 41 3 57 84 90 36 57 10 willow 21 53 3 9 17 61 paper birch 11 12 1 17 13 7 cherry 9 8 5 3 2 24 1 3 quaking aspen 8 4 36 1 23 2 mountain ash 4 17 5 5 3 juneberry 2 1 12 red maple 1 5 rare 1 1 1 3 n 31 3 2 3 3 3 6 4 alces vol. 51, 2015 portinga and moen – novel browse surveys 119 provides an estimate of effective browse density. a challenge to simulating foraging decision rules when following a foraging path is that humans find large feeding stations by sight, but moose likely use other senses as well. diet composition was statistically different among seasons and path types. the average combined diet in both winter and summer was best categorized as generalist because one genus did not account for >60% of the diet (shipley 2010). the two primary browsed species were hazel and paper birch in winter and mountain maple and willow in summer, hence, moose may forage in different areas in winter and summer. for example, available browse density estimated by twig counts was higher in winter than in summer, with hazel consumed commonly in winter but rarely in summer. use of gps locations may help distinguish seasonal differences in foraging locations and browse species availability. the diet composition was similar to that measured >3 decades previously in the region (peek et al. 1976). the top 5 summer species (percent of diet) were the same in both studies: mountain maple, willow, paper birch, cherry, and quaking aspen. mountain maple was ranked first in our study and fifth by peek et al. (1976), and quaking aspen had the opposite rankings. hazel, willow, and quaking aspen were 3 of the top 5 winter species in both studies. one difference was that paper birch and juneberry were included in our top 5, whereas peek et al. (1976) had balsam fir and red-osier dogwood. during both seasons the primary species consumed table 8. browse species selection in both seasons when data from all moose was combined. if the moose were simply browsing at random, we would expect the 95% confidence interval of the percent browsed to contain the percent available at large feeding stations. season species percent available at large feeding stations 95% confidence interval of percent browsed at large feeding stations selection winter juneberry 4.7 5.1 ≤ – ≥ 6.8 + red maple 3.3 3.8 ≤ – ≥ 5.3 + mountain maple 2.7 4.0 ≤ – ≥ 5.5 + paper birch 19.3 24.7 ≤ – ≥ 27.9 + red-osier dogwood 2.1 3.3 ≤ – ≥ 4.8 + quaking aspen 5.6 5.8 ≤ – ≥ 7.6 + cherry 3.0 2.7 ≤ – ≥ 4.0 0 willow 11.9 11.2 ≤ – ≥ 13.5 0 balsam fir 9.0 2.8 ≤ – ≥ 4.1 − hazel 36.8 26.3 ≤ – ≥ 29.5 − summer red maple 0.5 0.6 ≤ – ≥ 1.3 + mountain maple 27.6 34.6 ≤ – ≥ 38.2 + cherry 7.2 8.3 ≤ – ≥ 10.5 + mountain ash 4.2 8.6 ≤ – ≥ 10.8 + juneberry 3.3 2.2 ≤ – ≥ 3.4 0 paper birch 10.4 9.8 ≤ – ≥ 12.1 0 quaking aspen 8.1 6.1 ≤ – ≥ 8.1 0 willow 28.6 18.9 ≤ – ≥ 21.9 − 120 novel browse surveys – portinga and moen alces vol. 51, 2015 were consistent regardless of path type. because more twigs were counted on the large feeding station paths, they probably provided the better estimate of diet and species consumption rates. this study was unique because we collected data from individual free-ranging moose by using their gps locations to identify their foraging paths shortly after use. presumably each moose selected browse based on availability within the patch they occupied. individual consumption differences occurred in both winter and summer, and though previous studies have not provided for analysis and comparison of individual diet selection, individual differences in habitat selection by moose were documented in british columbia (gillingham and parker 2010). pooling the data from many foraging paths identified the generalized seasonal diets and the most important browse species in this region, and concurred with previous research. it also identified individual diet variation which suggests that moose adapt their diet based on the local composition and availability of browse species. we were able to simulate how a moose browsed in a patch using the large feeding station method. there was some subjectivity in choosing which large feeding station was closest (consecutive) when establishing the foraging path; however, a moose would face the same choice. contrasting browse measurements between simulated and actual foraging paths in the same patch would provide a good evaluation of our approach and potential differences. we offer that incorporating large feeding stations and the distance between adjacent large feeding stations is an efficient method to estimate browse availability at the patch level. acknowledgements we would like to thank the epa great lakes restoration initiative and the minnesota environment and natural resources trust fund (enrtf) for funding and n. bogyo of the 1854 treaty authority for winter field work assistance. references gillingham, m. p., and k. l. parker. 2010. differential habitat selection by moose and elk in the besa-prophet area of northern british columbia. alces 44: 41–63. goddard, j. 1968. food preferences of two black rhinoceros populations. african journal of ecology 6: 1–18. heinselman, m. 1996. the boundary waters wilderness ecosystem. university of minnesota press, minneapolis, minnesota, usa. hjeljord, ø., n. hovik, and h. b. pedersen. 1990. choice of feeding sites by moose during summer, the influence of forest structure and plant phenology. holarctic ecology 13: 281–292. ———, b. e. saether, and r. andersen. 1994. estimating energy intake of freeranging moose cows and calves through collection of feces. canadian journal of zoology 72: 1409–1415. lenarz, m. s. 2011. 2011 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. (mndnr) minnesota department of natural resources. 2011. minnesota moose research and management plan. minnesota department of natural resources, st. paul, minnesota, usa. moen, r. a., r. peterson, s. windels, l. frelich, d. becker, and m. johnson. 2011. minnesota moose status: progress on moose advisory committee recommendations. nrri technical report no. nrri/tr-2011/41. natural resources research institute, duluth, minnesota, usa. alces vol. 51, 2015 portinga and moen – novel browse surveys 121 novellie, p. a. 1978. comparison of the foraging strategies of blesbok and springbok on the transvaal highveld. south african journal of wildlife research 8: 137–144. neu, c. w., c. r. byers, and j. m. peek. 1974. a technique for analysis of utilization-availability data. journal of wildlife management 38: 541–545. peek, j. m., d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3–65. renecker, l. a., and r. j. hudson. 1985. estimation of dry matter intake of freeranging moose. journal of wildlife management 49: 785–792. ———, and ———. 1986. seasonal foraging rates of free-ranging moose. journal of wildlife management 50: 143–147. risenhoover, k. l. 1987. wintering foraging strategies of moose in subarctic and boreal forest habitats. ph. d. thesis, michigan technical university, houghton, michigan, usa. schwartz, c. c. 1992. physiological and nutritional adaptations of moose to northern environments. alces supplement 1: 139–155. senft, r. l., m. b. coughenour, d. w. bailey, l. r. rittenhouse, o. e. sala, and d. m. swift. 1987. large herbivore foraging and ecological hierarchies. bioscience 37: 789–799. shipley, l. a. 2010. fifty years of food and foraging in moose: lessons in ecology from a model herbivore. alces 46: 1–13. ———, s. blomquist, and k. danell. 1998. diet choices made by free-ranging moose in northern sweden in relation to plant distribution, chemistry, and morphology. canadian journal of zoology 76: 1722–1733. sikes, r. s., and w. l. gannon. 2011. the animal care and use committee of the american society of mammalogists. (2011). journal of mammalogy 92: 235–253. spalinger, d. e., and n. t. hobbs. 1992. mechanisms of foraging in mammalian herbivores: new models of functional response. the american naturalist 140: 325–348. telfer, e. s. 1969. twig weight-diameter relationships for browse species. journal of wildlife management 33: 917–921. ward, r. l. 2014. browse availability, bite size, and effects of stand age on species composition and browse density for moose in northeastern minnesota. m.s. thesis, university of minnesota, minneapolis, minnesota, usa. ———, and r. a. moen. 2014. effects of stand age on species composition and browse density in northeastern minnesota. nrri technical report no. 2014– 36. natural resources research institute, duluth, minnesota, usa. 122 novel browse surveys – portinga and moen alces vol. 51, 2015 a novel method of performing moose browse surveys study area methods regressions and estimating bite mass gps collars measuring browse availability statistical analyses results regressions bite size bite density at feeding stations browse density consumption rate diet composition browse species selection discussion acknowledgements references alces26_104.pdf alces vol. 45, 2009 heikkilä & tuominen moose in conserved forests in finland 49 the influence of moose on tree species composition in liesjärvi national park in southern finland risto heikkilä1 and marita tuominen2 1finnish forest research institute, vantaa research unit, p.o. box 18, fi-01301 vantaa, finland 2university of helsinki, faculty of forest ecology, p.o. box 27, fi-00014 helsinki, finland abstract: intensive forest management has promoted a rapid increase in finland’s moose (alces alces) population since the 1970s. the main objective of this study was to determine the role of moose browsing in modifying natural processes of protected forests that are influenced by high moose populations in adjacent managed forests. this study occurred in liesjärvi national park located in the mid-boreal vegetation region of finland. forest stands were sampled with line-plot sampling (50 m² plots at 100 m distances) in the older (oa; 1956) and newer (na; 2005) parts of the park. we found that long-term selective browsing in oa retarded the development of young stands in favor of norway spruce (picea abies) and low-growing broadleaf species. browsing in recent years was relatively intensive in na where young regeneration areas still existed from previous forest management. the most intensive browsing occurred on 18.6 % of trees in na and 3.1 % in oa; young palatable tree species were taller in na than oa. also, in oa the density of preferred aspen (populus tremula) and rowan (sorbus aucuparia) trees was relatively low in the height class that produces the dominant tree canopy. despite short-term intensive browsing, na appeared better able to recover to a natural forest state. fecal pellet groups associated with young scots pine (pinus sylvestris) and browsing of birch (betula spp.) and aspen indicated the importance and role of forage quantity and quality on winter range of moose. the amount of consumed new twig biomass was 20-fold greater in na compared to oa, indicating a difference in the size of the moose population and presumably habitat quality between the areas. the effect of browsing on different tree species was measured at the stand level in oa in an area restored with prescribed burning 11 years previous. comparative measurements in two exclosures and adjacent open areas indicated that regeneration in the burned area was browsed intensively and growth of young trees was retarded, except spruce. the major impacts of browsing on aspen and rowan identify the need for new approaches to maintain forest diversity. a crucial issue will be the contradiction between preferred and sustained high moose harvests and the desire for natural forest diversity in conservation areas. alces vol. 45: 49-58 (2009) key words: alces alces, browsing, tree species, diversity, moose, forest conservation. moose (alces alces) herbivory plays an essential role in the dynamics of natural forests (risenhoover and maass 1987, pastor and naiman 1992, persson et al. 2000). in intensively managed boreal forests of finland, natural or unmanaged ecosystems are maintained in relatively small conservation areas where natural processes are the sole disturbance factor. even during a long unmanaged history, the characteristics of conserved forests are probably affected to some degree by outside influences. for example, forest management over large areas has favored moose populations by creating abundant regeneration similarly as natural disturbance processes such as fire (linder et al. 1997, lavsund et al. 2003). continuous forest management more effectively maintains moose populations than the sporadic occurrence of forest fires occurring irregularly in natural forests. consequently, the moose density in natural forests adjacent to managed forests is often higher than expected relative to normal regeneration rate and turnover of a natural forest. moose in conserved forests in finland heikkilä & tuominen alces vol. 45, 2009 50 the impacts on vegetation by moose, expressed through selective browsing (bergström and hjeljord 1987), have become acute in the case of certain tree species like balsam fir (abies balsamea) and aspen (populus spp.) in north america (brandner et al. 1990, kay 1997) and fennoscandia (kouki et al. 2004). attention has been paid to the question of population overabundance and generally to the role of moose as a disturbance factor in managed forests (gill 1992, edenius et. al 2002). in the long term, high-density moose populations can damage forest habitats in the absence of predation or human control (mclaren et al. 2004). heikkilä et al. (2003) suggested that browsing can cause considerable change in the early successional habitats of managed and natural forests in finland. nature conservation preserves are often located in close proximity to managed forests that are occupied by moose. depending on when preserves are established, they may reflect characteristics of previous forest management and fluctuations in the moose population. restoration of managed forests for conservation purposes often requires treatment strategies that allow natural processes (e.g., burning). such work has occurred in both older and more recently conserved areas in finland. the effects of the traditional, annual concentration of moose on their winter range (lavsund 1987, andersen 1991) may be important in this respect, because young conserved areas may still retain the characteristics of managed forests preferred by moose. further, intensive browsing may occur in older conserved areas from moose seeking sanctuary from nearby hunted areas. it has been suggested that some hunting activity should be permitted in conserved forest areas in finland (ympäristöministeriö 2006). in this study we analyzed the effects of moose browsing on a conserved forest area currently being restored in liesjärvi national park in southern finland. it was hypothesized that moose browsing may alter tree species composition and forest diversity. we questioned whether possible changes were unnatural and conflicted with goals to maintain this natural conservation area. study area the study area was the liesjärvi national park established in southern finland in 1956 (fig. 1). part of the original 660 ha area had been under forest management resulting in middle-aged and old closed forest; no inspection was made to classify the structure of the forest afterward. without disturbance factors, norway spruce (picea abies) became dominant over less shade tolerant broadleaved species and scots pine (pinus sylvestris). shaded undergrowth trees were common with silver birch (betula pubescens) and downy birch (betula pubescens) more dominant fig. 1. the location of liesjärvi national park in southern finland. the geobotanical vegetation zones are after kalliola (1973: 1 = hemiboreal zone, 2 = southern boreal zone, 3 = middle boreal zone, and 4 = northern boreal zone. alces vol. 45, 2009 heikkilä & tuominen moose in conserved forests in finland 51 than aspen and rowan (sorbus aucuparia). new regeneration occurred mostly as shaded undergrowth in small natural gaps and at forest edges. in 1993 an area of about 3 ha was burned to restore a dry forest site dominated by scots pine. in 2005 the park was enlarged to 2,200 ha by adding a new area where intensive forest management had not occurred for about 20 years. despite initial restoration measures, the new area (na) was less natural than the original, older area (oa) relative to forest age structure, although it had denser young stages and an increased proportion of moist sites. the characteristics of a managed forest were still prominent in the forest stands of different age classes. the average density of moose in the area was 3.1-4.0 moose/1,000 ha (ruusila et al. 2006). hunter surveys (information from game management association of tammela) indicated that the local winter density of moose after hunting was 4.6 moose/1,000 ha (range 4.3-5.1) in 2000-2005. methods forest inventory data, browsing intensity, and pellet groups were measured in both the oa and na portions of the park in 2005. we established 50 m2 circular plots on parallel transects with a distance of 100 m between transects and plots. because sampling in the oa was intended to be relatively intensive, the plot size was reduced to 20 m2 to measure all trees within a plot; trees >6 m high and fecal pellet groups were measured in the 50 m2 plots. the height of all tree species ≤6 m was measured to an accuracy of +10 cm. the diameter at breast height (1.3 m) was measured on trees >6 m high. to assess forest structure, the abundance of tree species was calculated in 4 height categories: ≤ 2 m, 2.14 m, 4.1-6 m, and >6 m. five tree condition categories were described: healthy, lightly damaged (no effect on growth), moderately damaged (growth loss obvious), seriously damaged (retarded growth and development), and dead. browsing intensity, including shoot and bark damage, was scaled in 4 categories according to a visual estimate of lost biomass: 0-25%, 26-50%, 51-75%, and >75%. the most intensively utilized trees were classified as bushy deformed. fresh moose browsing points on side shoots were counted to estimate consumed biomass according to the bite diameter and weight calculations for tree species (heikkilä and härkönen 1993); new and old main stem breakages were counted. old and new bites and breakages were distinguished according to the point and color of browsing. new (lying on the forest litter) and old fecal pellet groups (at least 20 pellets) were counted to compare presence and habitat use of moose in the study areas (neff 1968, härkönen and heikkilä 1999). we also used 2 exclosures built in the oa 10 years earlier where a 3 ha area was restored by burning. data were collected both inside and outside of 9, 20 m² sample plots in the 25 m x 25 m exclosures. height of tree species was measured according to the landscape-level inventory. statistical analyses were performed using sas program version 9.3.1 (sas institute, 2005), and comparisons between areas were made with a non-parametric mann-whitney u-test. spearman rank correlation at the sample plot level was used to analyze habitat selection according to fecal pellet groups and characteristics of the forests. height differences between exclosures and open areas were analyzed with one-way anova (spss package). results norway spruce and downy birch dominated the tree species composition in oa, and spruce, scots pine, silver birch, and other broadleaf species were abundant in na. more willow, rowan, and pine occurred in na (table 1). the availability of palatable trees (excluding spruce) ≤6 m in height was 447 trees/ha moose in conserved forests in finland heikkilä & tuominen alces vol. 45, 2009 52 greater in oa than in na. the mean height of trees ≤6 m was greater in na than in oa for all species except grey alder (alnus incana) (table 2). the number of trees ≤4 m was 9060/ha in oa and 6060/ha in na; about one third was spruce in both areas. the total tree density in na was twice that in oa for trees in the 2.1-4 m height category (1633/ha vs. 826/ha); pines and broadleaf trees accounted for 76% of trees in na and 62% in the more spruce-dominated oa. the proportion of trees ≤2 m was considerably higher in oa than in na (88% vs. 70%), and their density in oa was twice that in na. moose browsing occurred on 25% of trees in oa, whereas 42% of trees were browsed in na (fig. 2). intensive browsing (>75%) occurred on 18.6% of trees in na and 3.1% in oa. browsing was 30% on scots pine in both areas. aspen, rowan, and willows were intensively browsed in na. bushy deformed trees totaled 224 trees/ha in oa, and included pine, aspen, downy birch, rowan, and willows. in na only 48 trees/ha were bushy deformed of which >90% were willows and the rest rowans. the mean bite diameter was 2.76 oa na u-value p-value pine 904 ± 162 1550 ± 146 6.7290 <0.0001 spruce 4655 ± 948 2334 ± 230 -1.8487 0.0645 silver birch 298 ± 70 343 ± 64 3.5385 0.0004 downy birch 2441 ± 626 1898 ± 238 2.1439 0.0320 aspen 159 ± 39 217 ± 55 1.3537 0.1758 willows 255 ± 165 441 ± 106 4.5307 <0.0001 rowan 928 ± 168 1490 ± 211 4.2995 <0.0001 juniper 57 ± 52 23 ± 14 1.4411 0.1495 grey alder 46 ± 23 188 ± 65 3.6821 0.0002 other spp 23 ± 14 table 1. mean density (± se) of tree species (trees/ha) in liesjärvi national park, southern finland, 2005. the park was established in 1956 and consisted of two areas with different histories of forest management and moose population density; the original area was designated oa (old area), and a new area added in 2005 was designated na (new area). oa na u-value p-value pine 109.8 ± 5.0 193.0 ± 5.6 -11.6159 <0.0001 spruce 91.7 ± 2.8 133.3 ± 3.0 19.6262 <0.0001 silver birch 188.8 ± 13.2 242.2 ± 8.4 -5.0225 <0.0001 downy birch 118.1 ± 2.8 213.4 ± 4.1 20.6175 <0.0001 aspen 76.4 ± 8.0 151.6 ± 7.6 -7.4651 <0.0001 willows 63.7 ± 2.3 157.6 ± 5.2 -14.2084 <0.0001 rowan 76.7 ± 3.6 141.2 ± 2.6 -18.0767 <0.0001 juniper 58.1 ± 2.9 161.6 ± 13.6 4.0430 <0.0001 grey alder 237.6 ± 35.7 181.3 ± 12.5 0.9246 0.3552 other spp 65.0 ± 13.0 table 2. the mean (± se) height (cm) of tree species ≤6 m high in liesjärvi national park, southern finland, 2005. the park was established in 1956 and consisted of two areas with different histories of forest management and moose population density; the original area was designated oa (old area), and a new area added in 2005 was designated na (new area). alces vol. 45, 2009 heikkilä & tuominen moose in conserved forests in finland 53 mm (± 0.19 se) in na and 2.03 mm (± 0.21 se) in oa. the lowest diameter occurred on downy birch (1.64 mm ± 0.11 se) and the highest on pine (3.52 mm ± 0.11 se). thicker than average bites were taken from rowan and pine in both areas, and from aspen and alder in na. the estimated consumption of new twig biomass was 2.44 kg/ha in na and 0.09 kg/ha in oa. the number of old and new stem breakages was higher in oa than na (table 3). the stem damage/tree was consistently higher in oa than na for all tree species except juniper (juniperus communis). rowan had the highest number of breakages per tree in both oa (9.3) and na (0.9). silver birch, willows, and aspen were also broken more than once in oa. in general, the ≤6 m high trees in oa were less seriously affected than those in na. the proportion of dead trees was 16% in na and 6.5% in oa. of species preferred by moose, 65% of aspen, rowan, and willows were affected in both oa and na; dead trees were 7% of the total in oa and 30% in na. the oa na u-value p-value new twig browsing/ha 909.0 ± 48.9 155.0 ± 9.5 2.8777 0.0040 old stem breakage/ha 5294.0 ± 256.2 2111.0 ± 159.0 -24.5731 <0.0001 new stem breakage/ha 67.0 ± 4.0 50.0 ± 2.0 5.6691 <0.0001 stem breakage/tree 2.9 ± 0.1 0.4 ± 0.0 12.5000 0.0110 pellet groups/ha 18.6 ± 6.3 80.3 ± 20.5 4.0442 <0.0001 table 3. damage associated with moose browsing and total number of fecal pellet groups in liesjärvi national park, southern finland, 2005. the park was established in 1956 and consisted of 2 areas with different histories of forest management and moose population density; the original area was designated oa (old area), and a new area added in 2005 was designated na (new area). fig. 2. proportions of 4 categories of moose browsing intensity on tree species in liesjärvi national park in southern finland. the park was established in 1956; the original area was designated oa (old area), and a new area added in 2005 was designated na (new area). moose in conserved forests in finland heikkilä & tuominen alces vol. 45, 2009 54 proportion of live aspen was 75% in oa (126 aspens/ha) and 62% in na (114 aspens/ha). most live trees 4.1-6.0 m tall (313/ha in oa and 260/ha in na) were spruce and downy birch (79%) in oa, and downy birch and pine (67%) in na. all aspen (2.1/ha) were dead and willows were absent in oa. in na 5.9 aspens/ha were live of which 4.4/ha were injured; no dead aspens were found. in oa 16.3 rowans/ha were healthy or injured, whereas 12.0 rowans/ha were injured and 3 rowans/ ha were dead in na. grey alder was either injured or dead in both oa and na. conifers were the dominant trees >6 m tall in both oa (81%) and na (70%). aspen, rowan, and willows accounted for 2% of trees >6 m in oa and 3.5% in na. in oa 13% of the 48 aspens/ha were dead; 4% of 33 aspens/ha were dead and 22% affected by bark stripping in na. the maximum stem diameter of aspen was ≤20 cm and that of rowan and willows ≤15 cm. the diameter of the largest conifers was 46-50 cm in oa and 36-40 cm in na. the abundance of fecal pellet groups indicated that moose used na more than oa during the previous winter (80.3 groups/ha ± 20.5 se vs. 18.6 ± groups/ha ± 6.3 se, mann-whitney u = 4.0442, p <0.0001). in oa the number of fecal pellet groups correlated positively with the number of pines <2.5 m tall (r = 0.21, p = 0.004), the number of stem breakages of silver birch (r = 0.318, p <0.0001) and aspen (r = 0.168, p = 0.023), and the total number of moose-affected pines and downy birches (r = 0.465, p <0.0001). in na the number of pellet groups correlated positively with the number of pines <2.5 m tall (r = 0.198, p = 0.02). the fresh browsing of pine correlated with pellet groups (r = 0.332, p <0.0001); fresh willow browsing was nearly significant (p = 0.08). the number of pellet groups increased with the number of stem breakages on pine (r = 0.210, p = 0.016) and silver birch (r = 0.216, p = 0.011), and with rowans that were seriously damaged (r = 0.179, p = 0.037) or dead (r = 0.250, p = 0.003). measurements within the exclosures indicated that moose reduced (p <0.05) the growth (height) of pine, silver birch, and aspen during the 11 year period post-burn (fig. 3). downy birch was uncommon and spruce was not browsed. aspen was 1 m higher in the exclosures, and rowan did not occur outside the exclosure. willows were absent inside the exclosures. silver birch and pine combined accounted for about 80% of the total sapling density per ha, about 7,500 inside and 10,500 outside the exclosures. discussion liesjärvi national park in finland provides an excellent example of how previous and current forest and moose management can influence natural forest succession in a conservation area. the management history of the original portion of the park (oa) was different from that added later (na), both in forest management and moose population density. the forests in oa developed with minimal disturbance in the 1950-1970s when the moose population density was relatively low thereby allowing tree species typical of moose forage to grow. the subsequent rapid increase in moose population density and the forest management history in na (added 0 50 100 150 200 250 pinus sylvestris picea abies betula pendula betula pubescens populus tremula inside outside height, cm *** *** *** fig. 3. height of tree species inside and outside two exclosures in a forest restoration treatment area in oa, 10 years post-burn, liesjärvi national park, southern finland. alces vol. 45, 2009 heikkilä & tuominen moose in conserved forests in finland 55 in 2005) created the structural differences evident in forests in oa and na. the longterm impact from browsing that occurred after the 1970s retarded the development of young trees in oa, and created a low-growing community of palatable tree species (table 2; risenhoover and maass 1987, abaturov and smirnov 2002). the recent intensive browsing in na (>75 % browsing of 18.6 % of trees vs. 3.1% in oa) is reflected in the high proportion of injured and dead trees. in this short time na experienced more damage than oa (table 2, fig. 2). however, the availability of palatable trees in na was greater than in oa, especially in the 2.1-4.0 m height class that contains relatively abundant moose forage (parker and morton 1978, heikkilä and härkönen 1998). in the absence of past disturbances other than browsing, the oa forest developed towards spruce dominance (table 1). this trend gradually reduced habitat value for moose that prefer highly available deciduous forage (saether and andersen 1990, ball et al. 2000) that is associated with high moose population densities. however, an abundant moose population affects tree species diversity in ways that are difficult to predict (mclaren et al. 2004). because the local management goal is to maintain moose populations at levels that provide widespread hunting opportunity, understanding browsing impacts by high populations is critical. our data indicate that browsing pressure within a conservation area needs to be related to management goals in surrounding managed forests to best ascertain the effects on biodiversity within conservation areas. the high pellet group density in na indicated that a concentrated winter population of moose may reduce and alter diversity of tree species (andersen 1991). generally, the pellet group density increased with browsing intensity which is consistent with earlier findings (heikkilä and härkönen 1993). in both oa and na the number of pellet groups correlated positively with the density of young pine that is palatable winter browse (lundberg et al. 1990). the positive correlation between pellet groups and stem breakage on aspen in oa indicated higher and more frequent use by moose than in conserved forests in koli national park in eastern finland where moose used aspen much less (härkönen et al. 2008). gap dynamics may keep aspen inaccessible to moose and completely avoid browsing damage (syrjänen et al. 1994, cumming et al. 2000, edenius et al. 2002). local moose population density may also influence whether aspen escapes browsing damage. the risk of reduced biodiversity is recognized in managed forests located in highdensity moose winter range (heikkilä and härkönen 1993). most concern is for rowan and especially aspen, one of the most threatened tree species (kouki et al. 2004). although rowan was relatively abundant in liesjärvi national park, only a few individuals grew beyond the browsing height of moose. because rowan does not resist intensive browsing well, it commonly remains low-growing (saether 1990). rowan disperses widely even from a few seed producing trees, whereas aspen reproduces mainly from suckering (zackrisson 1985). aspen was much less abundant than rowan in both oa and na, and only a few were >2 m-≤ 6 m high. no live aspen was found in the 4.1-6.0 m height class in oa indicating a lack of recruitment of dominant trees and risk for future forest diversity. there are several ways of enhancing the diversity of intensively browsed tree species beyond population management of large ungulates, for example, as in aspen communities of conserved forests in north america (suzuki et al. 1999, kaye et al. 2005, romme et al. 2005). the strategy of aspen to resist ungulate browsing presupposes favorable conditions for regeneration. small-scale restoration by burning at the stand level was attempted in liesjärvi national park to compensate for moose in conserved forests in finland heikkilä & tuominen alces vol. 45, 2009 56 the lack of natural disturbance. however, no young aspens in the burned area grew beyond moose browsing after 10 years. rowan, preferred by moose year-round, was also absent due to intensive browsing, and even low-density moose populations can influence tree species composition. our results suggest that a small-scale restoration needs to be supplemented with protection against moose browsing; partial fencing might ensure aspen regeneration (mclaren et al. 2004, edenius and ericsson 2007). one crucial issue is the conflict between a sustained high harvest desired by hunters and the need for natural high forest diversity. diversity may vary widely in conserved areas (kouki et al. 2004); for example, härkönen et al. (2008) reported that browsed aspens recover well at low moose density. however, moose populations often impact development of conserved forests by retarding primary succession and causing conifer dominance (davidson 1993). further, the absence of natural disturbance results in a closed spruce-dominated boreal forest (linder et al. 1997). high density moose populations require large-scale disturbances to create preferred regeneration habitat to maintain balance between browsing pressure and forage availability. our data indicate that the continuous selective browsing pressure by moose in liesjärvi national park gradually reduced forage diversity and availability. it is clear that potentially threatened species and the composition and availability of all forage trees need to be addressed in management plans because management practices to date have not prevented a critical drop in forest health and diversity in liesjärvi national park. a cooperative decision-making process among adjacent landowners and moose managers is needed to help establish and maintain natural development in recently conserved forest areas. acknowledgements we would like to thank mr. jorma sillanpää for the help in carrying out field work, and 2 anonymous referees for their valuable comments on the manuscript. references abaturov, b. d., and k. a. smirnov. 2002. effects of moose population density on development of forest stands in central european russia. alces supplement 2: 1-5. andersen, r. 1991. habitat deterioration and the migratory behavior of moose (alces alces l.) in norway. journal of applied ecology 28: 102-108. ball, j. p., k. danell, and p. sunesson. 2000. response of a herbivore community to increased food quality and quantity: an experiment with nitrogen fertilizer in a boreal forest. journal of applied ecology 37: 247-255. bergström, r., and o. hjeljord. 1987. moose and vegetation interactions in northwestern europe and poland. swedish wildlife research supplement 1: 213-228. brandner, t. a., r. o. peterson, and k. l. risenhoover. 1990. balsam fir on isle royale: effects of moose herbivory and population density. ecology 71: 155164. cumming, s. g., f. k. a. schmiegelow, and p. j. burton. 2000. gap dynamics in boreal aspen stands: is the forest older than we think? ecological applications 10: 744-759. davidson, d. w. 1993. the effects of herbivory and granivory on terrestrial plant succession. oikos 68: 23-35. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forests. silva fennica 36: 57-67. _____, and g. ericsson. 2007. aspen demographics in relation to spatial context and ungulate browsing: implications for conservation and forest management. biology and conservation 135: 293-301. alces vol. 45, 2009 heikkilä & tuominen moose in conserved forests in finland 57 _____, _____, and p. näslund. 2002. selectivity by moose vs. spatial distribution of aspen: a natural experiment. ecography 25: 289-294. gill, r. m. a. 1992. a review of damage by mammals in north temperate forests. 3. impact on trees and forests. forestry 65: 363-388. heikkilä, r., and s. härkönen. 1993. moose (alces alces l.) browsing in young scots pine stands in relation to the characteristics of their winter habitats. silva fennica 27: 127–143. _____, and _____. 1998. the effects of salt stones on moose browsing in managed forests in finland. alces 34: 435-444. _____, p. hokkanen, m. kooiman, n. ayguney, and c. bassoulet. 2003. the impact of moose browsing on tree species composition in finland. alces 39: 203-213. härkönen, s., k. eerikäinen, r. lähteenmäki, and r. heikkilä. 2008. does moose browsing threaten european aspen regeneration in koli national park, finland? alces 44: 31-40. _____, and r. heikkilä. 1999. use of pellet group counts in determining density and habitat use of moose alces alces in finland. wildlife biology 5: 233-239. kalliola, r. 1973. suomen kasvimaantiede. (the finnish botanical geography) werner söderström osakeyhtiö, porvoo, finland. (in finnish). kay, c. e. 1997. is aspen doomed? journal of forestry 95 (5): 4-11. kaye, m. w., d. binkley, and t. j. stohlgren. 2005. effects of conifers and elk browsing on quaking aspen forests in the central rocky mountains, usa. ecological applications 15: 1284-1295. kouki, j., k. arnold, and p. martikainen. 2004. long-term persistence of aspen – a key host for many threatened species – is endangered in old-growth conservation areas in finland. journal of nature conservation 12: 41-52. lavsund, s. 1987. moose relationships to forestry in finland, norway and sweden. swedish wildlife research supplement 1: 229-244. _____, t. nygren, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. linder, p., b. elfving, and o. zackrisson. 1997. stand structure and successional trends in virgin boreal forest reserves in sweden. forest ecology and management 98: 17-33. lundberg, p., m. åström, and k. danell. 1990. an experimental test of frequencydependent food selection: winter browsing by moose. holarctic ecology 13: 177-182. mclaren, b. e., b. a. roberts, n. djanchékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 45-59. neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. journal of wildlife management 32: 597-614. parker, g. r., and l. d. morton. 1978. the estimation of winter forage and its use by moose on clearcuts in northcentral newfoundland. journal of range management 31: 300-304. pastor, j., and r. j. naiman. 1992. selec-selective foraging and ecosystem processes in boreal forests. american naturalist 139: 690-705. persson, i.-l., k. danell, and r. bergström. 2000. disturbance by large herbivores in boreal forests with special reference to moose. annales zoologici fennici 37: 251-263. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royal forests. canadian journal of forest research 17: 357-364. moose in conserved forests in finland heikkilä & tuominen alces vol. 45, 2009 58 romme, w. h., m. g. turner, g. a. tuskan, and r. a. reed. 2005. establishment, persistence, and growth of aspen (populus tremuloides) seedlings in yellowstone national park. ecology 86: 404-418. ruusila, v., m. pesonen, r. tykkyläinen, a. karhapää, and m. wallén. 2006. hirvikannan koko ja vasatuotto vuonna 2005. (moose population size and calf production in 2005). riistantutkimuksen tiedote 211. 7 s. (in finnish). saether, b.-e. 1990. the impact of differ-the impact of different growth patterns on the utilization of tree species by a generalist herbivore, the moose alces alces: implications for optimal foraging theory. behavioural mechanisms of food selection. nato asi series, volume g 20: 323-340. _____, and r. andersen. 1990. resource limitation in a generalist herbivore, the moose alces alces: ecological constraints on behavioural decisions. canadian jour-canadian journal of zoology 68: 993-999. suzuki, k., h. suzuki, d. binkley, and t. j. stohlgren. 1999. aspen regeneration in the colorado front range: differences at local and landscape scales. landscape ecology 14: 231-237. syrjänen, k., r. kalliola, a. puolasmaa, and j. mattsson. 1994. landscape structure and forest dynamics in subcontinental russian european taiga. annales zoo-ian european taiga. annales zoo-annales zoologici fennici 31: 19-34. ympäristöministeriö (the ministry of environment). 2006. metsästys eteläisen suomen kansallispuistoissa. (hunting in the national parks of southern finland). ympäristöministeriön asettaman työryhmän raportteja 10/2006. (in finnish). zackrisson, o. 1985. some evolutionary aspects of the life history characteristics of broadleaved tree species found in the boreal forest. pages 17-36 in b. hägglund and g. peterson, editors. broadleaves in boreal silviculture an obstacle or an asset? swedish university of agricultural sciences, department of silviculture. report 14. alces29_267.pdf alces22_419.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces27_183.pdf alces26_154.pdf alces(23)_157.pdf alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces vol. 23, 1987 alces24_62.pdf alces(25)_118.pdf alces(25)_52.pdf alces29_219.pdf alces21_509epilogue.pdf alces vol. 21, 1985 alces26_44.pdf alces26_80.pdf fine-scale temperature patterns in the southern boreal forest: implications for the cold-adapted moose bryce olson1,2, steve k. windels1, mark fulton2, and ron moen3 1national park service, voyageurs national park, 360 highway 11 e, international falls, minnesota 56649; 2bemidji state university, 1500 birchmont drive ne, bemidji, minnesota 56601; 3natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811 abstract: moose (alces alces) respond to warm temperatures through both physiological and behavioral mechanisms. moose can reduce heat load via habitat selection when spatial and temporal variation exists within the thermal environment. we recorded operative temperatures (to) throughout the kabetogama peninsula of voyageurs national park, minnesota for 1 year to describe seasonal patterns in the thermal environment available to moose and identify physical and landscape characteristics that affect to in southern boreal forests. significant predictors of to varied by season and time of day and included vegetation cover type, canopy cover, and slope/aspect. vegetation cover type influenced to during summer and fall afternoons with additional variation during summer afternoons explained by percent canopy cover. slope/aspect was the main driver of to during winter and spring afternoons. slope position was not a significant predictor of temperature, likely because of low topographic relief in our study area. the tos were significantly warmer in open versus closed habitats during the day with the pattern reversed at night. our results can be used to test if moose display a behavioral response to to at various spatial and temporal scales. alces vol. 50: 105–120 (2014) key words: alces alces, aspect, canopy cover, cover type, forests, moose, operative temperature, topography. moose (alces alces) are adapted to cold environments (karns 2007) but conversely, are less tolerant of high ambient temperatures (ta). renecker and hudson (1986) estimated the upper critical temperatures (tuc) of moose as −5 °c in winter and 14 °c in summer, with open mouth panting occurring at 0 °c and 20 °c, respectively. a recent study estimated a slightly higher tuc in summer (17 °c; mccann et al. 2013). these estimates of tuc provide a lower limit of ta at which moose presumably employ physiological and behavioral mechanisms to reduce thermal stress. moose respond physiologically to high ta by reducing metabolic rate, flattening their pelage, and increasing respiratory rate to expel excess heat, but they cannot sweat (schwartz and renecker 2007). they also exhibit behavioral responses including higher use of conifer stands for thermal refuge and nocturnal activity (demarchi and bunnel 1995, dussault et al. 2004, broders et al. 2012). chronic exposure of moose to high ta has been correlated with reduced weight gain in norway (van beest and milner 2013), lower survival in northeastern minnesota (lenarz et al. 2009), population declines in northwestern minnesota (murray et al. 2006), and distribution shifts in china (dou et al. 2013). fine-scale differences in ta likely exist across space and time at the southern extent of moose range, and individual moose should exploit these differences to mitigate the effects of high ta on body 105 condition and ultimately fitness. previous studies of moose habitat selection and ta focused on forest cover type as the main driver of thermal conditions across the landscape (e.g., lowe et al. 2010). other factors affecting variability in the thermal landscape include elevation, canopy cover, slope and aspect, and position on slope (reifsnyder et al. 1971, chen and franklin 1997, danielson et al. 1997, chen et al. 1999, ellis and pomeroy 2007). habitat selection patterns relative to ta cannot be fully understood without a clear understanding of patterns in the thermal environment at different spatial and temporal scales. operative temperature (to) is an approximation of the convective and radiant heat transfer on the surface of an animal, making it a useful measure to interpret the thermal environment experienced by animals versus ta alone (dzialowski 2005). for example, animals experience different to in sunlight, wind, or under forest canopy at the same ta. it is easiest to estimate to with a black globe thermometer (vernon 1930, 1932, 1933; fig. 1) which consists of a matte black painted copper sphere containing a temperature logger that integrates ta, mean radiant temperature, and air movements into a single metric (bedford and warner 1934). our objectives were to identify physical and vegetative factors that influence to, and to characterize the thermal environment experienced by moose across different cover types in voyageurs national park (vnp) in northeastern minnesota. study area voyageurs national park (vnp) is situated on the southern limit of north american fig. 1. cross-section and attached black globe thermometer showing hobo u22 water temp pro v2 temperature loggers inserted into a 15 cm diameter copper toilet bowl float painted matte black. loggers were hung 0.75 m above ground and 15 cm from the trunk on the northeast side of a tree. 106 forest temperature patterns – olson et al. alces vol. 50, 2014 moose range, along the minnesota-ontario border (fig. 2). the climate is mid-continental with long cold winters and short cool summers. mean monthly temperatures range from −15 °c in january to 19 °c in july with an annual mean temperature of 3 °c (noaa 2010). first snowfall usually occurs in early november and final snowfall in early april. average annual precipitation is 61 cm, with an average annual snowfall of 183 cm (noaa 2010). we limited the study area to the 329 km2 kabetogama peninsula as this is where most moose in vnp currently reside (windels 2014, fig. 2). vegetation in the kabetogama peninsula is typical of the southern boreal and laurentian mixed conifer-hardwood regions (faber-langendoen et al. 2007). forest cover is a mosaic of quaking aspen (populus tremuloides), paper birch (betula papyrifera), balsam fir (abies balsamea), and jack (pinus banksiana), red (p. resinosa), and white pine (p. strobus). a variety of wetlands including bogs, fens, marshes, and swamps are interspersed across the landscape (faber-langendoen et al. 2007). geological features include thin and sandy topsoil with regions of exposed bedrock (ojakangas and matsch 1982). moose and white-tailed deer (odocoileus virginianus) are the only ungulate species in vnp; woodland caribou (rangifer tarandus) were extirpated by the early 1900s (cole 1987). moose density in the kabetogama peninsula ranges from 0.14–0.19 moose/km2 and has remained stable since the 1990s (windels 2014). beaver (castor canadensis) are abundant and contribute significantly to the spatial heterogeneity of the landscape (johnston and naiman 1990). a variety of human and other natural disturbances have created a diverse mosaic of fig. 2. location of the kabetogama peninsula study area in voyageurs national park, minnesota, usa. alces vol. 50, 2014 olson et al. – forest temperature patterns 107 vegetation in multiple seral stages. wildfires and extensive logging occurred in the 1920s and 1930s, followed by less intensive logging through the 1960s (gogan et al. 1997). major fires occurred throughout vnp in 1923 and 1936. suppression of fire followed until the late 1980s when the national park service implemented a wildland fire management plan, though most prescribed burns have been relatively small (national park service, unpublished data). methods to measure to across the kabetogama peninsula, we stratified our sampling design by landscape and vegetation characteristics identified within a 30-m × 30-m pixel matrix that matches with landsat imagery. we derived slope and aspect for each pixel using 30 m resolution shuttle radar topography mission data (version 1) (rabus 2003, rodriguez et al. 2005, 2006). we estimated slope using the slope function from erdas imagine (erdas inc. 2010). we categorized slope as either <10% (5.7°) or ≥10% (5.7°). we calculated aspect using the aspect function of erdas imagine (erdas inc. 2010) and classified each pixel into 1 of 2 categories: aspects between 315-45° (i.e., north) and aspects between 45-315° (i.e., east/south/west). we assumed that solar inputs would be lower on north facing slopes compared to east, south, and west facing slopes. therefore, we combined slope and aspect into flat, slopes facing north, and slopes facing east/south/west for analysis. detectable variation in ta as a function of elevation was not expected in vnp as most local relief is <30 m and maximum relief within the peninsula is only 81 m. therefore, elevation was not considered in our sampling design or subsequent modeling. we developed a canopy cover model using the methodology outlined in the great lakes inventory and monitoring network’s landscape dynamics protocol (kennedy and kirschbaum 2010, kennedy et al. 2010). percent cover of trees, shrub, and ground layer was estimated at 30-m pixels in arcmap (esri 2011) using high-resolution air photos taken in the spring (leaf-off) and summer (leaf-on) of 2008 with 0.15 and 1 m resolutions, respectively (kirschbaum and gafvert 2010). estimates of canopy cover in each pixel were related to the normalized burn ratio (van wagtendonk et al. 2004) calculated from the landsat image corresponding to that time period to create a regression model of canopy cover. we categorized canopy cover as open (i.e., no or few canopy-forming trees), variable cover (i.e., non-forested, discontinuous canopy), <70% forest cover, 70–80% forest cover, and >80% forest cover. we classified vegetation cover type using the national vegetation classification system subclass level (deciduous, evergreen, mixed, woodland, shrub, or herbaceous) developed for vnp (faberlangendoen et al. 2007). we developed a sampling matrix using the 3 sets of variables (vegetation cover type, canopy cover, slope/aspect). each unique combination of the 3 variables was given an identifying code using erdas imagine. a 30-m pixel raster map was created and converted to a polygon shapefile. areas for each polygon were calculated and polygons <1.07 ha (12 pixels) were deleted to avoid sampling very small patches. we sampled at an intensity of 1 temperature logger for every 333 ha in the study area. we randomly selected polygons to sample from each of 38 unique combinations of vegetation cover type, canopy cover class, and slope/aspect class. we located the sample point near the centroid of the polygon to allow a sufficient buffer between adjacent polygons. we used alternate sites if the selected polygon centroid was <30 m (1 pixel) from the edge, field reconnaissance found that site characteristics were different 108 forest temperature patterns – olson et al. alces vol. 50, 2014 from remotely-sensed data, or sites were otherwise inaccessible (e.g., flooding). black globe temperature loggers consisted of a data-logging thermocouple (onset computer corporation, bourne, massachusetts, usa) inserted into a copper toilet tank float painted matte black (fig. 1). we calibrated loggers for a minimum of 96 h to verify accuracy and resolution as compared to the stated equipment specifications of the logger (±0.21 °c from 0°−50 °c). all loggers were synchronized and programmed to record temperatures every 15 min for 1 year. at each sample point, we hung loggers 0.75 m above the ground and 15 cm from the trunk. loggers were placed on the northeast side of trees to minimize direct solar radiation during the warmest time of day (fig. 1). we used handheld field computers with gps to verify that logger placement in the field was consistent with identified cover type and location within the cover type polygon. we used real-time gis and measurements in the field to ensure we were within the identified cover type, canopy cover class, and slope/aspect category before deploying loggers. loggers were deployed from june 2010 to july 2011, with periodic downloads to reduce risk of data loss. data were screened to remove biased or failed measurements (e.g., faulty logger, damaged globe or logger, and snow-covered loggers). we deployed an additional set of loggers from august 2011 to january 2012 to test for differences in position on slope. we randomly deployed 3 loggers in each of 9 combinations of cover type (deciduous, evergreen, and mixed), canopy cover class (<70%, 70–80%, >80% forested canopy), and slope position (top, mid-slope, and base); slopes ranged from 17–47%. we analyzed data by season with full factorial repeated-measures anova with type iii sums of squares using spss 20 (ibm corporation 2011). we defined seasons as spring (1 march–31 may), summer (1 june–31 august), fall (1 september–30 november), and winter (1 december–28 february). each 15 min logger interval was treated as the response variable but controlled for repeated measures. post-hoc pairwise comparisons were made using the bonferroni method. significance levels were set to p = 0.05 for all tests. subsamples of the dataset were made to compare differences during the 3 warmest hours of the day (1300–1600 hr) and the 3 coldest hours of the day (0300–0600 hr). we used a similar statistical approach to test for differences in slope position for summer and winter only. based on model results, we assessed the availability of potential thermal refugia to moose across the kabetogama peninsula. we simulated moose home ranges by creating 25 random points within the study area and then buffering those points to approximate mean annual moose home ranges in vnp (48 km2 ± 33.5 sd) as reported by cobb et al. (2004). home ranges varied in size due to the highly variable shoreline of the kabetogama peninsula. within each simulated home range, we estimated the proportion of each habitat type (vegetation cover type, canopy cover class, slope/aspect) and used summary statistics to highlight availability of selected habitat features. results open habitat types (shrub and herbaceous) were significantly warmer than forested habitat types during the summer (maximum difference = 3.38 °c; fig. 3) with the greatest difference occurring in the afternoon (maximum difference = 8.10 °c; table 1, fig. 3). open habitat types were also the coolest at night (maximum difference = 2.66 °c; table 1, fig. 3). mean to during the summer did not differ among forested cover types for any of the time periods (daily, hot, cold). herbaceous cover types were warmer than forested cover types in the afternoon during the fall (maximum alces vol. 50, 2014 olson et al. – forest temperature patterns 109 difference = 7.43 °c; table 1, fig. 3). temperatures in shrub cover types were intermediate between herbaceous and forested cover types. the amount of canopy influenced to within forested cover types. areas with >80% canopy coverage were cooler than those with <70% cover (table 2, fig. 4); to in the 70–80% cover class was intermediate to these. slope/aspect influenced to only during the 3 coldest hours of day in summer. flat areas were cooler than east/south/west facing slopes (difference = 1.14 °c; table 3). north-facing slopes were intermediate between flat areas and east/south/west facing slopes. slope/aspect had no influence on to during the fall months. winter temperature was only differentiated by slope/aspect during the afternoon with northern facing slopes cooler than both flat and east/south/west categories (table 3, fig. 5). spring temperatures varied across slope/aspect categories during the afternoon hours. north-facing slopes were cooler than flat areas (table 3, fig. 5). to was not different among slope positions in summer (f2,21 = 0.287, p = 0.755) or winter (f2,21 = 0.606, p = 0.556). the majority of the kabetogama peninsula consists of forested cover types with >70 percent canopy cover and flat topography (table 4, fig. 6, 7). simulated home ranges varied in size from 23–57 km2 with a mean of 46 km2 (sd = 10.1 km2). potential summer refugia, such as high-canopy cover forests, were found in about 40% of simulated home ranges (table 4). however, north-facing slopes are relatively limited in the study area and the percentage of northfacing slopes in simulated home ranges was <10% (table 4, fig. 7). discussion vegetation cover type, percent canopy cover, and slope/aspect all influenced to, although differently depending on season and time of day. vegetation cover type had the strongest influence on to during summer months and fall afternoons. this was largely fig. 3. mean operative temperatures across vegetation cover types in summer over a 24-hour period, 1 june–31 august 2010, voyageurs national park, minnesota, usa. 110 forest temperature patterns – olson et al. alces vol. 50, 2014 table 1. mean daily (24-hour mean) operative temperatures (°c) across vegetation cover types in spring (1 march–31 may 2011), summer (1 june–31 august 2010), fall (1 september–30 november 2010), and winter (1 december 2010–28 february 2011), voyageurs national park, minnesota, usa. mean operative temperature for the 3 warmest (hot) and 3 coldest (cold) hours of the day are also shown. means followed by the same letter within a row are not significantly different from each other. vegetation cover type significance season deciduous evergreen mixed woodland shrub herbaceous f5,85 p value sprint daily 3.97 ± 0.74 ab 2.89 ± 0.78 a 4.28 ± 0.73 ab 3.53 ± 1.81 ab 5.26 ± 1.48 ab 6.70 ± 1.15 b 0.86 0.496 hot 10.47 ± 0.90 a 7.95 ± 0.95 b 10.29 ± 0.90 ab 9.23 ± 2.21 abc 14.19 ± 1.80 ac 16.63 ± 1.40 c 2.59 0.051 cold −1.45 ± 0.65 a −1.71 ± 0.68 a −1.25 ± 0.64 a −1.71 ± 1.59 a −1.64 ± 1.01 a −2.39 ± 1.30 a 0.14 0.934 summer daily 18.32 ± 0.33 a 18.76 ± 0.35 a 18.42 ± 0.33 a 19.33 ± 0.60 ab 21.25 ± 0.60 bc 21.70 ± 0.42 c 2.93 0.029 hot 23.03 ± 0.57 a 23.66 ± 0.60 a 23.13 ± 0.57 a 24.52 ± 1.04 a 29.84 ± 1.04 b 31.13 ± 0.73 b 7.10 <0.001 cold 13.76 ± 0.37 a 13.97 ± 0.39 ab 13.54 ± 0.37 ab 14.68 ± 0.67 b 12.88 ± 0.47 ab 12.02 ± 0.67 a 2.86 0.045 fall daily 5.56 ± 0.65 a 5.56 ± 0.69 a 5.66 ± 0.67 a 5.99 ± 1.18 a 6.38 ± 1.18 a 7.49 ± 0.83 a 0.32 0.862 hot 9.97 ± 0.82 ab 9.26 ± 0.87 ab 9.77 ± 0.85 ab 10.48 ± 1.49 ab 14.24 ± 1.49 ac 16.69 ± 1.05 c 3.00 0.026 cold 2.48 ± 0.57 a 2.78 ± 0.60 a 2.63 ± 0.59 a 2.93 ± 1.03 a 2.23 ± 0.73 a 1.52 ± 1.03 a 0.35 0.786 winter daily −13.37 ± 0.24 a −12.99 ± 0.25 a −13.41 ± 0.25 a −13.25 ± 0.44 a −14.58 ± 0.44 a −14.20 ± 0.36 a 1.70 0.166 hot −9.20 ± 0.40 ab −9.68 ± 0.43 a −9.72 ± 0.42 a −9.15 ± 0.73 ab −9.08 ± 0.73 ab −6.95 ± 0.60 b 2.14 0.091 cold −16.17 ± 0.33 a −15.59 ± 0.35 a −16.02 ± 0.34 a −16.10 ± 0.60 a −17.22 ± 0.49 a −17.30 ± 0.60 a 1.19 0.324 a l c e s v o l . 5 0 , 2 0 1 4 o l s o n e t a l . – f o r e s t t e m p e r a t u r e pa t t e r n s 1 1 1 table 2. mean daily (24-hour mean) operative temperatures (°c) across varying amounts of canopy cover by season in spring (1 march–31 may 2011), summer (1 june–31 august 2010), fall (1 september–30 november 2010), and winter (1 december 2010–28 february 2011), voyageurs national park, minnesota, usa. mean operative temperature for the 3 warmest (hot) and 3 coldest (cold) hours of the day are also shown. means followed by the same letter within a row are not significantly different from each other. canopy cover significance season open variable <70% 70–80% >80% f 4,85 p value spring daily 6.70 ± 1.15 a 4.40 ± 1.17 a 4.47 ± 0.75 a 3.14 ± 0.77 a 3.47 ± 0.72 a 0.59 0.561 hot 16.46 ± 1.35 a 11.67 ± 1.38 a 10.99 ± 0.92 ab 8.66 ± 0.94 bc 9.00 ± 0.88 bc 1.38 0.265 cold −1.644 ± 1.01 a −2.05 ± 1.03 a −1.27 ± 0.66 a −1.87 ± 0.68 a −1.31 ± 0.64 a 0.25 0.783 summer daily 21.70 ± 0.42 a 20.29 ± 0.42 a 19.00 ± 0.35 b 18.54 ± 0.34 bc 17.96 ± 0.31 c 2.48 0.093 hot 30.93 ± 067 a 27.09 ± 0.67 b 24.39 ± 0.61 c 23.24 ± 0.59 cd 22.18 ± 0.54 d 3.71 0.031 cold 12.88 ± 0.47 a 13.35 ± 0.47 a 13.49 ± 0.39 a 13.95 ± 0.38 a 13.83 ± 0.35 a 0.38 0.688 fall daily 7.48 ± 0.83 a 6.18 ± 0.83 a 5.57 ± 0.69 a 5.73 ± 0.67 a 5.48 ± 0.65 a 0.04 0.965 hot 16.39 ± 1.01 a 12.36 ± 1.01 a 10.25 ± 0.88 b 9.79 ± 0.85 b 8.95 ± 0.81 b 0.62 0.541 cold 2.23 ± 0.73 a 2.23 ± 0.73 a 2.40 ± 0.61 a 2.69 ± 059 a 2.81 ± 0.56 a 0.13 0.879 winter daily −14.20 ± 0.36 a −13.92 ± 0.31 a −13.45 ± 0.26 a −13.28 ± 0.25 a −13.07 ± 0.24 a 0.57 0.569 hot −7.11 ± 0.53 a −8.82 ± 0.46 a −8.94 ± 0.43 a −9.70 ± 0.42 b −10.02 ± 0.40 b 1.75 0.184 cold −17.22 ± 0.49 a −16.70 ± 0.42 a −16.29 ± 0.35 a −15.98 ± 0.34 a −15.50 ± 0.33 a 1.39 0.259 1 1 2 f o r e s t t e m p e r a t u r e pa t t e r n s – o l s o n e t a l . a l c e s v o l . 5 0 , 2 0 1 4 fig. 4. mean operative temperatures across varying amounts of canopy cover in summer over a 24-hour period, 1 june–31 august 2010, voyageurs national park, minnesota, usa. table 3. mean daily (24-hour mean) operative temperatures (°c) across slope/aspect categories in spring (1 march–31 may 2011), summer (1 june–31 august 2010), fall (1 september–30 november 2010), and winter (1 december 2010–28 february 2011), voyageurs national park, minnesota, usa. mean operative temperature for the 3 warmest (hot) and 3 coldest (cold) hours of the day are also shown. means followed by the same letter within a row are not significantly different from each other. slope / aspect significance season flat east/south/west north f2,85 p value spring daily 4.15 ± 0.39 a 4.48 ± 0.88 a 2.87 ± 0.91 a 0.61 0.551 hot 11.14 ± 0.48 a 11.05 ± 1.07 ab 7.27 ± 1.10 b 3.36 0.045 cold −1.79 ± 0.35 a −1.07 ± 077 a −1.57 ± 0.80 a 0.20 0.816 summer daily 18.79 ± 0.16 a 19.01 ± 0.37 a 18.35 ± 0.40 a 2.16 0.125 hot 24.73 ± 0.28 a 24.03 ± 0.65 ab 22.30 ± 0.69 b 1.76 0.181 cold 13.03 ± 0.18 a 14.17 ± 0.42 b 14.11 ± 0.45 ab 4.92 0.011 fall daily 5.32 ± 0.33 a 6.34 ± 0.76 a 5.56 ± 0.79 a 1.54 0.225 hot 10.60 ± 0.41 a 10.92 ± 0.96 a 8.55 ± 0.99 a 1.50 0.233 cold 1.86 ± 0.29 a 3.08 ± 0.67 b 3.07 ± 0.69 ab 2.60 0.084 winter daily −13.41 ± 0.12 a −12.84 ± 0.28 a −13.74 ± 0.29 a 2.45 0.097 hot −8.89 ± 0.21 a −8.35 ± 0.47 a −11.22 ± 0.49 b 10.24 <0.001 cold −16.18 ± 0.17 a −15.84 ± 0.39 a −15.99 ± 0.40 a 0.041 0.959 alces vol. 50, 2014 olson et al. – forest temperature patterns 113 fig. 5. mean operative temperatures across slope/aspect categories in winter over a 24-hour period, 1 december 2010–28 february 2011, voyageurs national park, minnesota, usa. table 4. mean, standard deviation (sd), minimum (min), and maximum (max) percentage of vegetation cover type, canopy cover, and slope/aspect categories in 25 simulated moose home ranges in voyageurs national park, minnesota, usa. simulated home range variable mean sd min max vegetation cover type evergreen 13 5 6 23 deciduous 34 11 20 58 mixed 27 7 14 48 woodland 4 2 1 7 shrub 6 4 2 16 herbaceous 16 8 8 36 canopy cover class high 41 6 29 52 med 23 6 13 34 low 10 4 4 15 variable 10 3 5 16 open 16 8 8 36 slope / aspect east/south/west 8 4 2 13 flat 87 5 79 96 north 5 2 2 8 114 forest temperature patterns – olson et al. alces vol. 50, 2014 driven by open versus forested habitats, although we detected small but significant differences in to within closed forest habitats, similar to other studies (e.g., mcgraw et al. 2012). amount of canopy cover significantly affected to only during afternoons in the summer months. forest type and the amount of canopy cover effectively combine to reduce the amount of solar radiation that reaches the forest floor and therefore can reduce heat loading from direct solar radiation (demarchi and bunnel 1993). vegetation volume, hence canopy cover, is greatest in summer months, likewise solar angle is most direct during summer afternoons. areas with thick vegetation and dense canopy cover may serve as ideal thermal refuge for moose during the day (demarchi and bunnell 1995, dussault et al. 2004, van beest et al. 2012). although some variation exists in the amount of forested habitat types with high canopy cover within our simulated home ranges, these habitat types do not seem limited in the study area. open cover types were cooler than forested cover types during the 3 coldest hours of the day. dense vegetation and canopy cover actually retain heat within forested cover types while open cover types release more heat at night (chen et al. 1993). moose in central norway use open habitat types at night and older forested stands during daytime (bjørneraas et al. fig. 6. spatial distribution of potential summer thermal refugia across the kabetogama peninsula in voyageurs national park, minnesota, usa. forested areas with >80% canopy cover are coolest during the day (black pixels) while herbaceous and shrub cover types (gray pixels) are coolest at night. all other habitat types are shown as white. inset shows fine-scale juxtaposition of “cool” and “hot” habitats at 30-m pixel resolution. alces vol. 50, 2014 olson et al. – forest temperature patterns 115 2011). the availability of open cover types may be limited for some moose in the kabetogama peninsula. we defined 4 equal seasons in our models based on calendar months rather than the timing of leaf phenology which varies annually in response to weather events, disease, drought, and other factors (lechowicz 1984). as a consequence, our ability to detect significant differences may have been diminished for some variables, specifically canopy cover. future studies of to should consider incorporating important predictor variables that may change at relatively fine time scales. also, canopy cover estimates were based on the leaf-on period, and may not accurately reflect the true amount of canopy cover during leaf-off periods for all cover types with a deciduous tree component. the majority of the study area was flat and cooler at night than east/south/west facing slopes during summer, likely due to differences in radiant heat loss. slope/aspect was the only significant influence on to during winter months as well as spring afternoons. slope/aspect may have a stronger effect on to during winter months when the solar angle is at its lowest. these environments may serve as thermal refugia on warm days in winter and early spring as topographic exposure can influence maximum daily temperatures (bolstad et al. 1998). additionally, radiation received on flat and south facing slopes may be reflected to the body by the high reflectivity of snow in winter. more fig. 7. spatial distribution of potential winter and spring thermal refugia (i.e., northfacing slopes; black pixels) across the kabetogama peninsula in voyageurs national park, minnesota, usa. areas with flat aspect are shown for comparison (gray pixels). all other habitat types are white. inset shows fine-scale juxtaposition of north-facing slopes with other aspects at 30-m pixel resolution. 116 forest temperature patterns – olson et al. alces vol. 50, 2014 northerly locations may realize increased effect of slope/aspect, as well as areas with greater topographic relief. moose strongly selected north-facing slopes in southwestern alberta due to increased shade and browse availability (telfer 1988). north-facing slopes make up less than 5 percent of our study area and across most simulated home ranges, suggesting this type of seasonal thermal refugia may be limited for moose in our study area. we did not detect an effect of slope position on to, presumably because of the low topographic relief in the study area (danielson et al. 1997). areas with larger elevational gradients than our study area should include elevation as a variable due to the adiabatic lapse or rate change in temperature of an air mass as it changes with altitude (american meteorological society 2000). in certain studies elevation was the single strongest driver of temperature difference (e.g., lookingbill and urban 2003). moose use aquatic habitats for a variety of reasons including foraging, sodium acquisition, insect relief, and thermoregulation (peek 2007). aquatic habitats in our study area contain little to no canopy cover and related to regimes are likely similar to that of open habitat types. although moose using shallow, aquatic habitats during daytime may be exposed to direct solar radiation, they could mitigate heat loading by submerging in water. thermal variability exists at relatively fine scales across our study area due primarily to the fine mosaic of vegetation cover types, canopy coverage, and site aspect (fig. 6, 7). we detected maximum differences in mean to of ≤9 °c across all habitat types during the warmest parts of summer days. within forested habitat types, there was >2 °c difference across canopy cover categories in summer. slope/aspect accounted for as much as a 4 °c difference in to during winter and spring. even small differences in the thermal environment may be relevant for achieving individual heat balance (renecker and hudson 1990). the availability of thermal refugia will be of greater importance at the southern edge of moose range as mean annual temperature continues to rise with climate change (ipcc 2007). behavioral responses to high ta include specific microhabitat use and activity shifts in other parts of moose range (dussault et al. 2004, broders et al. 2012, van beest et al. 2012). to mitigate the effects of increasing ta, managers should promote a variety of habitat types to provide adequate thermal refugia within a typical home range while meeting other life history requirements of moose (peek 2007). acknowledgements we thank b. severud, j. warmbold, c. eckman, d. morris, and n. walker for assistance with field work and a. kirschbaum for remote sensing support. this project was funded by voyageurs national park, a grant from the usgs-nps natural resource preservation program, a grant from the u.s. national park service’s great lake research and education center, bemidji state university, the natural resources research institute at the university of minnesota duluth, and the environment and natural resources trust fund. references american meteorological society. 2000. glossary of meteorology. second edition. allen press, new york, new york, usa. bedford, t., and c. g. warner. 1934. the globe thermometer in studies of heating and ventilation. journal of hygeine 34: 458. bolstad, p. v., l. swift, f. collins, and j. regniere. 1998. measured and predicted air temperatures at basin to regional scales in the southern appalachian mountains. agricultural and forest meteorology 91: 161–176. alces vol. 50, 2014 olson et al. – forest temperature patterns 117 bjørneraas, k., e. j. solberg, i. herfindal, b. v. moorter, c. m. rolandsen, j. tremblay, c. skarpe, b. saether, r. eriksen, and r. astrup. 2011. moose alces alces habitat use at multiple temporal scales in a human-altered landscape. wildlife biology 17: 44–54. broders, h. g., a. b. coombs, and j. r. mccarron. 2012. ecothermic responses of moose (alces alces) to thermoregulatory stress on mainland nova scotia. alces 48: 53–61. chen, j., and j. f. franklin. 1997. growingseason microclimate variability within an old-growth douglas fir forest. climate research 8: 21–34. ———, ———, and t. a. spies. 1993. contrasting microclimates among clearcut, edge, and interior old-growth douglasfir forest. agricultural and forest meteorology 63: 219–237. ———, s. c. saunders, t. r. crow, r. j. naiman, k. d. brosofske, g. d. mroz, b. l. brookshire, and j. f. franklin. 1999. microclimate in forest ecosystem and landscape ecology. bioscience 49: 288–297. cobb, m. a., p. j. p. gogan, k. d. kozie, e. m. olexa, r. l. lawrence, and w. t. route. 2004. relative spatial distribution and habitat use patterns of sympatric moose and white-tailed deer in voyageurs national park, minnesota. alces 40: 169–191. cole, g. f. 1987. changes in interacting species with disturbance. environmental management 11: 257–264. danielson, e. w., j. levin, and e. abrams. 1997. meterology. mcgraw-hill, new york, new york, usa. demarchi, m. w., and f. l. bunnell. 1993. estimating forest canopy effects on summer thermal cover for cervidae (deer family). canadian journal of forestry research 23: 2419–2426. ———, and ———. 1995. forest cover selection and activity of cow moose in summer. acta theriologica 40: 23–36. dou, h., g. jiang, p. stott, and r. piao. 2013. climate change impacts population dynamics and distribution shift of moose (alces alces) in heilongjiang province of china. ecological research 28: 625–632. dussault, c., j. p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioral responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. dzialowski, e. m. 2005. use of operative and standard operative temperature models in thermal biology. journal of thermal biology 30: 317–334. ellis, c. r., and j. w. pomeroy. 2007. estimating sub-canopy shortwave irradiance to melting snow on forested slopes. hydrological process 21: 2581–2593. esri. 2011. version 9.3 user manual. environmental systems research institute, redlands, california, usa. erdas inc. 2010. erdas field guide. erdas, inc., atlanta, georgia, usa. faber-langendoen, d., n. aaseng, k. hop, m. lew-smith, and j. drake. 2007. vegetation classification, mapping, and monitoring at voyageurs national park, minnesota: an application of the u.s. national vegetation classification. applied vegetation science 10: 361– 374. gogan, p. j. p., k. d. kozie, e. m. olexa, and n. s. duncan. 1997. ecological status of moose and white-tailed deer in voyageurs national park, minnesota. alces 33: 187–201. ipcc. 2007. climate change 2007: the physical science basis. contribution of working group i to the fourth assessment report of the intergovernmental panel on climate change. geneva, switzerland. johnston, c. a., and r. j. naiman. 1990. aquatic patch creation in relation to beaver population trends. ecology 71: 1617–1621. 118 forest temperature patterns – olson et al. alces vol. 50, 2014 karns, p. d. 2007. population distrubution, density and trends. pages 125-139 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. second edition. smithsonian institution, washington d.c., usa. kennedy, r., and a. kirschbaum. 2010. standard operating procedure #4: developing landtrendr maps of change. in landsat-based monitoring of landscape dynamics in the national parks of the great lakes inventory and monitoring network (version 1.0). national resource report nps/glkn/nrr2010/221. national park service, fort collins, colorado, usa. ———, ———, u. gafvert, p. nelson, z. yang, w. cohen, e. pfaff, and b. gholson. 2010. landsat-based monitoring of landscape dynamics in the national parks of the great lakes inventory and monitoring network (version 1.0). national resource report nps/glkn/nrr-2010/221, national park service, fort collins, colorado, usa. kirschbaum, a. a., and u. b. gafvert. 2010. landsat-based monitoring of landscape dynamics at voyageurs national park, 2002–2007. natural resources technical report nps/glkn/nrtr2010/356. national park service, fort collins, colorado, usa. lechowicz, m. j. 1984. why do temperate deciduous trees leaf out at different times? adaptation and ecology of forest communities. the american naturalist 124: 821–842. lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–510. lookingbill, t. r., and d. l. urban. 2003. spatial estimation of air temperature differences for landscape-scale studies in montane environments. agricultural and forest meteorology 114: 141–151. lowe, s. j., b. r. patterson, and j. a. schaefer. 2010. lack of behavioral responses of moose (alces alces) to high ambient temperatures near the southern periphery of their range. canadian journal of zoology 88: 1032–1041. mccann, n. p., r. a. moen, and t. r. harris. 2013. warm-season heat stress in moose (alces alces). canadian journal of zoology 91: 893–898. mcgraw, a. m., r. a. moen, and l. g. overland. 2012. effective temperature differences among cover types in northeast minnesota. alces 48: 45–52. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. noaa (national oceanic and atmospheric administration). 2010. climatological data for international falls, minnesota. national climatic data center, ashville, north carolina, usa. ojakangas, r. w., and c. l. matsch. 1982. minnesota’s geology. university of minnesota press, minneapolis, minnesota, usa. peek, j. m. 2007. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. second edition. smithsonian institution, washington d.c., usa. rabus, b. 2003. the shuttle radar topography mission—a new class of digital elevation models acquired by spaceborne radar. journal of photogrammetry and remote sensing 57: 241–262. reifsnyder, w. e., g. m. furnival, and j. l. horowitz. 1971. spatial and temporal distribution of solar radiation beneath forest canopies. agricultural meteorology 9: 21–37. alces vol. 50, 2014 olson et al. – forest temperature patterns 119 renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology. 64: 322–327. ———, and ———. 1990. behavorial and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspen-dominated forests. alces 26: 66–72. rodriguez, e., c. s. morris, and j. e. belz. 2006. a global assessment of the srtm performance. photogrammetric engineering and remote sensing 72: 249–260. ———, ———, ———, c. chapin, j. m. martin, w. daffer, and s. hensley. 2005. an assessment of the srtm topographic products. technical report jpl d-31639, jet propulsion laboratory, pasadena, california, usa. schwartz, c. c., and l. a. renecker. 2007. nutrition and energetics. pages 441-478 in a.w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. second edition. smithsonian institution, washington d.c., usa. telfer, e. s. 1988. habitat use by moose in southwestern alberta. alces 24: 14–21. van beest, f. m., and j. s. milner. 2013. behavioural responses to thermal conditions affect seasonal mass change on a heat-sensitive northern ungulate. plos one 8: 1–10. ———, b. van moorter, and j. m. milner. 2012. temperature-mediated habitat use and selection by a heat-sensitive northern ungulate. animal behavior 84: 723–735. van wagtendonk, j. w., r. r. root, and c. h. key. 2004. comparisons of aviris and landsat etm+ detection capabilities for burn severity. remote sensing of environment 92: 397–408. vernon, h. m. 1930. the measurement of radiant heat in relation to human comfort. journal of physiology 70: 15. ———. 1932. the measurement of radiant heat in relation to human comfort. journal of industrial hygiene. 14: 95. ———. 1933. the estimation of solar radiation in relation to its warming effect on the human body. quarterly journal of the royal meterological society 59: 239. windels, s. k. 2014. 2014 voyageurs national park moose population survey report. natural resource data series nps/voya/nrds-2014/645. national park service, fort collins, colorado, usa. 120 forest temperature patterns – olson et al. alces vol. 50, 2014 fine-cale temperature patterns in the southern boreal forest: implications for the cold-dapted moose study area methods results discussion acknowledgements references alces24_34.pdf alces27_79.pdf alces24_102.pdf alces22_277.pdf alces vol. 22, 1986 rodgersar text box alces 22 (1986) alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces24_218.pdf alces28_175.pdf alces22_345.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces21_393.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces28_15.pdf alces21_475.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces22_181.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 improving moose population estimates in russia: accounting for distance between residential areas and track sightings vladimir m. glushkov russian institute of game management and fur farming, russian academy of agricultural sciences, kirov, russia. abstract: moose (alces alces) population density in the kirov region of russia is often overestimated when using the relationship between the distance between a residential area and the initial sighting of moose tracks. this paper presents a modified approach to provide better estimates when using this techinique. statistically valid density estimation techniques, standardization of estimation points and routes, landscape characteristics, and time have been addressed in the new approach. moose density is estimated once annually based on the distance to the first track, and annual surveys should maintain alike protocol. this improved method will provide more accurate population density estimates critical to prevent regional overharvest of moose. alces vol. 49: 149–154 (2013) key words: alces alces, density, distribution, moose, population estimate, russia, track surveys. introduction & background spatial distribution of individuals within a population is generally described as 3 types – equal, occasional, and grouped (odum 1986) – that can be affected by regional and temporal influences (naumov 1963). in estonia, moose (alces alces) distribution changes seasonally; moose in summerautumn are evenly distributed but in winter their distribution is sporadic or a “focal type of distribution” (ling 1977). likewise, moose distribution differs between summer and winter in the northeast portion of european russia (i.e., kirov region; glushkov 1982). this seasonal difference is caused by november migration related to forage deficiency on summer range (yazan 1972), as well as increased moose hunting that occurs after snow cover (glushkov 1997, 2001). the relationship between snow cover and increased harvest has not been considered previously relative to population density estimates (i.e., ecological density; bubenik 1965) that are based upon the distance between a residential area and the initial sighting of moose tracks. the unique spatial distribution caused by this relationship is neglected in typical winter route censuses (wrc), creating error in abundance estimates (glushkov 2004) and potential overharvest of moose that threatens population stability (glushkov et al. 2012). this paper provides the rationale for a modified approach to account for this relationship when calculating a population estimate. previous studies provide baseline information about seasonal moose distribution in the kirov region of russia (glushkov 1977). group size is larger in winter (2.8 ± 0.9) than in summer (2.0 ± 0.6), and dispersion:density ratios of 2.4 in november versus 5.1 in march (measured from aerial surveys within 1 min flight range of 60 ha corresponding author: vladimir m. glushkov, russian institute of game management and fur farming, russian academy of agricultural sciences, kirov, russia email: v.m.glushkov@yandex.ru 149 plots [n = 970]) confirms the more uneven winter distribution of moose. these early (november) and late winter (march) data (1976–1985) were used to construct a graph of sighting frequency in plots with varied moose density that described the character and variance of seasonal moose distribution in the kirov region (fig. 1). there were fewer unoccupied plots (0 moose) and plots occupied by ≥4 moose in november than in march when there were fewer plots with 2–3 moose. these seasonal differences are statistically different, and specific to both particular areas (χ2 = 42.7–171.4) and the kirov region as a whole (χ2 = 118.1). these data made it possible to classify summerautumn distribution of moose as “occasional” and winter distribution as “grouped with cluster formation” (glushkov 2001). the distance between human settlements and the initial observation of a moose track was measured during helicopter surveys; the area between was assumed absent of moose. this distance was compared to the sighting frequency of animals and tracks in occupied habitat. in the southern area of the region the correlation was not as strong (r = − 0.50, tr = 2.15) as in the north (r = – 0.65, tr = 3.42). similar tests were conducted with data from 288 terrestrial straight-line survey routes in 27 regional districts (2595 km total length with 288–200 ha sample plots; november 1996); the distance to the initial moose track and the population estimate was inversely related (corr. coeff. = −0.35; p = 0.002). in 10 of 15 districts surveyed, the average distance to the sighting of the first track was >7 km, and in 2 districts it was ∼9 km; tracks were first observed at a distance of 14, 16, and 25 km on the other 3 routes. because physical ability limits the intensity and extent of a terrestrial survey, an equation was developed (glushkov 1999) to calculate the probable distance to the initial track encountered (l) from the length of the route travelled where no tracks were encountered (r0): l ¼ 0.816r0þ2.98 ð1þ the histogram depicting the distribution of theoretical frequencies of plots with various moose densities indicated that the proportion of plots with 2 animals was underestimated 4 times and that of unoccupied plots was overestimated 2 times. in general, the equation to estimate population density (p) from distance (x) had little practical value (p = 9.17 − 0.54 x). the error in density estimates was presumably due to insufficient area in sample plots. a subsequent survey (1999) was carried out on 14 routes with larger sample plots (700–2200 ha) at the end of each route. the following describes this new survey approach that provides more reliable population density estimates. results and discussion figure 2 depicts the relationship between moose population density (d) and the distance (x) from a residential area to the initial (recent) track. the predictive equation was formulated with a logarithmic density fig. 1. the relationship between the sighting frequency and the number of observed moose per plot in early (november) and late (march) winter in the kirov region, russia. 150 moose population estimates in russia – glushkov alces vol. 49, 2013 function that is considered reasonable and acceptable for sample estimates based on the inequality criterion of dispersion and the mean-square deviation of p (draper and smith 1986). d ¼ � 7.7023 lnðxþ þ 20.635; r2¼ 0.8981, p < 0.001 ð2þ a verification of this equation was attempted in 2003–2004 within an experimental hunting farm (63,000 ha) by comparing the “known” moose population estimate with an estimate derived from survey routes and sample plots; the estimate was comparable and deemed satisfactory. however, this comparison is general at best because there was no differentiation between seasonal estimates (early and late winter), and no method to evaluate extrapolation across a larger area. in an attempt to verify the method in practice, and to achieve necessary reduction of the dispersion value, the number of paired observations would need to increase to 40 based on equation (2) and the value of coefficient of determination. in general, the experimental estimates were not contradictory of the hypothesis that moose density is directly related to the distance from a residential area due to their anthropophobic behavior as a result of intensive hunting. this new method is more elementary and easier to implement at the beginning of winter to estimate moose density at both the district and regional scale. its use is intended for determining abundance trends and setting seasonal harvest quotas (glushkov and buldakov 1997). specification of the starting route point, radial direction, and reference to a sample plot at the end of the route removed some associated drawbacks of the traditional wrc method. the independence of the “distance” parameter from weather conditions increases not only accuracy but also comparability of estimates. a relatively even population distribution in early winter predetermines reduction of the estimate error, and defines the “native population which inhabits a given area during summer, autumn, and early winter and is subject to hunting”, a definition critical to determine harvest level. the estimate makes it possible to determine, apart from ecological density, an area that is actually used by moose during early winter (extrapolation area), and animal numbers at the district and regional levels. comparability theory (yurghenson 1970) can be used as the basis to extrapolate population density estimates provided that data are available in a particular region to estimate density in subsequent years. it is possible to use equation (2) initially while simultaneously measuring and calculating plot estimates to improve the population estimate. if necessary, a locally specific equation can be developed from a single estimate from the plots and sample routes; calculations of the average r value and extrapolation areas are provided in glushkov (2001). application of this new method utilizes gis technology and requires preparatory work to organize permanent estimation fig. 2. the relationship between moose population density and the distance to the first track sighting in the kirov region, russia. alces vol. 49, 2013 glushkov – moose population estimates in russia 151 points and placement of routes and plots. the area unused by animals and the extrapolation area are determined with gis technology. an estimation point can be any “standard” residential area – a village, a workers’ settlement, or a farm enterprise with ≥30 people; all are recorded in reference books, marked on maps, and have a post office and permanent approach roads. the principle criteria for selecting these areas are that they are dead-end locations on a year-round motorway and representative of the surveyed lands within the district. the number of estimation points in a district depends on the total area, % forest cover, land cover diversity, and the number of settlements and their distribution. ideally, 4 radial routes with plots at the end must cover the study area completely (see fig. 3); the inner circle corresponds to the anthropogenic zone with zero moose density. moose population density within the ring with sample plots is equal to the density over the whole habitation area (outer ring, fig. 3). in districts with large forest area and lack of human settlements, the width of the ring and the area of land used by animals (extrapolation area) can be correspondingly large. in districts with small fragmented forests and densely populated settlements, the habitation area around proximate human settlements decreases by the value of the overlapping anthropogenic area; i.e., the extrapolation declines. the area standards for one estimation point are 100,000 ha for districts with forest cover >65%, and 60,000 ha otherwise. however, these standards may require a design compromise due to conflict with statistical requirements and the predetermined error value of the estimation data. for example, in densely populated districts with little forest cover, fragmented forests are often isolated by farming lands, settlements, and other man-made features. in this case, an estimation point can be a settlement which is located near a relatively big forest. the width of the forest along the line to the nearest settlement should be ≥2x the average distance to the first moose track; smaller forests and forests located in anthropogenic zones are not subject to estimation. the routes from such estimation points should be oriented into the forest not the cardinal directions (fig. 4). reducing the number of routes and plots to 1–2 per estimation point requires an adequate increase in the number of estimation points. standard route lengths are necessary to carry out the first estimate that is used for further corrections (table 1). the route length is subsequently corrected from the distances to the first track in the experimental estimates. a route is travelled one-fold, once a year, preferably by vehicle. choosing the size and the shape of sample plots is particular to the size of compartments, configuration of forests, and availability of access routes. it is best to use rectangular plots of fig. 3. the principle scheme for establishing 4 radial survey routes with sample plots to estimate moose density in the kirov region, russia. the center typically represents human settlement with zero moose density; density is extrapolated for the area of concentric rings. 152 moose population estimates in russia – glushkov alces vol. 49, 2013 2x4 dimension or plots of other shapes in areas >800 ha (agafonov et al. 1988). the new method of field data collection and subsequent calculation of moose population estimations described here will provide more reliable population estimates than with previous approaches. this is critical in the kirov region of russia that has experienced population overestimates and subsequent overharvest of moose. a coordinated strategy of using better population estimates, and measuring calf survival and non-harvest mortality, including poaching, will benefit regional moose management in russia (glushkov 2009). references agafonov, v. a., s. a. korytin, and i. n. solomin. 1988. winter estimate of game animals on circular routes: rational methods of game animals study. pages 17–25 in russian state research institute of game management and fur farming, kirov, russia. (in russian.) bubenik, a. b. 1965. population density of game animals, feed capacity of game habitat, and forest damage caused by animals. moose biology and hunting 2: 265–280. moscow, russia. (in russian.) draper, n., and h. smith. 1986. applied regression analysis. mir, moscow, russia. (translated from 1981 english version.) glushkov, v. m. 1977. on the method of arial moose count. hunting and game management 12: 14–15. (in russian.) ———. 1982. the moose of vyatka forests. hunting and game management 1: 16– 18. (in russian.) ———. 1997. extrapolation area estimates for moose and wild boar. pages 81-83 in applied ecology issues. proceedings of scientific conference devoted to 75th anniversary of russian state research institute of game management and fur farming. kirov, russia. (in russian.) ———. 1999. moose. pages 117–163 in management of game animals. proceedings of russian state research, institute of game management and fur farming, russian academy of agricultural sciences. kirov, russia. (in russian.) ———. 2001. moose: ecology and management of populations. russian academy fig. 4. a depiction of how survey routes are distributed in irregular fragmented forests; forest width to the nearest settlement should be >2x the average distance to the first moose track. multiple estimation points may be required to achieve a sufficient number of survey routes in an area. table 1. the stand and length of radial routes required in varying proportions of forest cover to estimate moose population density in the kirov region, russia. % forest cover route length (km) 80–100 12 70–79 10 55–69 8 40–54 7 25–39 5 <25 4 alces vol. 49, 2013 glushkov – moose population estimates in russia 153 of agricultural sciences, kirov, russia. (in russian.) ———. 2004. on route standardization for moose estimates. game management bulletin: 195–200. (in russian.) ———. 2009. improving population management and harvest quotas of moose in russia. alces 45: 43–47. ———, and a. l. buldakov. 1997. principles of ungulate estimate with the combined method. pages 83–85 in issues of applied ecology, institute of game management and fur farming. proceedings of scientific conference devoted to 75th anniversary of russian state research institute of game management and fur farming, kirov, russia. (in russian.) ———, m. g. dvornikov, v. v. kolesnikov, v. g. safonov, a. a. sergeyev, m. s. shevnina, and v. v. shiryaev. 2012. factors obstructing wild ungulate management in russia. pages 76–83 in theoretical and applied ecology. (in russian.) ling, h. 1977. experience of system analysis of population structure and dynamics. pages 7–105 in proceedings of tartu university, issue 408. university of tartu, tartu, estonia. naumov, n. p. 1963. animal ecology. moscow, russia. (in russian.). odum, e. 1986. ecology. volume 2. moscow, russia. (translated from english.) smirnov, v. s. 1964. mammal estimation methods. proceedings of the biology institute of ural academy of sciences. (in russian.) yazan, p. 1972. game animals of pechora taiga. research institute of game management and fur farming. kirov, russia. (in russian). yurghenson, p. b. 1970. distribution theory, spatial analysis, and applied ecology of animals. bulletin of moscow society of nature explorers, biology department, moscow, russia. (in russian.) 154 moose population estimates in russia – glushkov alces vol. 49, 2013 improving moose population estimates in russia: accounting for distance between residential areas and track sightings introduction & background results and discussion references alces(23)_preface.pdf alces vol. 23, 1987 alces27_93.pdf alces21_77.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 body temperature of captive moose infested with winter ticks edward m. addison1,2, robert f. mclaughlin1,3, and peter a. addison1,4 1wildlife research and development section, ontario ministry of natural resources, 300 water street, 3rd floor north, peterborough, ontario, canada k9j 8m5. abstract: eighteen captive moose calves (alces alces) were divided into 3 groups that represented 3 levels of winter tick (dermacentor albipictus) infestation (0, 21,000, and 42,000 ticks). a total of 321 body temperatures (tb) were taken on 19 occasions between late november and mid-april. the mean tb of individuals was 38.2 ± 0.4 °c, ranging from 38.0–38.3 °c, and was not different among the control and infested groups (p = 0.816), but varied temporally (p < 0.001) with a significant interaction effect between treatment and time (p = 0.041); these temporal differences are unexplained. the tbs measured in this study are some of the lowest reported for moose and presumably represent the resting tb of free-ranging moose, more so than those measured after pursuit, restraint, and/or immobilization during capture. this was not a definitive test of the effects of tick infestation on wild moose because the captive moose consumed a high quality diet throughout winter and surprisingly low numbers of ticks remained on the animals in mid-april. alces vol. 50: 81–86 (2014) key words: body temperature, alces alces, dermacentor albipictus, moose, tb, winter tick. premature hair loss by moose (alces alces) in winter that is associated with infestations of winter tick (dermacentor albipictus) is well documented (e.g., addison et al. 1979, samuel and barker 1979, samuel 1991) including by mclaughlin and addison (1986) studying the same captive moose reported here. this hair loss might influence body temperature (tb) that is reflective of increased energetic cost and stress in moose. the typical tb of moose reported in the literature is usually measured on individuals that were pursued, restrained, and/or immobilized during capture. many of these values may reflect higher than resting tb since excitability raises tb in moose (franzmann et al. 1984). objectives of this study were to assess the possible effects of winter tick infestation on tb of moose, and to obtain tb from captive animals that more accurately represent resting tb of unstressed free-ranging moose. importantly, animals in this study were young-of-the-year, exceptionally tractable, and readily accepted the measurement procedure. because technological advances in telemetry now allow tb to be measured in free-ranging moose, these data are also valuable for related comparisons. methods the experiments were conducted in algonquin provincial park, ontario (45° 30′ n, 78° 35′ w) where 13 of 18 calves were captured at <2 weeks of age in may 1982; 5 calves were from other areas in central and northeastern ontario (addison and mclaughlin 1993). male and female calves were paired in each of 6 adjacent pens 2present address: ecolink science, 107 kennedy street west, aurora, ontario, canada l4g 2l8 3r.r. #3, penetanguishene, ontario, canada l0k 1p0 4northwest region, regional operations division, ontario ministry of natural resources, 173 25th sideroad, rosslyn, ontario, canada p7k 0b9 81 (29.6 × 16.5 m) located within a mixed forest stand with little undergrowth and a partial canopy (50% in summer) of white pine (pinus strobus), white birch (betula papyrifera), trembling (populus tremuloides) and big tooth aspen (p. grandidentata). calves were weaned as described by addison et al. (1983) and from late october to the end of the experiment were fed ad libitum a ruminant ration containing 16% crude protein, 2.5% crude fat, and 6% crude fiber (united cooperative of ontario, mississauga, ontario, canada). husbandry of moose and experimental design for this study were as described in addison et al. (1994) with all animals assumed born on 15 may 1982. the 18 calves were divided into 3 treatment groups: moose with no winter ticks (n = 5; 2f:3m), moose infested with 21,000 larval winter ticks (n = 7; 3f:4m), and moose infested with 42,000 larval winter ticks (n = 6; 3f:3m). larval ticks were applied between mid-september and midoctober 1982, and all moose were euthanized at the end of the experiment (18–28 april 1983). the hair was dissolved and hides checked for ticks as described by addison et al. (1979). for months prior to the application of ticks, the study animals were attracted with food to a monitoring station where they stood quietly while we measured weight and took linear measurements. the tb was measured by inserting a standard, large animal mercury thermometer into the rectum. for 16 moose, tb was usually measured every 5–9 days from 24 november 1982 to 14 april 1983, except for a 2-week period from late january to mid-february 1983 (table 1). fewer data were available from the 2 other moose that were sacrificed prior to the completion of this study. the mean tb of individuals within and between sampling times was calculated using all 18 moose; however, data were missing for certain individuals on particular dates. all data for 3 moose with missing rectal temperatures for ≥3 of the 19 dates were removed from statistical analysis. further, because measurements were missing from 3 moose on one of the 19 dates, all were removed from the analysis for this date. after removing these data, each treatment group was comprised of 5 moose with measurements from 16 dates, for a total of 80 measurements per treatment. we tested for treatment effect (among groups), temporal effect, and an interaction effect between treatment and time using a two-factor anova with repeated measures of tb with the aov function in r (r core team 2013). results female and male calves respectively weighed 161 ± 8 and 178 ± 5 kg in midnovember, and 200 ± 17 and 218 ± 20 kg at the end of the experiment when 11 months old (addison et al. 1994). the 5 control moose harboured 0, 0, 4, 21, and 85 winter ticks at the conclusion of the experiment; the animal harbouring 85 ticks had limited (5%) hair loss (mclaughlin and addison 1986). in contrast, 1179–8290 ticks were recovered from the infested moose at the end of the experiment. hair loss was estimated at 23–44% in 8 of 10 infested animals, and 2 and 4% in the other 2 moderately infested moose (see mclaughlin and addison 1986). the mean tb (n = 321) was 38.2 ± 0.4 °c ranging from 36.8–40.7 °c. individual mean tb ranged from 38.0–38.3 °c with >99% of individual measurements from 36.8 – 39.4 °c (fig. 1). mean tb was not different among treatment groups (f2,12 = 0.207, p = 0.816), but did vary over time (f15,180 = 6.385, p < 0.001). there was a significant interaction effect between treatment and time (f30,180 = 1.561, p = 0.041) indicating that tb of treatment groups varied temporally. however, no discernible relationship existed as the mean tb of groups did not change in similar direction in all periods (fig. 2). 82 body temperatures of moose – addison et al. alces vol. 50, 2014 table 1. mean rectal body temperature (tb, °c) of standing captive calf moose exposed to 3 levels of winter tick loads (0, 21,000, 42,000 larvae) in fall 1982, algonquin provincial park, ontario; moose sample size is in parentheses. winter tick infestation level date 0 21,000 42,000 24 nov 1982 38.2 ± 0.2 (5) 38.3 ± 0.4 (6) 37.7 ± 0.5 (6) 30 nov 38.2 ± 0.1 (5) 38.0 ± 0.4 (7) 37.9 ± 0.5 (6) 5 dec 38.3 ± 0.2 (5) 38.0 ± 0.3 (6) 38.0 ± 0.6 (6) 12 dec 37.9 ± 0.4 (5) 37.8 ± 0.2 (7) 38.0 ± 0.3 (6) 19 dec 38.0 ± 0.3 (5) 38.4 ± 0.5 (6) 38.2 ± 0.7 (6) 26 dec 37.9 ± 0.1 (5) 38.0 ± 0.1 (7) 37.8 ± 0.1 (6) 3 jan 1983 38.5 ± 0.6 (4) 38.5 ± 1.0 (7) 38.3 ± 0.3 (6) 12 jan 38.4 ± 0.2 (5) 38.4 ± 0.3 (7) 38.4 ± 0.1 (6) 17 jan 37.7 ± 0.7 (4) 37.8 ± 0.6 (7) 38.0 ± 0.4 (6) 25 jan 38.4 ± 0.3 (5) 38.6 ± 0.1 (7) 38.4 ± 0.2 (6) 31 jan 37.9 ± 0.2 (5) 38.2 ± 0.5 (7) 38.1 ± 0.2 (6) 15 feb 38.1 ± 0.2 (5) 38.4 ± 0.1 (7) 38.3 ± 0.2 (6) 1 mar 38.1 ± 0.3 (5) 37.9 ± 0.3 (5) 38.0 ± 0.4 (5) 7 mar 38.1 ± 0.1 (5) 38.1 ± 0.3 (6) 38.3 ± 0.3 (5) 15 mar 38.3 ± 0.2 (5) 38.3 ± 0.4 (6) 38.6 ± 0.2 (5) 23 mar 38.4 ± 0.2 (5) 38.5 ± 0.3 (6) 38.6 ± 0.3 (5) 28 mar 38.3 ± 0.4 (5) 38.5 ± 0.2 (6) 38.5 ± 0.3 (5) 5 apr 38.3 ± 0.3 (5) 38.5 ± 0.4 (6) 38.5 ± 0.3 (5) 14 apr 38.5 ± 0.5 (5) 38.4 ± 0.5 (6) 38.4 ± 0.5 (5) 0 0.05 0.1 0.15 0.2 0.25 0.3 37 37.2 37.4 37.6 37.8 38 38.2 38.4 38.6 38.8 39 39.2 39.4 39.6 39.8 40 40.2 40.4 40.6 40.8 body temperature (°c) p ro po rt io na l f re qu en cy heavily infested moderately infested not infested figure 1. the frequency distribution of body temperature as measured in captive calf moose in ontario, 1982–1983. alces vol. 50, 2014 addison et al. – body temperatures of moose 83 discussion although one could postulate that hair loss is one factor influencing tb, the lack of difference in tb among the treatment groups was not surprising. the number of recovered ticks was relatively low in contrast with tick loads measured on heavily infested wild moose (samuel and barker 1979, samuel 2004). the number of larval winter ticks applied was an a priori estimate of the maximum numbers of ticks that would allow for the parasitic phases of the tick-moose cycle to be completed, while maintaining accept‐ able standards for the humane treatment of experimental animals, an objective that was achieved. for example, although re‐ duced pericardial and abdominal fat reservoirs occurred in the infested versus control moose (mclaughlin and addison 1986), we presume that all moose retained sufficient tissue reservoirs for adequate thermoregulation. further, given the wide range in volume of hair loss reported within a single treatment group (i.e., 2–24% in moderately infested moose; mclaughlin and addison 1986) and the limited number of moose per treatment group, it would be difficult to detect treatment differences. the negligible to limited seasonal variation in tb is consistent with previous reports of seasonal variation in tb in moose (franzmann et al. 1984) and wapiti (cervus elaphus) (parker and robbins 1984). on a cautionary note, the tbs measured in this study should not be considered representative of those of heavily infested wild moose with extensive hair loss. the captive moose received higher quality, more accessible food throughout winter compared to freeranging moose, and seldom experienced ambient temperatures considered thermally stressful (renecker and hudson 1986, addison and mclaughlin 2014). the significant interaction effect bet‐ ween treatment and time indicated that tb of treatment groups varied over time, but did so in different directions. environmental factors that might influence temporal differences remain unclear, but could include effects of handling during measurements as 36.0 36.5 37.0 37.5 38.0 38.5 39.0 39.5 11/5/82 11/25 12/15 1/4/83 1/24 2/13 3/5 3/25 4/14 5/4 b od y t em pe ra tu re ( °c ) heavy moderate none poly. (heavy) poly. (none) poly. (moderate) figure 2. the relationship between time and body temperature (rectal) of captive calf moose in 3 treatment groups of winter tick infestation: heavy (42,000), moderate (21,000), none. although treatment and time were statistically related, the temporal trend differed among treatment groups; ontario, 1982–1983. 84 body temperatures of moose – addison et al. alces vol. 50, 2014 higher tb occurs with increased excitability in immobilized moose (franzmann et al. 1984). however, we recognized no overt excitability in the study animals during measurements and the differences may simply reflect normal variation. the tbs measured in this study were lower than any reported from healthy moose, and likely reflect the psychical state of our calm tractable moose, or conversely, the more stressful condit‐ ions associated with measurements of freeranging moose. the upper end of the range of tb was consistent with data from prior studies (e.g., franzmann et al. 1984) and most similar to those of captive moose (38.0–39.7 °c) that were not immobilized (renecker and hudson 1986). seal et al. (1985) reported a mean tb of 38.6 °c for free-ranging moose immobilized from the ground as they approached mineral licks. in contrast, higher tb was reported for wild moose pursued and restrained (x̄ = 39.3 °c, range = 38.0– 40.4 °c), or pursued and immobilized (x̄ = 40.5 °c, range = 38.0–42.8 °c, roussel and patenaude 1975; x̄ = 39.1–39.7 °c, delvaux et al. 1999). most tbs of moose have been measured in adults and not young-of-the-year as reported here. differences in size and age likely have little if any influence on tb since in most ungulates tb varies little relative to body mass, and if variable, young animals generally have higher tb than adults (parker and robbins 1985). in summary, there was no evidence that presence of winter ticks as applied in this experiment had any direct influence on tb of moose. the tbs measured in our highly tractable animals were the lowest reported for healthy moose, and consistent with the view that level of excitability influences tb. importantly, they provide the baseline tb for resting moose that is important for metabolic modeling and comparison with tb measured via telemetry of free-ranging moose. acknowledgements we appreciate d. j. h. fraser for his coordination of many early aspects of this study. special thanks go to a. rynard, a. macmillan, m. jefferson, v. ewing, and d. bouchard for their steadfast assistance in collection of data and moose husbandry under adverse conditions. additional assistance was provided by c. pirie, m. a. mclaughlin, d. carlson, d. joachim and p. methner, and l. smith, k. paterson, k. long, a. jones, s. gadawski, s. fraser, d. fraser, and l. berejikian assisted in the earlier care of calves. we appreciate the assistance of c. d. macinnes and g. smith and staff for their administrative support. thanks to a. r. rodgers for advice with statistics. field work was conducted at the wildlife research station in algonquin park where r. keatley, p. c. smith and staff were of great help. thank you to m. lankester and anony‐ mous reviewers whose valuable suggestions were incorporated into the manuscript. references addison, e. m., f. j. johnson, and a. fyvie. 1979. dermacentor albipictus on moose (alces alces) in ontario. journal of wildlife diseases 15: 281–284. ———, and r. f. mclaughlin. 1993. seasonal variation and effects of winter ticks (dermacentor albipictus) on consumption of food by captive moose (alces alces) calves. alces 29: 219–224. ———, and ———. 2014. shivering by captive moose infested with winter ticks. alces 50: 87–92. ———, ———, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and uninfested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469– 1476. alces vol. 50, 2014 addison et al. – body temperatures of moose 85 ———, ———, and d.j.h. fraser. 1983. raising moose calves in ontario. alces 18: 246–270. delvaux, h., r. courtois, l. breton, and r. patenaude. 1999. relative efficiency of succinylcholine, xylazine, and carfentanil mixtures to immobilize free-ranging moose. journal of wildlife diseases 35: 38–48. franzmann, a. w., c. c. schwartz, and d. c. johnson. 1984. baseline body temperatures, heart rates, and respiratory rates of moose in alaska. journal of wildlife diseases 20: 333–337. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)-induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. parker, k. l., and c. t. robbins. 1984. thermoregulation in mule deer and elk. canadian journal of zoology 62: 1409–1422. ———, and ———. 1985. thermoregulation in ungulates. pages 161–182 in r. j. hudson and r. g. white, editors. bioenergetics of wild herbivores. crc press, boca raton, florida, usa. r core team. 2013. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. . renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. roussel, y. e., and r. patenaude. 1975. some physiological effects of m99 etorphine on immobilized free-ranging moose. journal of wildlife management 39: 635–636. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m. 1991. grooming by moose (alces alces) infested with the winter tick, dermacentor albipictus (acari): a mechanism for premature loss of winter hair. canadian journal of zoology 69: 1255–1260. ———, and m. j. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869) on moose alces alces (l.) of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. seal, u. s., s. m. schmitt, and r. o. peterson. 1985. carfentanil and xylazine for immobilization of moose (alces alces) on isle royale. journal of wildlife diseases 21: 48–51. 86 body temperatures of moose – addison et al. alces vol. 50, 2014 http://www.r-project.org/ http://www.r-project.org/ body temperature of captive moose infested with winter ticks methods results discussion acknowledgements references alces29_197.pdf alces27_140.pdf using a double-count aerial survey to estimate moose abundance in maine lee e. kantar1 and rod e. cumberland2 1maine department of inland fisheries and wildlife, bangor, maine 04401, usa; 2maritime college of forest technology, fredericton, new brunswick e3c 2g6, canada abstract: management goals and objectives for moose (alces alces) in maine are centered on providing hunting and wildlife viewing opportunity. robust population estimates of moose are critical to assure that harvest rates are appropriate and biologically sustainable while also addressing values of other user groups. the maine department of inland fisheries and wildlife most recently used the relationship between moose sightings by deer hunters and moose abundance to produce density indices within wildlife management districts (wmd). due to the marked decline of deer hunters in much of northern maine that invalidates use of this technique, we tested a double-count aerial survey method to estimate moose abundance in 9 northern wmds. density estimates ranged from 0.4–4.0 moose/ km2, sightability was high (>70%) for all size moose groups (1–≥3 moose), and moose were well distributed across the landscape in early winter. the density estimates tracked closely with trends in moose sighting rate by moose hunters, harvest level, and hunter success rate in the survey area, and were consistent with jurisdictions in eastern canada that also have low levels of predation and a preponderance of younger-aged forests. the double-count aerial survey is considered the preferred method to estimate population density, whereas hunter sighting indices would be most useful to track temporal population changes within a wmd. alces vol. 49: 29–37 (2013) key words: aerial survey, alces alces, double-count survey, moose, population estimate, maine. in maine, management goals and objectives for moose (alces alces) specify providing hunting and wildlife viewing opportunity, as well as an adequate age structure of bulls (morris 2002). based on a planning process that involves public input, the maine department of inland fisheries and wildlife (mdifw) created 3 categories (zones) to manage moose: recreation, compromise, and road safety zones. the primary goal in recreation zones is to maximize hunting and viewing opportunities by maintaining high moose density without affecting habitat quality (i.e., 55–65% of carrying capacity). in compromise management zones, the goal is to balance public concern about moose-vehicle collisions with its desire to hunt moose; further, it is stipulated that moose populations in such zones must be reduced by 1/3 from the population estimate calculated in the year 2000 that was based upon the relationship between moose abundance and moose observations by deer hunters (bontaites et al. 2000). objectives in road safety zones are broadly stated as reducing the population as necessary to lower the danger or frequency of moose-vehicular collisions. moose are highly prized for hunting and one of the most sought after wildlife species to view, yet moose also create numerous negative impacts. obtaining robust estimates of abundance is vital to allocating moose hunting opportunity (permit levels) and managing populations at publicly derived population objectives. ostensibly, providing appropriate levels of harvest to manage moose abundance involves risk (i.e., over or under harvest) in terms of possible effects 29 to the target population, as well as accountability to multiple groups that value and use the moose resource. therefore, it is imperative to have adequate data to support management decisions, particularly harvest levels. the mdifw has attempted various techniques to estimate abundance including transect counts from fixed wing aircraft, line-track intercept techniques, a modified gasaway survey, and forward looking infrared (flir). maine had adopted a technique that uses the relationship between moose abundance and moose observations by deer hunters based on work in adjacent new hampshire (bontaites et al. 2000). this relationship was predicated on aerial survey data using flir technology and having an adequate sample of moose sightings by deer hunters. this technique has been the cornerstone of estimating moose abundance in maine, yet model assumptions were not tested in maine, and the recent marked decline in deer and deer hunters in northern parts of maine presumably invalidates the sighting data requirement. therefore, we used a double-count aerial survey, developed originally for white-tailed deer (odocoileus virginianus) in quebec (potvin et al. 1992) and adapted for moose in new brunswick (cumberland 2012), to estimate moose abundance in 9 northern wildlife management districts (wmd) with presumed high density and 1 wmd with low density. study area northern wmds with the highest moose density based on hunter sighting rates, highest harvest rates, and permit allocations (mdifw data) were prioritized for the double-count surveys. surveys were flown in aroostook county, maine (17,687 km2) that is comprised mostly of farmland (>131,100 ha) in its eastern part, with ∼76,000 ha of actual cropland (maine department of agriculture, food and rural resources 2003), and in northern portions of adjacent franklin, hancock, penobscot, piscataquis, somerset, and washington counties. in total, the study area included wmds 1–6, 8, 11, and 19 and comprised ∼32,950 km2 (fig. 1). forested areas are dominated by spruce (picea spp.), balsam fir (abies balsamea), northern white cedar (thuja occidentalis), and white pine (pinus strobus), with mixed hardwoods of aspen (populus spp.), birch (betula spp.), beech (fagus grandifolia), and maple (acer spp.). other species highly palatable to moose include red-osier dogwood (cornus stolonifera) and willow (salix spp.). methods using a geographical information system (gis), each wmd was divided into blocks that were 259 km2 and oriented either s-n or e-w depending on the unique shape/ orientation of the wmd; survey blocks were uniformly 15 x 24 km rectangles. the maine land cover dataset (2004) was used to select 7 habitat variables: 1) crops/grasslands/blueberry barrens, 2) deciduous forest, 3) evergreen (coniferous) forest, 4) mixed forest, 5) recent cuts/regenerating forest/ scrub-shrub, 6) wetlands, and 7) partial cuts. urban centers, lakes, or non-permeable surfaces were not included in the gis analysis. we calculated the percentage of each habitat variable within each survey block and the total percentage of each habitat variable within all blocks in a wmd. we performed a linear regression analysis to prioritize survey blocks based on the strongest relationships between each habitat variable and the habitat composition within a survey block. we selected the 1 survey block within each wmd that was most representative of the zone's overall habitat; an exception was in mountainous wmd 8 where safety concerns dictated the specific survey block. further, 3 survey blocks in 30 double count aerial survey – kantar and cumberland alces vol. 49, 2013 wmd 4 were flown the same winter (2012) to investigate consistency of density estimates in the largest wmd (>5200 km2) and to survey areas receiving high hunting pressure. as an initial evaluation of the relative reliability of the technique at lower moose density, we surveyed wmd 11 where density was considered low based on sighting and harvest rate. universal transverse mercator (utm) points delineated the survey blocks and transects spaced 1 km apart systematically set at the edge of a block. for each survey block, flight maps included 7, ∼15 km fig. 1. location of double-count aerial surveys of moose in maine wildlife management districts 1–6, 8, 11 and 19 in maine, usa. alces vol. 49, 2013 kantar and cumberland – double count aerial survey 31 transects that were oriented either n-s or e-w to account for as much habitat variation as possible, yet ensure that variation among transects was minimal and precision maximized (caughley et al. 1976). a bell jet ranger 206 was outfitted with bubble windows for the front and rear observer to increase and improve viewing angle from the helicopter skid. a radar altimeter was installed to ensure that height above ground was maintained at 60 m regardless of topography. flights were conducted and maintained at 60 m above ground level at speeds of 56–72 km/h; they occurred on days with no precipitation and low wind speed (<24 km/h). we flew when snow conditions would not impede moose movement (<61cm) and ambient temperature was cold (<−12 °c; quayle et al. 2001). transects widths were initially calibrated at 60 m above ground level while the aircraft hovered over a 60 m width delineated on the ground. observers sighted the width from the edge of the helicopter skid out to 60 m by affixing a tape line across the bubble window. a modified communications system was used to isolate the 2 observers seated fore and aft on the port side of the helicopter. to ensure independence, moose observations were communicated separately by each to the recorder who sat behind the pilot; observers could not hear or speak to each other (gauthier and cumberland 2000). the recorder initiated and ended each flight transect, relayed all information to the observers and pilot, and was responsible for following the flight lines and checking navigation. transect coordinates were entered into a mobile gps unit and the onboard gps prior to flight departure. the flight crew remained constant for all flights. the recorder cued the 2 observers at the start and end of each transect. observers independently sighted moose within the 60 m transect width and relayed the number of moose to the recorder. if the recorder was unable to determine if both observers counted the same moose, they immediately asked questions concerning the location, habitat, and direction of movement. the recorder then determined independently whether the 2 observers identified the same moose, thereby verifying and subsequently recording the observation data. thus, observations could be categorized as by the primary observer only, the secondary observer only, or by both observers; these data served as the mark-resight elements. estimates of abundance (moose/km2) and sighting probabilities were calculated following rivest et al. (1995) and cumberland (2012). to evaluate the relationship between the double-count survey and other population indices in maine, we compared our density estimates with trends in moose sightings by moose hunters during the october season, harvest success rates, and permit allocations over a fairly stable hunt framework period (2003–2010). as a general assessment of the technique, we reviewed moose sightings and group size on transects and among survey blocks to characterize moose distribution within survey blocks. results a total of 294 observations (≥1 moose/ observation) occurred in the 12 surveys. the observation rate ranged from 4–43 per survey block, averaging 24.5 (table 1). across all flights, single moose represented 50.5%, 2 moose 37.7%, and ≥3 moose 11.8% of all observations (table 1). single moose represented ≥50% of observations in 75% of the surveys; the largest group size was 5 moose. across all surveys, the sighting probability of a single moose by the forward (0.72, range = 0.47–1.00) and rear observer (0.84, range = 0.60–1.00) was not different (p = 0.07) (table 2), although the rear observer was 12% higher. for observations of ≥2 moose, the sighting probability of the forward (0.88, range = 0.61–1.00) 32 double count aerial survey – kantar and cumberland alces vol. 49, 2013 and rear observers (0.89, range = 0.68–1.00) were nearly identical and also not different (p = 0.973) (table 2). in the winter of 2010–11, flights were conducted on 28 and 31 january and 1 february 2011 in wildlife management districts 2, 3, and 6; density estimates were 3.0 ± 0.2, 2.7 ± 0.6, and 1.2 ± 0.3 moose/km2, respectively (table 3). favorable conditions occurred earlier in 2011; flights were conducted on 12 and 16 december, and continued on 8, 9, 11, 22, and 26 january, and 8 and 15 february; surveys occurred in wmds 1, 2, 4 (3 flights), 5, 8, 11, and 19. the density estimates ranged from 0.4 ± 0.2 (wmd 11 with presumed low density) to 4.0 ± 0.7 moose/km2 (wmd 4). no change in the density estimate occurred from 2011 (3.0 ± 0.2) to 2012 (3.1 ± 0.6 moose/km2) in wmd 2 that was flown both winters. density estimates for all surveys are provided in table 3. from 2003–2010, the highest sighting rates occurred in the northern tier of the state, ranging from 1.3 (wmd 11) to 6.9 (wmd 2), and declining north to south as moose habitat declines in quality (table 3). we used linear regression to test whether these average sighting rates were related to the density estimates in the surveyed wmds; the relationship was significant with a moderate r2 (n = 13, f = 5.10, p = 0.045, r2 = 0.32). discussion the predominance (>85%) of observations as single or 2 moose indicated that conducting flights under optimal conditions (<30 cm snow and <∼−17 °c) helped ensure that moose were distributed spatially and not clumped. in deeper snow conditions or in late winter when temperatures have ameliorated, moose typically spend more time in coniferous cover and/or in larger groups. we saw no evidence of these behaviors during our flights. although habitat was not described at the point of observation, forest stands with >50% canopy closure were uncommon and flight data indicated that moose were observed in multiple habitats (e.g., hardwood, coniferous, and mixed stands). most forest stands were commercial and in early to mid-seral stages that provide preferred moose browse and conditions for optimal sightability. taken in whole, the low group size and continual observations along transects in varied habitats support our assumption that moose were not clumped in distribution. table 1. observations (n = 294), group size, and proportional frequency of group size in double-count aerial surveys of moose in 9 wildlife management districts, maine, usa, winters 2011 and 2012. wmd total observations group = 1 (%) group = 2 (%) group = ≥3 (%) largest group size 1 28 50.0 42.9 7.1 4 2y1* 32 53.1 40.6 6.3 5 2y2 30 50.0 40.0 10.0 5 3 25 44.0 44.0 12.0 4 4h12** 34 61.8 35.3 2.9 3 4h7 43 53.5 37.2 9.3 3 4h3 35 57.1 34.3 8.6 3 5 11 18.2 54.5 27.3 3 6 13 53.8 38.5 7.7 3 8 18 55.6 33.3 11.1 4 11 4 75.0 0.0 25.0 4 19 21 33.3 52.4 14.3 3 mean ± se 24.5 ± 3.3 50.5 ± 4.1 37.7 ± 4.0 11.8 ± 2.1 y1* = 2011, y2 = 2012. h12**, h7, h3 = survey block identification. alces vol. 49, 2013 kantar and cumberland – double count aerial survey 33 video recording during flights indicated minimal detection of moose directly under the helicopter; rather, tracks and beds were recorded in the footage. moose typically responded to the helicopter by moving away, thereby providing high sightability along a wide continuum of forest cover types and stand conditions. the sighting probability of single and groups of moose was moderate to high (∼70–80%) compared to other aerial techniques; potvin and breton (2005) found that reliable estimates could be calculated with sighting probabilities as low as 45%. cumberland (2012) provided a complete synopsis of the performance of this technique and the relationship of sighting probabilities to the mark-resight model. trends in moose sightings by moose hunters were presented to provide context about moose abundance in northeastern maine (table 3). long-term, high harvest success rates and high sighting rates provide indirect support of the density estimates reported here; in general, sighting rates and density estimates were correlated. sighting rates by moose hunters in october serve as the most reliable and least biased index to moose abundance across the state because of high visibility due to leaf drop, and bulls are less receptive to calling and rutting behavior. while sightings by deer hunters can provide a reliable index to moose abundance (see ericsson and wallin 1999, solberg and saether 1999, bontaites et al. 2000), loss of deer hunters in northern maine has led to inadequate sample sizes to make inferences at the wmd level. the estimated moose density (2.7–4.0 moose/km2) in wmds 1–4 was moderately high relative to many other north american populations (range 0.04–9.3 moose/km2; karns 2007). population increases in the northeastern united states have been facilitated by low predation risk, low deer densities (i.e., low transmission rates of parelaphostrongylus tenuis), and optimal habitat conditions (i.e., clear-cutting, wetland habitat, farmland reverting to forest) (karns 2007). in the cape breton highlands table 2. sighting probabilities for single and groups of moose by 2 observers (front, rear) during double-count aerial surveys in 9 wildlife management districts (wmd) in northern maine, usa, winter 2011 and 2012. no difference (p >0.05) in sightability was found between the 2 observers for either single or groups of moose. single group wmd front rear front rear 1 1.00 0.79 0.86 0.68 2 0.75 0.69 0.96 0.76 2 0.63 0.91 0.97 0.94 3 0.78 0.88 0.78 0.78 4 0.68 0.94 0.84 1.00 4 0.90 0.86 1.00 0.95 4 0.47 0.67 0.84 1.00 5 0.50 1.00 1.00 0.81 6 0.60 0.60 1.00 0.85 8 0.67 0.86 0.74 1.00 11 0.67 1.00 1.00 1.00 19 1.00 0.86 0.61 0.85 mean 0.72 0.84 0.88 0.89 se 0.05 0.04 0.04 0.03 table 3. ranking of average moose hunter sighting rates (moose/10 h; 2003–2010) and density estimates from double-count aerial surveys of moose in 9 wildlife management districts in maine, usa, winters 2011 and 2012. sighting rates and density estimates were correlated (p <0.05). wmd rank sighting rate density estimate 2 1 6.9 3.0, 3.1 1 2 5.3 2.7 5 4 4.5 1.4 4 5 4.2 2.9, 3.4, 4.0 8 7 3.9 1.7 3 8 3.7 2.7 6 10 2.7 1.2 19 11 2.0 2.4 11 15 1.3 0.4 34 double count aerial survey – kantar and cumberland alces vol. 49, 2013 of nova scotia, where predation is low and hunting is either not allowed or access is difficult, moose density has been as high as 20 moose/km2 (smith et al. 2010). similarly, moose density in newfoundland has ranged from 0.80–6.13/km2 in a system where predation is low (mclaren and mercer 2005); timmerman and rodgers (2005) outlined similar dynamics (i.e., low predation, modern forestry, closely managed hunting) in other moderate to high moose populations. arguably, both moderate-high population density and similar dynamics exist currently in maine. from a habitat standpoint, much of maine's commercial forestlands are able to sustain and produce moderate to high moose density, and historic and current land use, moose sighting rates, and hunter success rates support this assertion. during 1970–1980s, a widespread spruce budworm (choristoneura fumiferana) outbreak occurred across much of maine's commercial forestlands (department of conservation 2005). in subsequent years, extensive salvage cutting increased the intensity of forest harvesting and clear-cuts, and creation of extensive road systems across the commercial forest landscape, utilization of smaller diameter wood, and shorter rotations all produced a landscape with increased carrying capacity for moose populations. forestry practices within northern maine have created a widespread patchwork of early to middle seral stage forestlands characterized by small patch size (k. legaard, university of maine, unpublished data). while the amount of browse available during the growing season has declined since the 1990s (mdifw data), it is unlikely that the moose population is limited by quantity or quality of habitat. further, annual mortality of adult moose appears to be relatively low in the absence of a significant predator and conservative hunting permit allocations. the role of parasitism (i.e., dermacentor albipictus) in overwinter calf survival, and both the frequency and distribution of winter die-offs across maine are unknown, but likely have measurable influence on the population trajectory due to mortality of calves and reduced productivity of yearling females as in adjacent new hampshire (musante et al. 2010). overall, the estimates in all wmds appear reasonable given the combination of these attributes, current habitat conditions, low annual mortality, and historic permit allocations. management implications reliable measures of abundance are critical to management of moose given their value to multiple stakeholders and the mdifw mandate to manage species for the good of all maine citizens. sighting rates appear to be most useful to track temporal population changes within a wmd, but less meaningful as an indication of absolute abundance. unlike aerial surveys, sighting rates are likely influenced by variation in habitat conditions, size and juxtaposition of forest harvest classes, and road density. estimates of abundance provide a starting point for understanding population dynamics, determining sustainable harvest, and establishing stakeholder confidence. aerial survey work is likely more acceptable and better grasped by a diverse public that often is suspicious of “bureaucracy” and weary of models and academic frameworks. implementation of a double-count aerial survey that provides reliable estimates of moose populations in maine is a significant step forward for mdifw. while this technique is most aptly applied to areas of moderate to high moose density, it will likely be applicable in the majority of maine's management zones. this technique may be coupled with sex and age composition counts to provide further assessments of population dynamics to aid future management of moose in maine. alces vol. 49, 2013 kantar and cumberland – double count aerial survey 35 acknowledgements l. kantar thanks maine forest service ranger pilots c. blackie, l. mazzei, j. knight, and j. crowley, consummate professionals whose dedication, expertise, and commitment kept us safe and on line. i especially thank our exceptional flight crew including k. marden, s. mclellan, a. starr, and d. hentosh. without the flexibility and dedication of all these individuals this work would not have been possible. we thank dr. p. pekins for editorial advice and critical review. references bontaites, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69–76. caughley, g., r. sinclair, and d. scottkemmis. 1976. experiments in aerial survey. journal of wildlife management 40: 290–300. cumberland, r. e. 2012. potvin doublecount aerial survey in new brunswick: are results reliable for moose? alces 48: 67–77. department of conservation maine forest service. 2005. the 2005 biennial report on the state of the forest and progress report on sustainability standards. report to the joint standing committee of the 122nd legislature on agriculture, conservation and forestry. maine department of conservation, augusta, maine, usa. ericsson, g., and k. wallin. 1999. hunter observations as an index to moose alces alces population parameters. wildlife biology 5: 177–185. gauthier, p., and r. e. cumberland. 2000. a new brunswick manual for aerial surveys of white-tailed deer populations utilizing the potvin mark-recapture technique. new brunswick deer technical report no. 5. fredericton, new brunswick, canada. karns, p. d. 2007. population distribution, density and trends. pages 125–139. in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose (2nd edition). university press of colorado, boulder, colorado, usa. maine department of agriculture, food and rural resources. 2003. aroostook county: 2003–2004 county profile of maine agricultural enterprises. (accessed april 2010). maine office of geographic information system. 2004. maine land cover dataset. (accessed october 2010). mclaren, b. e., and w. e. mercer. 2005. how management unit license quotas relate to population size, density, and hunter access in newfoundland. alces 41: 75–84. morris, k. i. 2002. moose management system. maine department of inland fisheries and wildlife, augusta, maine, usa. musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. potvin, f., l. breton, l. rivest, and a. gingras. 1992. application of a doublecount aerial survey technique for whitetailed deer, odocoileus virginianus, on anticosti island, quebec. canadian field-naturalist 106: 435–442. ———, and ———. 2005. from the field: testing 2 aerial survey techniques on deer in fenced enclosures – visual double-counts and thermal infrared sensing. wildlife society bulletin 33: 317–325. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–54. 36 double count aerial survey – kantar and cumberland alces vol. 49, 2013 http://www.maine.gov/agriculture/mpd/farmland/facts/aro1103.pdf http://www.maine.gov/agriculture/mpd/farmland/facts/aro1103.pdf http://www.maine.gov/agriculture/mpd/farmland/facts/aro1103.pdf http://geolibportal.usm.maine.edu/geonetwork/srv/en/metadata http://geolibportal.usm.maine.edu/geonetwork/srv/en/metadata rivest, l. p., f. potvin, h. crepeau, and g. daigle. 1995. statistical methods for aerial surveys using the double-count technique to correct visibility bias. biometrics 51: 461–470. smith, c., k. beazley, p. duinker, and k. a. harper. 2010. the impact of moose (alces alces andersoni) on forest regeneration following a severe spruce budworm outbreak in the cape breton highlands, nova scotia, canada. alces 46: 135–150. solberg, e. j., and b. e. saether. 1999. hunter observations of moose, alces alces, as a management tool. wildlife biology 5: 107–117. timmerman, h. r., and a. r. rodgers. 2005. moose: competing and complementary values. alces 41: 85–120. alces vol. 49, 2013 kantar and cumberland – double count aerial survey 37 using a double-ount aerial survey to estimate moose abundance in maine study area methods results discussion management implications acknowledgements references alces28_59.pdf moose habitat in massachusetts: assessing use at the southern edge of the range david w. wattles1 and stephen destefano2 1massachusetts cooperative fish and wildlife research unit, department of environmental conservation, university of massachusetts, amherst, massachusetts, 01003 usa; 2u. s. geological survey, massachusetts cooperative fish and wildlife research unit, university of massachusetts, box 34220, amherst, massachusetts, 01003 usa. abstract: moose (alces alces) have recently re-occupied a portion of their range in the temperate deciduous forest of the northeastern united states after a more than 200 year absence. in southern new england, moose are exposed to a variety of forest types, increasing development, and higher ambient temperatures as compared to other parts of their geographic range. additionally, large-scale disturbances that shape forest structure and expansive naturally occurring shrub-willow communities used commonly elsewhere are lacking. we used utilization distributions to determine third order habitat selection (selection within the home range) of gps-collared moose. in central massachusetts, forests regenerating from logging were the most heavily used cover type in all seasons (48 63% of core area use). habitat use of moose in western massachusetts varied more seasonally, with regenerating forests used most heavily in summer and fall (57 and 46%, respectively), conifer and mixed forests in winter (47 65%), and deciduous forests in spring (41%). this difference in habitat selection reflected the transition from northern forest types to more southern forest types across the state. the intensive use of patches of regenerating forest emphasizes the importance of sustainable forest harvesting to moose. this study provides the first assessment of habitat requirements in this southern portion of moose range and provides insights into re-establishment of moose in unoccupied portions of its historic range in new york and pennsylvania. alces vol. 49: 133–147 (2013) key words: alces alces, forest regeneration, habitat use, massachusetts, moose. moose (alces alces) have recently recolonized a portion of their historic range in the temperate deciduous forest of southern new england after more than 200 years absence (vecellio et al. 1993, wattles and destefano 2011). this environment is unique in moose range and provides a number of potential challenges, including forest types that differ from most of the geographic range (westveldt et al. 1956, degraaf and yamasaki 2001, franzmann and schwartz 2007), a thermal environment that could reduce fitness and survival (renecker and hudson 1986, boose 2001, murray et al. 2006, lenarz et al. 2009, 2010), and high levels of human development (u. s. census bureau 2000, destefano et al. 2005). habitat use and diet have been studied throughout much of moose range (franzmann and schwartz 2007), including elsewhere in the northeastern united states (crossley and gilbert 1983, leptich and gilbert 1989, garner and porter 1990, miller and litvaitis 1992, thompson et al. 1995, corresponding author: david wattles, massachusetts cooperative fish & wildlife research unit, 225 holdsworth natural resources center, university of massachusetts, amherst, ma 01003. 617-2567055, (fax) 413-545-4358, email: dwattles@eco.umass.edu 133 scarpitti et al. 2005, scarpitti 2006). however, similar information has been lacking in the transitional forests of southern new england. the recolonization of southern new england saw moose push the southern extent of their range from spruce-fir and northern hardwood forests into transitional and more southerly forest types, which lack many of the plant species preferred by moose further north. massachusetts provides a unique environment to examine the effects of this transition in use of forest types and habitat selection over a relatively small area. the objectives of this study were to 1) determine how moose use the temperate deciduous forest of southern new england, 2) compare habitat use among seasons, and 3) assess whether suitable habitat exists to support long-term occupation of southern new england. study area the study area was located in central and western massachusetts, usa (fig. 1). topography is dominated by glaciated hills underlain by shallow bedrock. glacial activity created abundant small stream valleys, lakes, ponds, and wetlands whose size and nature vary with beaver (castor canadensis) activity. the central and western sections of the study area are separated by the connecticut river valley which runs n-s through westcentral massachusetts. elevation ranges from 100 m above sea level in the connecticut river valley to 425 m in the hills of central massachusetts, and 850 m in the berkshire hills of western massachusetts. the western two-thirds of massachusetts was >80% mixed deciduous, secondor multiple-growth forest, much of it resulting from regeneration of farm fields abandoned fig. 1. area used to study moose-habitat relationships in northeastern usa, specifically westcentral massachusetts depicted in blow-up with dashed line, and bordered by southern vermont and new hampshire. figure also depicts the forest types of massachusetts (after westveldt et al. 1956 and degraaf and yamasaki 2000). 134 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 in the mid-to-late 1800s (hall et al. 2002). with the exception of wetlands and smallscale logging, the undeveloped portion of the massachusetts landscape was nearly 100% closed canopy mixed-coniferousdeciduous forest. massachusetts represents a forest transition zone, where forests shift from those common in northern new england to more southern forest types. moose transition across 4 forest types in massachusetts, including spruce-fir-northern hardwoods, northern hardwoods-hemlock (tsuga canadensis)-white pine (pinus strobus), transition hardwoods-white pine-hemlock, and central hardwoods-hemlock-white pine (fig. 1). the spruce-fir-northern hardwoods type is dominated by spruce (picea spp.), balsam fir (abies balsamea), american beech (fagus grandifolia), birch (betula spp.), trembling aspen (populus tremuloides), eastern hemlock, and maple (acer spp.). in the northern hardwoods forest, white pine and hemlock largely replace spruce and fir. transition hardwoods-white pine-hemlock forests contain most of the species in the northern hardwoods type; in addition, oaks (quercus spp.) and hickories (carya spp.) become increasingly common. in the central hardwoods-hemlock-white pine forest, beech, sugar maple (a. saccharum), and yellow birch (b. alleghaniensis) are rare, replaced by oaks and hickories. transitions between forest types can be gradual or distinct depending on localized physiography, climate, bedrock, topography, land-use history, and soil conditions, resulting in a patchwork of forest types and species groups (westveldt et al. 1956, degraaf and yamasaki 2001). early successional habitat was created primarily through timber harvest practices, and occasionally through wind and other weather events. from 1984 to 2000, about 1.5% of the forest was logged annually, consisting of small (mean = 16.5 ha) cuts of moderate intensity (removal of 27% of timber volume) widely distributed on the landscape (kittredge et al. 2003, mcdonald et al. 2006). the pattern of forest harvest, glaciation, and transitional forest types provided a patchy mosaic of well interspersed forest types, age classes, and wetlands. july was the warmest month when mean daily temperature was 21.1 °c, and january the coldest when mean daily temperature was −6.1 °c. mean annual precipitation was 107 cm in central areas and 124 cm in western areas, with all months receiving 7–11 cm and 8–12 cm, respectively (the weather channel 2011a, 2011b). the average date of last frost in the region was 15 may; the average day of first frost was 1 october and 15 september in central and western areas, respectively (degraaf and yamasaki 2001). maximum snow depth was typically greater in western massachusetts than central areas and reaches depth in both areas (50–70 cm) that can restrict moose movement (coady 1974). methods study animals and gps telemetry adult (>1 yr old) moose were captured by locating, stalking, and darting them from the ground in state forests, wildlife management areas, and other conservation areas between march 2006 and november 2009. moose were immobilized using either 5 ml of 300 mg/ml or 3 ml of 450 mg/ml xylazine hydrochloride (congaree veterinary pharmacy, cayce, south carolina, usa; mention of trade names does not imply endorsement by the u. s. government) administered from a 3 or 5 cc type c pneudart dart (pneudart, inc., williamport, pennsylvania, usa). tolazolene (100 mg/ml) at a dosage of 1.0 mg/kg was used as an antagonist. moose were fitted with gps collars; either ats g2000 series (advanced telemetry systems, inc., isanti, minnesota, usa) or telonics twg-3790 gps collars (telonics, inc., mesa, arizona, usa). alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 135 we programmed the collars to attempt a gps fix as frequently as possible while allowing the battery life to extend for at least 1 year; depending on the collar, a gps fix was attempted every 135, 75, or 45 min. collars were equipped with very high frequency (vhf) transmitters, mortality sensors, and release mechanisms that opened the collars either at a low battery state or a preprogrammed date. capture and handling procedures were approved by the university of massachusetts institutional animal care and use committee, protocol numbers 25-02-15, 28-02-16, and 211-02-01. seasons we defined the length and timing of seasons based on several ecological factors including timing of the growing season of vegetation, weather (including temperature and snow conditions), and the moose reproductive cycle (table 1). the transition between seasons can vary by several days to several weeks depending on weather conditions and other factors. if movements were identified in the location data for an animal that obviously demonstrated a change in season (e.g., a large increase in movements at the end of the winter when snow had melted, or at the end of summer indicating the beginning of rutting behavior), the data were truncated at that point and included in the following season. habitat availability and core area habitat use we compared vegetation and land cover types in the home range cores of moose to that available in larger mcp home ranges (third-order habitat selection; johnson 1980). we used a fixed kernel density estimator (kde) (worton 1989) and the kernel density estimation tool in hrt: home range tools for arcgis (rogers et al. 2007) to calculate utilization distributions (ud). we then used the create minimum convex polygons tool in hawth's tools (beyers 2006) to calculate 100% minimum convex polygon (mcp) home ranges (mohr 1947). all geographic information system (gis) work was performed in arcgis 9.3 (esri 2008). table 1. seasons used for calculating home-range, movements, and core-area habitat analyses. season dates vegetation/ browse temperaturea movement season length (d) spring 16 april – 31 may growing season; bud-break-leaf out cool-hot not snow restricted, potentially temperature restricted 46 calving (females) 8–13 may – 15 june growing season; bud-break-leaf out cool-hot restricted by newborn calf 30 summer 1 june – 30 aug growing season; full leaf out hot restricted by temperature 92 fall 1 sept – 31 oct leaf out to leaf off hot-cool rut and temperature influenced 61 early winter 1 nov – 31 dec dormant season; woody/evergreen warm-cold not snow restricted, potentially metabolism restricted 61 late winter 1 jan – 15 april dormant season; woody/evergreen cold-warm potentially snow and metabolism restricted 107 atemperature ranges describing typical temperatures experienced during a season; cold ≤0°c, cool >0°c and <14°c, warm ≥14°c and <20°c, hot ≥20 °c. 136 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 the kernel bandwidth or smoothing factor (h) is known to have the greatest effect on uds (worton 1989). a large h oversmooths the data, resulting in a more biased ud that encompasses unused habitats, while a small h under-smooths the data, resulting in a fragmented ud (fieberg 2007). there is lack of agreement on the best method for calculating h (powell 2000, hemson et al. 2005, gitzen et al. 2006, fieberg 2007, kie et al. 2010); therefore, we used 2 values (80 m and 30 m) of h to calculate uds. we used the 50-percent isopleth of the 80 m ud to identify home range cores. however, the 80 m bandwidth still resulted in over-smoothed uds with large buffers around gps locations that incorporated unused habitat. as a result we used a second h value of 30 m, based on the median distance between gps locations for our most intensively sampled animals, approximating within-patch movement of the animals. the resulting uds incorporated little unused habitat and were used to assess habitat use within the core areas calculated with the 80 m h. we classified habitats into 8 categories: coniferous forest (mostly coniferous with minimal deciduous component), deciduous forest (mostly deciduous with minimal coniferous component), mixed forest (mixed deciduous and coniferous), regenerating forest (logged areas <20 years old and powerline right-of-ways), wooded wetlands (conifer, mixed, and deciduous wooded wetlands), other wetlands (grassy fens, shrub swamps, bogs, deep wetlands, and open water), open (e.g., fields and meadows), and developed. we set the age restriction of regenerating forest at 20 years because, while logged areas >20 years may still provide browse, these stands more closely resembled mature forest. in addition, older harvests were difficult to distinguish or map accurately. open and developed were absent from almost all core area habitat use and were later dropped from the analysis. we manually digitized cover and land use within the cores in arcgis 9.3 (esri 2008) using a compilation of available gis base-layers from the massachusetts office of geographic information (massgis; massgis 2011) and other sources, including 2005 and 2009 orthophotos, department of environmental protection wetland layers, forest harvest information from the massachusetts department of conservation and recreation (dcr) and harvard forest (mcdonald et al. 2006), 2003 and 2009 national agricultural imagery program (naip) satellite imagery, and mid-1990s black and white orthophotos, as well as state wetland layers for vermont and new hampshire. we assessed habitat availability within the home range by generating sets of 250 random points within 100% annual mcp home ranges using the generate random points tool in hawth's tools (beyers 2006, wattles 2011) and manually classifying cover and land use. we calculated use:availability ratios by comparing cover and land use within 30 m ud core areas (used) to mcp home ranges (available) (aebischer et al. 1993). use:availability ratios >1 indicated a cover type was used more than available; a ratio <1 indicated use was less than available. calving sites were identified based on large decreases in daily movement of cows followed by a concentration of gps locations during the calving season (may– june) (poole et al. 2007). analyses we used analysis of variance (anova) to analyze the differences in habitat availability, core area habitat use, and use:availability ratios within and between sexes, seasons, and portions of the study area. we used type iii anova to account for unequal sample sizes among groups and seasons. we performed pairwise comparisons using alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 137 tukey's contrasts with adjusted p-values using the single-step method. significance level for all analyses was set at 0.05. we used r, version 2.12.2 (r development core team 2005) for all statistical analyses. results capture and deployment of gps collars we deployed gps collars on 26 adult moose (7 females and 19 males); 5 were excluded due to mortality, suspected infection with brainworm (parelaphostrongylus tenuis), or collar failure. data analysis included 5 females and 8 males in central and 8 males in western massachusetts. nine moose were recaptured and recollared when the batteries in their initial gps collars ran low. we obtained 127,408 locations from the 21 moose with an overall fix rate of 85%. seasonal data for any animal were included in the analyses only if data were obtained across the entire season. the median number of locations per animal per season ranged from 402 in spring to 1,015 in late winter. the minimum number of locations was 281 for one animal in spring. home range core area habitat use habitat use within seasons. regenerating forest was used more than all other cover types by both central males and females during all seasons (proportion of use 0.48 to fig. 2. mean proportional seasonal core area habitat use for female (n = 5, 5, 4, 5, and 5 individuals for spring, summer, fall, early winter, and late winter, respectively) and male (n = 7, 7, 7, 6, and 7 individuals for spring, summer, fall, early winter, and late winter, respectively) moose in central massachusetts and male moose in western massachusetts (n = 7, 6, 4, 8, and 7 individuals for spring, summer, fall, early winter, and late winter, respectively). error bars represent standard errors of the means. 138 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 0.63; p ≤ 0.006), with the exception of wooded wetlands, mixed forest, and conifer forests during spring by females (fig. 2). no other differences in seasonal core area habitat use were significant for central males or females. both central males and females showed selection for regenerating forest during all seasons (table 2), except for females during spring. central males also showed selection for wooded wetlands during fall. all other habitat types were either used in proportion to or less than their availability. the lack of selection of regenerating forests by females in spring was likely because calving areas dominated female spring habitat use; calving sites varied among individuals and included wooded wetlands, mature mixed and conifer-dominated mixed stands, and mixed and conifer shelter cuts. the importance of vegetative cover type varied with season for western males (table 2, fig. 2), that selected for deciduous forest and used it (proportional use = 0.41) more than all other habitat types except regenerating forest during spring (p ≤ 0.03). regenerating forest (0.22) was also used more than other wetlands at this time of year (p = 0.03). during summer (0.57) and fall (0.46) regenerating forest use was greater than all other habitat types (p ≤ 0.02); however, it was used more than its availability only during summer. no other habitat types were used more than their availability at any other time of year. forest table 2. p-values for anova of use:availability ratios. dark gray indicates use:availability >1, light gray <1, and white use not significantly different than availability. spring summer fall early winter late winter females coniferous 0.688 0.329 0.066 0.027 0.611 mixed 0.219 <0.001 0.001 0.030 0.070 deciduous 0.008 0.045 0.228 0.368 0.561 regenerating 0.089 0.008 0.010 0.021 0.028 wooded wetland 0.445 0.213 0.523 0.113 0.090 other wetland 0.061 0.958 0.956 0.004 0.006 central males coniferous 0.508 0.458 0.035 0.940 0.898 mixed <0.001 <0.001 <0.001 0.786 0.145 deciduous 0.687 0.059 0.088 0.072 0.066 regenerating 0.004 0.002 <0.001 0.039 0.003 wooded wetland 0.077 0.063 0.002 0.358 0.004 other wetland 0.007 0.190 0.482 0.001 0.045 western males coniferous 0.037 0.024 0.002 0.089 0.188 mixed 0.369 <0.001 0.165 0.393 0.053 deciduous 0.014 0.165 0.809 0.300 0.505 regenerating 0.885 <0.001 0.083 0.249 0.360 wooded wetland 0.677 0.988 0.181 0.027 0.049 other wetland 0.139 0.609 0.475 0.722 0.019 alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 139 types with a conifer component (mixed and coniferous forest) combined to be most used in early (0.47) and late winter (0.65). high use of regenerating forest continued in early winter (0.32). habitat use among seasons. there were no differences in the use of various vegetative cover types by females among seasons. central males used more mixed forest in early winter than fall (p = 0.046). central males used wooded wetlands more during fall than early and late winter (p < 0.001), and more during both spring and summer than in late winter (p ≤ 0.02). western males used less conifer forest in their home range cores during spring, summer, and fall than in early winter (p ≤ 0.03), and less in fall than late winter (p ≤ 0.04). similarly, they used more mixed forest in late winter cores than all other seasons, but only significantly more in fall (p≤ 0.01). western males use regenerating forests more during summer than spring (p = 0.02) or late winter (p = 0.01), while use of deciduous forest was greater during spring than summer and early or late winter (p ≤ 0.04). wooded wetland use was greater during fall than early and later winter (p ≤ 0.03). habitat use based on gender and region. there were no differences in seasonal core area habitat use between central males and females. however, western males used deciduous forest more than central males or females during spring and fall (p ≤ 0.01), and more coniferous forest during early winter (p ≤ 0.02). western males also used more mixed forest than central males during late winter (p = 0.03), but less regenerating forest (p = 0.04) than central males during spring, and less regenerating forest than either central males or females in late winter (p ≤ 0.01). discussion not all areas within a home range hold equal importance to the animal. if food and other resources are unevenly distributed, areas of higher densities of critical resources should be more important than areas with lower levels of that resource (powell 2000). if animals focus their use in some portion of the home range where resources are concentrated, those areas represent centers of activity or cores of the home range (hayne 1949, kaufmann 1962, samuel et al. 1985, powell 2000). due to the concentrated use of these areas, home range cores may be critically important to an individual's survival and reproductive success. identifying home range core areas and core area habitat can provide important insights into the ecology of a species and its survival strategies. this is particularly important for managers in southern new england, where moose have only recently re-established after many decades of absence and where habitat differs from much of the rest of their geographic range. the typical annual pattern of habitat use by moose reflects the seasonal availability of resources (peek 2007). sites that optimize forage quantity and quality vary by forest type and season and are a main driver of the vegetative cover types that moose select (telfer 1988, westworth et al. 1989, mccracken et al. 1997, poole and stuartsmith 2005, peek 2007). as a result, habitat use follows a familiar pattern across their geographic range (peek 2007). for example, moose in our study were extremely reliant on young, regenerating forest for browse (phillips et al. 1973, pierce and peek 1984, bangs et al. 1985, mccracken et al. 1997, poole and stuart-smith 2005, peek 2007, gillingham and parker 2008); used wetlands for thermal cover (renecker and hudson 1986) and some summer forage (ritcey and verbeek 1969, jordan et al. 1973, crossley 140 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 and gilbert 1983, morris 2002); browsed conifers such as balsam fir (where available) and hemlock in winter (crossley and gilbert 1983, thompson et al. 1995); and used conifers as cover during warm periods (schwab and pitt 1991, dussault et al. 2004) and periods of deep snow (peek et al. 1976, monthey 1984, thompson et al. 1995). however, we found large differences in the availability and distribution of some vegetative cover types compared to the rest of the species range, which resulted in differences in habitat selection. probably the most important difference was the amount and distribution of early successional forest habitat and the processes that create these habitats. while this cover type was heavily used by moose, large disturbances that create it – either natural (fire, wind, insects) or human-caused (logging) – are rare and becoming rarer in southern new england. the amount and distribution of timber harvesting activities is minimal as compared to many other regions, and large-scale natural processes such as flooded river deltas, sup-alpine and riparian shrub communities, and avalanche corridors do not exist. in addition, some key woody species such as willows (salix spp.), aspen, mountain-ash (sorbus americana), and other shade-intolerant species, all of which provide high quality browse for moose in more northern regions, are not abundant in southern new england. with the exception of wetlands and small-scale logging, the undeveloped portion of the massachusetts landscape is nearly 100% closed canopy mixedconiferous-deciduous forest. as a result, moose use the various cover types of closed canopy forest, small wetlands, and patches of young forest created by logging. additionally, while wetlands that supported aquatic vegetation were used throughout spring, summer, and fall, and these sites likely provided critical nutrients, their importance as feeding sites was relatively low compared to regenerating forests in our study area and to wetlands elsewhere in moose range (jordan et al. 1973, crossley and gilbert 1983, ritcey and verbeek 1989, morris 2002). similarly, while roadside salt licks are commonly used by moose in northern new hampshire (miller and litvaitis 1992, scarpitti et al. 2005), we saw no indication of the use of roadside wetlands that would indicate their use as salt licks. we also saw clear differences in forest cover use between central and western massachusetts, and by extension between the forest types of southern and northern new england. the most important factor was likely the transition across the state from spruce-fir-northern hardwoods and northern hardwoods-hemlockwhite pine forest to the transition hardwoodwhite pine-hemlock and central hardwoodhemlock-white pine forest types, and the associated changes in plant communities and structure. the forests in the berkshire mountains of western massachusetts are similar to forests in southern vermont and new hampshire (degraaf and yamasaki 2001), and use of these forests reflected many similar habitat patterns that have been reported in northern new england (crossley and gilbert 1983, leptich and gilbert 1989, thompson et al. 1995, scarpitti 2006). conifer and mixed-coniferous-deciduous stands, with balsam fir and hemlock, were important cover types during winter in western massachusetts, as in northern new england (crossley and gilbert 1983, thompson et al. 1995, scarpitti 2006). balsam fir occurred in the spruce-fir-northern hardwood forests at the highest elevations in western massachusetts, but it was absent in central massachusetts and lower elevations in western massachusetts. with the absence of balsam fir, hemlock was the only conifer that was a large portion of the winter diet of moose; white pine was avoided (faison et al. 2010). while the use of stands of hemlock and mixed stands with hemlock alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 141 and deciduous shrubs and saplings increased in central massachusetts during winter, the lack of balsam fir increased reliance on high-density regenerating stands of hardwoods. additionally, typically less restrictive snow conditions in central massachusetts (20–60 cm in late winter) may have played a role in the increased use of regenerating stands in late winter, while deep snow (80–110 cm in late winter) in western massachusetts may have forced moose into the shelter of spruce-fir stands. similarly, western males used deciduous forests more in spring and fall compared to moose in central massachusetts. favored deciduous species, such as hobblebush (viburnum lantanoides), striped maple (a. pensylvanicum), beech (early in the growing season), and aspen were less common in central massachusetts, and the reduced availability of these key species seemed to limit use of deciduous forest in central compared to western areas. the dominant habitat type used by moose throughout the state was regenerating forest created by logging. in central massachusetts moose used areas of forest regeneration intensively in all seasons. while use of regenerating stands in western massachusetts was more variable, moose still concentrated in these sites, especially during summer. early seral stage forest stands provided a concentrated source of abundant browse during the growing season (mcdonald et al. 2008), which allow moose to maximize their forage intake without moving over large areas (belovsky 1981, wickstrom et al. 1984). the use and selection of these sites during summer (≥57% of home range core areas by all groups) suggests that moose relied on regenerating forests to provide the forage required to gain weight at this critical time of year (belovsky and jordan 1978, van ballenberghe and miquelle 1990). the recent pattern of logging in massachusetts appeared to be favorable to moose. harvest sites on state and private lands were widely distributed, with <2% of the forested landscape logged annually (kittredge et al. 2003, mcdonald et al. 2006). this resulted in new patches of early successional habitat within a matrix of mature and maturing forest. the importance of thermal cover for moose in and around forest harvests and burns has been well documented (mcnicol and gilbert 1980, girard and joyal 1984, bangs et al. 1985, masterbrook and cummings 1989, thompson et al. 1995). the small size (mean = 16.5 ha) and moderate harvest intensity (27% of timber volume harvested) of forest harvest units in massachusetts (kittredge et al. 2003) resulted in short distance to edge, which provided both browse and cover in close proximity. shelterwood cuts were commonly applied to harvest units, resulting in cover from solar radiation along with browse, with the added advantage that vegetation growing in shade tends to be more nutritious and has lower secondary compound levels than growth in direct sunlight (hjeljord et al. 1990, schwartz and renecker 2007). the intense use of regenerating forests is similar to habitat use in northern new england and elsewhere (peek et al. 1976, joyal and scherrer 1978, crossly and gilbert 1983, monthey 1984, leptich and gilbert 1989, thompson et al. 1995, scarpitti 2006). however, both leptich and gilbert (1989) and miller and litvaitis (1992) found that only females selected for cut-over areas during summer in northern new hampshire and northern maine, with males selecting upland hardwoods; scarpitti (2006) found selection for regenerating stands only during winter. the high concentration of browse found in regenerating stands mimicked the permanent shrub communities used by moose in other portions of their range, including delta floodplains, tundra and subalpine areas, aspen parklands, and stream valley shrub communities, as well as the 142 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 transitory early successional habitats created by fire and insect outbreaks (phillips et al. 1973, pierce and peek 1984, bangs et al. 1985, mccracken et al. 1997, poole and stuart-smith 2005, peek 2007, gillingham and parker 2008). the clear importance of early successional forest as foraging habitat for moose, however, should not take away from that fact that moose used a mix of cover types and age classes to meet their annual habitat needs in southern new england. mature coniferous, mixed, and deciduous stands were seasonally important foraging sites. additionally, moose used mature forests and a variety of wetlands as thermal shelters during periods of high temperature, and mature coniferous and mixed stands during periods of deep snow. while moose now occupy most suitable habitat in massachusetts and connecticut, additional habitat may exist in unoccupied portions of its historic range in new york and pennsylvania. forest types transition in new york and pennsylvania in a similar way as in massachusetts, from spruce-fir and northern hardwood forests to transitional and central hardwood forests, and suitable habitat likely exists for moose in the forests types of southern new york and pennsylvania. however, different state management goals (wattles and destefano 2011), greater amounts of agriculture, the highly developed mohawk river valley, and high temperatures may prevent or slow the further expansion of moose in this region. management implications the year-round intensive use of regenerating forests by moose in massachusetts underlies the importance of early successional forest. the recent pattern of logging that continually created new patches of young forest seemed to be favorable for moose. however, recently adopted plans by the dcr (the agency that manages the state forest system and public watersheds in massachusetts) that restrict or eliminate logging on some state lands could have a negative impact on moose and other wildlife that use or require early successional forest. moose rely on these sites of high forage density to gain weight for winter and support lactation of calves. a reduction in logging would result in a loss of this cover type over time from some of the largest tracts of conservation land and would force moose to forage for lower density browse in mature forest stands. this could result in higher energy expenditures to obtain the same amount of food, which may be particularly harmful for a species living in an environment at the extremes of its temperature tolerances. that heat stress has been implicated in the recent declines in moose populations elsewhere along the southern edge of the species’ range (murray et al. 2006, lenarz et al. 2009, 2010) demonstrates the importance of energy balance for moose living in these environments. management of moose habitat on a landscape scale in massachusetts should ensure the protection of large blocks of forested habitat that support a mix of age classes and forest cover types, including mature stands of coniferous, mixed-coniferousdeciduous, and deciduous forests, patches of early successional forest, and a variety of wetlands. the mix of cover types, age classes, and wetlands that currently occur in the temperate deciduous forests of massachusetts and southern new england appear to provide suitable habitat for long-term occupation by moose. in general, moose are relatively widely dispersed, actively reproducing, and present at low density in almost all forest types in central and western massachusetts. the absence of major predators and hunting undoubtedly influence the population dynamics of moose in massachusetts. the differences in the distribution, structure, and landscape alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 143 configuration of key habitat components, along with large levels of development and a potentially thermally stressful environment will likely combine to limit the distribution and density of moose in southern new england. acknowledgments the massachusetts division of fisheries and wildlife (mdfw) through the federal aid in wildlife restoration program (w-35-r) provided funding and support for this research. the massachusetts dcr, u.s. geological survey, university of massachusetts-amherst, and safari club international provided additional funding and logistical support. capture of moose would not have been possible without the assistance of field technician k. berger, the massachusetts environmental police and mdfw personnel, and other technicians and volunteers. we thank r. deblinger for logistic support and t. k. fuller and j. mcdonald for reviewing the manuscript. references aebischer, n. j., p. a. robertson, and r.e. kenward. 1993. compositional analysis of habitat use from animal radio-tracking data. ecology 74: 1313–1325. bangs, e. e., s. a. duff, and t. n. bailey. 1985. habitat differences and moose use of two large burns on the kenai peninsula, alaska. alces 21: 17–35. belovsky, g. e. 1981. optimal activity times and habitat choice of moose. oecologia 48: 22–30. ———, and p. a. jordan. 1978. the timeenergy budget of a moose. theoretical population biology 14: 76–104. beyers, h. 2006. hawth's analysis tools for arcgis. version 3.27. . boose, e. 2001. fisher meteorological station (since 2001). harvard forest data archive: hf001. petersham, massachusetts, usa. coady, j. w. 1974. influence of snow on behavior of moose. canadian fieldnaturalist 101: 417–436. crossley, a., and j. r. gilbert. 1983. home range and habitat use of moose in northern maine. transactions of the northeast section of the wildlife society 40: 67–75. degraaf, r. m., and m. yamasaki. 2001. new england wildlife: habitat, natural history, and distribution. university press of new england, hanover, new hampshire, usa. destefano, s., r. d. deblinger, and c. miller. 2005. suburban wildlife: lessons, challenges, and opportunities. urban ecosystems 8: 131–137. dussault, c., j-p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioural responses of moose to thermal conditions in the boreal forest. ecoscience 3: 321–328. environmental systems research institute, inc. (esri). 2008. arcgis 9.3. redlands, california, usa. faison, e. k., g. motzkin, d. r. foster, and j. e. mcdonald. 2010. moose foraging in the temperate forests of southern new england. northeastern naturalist 17: 1–18. fieberg, j. 2007. kernel density estimators of home range: smoothing and the autocorrelation red herring. ecology 88: 1059–1066. franzmann, a. w., and c. c. schwartz. 2007. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. garner, d. l., and w. f. porter. 1990. movement and seasonal home ranges of bull moose in a pioneering adirondack population. alces 26: 80–85. gillingham, m. p., and k. l. parker. 2008. the importance of individual variation in defining habitat selection by moose in northern british columbia. alces 44: 7–20. 144 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 http://www.spatialecology.com/htools http://www.spatialecology.com/htools girard, f., and r. joyal. 1984. l'impact des coupes a blanc mecanisees sur l'original dans le nord-ouest du quebec. alces 20: 3–25. gitzen, r. a., j. t. millspaugh, and b. j. kernohan. 2006. bandwidth selection for fixed-kernel analysis of animal utilization distributions. journal of wildlife management 70: 1334–1344. hall, b., g. motzkin, d. r. foster, m. syfert, and j. burk. 2002. three hundred years of forest and land-use in massachusetts, usa. journal of biogeography 29: 1319–1335. hayne, d.w. 1949. calculation of size of home range. journal of mammalogy 30: 1–18. hemson, g., p. johnson, a. south, r. kenward, r. ripley, and d. macdonald. 2005. are kernels the mustard? data from global positioning system (gps) collars suggests problems for kernel home-range analyses with least-squares cross-validation. journal of animal ecology 74: 455–463. hjeljord, o., n. hovik, and h. b. pedersen. 1990. choice of feeding sites by moose during summer, the influence of forest structure and plant phenology. holarctic ecology 13: 281–292. johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61: 65–71. jordan, p. a., d. b. botkin, a. s. dominski, h. s. lowendorf, and g. e. belovsky. 1973. sodium as a critical nutrient for the moose of isle royale. proceedings of the north american moose conference and workshop 9: 13–42. joyal, r., and b. scherrer. 1978. summer movements and feeding by moose in western quebec. canadian field-naturalist 92: 252–258. kaufmann, j. h. 1962. ecology and social behavior of the coati, nasua nifica on barro colorado island panama. university of california publications in zoology 60: 95–222. kie, j. g., j. matthiopoulos, j. fieberg, r. a. powell, f. cagnacci, m. s. mitchell, j-m. gaillard, and p. r. moorscroft. 2010. the home-range concept: are traditional estimators still relevant with modern telemetry technology. philosophical transactions of the royal society b 365: 2221–2231. kittredge, d. b., jr., a. o. finley, and d. r. foster. 2003. timber harvesting as ongoing disturbance in a landscape of diverse ownership. forest ecology and management 180: 425–442. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. ———, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–510. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53: 880–885. massachusetts office of geographic information (massgis). 2011. (accessed january 2011). masterbrook, b., and h. g. cummings. 1989. use of residual strips of timber by moose within cutovers in northwestern ontario. alces 25: 146–155. mccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. wildlife monographs 136: 3–52. mcdonald, r. i., g. motzkin, m. s. bank, d. b. kitteridge, j. burke, and d. r. foster. 2006. forest harvesting and alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 145 http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/layerlist.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/layerlist.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/layerlist.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/layerlist.html http://www.mass.gov/anf/research-and-tech/it-serv-and-support/application-serv/office-of-geographic-information-massgis/datalayers/layerlist.html land-use conversion over two decades in massachusetts. forest ecology and management 227: 31–41. ———, ———, and d. r. foster. 2008. the effect of logging on vegetation composition in western massachusetts. forest ecology and management 225: 4021–4031. mcnicol, j. g., and f. f. gilbert. 1980. late winter use of upland cutovers by moose. journal of wildlife management 44: 363–371. miller, b. k., and j. a. litvatis. 1992. habitat segregation by moose in a boreal forest ecotone. acta theriologica 37: 41–50. mohr, c. o. 1947. table of equivalent populations of north american small mammals. american midlands naturalist 37: 223–249. monthey, r. w. 1984. effects of timber harvesting on ungulates in northern maine. journal of wildlife management 48: 279–285. morris, k. i. 2002. impact of moose on aquatic vegetation in northern maine. alces 38: 213–218. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monograph 166: 1–30. peek, j. m. 2007. habitat relationships. pages 351–375 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. ———, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3–65. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37: 266–278. pierce, d. j., and j. m. peek. 1984. moose habitat use and selection patterns in north-central idaho. journal of wildlife management 48: 1335–1343. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammalogy 88: 139–150. ———, and k. stuart-smith. 2005. finescale winter habitat selection by moose in interior montane forests. alces 41: 1–8. powell, r. a. 2000. animal home ranges and territories. pages 65-110 in l. boitani and t. k. fuller, editors. research techniques in animal ecology. columbia university press, new york, new york, usa. r development core team. 2005. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. . renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. ritcey, r. w., and n. a. m. verbeek. 1969. observations of moose feeding on aquatics in bowron lake park, british columbia. canadian field-naturalist 83: 339–343. rogers, a. r., a. p. carr, h. l. beyer, l. smith, and j. g. kie. 2007. hrt: home range tools for arcgis. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. samuel, m. d., d. j. pierce, and e. o. garton. 1985. identifying areas of concentrated use within home range. journal of animal ecology 54: 711–719. scarpitti, d. l. 2006. seasonal home range, habitat use, and neonatal habitat characteristics of cow moose in northern new hampshire. m.s. thesis, university of 146 habitat in massachusetts – wattles and destefano alces vol. 49, 2013 http://www.r-project.org http://www.r-project.org new hampshire, durham, new hampshire, usa. ———, c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. schwab, f. e., and m. d. pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071–3077. schwartz, c. c., and l. a. renecker. 2007. nutrition and energetics. pages 441-478 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. telfer, e. s. 1988. habitat use by moose in southwestern alberta. alces 31: 153–166. the weather channel. 2011a. monthly averages for petersham, massachusetts. (accessed january 2011). ———. 2011b. monthly averages for windsor, ma. (accessed january 2011). thompson, m. e., j. r. gilbert, g. j. matula, and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31: 233–245. u. s. census bureau. n.d. census. 2000. population housing units, area and density for states: 2000. (accessed june 2010). van ballenberghe, v., and d. g. miquelle. 1990. activity of moose during spring and summer in interior alaska. journal of wildlife management 54: 391–396. vecellio, g. m., r. d. deblinger, and j. e. cardoza. 1993. status and management of moose in massachusetts. alces 29: 1–7. wattles, d. w. 2011. status, movements, and habitat use of moose in massachusetts. m.s. thesis, department of environmental conservation, university of massachusetts, amherst, massachusetts, usa. ———, and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. westveldt, m. r., r. i. ashman, h. i., baldwin, r. p. holdsworth, r. s. johnson, j. h. lambert, h. j. lutz, l. swain, and m. standish. 1956. natural forest vegetation zones of new england. journal of forestry 54: 332–338. westworth, d., l. brusnyk, j. roberts, and h. veldhuzien. 1989. winter habitat use by moose in the vicinity of an open pit copper mine in north-central british columbia. alces 25: 156–166. wickstrom, m. l., c. t. robbins, t. a. hanley, d. e. spalinger, and s. m. parish. 1984. food intake and foraging energetic of elk and mule deer. journal of wildlife management. 48: 1285–1301. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70: 164–168. alces vol. 49, 2013 wattles and destefano – habitat in massachusetts 147 http://www.weather.com/weather/wxclimatology/monthly/graph/usma0497 http://www.weather.com/weather/wxclimatology/monthly/graph/usma0497 http://www.weather.com/weather/wxclimatology/monthly/graph/usma0497 http://www.weather.com/weather/wxclimatology/monthly/graph/usma0497 http://www.weather.com/weather/wxclimatology/monthly/graph/usma0497 http://www.census.gov/population/www/censusdata/density.html http://www.census.gov/population/www/censusdata/density.html http://www.census.gov/population/www/censusdata/density.html moose habitat in massachusetts: assessing use at the southern edge of the range study area methods study animals and gps telemetry seasons habitat availability and core area habitat use analyses results capture and deployment of gps collars home range core area habitat use discussion management implications acknowledgments references alces28_223.pdf alces21_139.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces(25)_15.pdf alces20_1preface.pdf alces vol. 20, 1984 alces27_1.pdf alces20_61.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces21_299.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces24_133.pdf alces22_449_distinguishedmoosebio.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces26_108.pdf alces(25)_172.pdf alces vol. 45, 2009 sipko – reintroduction of large herbivores in russia 35 status of reintroductions of three large herbivores in russia taras p. sipko institute of ecology and evolution ras, leninskii pr. 33, 119071 moscow, russia. abstract: reintroductions of muskoxen (ovibus moschatus), european bison (bison bonasus), and moose (alces alces) have occurred recently in russia. although the process of capturing and moving muskoxen was problematic in remote areas, the reintroduction of animals from canada and the usa successfully restored this extirpated species, and the current population in northern russia serves as a source for further transplants. european bison populations were stagnant and suffered from inbreeding in russia prior to reintroduction of captive animals from throughout europe. the population in orlovskoye polesie national park has experienced population growth with improved genetic potential. of concern is that reintroductions in other areas of russia were unsuccessful and the global population of european bison is not improving. moose from the penzhina river area in russia were successfully reintroduced to the kamchatka peninsula where they were absent for >400 years. the population is growing and dispersing across the peninsula from the transplant sites, and is among the largest physically in eurasia. alces vol. 45: 35-42 (2009) key words: alces alces, bison bonasus, population, reintroduction, restoration, ovibus moschatus, russia. the primary goal of reintroducing large herbivores in russia is to restore biological diversity in northern ecosystems. a secondary goal is to provide a dependable and renewable food supply for residents of northern russia. in this paper i summarize reintroduction efforts with muskoxen (ovibus moschatus), european bison (bison bonasus), and moose (alces alces) that were undertaken for different ecological reasons and circumstances. prior to the reintroduction efforts muskoxen were extirpated, the resident population of european bison was stagnant and suffered from inbreeding associated with a small founder population, and moose, although increasing in certain areas of russia, were absent for centuries from the proposed reintroduction area. shorter and earlier descriptions of these efforts can be found in "re-introduction news" a newsletter of the iucn (sipko et al. 2006, sipko and gruzdev 2006, sipko and mizin 2006). reintroduction of muskoxen in northern russia background and approach a considerable part of russia’s landmass borders the arctic ocean and has severe climatic conditions including long periods of cold temperatures. remains of muskoxen discovered on the taimyr peninsula were 2000-4000 years old indicating that they inhabited the region within relatively recent geological time (vereshagin and barishnikov 1985). previous research indicated that this region was capable of supporting >2 million muskoxen without damage to the fragile northern ecosystem. reintroduction of muskoxen into suitable habitat was expected to have a positive impact on the ecological community because of increased utilization of vegetation resulting in faster turnover of energy at all trophic levels. because yakushkin (1998) found that immature male muskoxen dispersed up to 800 km from reintroduction of large herbivores in russia – sipko alces vol. 45, 2009 36 their natal area, the plan called for herds to be reintroduced within 600-700 km of each other to encourage genetic interchange. sites along the shoreline of the arctic ocean were selected for the first reintroduction that occurred in 1974 when 10 animals were delivered from canada (banks island) to the eastern part of the taimyr peninsula. it was successful and eventually muskoxen spread north, east, and south (putorana plato) of the taimyr peninsula. the population was estimated at 2,500 in 2002 (sipko et al. 2003), and was nearly 4,000 by 2005. a second reintroduction on vrangel island in 1975 used 20 muskoxen obtained from the usa (nunivak island, alaska). population growth was slow because mortality was high in the initial acclimatization period. by 2003, the population was 750 animals (gruzdev and sipko 2003), and recent estimates indicate that the population has stabilized at 800-850. additional reintroductions were achieved by relocating muskoxen from the vrangel island and taimyr peinsula populations (table 1). the objective was to capture and relocate muskoxen that were 0.3-3.5 years old, however, most captured animals were 0.5 years old. vrangel island is a nature preserve and only vehicles with low-pressure tires and snowmobiles are allowed. once located, muskoxen were surrounded by people with dogs and selected animals were chemically immobilized with a dart gun. sedated animals were isolated from the herd and placed into containers for transportation to a holding enclosure. after the capture quota was met, a helicopter transported them to another enclosure on the mainland where they were placed in individual containers, loaded onto an airplane, and transported to the reintroduction site or temporary holding enclosure with local transportation equipment. the methods of capture, handling, and containment on taimyr island were the same as at vrangel island. a helicopter was used to locate and deliver muskoxen either to a temporary holding enclosure or directly to the reintroduction site. they were usually kept in the temporary enclosure for a period before release. discussion the initial reintroduction of muskoxen to northern russia in the 1970s proved to be successful, and set the stage for further reintroductions in other parts of russia (sipko et al. 2007). in addition to those captured for subsequent reintroductions, 81 muskoxen were also captured for zoological parks and domeslocation region year number 1st breeding 2008 population east taimyr peninsula krasnoyarsk 1974, 1975 30 1975 ~6500 wrangel island chukotka 1975 20 1977 ~ 800 bulun yakutia 1996 24 1999 >300 anabar yakutia 1997, 2000 41 2000 >150 begichev island yakutia 2001, 2002 25 2003 >50 allaikhov yakutai 2000 11 2004 64 taas-yrach yakutai 2001-2003 18 2004 12 tamma yakutai 2002, 2003 22 2004 0 polar ural yamal 1997, 1998, 2001, 2003 63 1999 108 kolima magadan 2004 22 none 20 total 284 >8000 table 1. a summary of the location, history, and status of reintroduced herds of muskoxen in northern russia. alces vol. 45, 2009 sipko – reintroduction of large herbivores in russia 37 tication experiments. regular surveillance of muskoxen is hampered by the remoteness of the reintroduction areas. surveys conducted in yakutia in 2005 indicated that 347 muskoxen were in 4 reintroduction areas, a >3x population increase in 10 years. the fastest growth rate occurred in the allaikhov herd where 18 calves were produced in 3 years. the muskoxen population in the bulun region split into 2 nearly equal sized herds; one herd dispersed 120 km west to the delta of the lena river where it resides currently. reintroductions of muskoxen in northern russia are problematic. difficult working conditions, remote locations, and numerous animal transfers with different modes of transportation meant that animals needed skilled animal care specialists to accompany them. the overall mortality rate was 10-15% during the process of capture and containment prior to release. muskoxen from the taimyr peninsula adapted well to relocation sites in central siberia, and those from vrangel island established viable herds in eastern russia. presumably, the introduction of muskoxen from 2 separate populations will have a positive impact on genetic variability that influences survival, productivity, and stability of new herds. a program for additional introductions is underway, and plans are in development to introduce muskoxen to the mountain ranges of northern asia. reintroduction of european bison in central russia background and approach the global population of european bison has not expanded in the past 15 years, remaining at approximately 3,000. small herds scattered in free-roaming populations and captive-rearing facilities typically consisted of 5-7 ancestors from the 12 original animals that were founders of all contemporary bison (belousova 1993). these circumstances caused high inbreeding coefficients in the lowland populations (44%) and moderate coefficients (26%) in the lowland-caucasian line (olech 1998). recent studies indicated that inbreeding occurred over a much longer period of time, thus, the actual inbreeding coefficient is probably much higher. as a result, phenotypic expressions of inbreeding depression are evident in certain populations (sipko 2002). it is estimated that an effective population size of 500 individuals is required to preserve genetic polymorphism that enables a population to adapt and evolve in a constantly changing environment to prevent extinction (soule and wilcox 1980). an effective population must have an adequate sex ratio and sufficient mature and sexually active animals that comprise 25-35% of the population. thus, establishing a population of 1,500-2,000 should ensure long-term viability and survival of a species. however, at least 2 geographically isolated populations are deemed necessary to reduce the risk of disease or unforeseen events that might decimate a population. russia has sufficient geographical area of suitable habitat to accommodate multiple distinct populations of european bison. two areas of appropriate size and ecological conditions were selected for the reintroductions; importantly, they also offered protection as designated wildlife reserves. one area was in the european broadleaf forest of the bryansk-oryol-kaluga region in the european (i.e., central) part of russia. it was a large, contiguous forest tract extending from the boundary of ukraine along the desna river (black sea basin) northeast to the oka river (caspian sea basin), and had few natural or artificial barriers to impede bison movement; railways and highways bisected the region in only one location. the forest was 30-50 km wide and stretched >400 km. the eastern section was known as the oka defense line of the state of moscow during the 14th-17th centuries. these types of frontier forests acted as a line of defense reintroduction of large herbivores in russia – sipko alces vol. 45, 2009 38 until the middle of the 18th century and were strictly protected from harvesting and use by people. as a result, this tract of contiguous broadleaf forest was one of only a few areas that remained largely intact in its natural state, and is a protected natural area. designated sections that contributed to this forest tract included 1) the desnaynskostarogutsky national park (ukraine) with 162 km2 area situated on the boundary of ukraine and russia, 2) the bryansky les biosphere reserve with 1230 km2 in the bryansk region of russia, and 3) the adjacent protected areas of the oryol and kaluga regions in the north including orlovskoye polesie national park (777 km2), kaluzhskie zaseki state nature reserve (185 km2), and ugra national park and biosphere reserve (986 km2). the second area used to reintroduce bison was the ust-kubenskoye hunting facility and surrounding region located about 400 km north of moscow in the vologda region (vologodskai oblast) on the 590 n latitude parallel. the landscape is a series of raised terrace plains at 110-200 m elevation within the severnai dvina river drainage; 64% was forest dominated by conifers (55%, mostly abies spp.). the ustkubenskoye hunting facility is 260 km2 with the russkii sever national park (1664 km2) at the western boundary. the southern border begins at the shore of lake kubenskoe, and the northern and western boundaries adjoin 8,000 km2 of federal forest lands. this area was considered optimal because european bison have survived there long-term without human support. further, resident animals have shown evidence of twinning that is uncommon in bison, suggesting high habitat quality. logging has produced large areas of secondgrowth forest with a shrub-layer suitable for foraging, and most agricultural areas were abandoned and these regenerating lands also provide rich food resources for bison. bison used in the reintroduction were from captive breeding centers in russia and west europe (table 2) to potentially enhance their future genetic viability. their gene pools were quite distinct because bison from russia and west europe have been isolated for almost 100 years. bison were transported in group or individual containers to temporary enclosures and released 1-2 months later. location region year number 2008 population comments cherga mountains altay 1982-1984 12 34 orlovskoye pollesye national park (opnp) oryol 1996-2001, 2006 75 143 successful kaluzhskie zaseki state nature reserve kaluga 2001 0 na dispersal from opnp petrovskoe hunting facility kaluga 2,007 9 9 ust-kubenskoye hunting facility vologda 1991, 1994 5 24 vilikoozerskoe hunting facility vladimir 1989, 1994, 2002, 2004, 2007 25 15 4 bison present at 2nd transplant muromskij sanctuary vladimir 2001-2004 13 16 sknjatinskoe hunting facility tver 1986, 1991 33 3 unsuccessful branskij les state nature reserve bryansk 1999-2000 11 0 unsuccessful total 183 >225 table 2. a summary of the location, history, and status of reintroduced herds of european bison in northern russia. alces vol. 45, 2009 sipko – reintroduction of large herbivores in russia 39 discussion our method of transporting bison over long distances in individual containers proved successful. based on our experience, using large containers containing several animals, as done with cattle, was problematic and should be avoided if possible. bison often injured each other during group transport, resulting in high injury and mortality rates during the reintroduction effort. the bison population in the orlovskoye polesie national park has the greatest genetic potential compared to other bison groups in the world (table 3). the population is growing rapidly with 20 calves born in 2005. animals have dispersed to areas adjacent to the park and regularly appear in the kaluzhskie zaseki state nature reserve. it appears that 3 separate herds have formed from the original population. additional releases into these areas are needed in order to quickly establish optimally sized populations. the region situated between the volga and oka rivers contains a large bison population. however, the region has much industry and transport and communication lines and this lowland area has little coniferous forest. the vilikoozerskoe hunting facility, muromskijj sanctuary, and the sknjatinskoe hunting facility located here have insufficient area for further expansion of the population. slow population growth and numerous mortalities are evident, and further reintroductions are not considered worthwhile. the release of bison into the bryansky les state nature reserve proved to be unsuccessful. long migration patterns and poaching in the russia-ukraine border areas resulted in their demise. there is need to supplement the bison population in the ust-kubenskoye hunting facility. also, the introduction of captive bison from the netherlands into the bukovina population in east carpathian had limited success. the introduced males were unable to compete during the rut with local bulls native to the rugged mountain conditions. it is concerning that european bison numbers are not increasing worldwide. the european bison pedigree book (2002) noted that overall growth was weak with 172 births and 112 deaths overall. arguably, there are insufficient animals to successfully establish new viable populations (sipko and kazmin 2004). a new reintroduction effort in yakutia will focus on a captive breeding and release program. in 2006, 30 wood bison were donated and transported from alberta, canada and relocated in a fenced enclosure 120 km from the city of yakutsk. successful births occurred in 2008 (6) and 2009 (7); 4 of the original animals have died. the long term plan is to establish a free-ranging population in yakutia through gradual release of young bison. reintroduction of moose to the kamchatka peninsula background and approach the kamchatka peninsula has been occupied and developed by russians since the 17th century with no evidence of moose country source number russia prioksko-terrasnyjj zapovednik 21 okskijj zapovednik 24 zoo rostov na donu 1 zoo st. peterburg 1 belarus belovezhskaja pushha 2 netherlands natuurpark lelystad 14 germany springe 6zoo dortmund tierpark chemnitz switzerland tierpark dahlholzly 4 wildpark langenberg finland zoo helsinki 1 belgium han-sur lesse 1 total 75 table 3. origin of european bison transplanted in the orlovskoye pollesye national park, russia. reintroduction of large herbivores in russia – sipko alces vol. 45, 2009 40 inhabiting this region during that time. the wildlife of kamchatka is low in diversity when compared to the mainland and both lynx (felis lynx) and squirrel (sciurus vulgaris) appeared only in the 20th century (valentsev and mosolov 2004). however, archeological evidence indicates that moose were present during the 11th-16th centuries in southern and eastern areas of the peninsula (vereschsgin and nikolaev 1979); this information lead to the interest of reintroducing moose on the kamchatka peninsula. moose have continuously inhabited northeastern mainland russia, but populations have been low in recent centuries. growth of these populations has been documented only since the mid-20th century. the human population has been localized in small settlements leaving vast tracts of land without hunting or poaching pressure. further, the whole region was involved in a wolf extermination (poisoning) program. as a consequence, the moose population in the mountain taiga sector of the penzhina river basin expanded to 2000 animals by 1974 (fil 1975), thus was considered a suitable donor population for the reintroduction. after explorating the peninsula, the kamchatka river valley was considered most suitable for moose. to the east, west, and south were mountain ranges that protected this area from deep snow, and food resources in the valley appeared adequate for moose. although the southern kamchatka region had considerable food resources for moose, deep snow and high risk of poaching were considered problematic. snow depth frequently reaches 120 cm but is not uniform throughout the region because of the local effect of wind. because many rivers do not freeze due to frequent thaw cycles and volcanic heat sources, the southern area of the kamchatka peninsula was considered to have highest potential as moose habitat. the reintroduction occurred in 2 stages, each with unique characteristics. the first stage was in 1977-1988 when 63 moose were captured along the tributaries of the penzhina and belaya rivers (pavlov 1999) and subsequently moved to the kamchatka river valley in the central part of the peninsula. captures were selective for 9-month old calves that were caught with the aid of a helicopter; they were pushed toward the treeless part of the bottomland and immobilized by darting from the helicopter. they were transported by helicopter to an enclosure where they remained 5-7 days, after which they were immobilized and placed in individual containers and transported by helicopter to the release site; transport took approximately 9 h. moose were kept in an enclosure at the release site for 15 days to allow them to recover and acclimatize to the local environment; they were released immediately if snow depth exceeded 60 cm. the first birth was documented in 1979. the second stage of the reintroduction occurred in march and april, 2004-2005 (table 4) when 26 moose were captured, transported, and released in the southern part of the kamchatka peninsula. most were 11 months old; 4 animals were 2 years of age. captures occurred in the kamchatka river basin where tall forest cover reduced the effectiveness of the helicopter. when moose were detected, the helicopter drove them towards openings in the forest where snowmobiles were used to overtake the animals. they were immobilized, hobbled, and transported by sled to an individual holding container where they remained until fully recovered (1-3 h), then loaded into the helicopter and transported to the reintroduction site; transportation time was 5-6 h. they were immediately released in the golygina and udochka river valleys. most moose were released on thermal terrain that was snow-free during winter because of volcanic heat. a cow moose with accompanying calf was observed in autumn 2005. discussion a stable moose population appears to be established and a few animals have even alces vol. 45, 2009 sipko – reintroduction of large herbivores in russia 41 dispersed to the western coast of kamchatka. a census conducted in the central kamchatka region in 2004 estimated the population at 1698-1775 animals (sipko et al. 2004). deep snow exceeding 150 cm in winter 2004-2005 caused considerable concern about winter mortality of released moose. however, they overwintered successfully with only a single mortality to a bear (ursus arctos) the following spring. this successful reintroduction of moose is important for developing tourism and hunting on the kamchatka peninsula. steady population growth has been realized in the southern area of the peninsula where twinning predominates. physical measurements of 32 moose indicate that individuals are realizing larger growth than their donor population from central kamchatka. it is presumed that there is optimal forage production and availability on the volcanic substrate, and that moose are using accessible marine grass (seaweed). however, deep snow is of concern relative to selective and heavy browsing pressure during winter. moose inhabiting the kamchatka peninsula are among the largest of all eurasian specimens, and this seems characteristic of the local population. body weight of large bull moose ranges from 600-750 kg, outside spread of antlers are 161.5-181 cm (n = 6; the largest bull was killed in 2002 in ust-kamchatskyi district), and 11 month-old calves weigh 220-325 kg (n = 10). the large size of these animals posed problems during immobilization, handling, and transport. increased drug volumes created considerable difficulty during recovery from drug-induced shock, and their large size caused handling problems during translocations by helicopter. moose appeared to be more sensitive to capture and transportation problems than muskoxen and european bison; much time and effort was spent in recovering them to a normal physical state. one important factor lending success to the second stage of the reintroduction was that the moose had improved resistance to capture and handling stress. in previous efforts capture mortality often exceeded 50%, whereas mortality was insignificant in the second stage of the reintroduction. acknowledgements my most sincere thanks are extended to v. g. tikhonov and s. s. egorov for the source year release area number of moose 2008 population total female penjinskii: palmatkina, essoveem, chichill, and belaya (white) rivers 1977 milkovskii 4 3 1978 milkovskii 9 5 1979 milkovskii 12 5 1980 milkovskii 12 4 1981 milkovskii 26 14 1982 milkovskii and elizovskii ~2000 chukotka: anadyr river valley 1988 smirnihovskii: sakhalin island region 10 7 10-15 kamchatka peninsula: milkovskii area 2004 ust’-bolsheretskii 11 6 2005 ust’-bolsheretskii 15 7 >45 table 4. a summary of the location, history, and status of moose reintroductions on kamchatka peninsula, russia. reintroduction of large herbivores in russia – sipko alces vol. 45, 2009 42 pleasure of the many days of joint work during the reintroduction of muskoxen and bison in yakutia. references belousova, i. p. 1993. influence of inbreeding on viability of european bison in russian breeding centers. pages 29-43 in k voprosu o vozmozhnosti sokhraneniya zubra v rossii. onti pnts ran, push-onti pnts ran, pushchino, russia. (in russian with english summary). fil, v. i. 1975. penszinskii the moose. journal of hunting and hunting economy 3: 12-13. (in russian). gruzdev, a. r., and t. p. sipko. 2003. productivity and demography of muskoxen on vrangel island. rangifer report 11: 30. olech, w. 1998. the inbreeding of european bison population and its influence on viability. 49th eaap meeting, warsaw, poland, august 24-27. abstract only. pavlov, m. p. 1999. acclimatization of the hunting-trade in animals and birds in the ussr. volume 3. moscow, russia. (in russian). sipko, t. p. 2002. zubr (bison bonasus l.). pages 386-405 in a population and genetic analysis: questions of a modern hunting economy. centrohotkontrol, moscow, russia. (in russian with english summary). _____, v. i. fil, and a. r. gruzdev. 2004. moving moose in kamchatka. the siberian zoological conference. september 15-22, 2004, novosibirsk, russia. abstract only. (in russian). _____, _____, and _____. 2006. re-introduction of moose into kamchatka, russia, re-introduction news. iucn/ssc. 25: 26-27. _____, and a. r. gruzdev. 2006. reintroduction of muskoxen in northern russia. re-introduction news. iucn/ ssc. 25: 25-26. _____, _____, and k. n. babashkin. 2003. demography and productivity of muskoxen in taimyr. rangifer report 7: 40-41. _____, _____, v. g. tikhonov, and s. s. egorov. 2007. capturing and reintroduction of muskoxen in the north russia. rangifer report 11: 32. _____, and v. d. kazmin. 2004. modern problems of european bison protection and their solution in russia. pages 123128 in proceedings of european bison conservation. mammals research institute pas, centre of excellence bioter, bialowieza, poland. _____, and i. a. mizin. 2006. re-introduction of european bison in central russia. re-introduction news iucn/ssc. 25: 27-28. soule, m. e., and b. a. wilcox. 1980. conservation biology: an evolutionaryecological perspective. sinauer associates publishers, inc., sunderland, massachusetts, usa. valentsev, a. s., and v. i. mosolov. 2004. lynx in kamchatka peninsula. proceedings of kamchatka branch of pacific institute of geography, far eastern division, russian academy of sciences. petropavlovskkamchatskii pechatnyi dvor publishing house. 5: 10-27. (in russian). vereshchagin, n. k., and g. f. barishnikov. 1985. extinction of mammas in northern eurasia. pages 3-38 in mammals of northern eurasia. l.: zoological institute an, ussr. (in russian). _____, and a. i. nikolaev. 1979. animals hunted by neolithic tribes on the shores of kamchatka. bulletin (moip) moscow communities verifiers nature. branch biology. 84: 40-44. (in russian). yakushkin, g. d. 1998. muskoxen in taimyr. north russian academy of agrarian sciences, novosibirsk, russia. (in russian). alces21_215.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces24_164.pdf alces22_433.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces20_245.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces26_163conferenceworkshop.pdf alces20_129.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces24_69.pdf alces29_225.pdf alces21_preface.pdf alces vol. 21, 1985 alces26_51.pdf alces20_307distmoosebio.pdf alces vol. 20, 1984 alces(25)_133.pdf alces19_118.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 45, 2009 glushkov moose population management in russia 43 improving population management and harvest quotas of moose in russia vladimir m. glushkov research institute of game management and fur farming, kirov, russia. abstract: annual harvest quotas for moose and other game species in russia have been based on population estimates derived from traditional winter track counts and hunter surveys. this labor-intensive approach has failed to account for evident changes in population density of moose. specifically, regional differences in survival and mortality data and the impact of increased poaching are not measured or included in population estimates, and overharvest of moose occurs. i propose implementing a standard management approach similar to that used in other countries with moose populations that includes population trend analyses, productivity and mortality data, and a regional management approach. such changes will improve the professional management of moose and other game species in russia. alces vol. 45: 43-47 (2009) key words: alces alces, harvest, management strategy, moose, poaching, population estimate, russia. in russia the population size of most game species including moose (alces alces l., 1758) is estimated from winter track counts along established census routes. these annual counts are conducted in the latter half of winter after the hunting season, and include approximately 45,000 routes, each about 10 km long. the necessity and continuation of this large-scale annual effort and traditional approach stem from the desire to ensure an adequate harvest quota; however, this approach and related data sets have not recognized annual fluctuations in the moose population that are critically important when setting harvest quotas (glushkov 1995). hunter surveys (about 10,000 questionnaires) conducted throughout russia failed to reveal abnormal causes or rates of mortality that could cause population fluctuations in the moose population (glushkov et al. 1989). the relative estimates of moose populations received from hunters at the start of winter under the program of “the harvest service of vniioz” also failed to show any annual fluctuations in the population (fig. 1). an analysis performed with a large sample size of biological data (2045 jaws from harvested moose, 555 female reproductive tracks, observation of 1360 family groups) collected in forests in the south taiga of the european part of russia (kirov region) showed no dynamic changes in birth rate and natural mortality (glushkov 1999). a decline in fecundity of sub-adult females was noted only when population density in a local area approached its maximum value (3.1-3.4 moose/1000 ha forest; 1981-1990), or when a decline in the proportion of females/litter (statistically significant only for females in the 4th age class) reduced the rate of population growth from 0.041 to 0.000. however, this decline was not only the result of self-regulation, but also of poaching that doubled from 1 to 2.1 moose per poaching incident (versus 1 moose/license). however, this population decline that started in 1987 is not reflected in the census data or hunter survey results. a comparison of the data from the population estimates of the census and the hunter surveys at the beginning of winter revealed that both provided similar conclusions about the moose population growth rate; that is, stable moose population management in russia glushkov alces vol. 45, 2009 44 and/or slow one-way growth (growth rate was 0.026). according to odum (1986), this type of growth rate is sigma-shaped (logistic) and is regulated directly by factors that are population density dependent. such growth is described simply by the logistic equation: dn/dt = rn (k – n)/k as the population reaches the upper asymptote k, growth rate (dn/dt) decreases and approaches zero. harvest and lower birth rate further reduce the growth rate and prevent the density of exploited moose populations from reaching their maximum level; this would be illustrated by a low angle or minimal slope of the population growth rate curve. the population estimates estimated from the winter census routes (fig. 1) show some irregularity as the population increased slowly (average growth rate was 0.048). this growth curve was similar to those depicting population changes caused by natural conditions for various species of birds (williamson 1975). migration of moose in the kirov region was reduced during years with late snow cover; presumably there is a relationship between fluctuating population estimates from the winter census data and the occurrence and intensity of migration. a number of aerial surveys were carried out in early and late winter during a 5-year period (1981-1985) that confirmed this assumption. further, it was also evident that the early winter aerial surveys could not identify small population growth rates (0.041 or 4.1 % per year) determined afterward from demographic tables; the aerial censuses indicated stable populations. data from moose populations in finland and canada confirmed 2 fundamentally important features concerning the type of population growth of a given species (i.e., stability and one-way direction; glushkov 2001), but those populations were characterized by higher growth rates and more measurable response to harvest regulations than those in russia. this occurs for two primary reasons: 1) the use of selective harvesting that produces a highly productive population, and 2) the absence of poaching that allows effective use of selec0 5 10 15 20 25 30 35 19 70 19 72 19 74 19 76 19 78 19 80 19 82 19 84 19 86 19 88 19 90 y e ar m o o se p o p u la ti o n (1 00 0s ) 0 0. 5 1 1. 5 2 2. 5 3 3. 5 4 re la ti ve p o p u la ti o n e st im at e l ate w inter es timate early w inter es timate fig. 1. moose population estimates in the kirov region of russia during early winter (relative scale) and from track counts in late winter (1000s of moose), 1970-1990. alces vol. 45, 2009 glushkov moose population management in russia 45 tive harvest quotas to manage populations effectively. analysis of our moose population growth rates and population responses of other hunted species indicated that all species with logistic growth rates (density dependent) are regulated by common internal (population) mechanisms. these include: 1) slow, difficult-to-measure population growth rates, 2) anthropogenic factors that cause measurable population decline (i.e., hunting mortality), 3) extended periods of population growth and decline, and 4) slow recovery when special conservation and bio-technical management techniques are required to restore the population. for conventional purposes, i described such species as “controlled” (glushkov 2008) in contrast to species with fluctuating (trigger) growth rates. it is clear that the continued existence of “controlled” species depends substantially on the intensity of hunting, and implementing conservation and biotechnical measures. effective harvest regulations provide the most reliable tool to control and reduce mortality to conserve hunted populations with logistic growth rates. the harvest quota for moose is set by comparing the difference between birth and natural mortality rates. however, annual juvenile mortality is influenced by variable environmental conditions that affect food resources, weather, and predation. because high calf mortality occurs in the first 3 months, harvest rates must account for calf survival not the actual birth rate (i.e., that is the growth rate at start of winter). this approach will provide the best estimate of the number of animals available for harvest. because natural winter mortality is much lower than calf mortality, it is often ignored when setting harvest quotas. however, in russia the rate of non-selective harvest has resulted in a negligible growth rate because it represents the difference between the population growth rate prior to the hunting season and winter mortality due to poaching. for example, the average population growth rate in the kirov region is 0.190 at the beginning of winter. if this rate were effectively reduced to compensate for winter predation (0.02), natural mortality from diseases, wounds, other unknown causes (0.015-0.020), poaching (0.08), and a population reserve for increased reproduction (0.02), the non-selective harvest should not exceed 5.5% (0.190 – 0.135). however, this broad calculation does not account for partial replacement of certain mortality factors; the overall mortality rate could be lower than the sum of the rates of individual mortality factors (glushkov 2002). therefore, the integrity of the harvest quota is principally dependent upon the accurate estimate of the autumn population and calf survival prior to the hunting season; the relative importance of winter mortality due to poaching and natural factors is magnified by errors in this estimate. ineffective and harmful harvest quotas in russia occur because of inaccurate population estimates, erroneous documentation about migration, lack of regional population growth rates and related mortality (e.g., poaching) data, and the temporal nature and population response to these factors. the introduction of a selective harvest system could increase both the birth and population growth rates of moose and help nullify their current, stagnant growth rate. however, its implementation and resultant change in harvest quotas will be difficult in the current system, and will require adaptive economics, harvest, and scientific management of moose in russia. this task may be simplified somewhat by following the example of foreign game biologists. rather than depend entirely upon absolute, quantitative population estimates from annual data, they typically analyze trends in annual population data to assess and set harvest quotas. for example, harvest quotas are set relative to the previous year’s quota by assessing special indices of population density (responses to the current level of harvest) on moose population management in russia glushkov alces vol. 45, 2009 46 a regional basis. proposed changes in harvest quotas could be delayed 2 years to reassess the status of the current quota and to reduce the potential negative impact of a changed harvest quota on the structure and productivity of the population. the ability of a selective harvest management system to increase population growth rates and harvest has been documented many times. for example, i used examples from scandinavian countries to illustrate moose harvest rates of 35%, or about 5x that in russia. management problems in these countries are also quite different; the scandinavians typically manage their moose population to maximize harvest, avoid overpopulation, and prevent agricultural and forestry damage. in russia, we strive to reduce poaching and hope to restore our moose population to a level sustainable with the natural productivity of the landscape. conclusions 1) the current method of setting moose harvest quotas is principally flawed because of error in estimating regional populations and mortality rates, and the regional and temporal variation of these parameters. harvest quotas based on erroneous and incomplete data reduce the efficiency and economics of moose management, and for both practical and scientific purposes, an improved method is needed to set moose harvest quotas. 2) an improved strategy in setting regional harvest quotas would mimic common approaches in other countries that include an analysis of the population response to the previous year’s harvest quota. if this system was introduced in russia, federal managers should focus on strategic elements including the overall harvest quota and structure, and implementing management changes; tactical elements such as regional/local harvest quotas should be determined by regional management branches. 3) population assessment of moose and other game species at the onset of winter should be done with annual population trend/index data. each administrative district should have one game biologist responsible for such analyses; such an approach will reduce laborious fieldwork and overall costs substantially. 4) biological assessment of population dynamics will need to improve. moose populations need to be managed regionally in order to address variable population growth rates and environmental conditions. standardized methods are needed to index/census populations of game species in order to produce reliable population density estimates. administrative protocols need to be adopted to guide population monitoring efforts. 5) the current system employed to set the moose harvest quota in russia has many weak components including lack of specific population dynamic information and laborious annual fieldwork to estimate population density. i propose a more simplified procedure of calculating harvest quotas for moose by using better estimates of population density, calf survival, mortality factors including the rate of poaching, and establishing population trend analyzes. these changes will make management of moose and other game species more professional and accurate, and provide an improved practical approach in conservation efforts with these valuable species. references glushkov, v. m. 1995. method of winter route census as a factor of irrational use of resources of wild ungulates. pages 5556 in hunting science and nature use. kirov, russia. (in russian). _____. 1999. moose. pages 117-163 in management of game animal populations. collected scientific papers of vniioz. kirov, russia. (in russian). _____. 2001. moose: ecology and management of populations. kirov, russia. (in russian). _____. 2002. ecological bases of population alces vol. 45, 2009 glushkov moose population management in russia 47 management. pages 115-119 in problems in recent hunting science. proceedings of scientific and practical conference, december 5-6, 2002. centrokhotcontrol, moscow, russia. (in russian). _____. 2008. okhota i okhotnichie khozyastvo (is it a rate or a quota)? 12: 1-2. (in russian). _____, v. n. piminov, and b. p. ponomaryev. 1989. winter mortality and reserves of wild ungulate harvesting. pages 81-92 in management of populations of wild ungulates. collected scientific papers of vniioz. kirov, russia. (in russian). odum, e. 1986. ecology, vol. 2. moscow, russia. (translated from english). williamson, m. 1975. the analysis of biological populations. mir, moscow, russia. (translated from 1972 english version). alces19_238.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces19_273distmoosebio.pdf alces vol. 19, 1983 alces22_303.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces18_25.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces21_493.pdf alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces vol. 21, 1985 alces19_36.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces18_116.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 in memoriam warren b. ballard jr. warren b. ballard, jr.—beloved husband of heather a. whitlaw and widely published author, editor, and nationally recognized professor at texas tech university's department of natural resources management—passed away peacefully at his lubbock, texas home on january 12, 2012. during warren's long career he produced more than 200 peer-reviewed journal articles and raised over $5.9 million in grant, contract, and research funding, which supported more than 60 graduate students. “his legacy lives on in the students, faculty, and research projects he touched,” said michael galyean, dean of texas tech's college of agricultural sciences and natural resources. warren was born on april 28, 1947 in boston, massachusetts. the family moved to albuquerque, new mexico in the early 1950s, where warren attended st. pius x high school. he earned a bsc. in fish and wildlife management from new mexico state university, his msc in environmental biology from kansas state university, and phd in wildlife science from the university of arizona. after completing his msc in 1971, warren worked for the u.s. fish and wildlife service and then from 1973–1990 as a wildlife biologist and research scientist with the alaska department of fish and game. his groundbreaking research on predator-prey relationships, wolf ecology, and ungulate populations is still widely recognized and considered foundational research. he then worked as a consultant with the national park service and lgl alaska (an ecological research company) while getting his phd. he soon became the director of the new brunswick cooperative fish and wildlife unit (1993–1996) and on june 7, 1995 married the love of his life, fellow wildlife biologist heather whitlaw. warren served as research supervisor with the arizona game and fish department (1996– 1998) before joining the texas tech faculty in 1998 where he achieved the rank of full professor in 2003, was named the bricker chair in wildlife management in 2006, and horn professor in 2008. “horn professors … represent the very best of our faculty,” said texas tech president guy bailey. i a member of more than seven professional societies, warren ballard served as editor of four international journals including the wildlife society bulletin and alces from 1998–2001. he was an associate editor of alces for many years as well as wildlife monographs and the wildlife society bulletin. the quality of ballard's research has been recognized with 22 awards. in 1989, he was honored by his peers with the distinguished moose biologist award. in 2002, he earned the chancellor's council distinguished research award at texas tech as well as a special service recognition award from the wildlife society. ballard was recognized as the outstanding researcher in the college of agricultural sciences and natural resources at texas tech—four times. he became a tws fellow in 2005, and has won various society awards for best professional article, monograph, and publications. warren's legacy in the moose world is certain. his two chapters on population dynamics and predator-prey relationships (coauthored with victor van ballenberghe) in “ecology and management of north american moose” stand as seminal summary works. the moose world is indebted to warren ballard and the species is better off for his being—he will be missed. ii alces22_361.pdf alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces vol. 22, 1986 alces27_100.pdf alces vol. 45, 2009 härkönen et al. – deer ked dermatitis in finland 73 deer ked (lipoptena cervi) dermatitis in humans – an increasing nuisance in finland sauli härkönen1,4, maria laine2, martine vornanen2, and timo reunala2,3 1finnish forest research institute, joensuu research unit, p.o. box 68, fi-80101 joensuu, finland; 2university hospital, p.o. box 2000, fi-33521 tampere, finland; 3university of tampere, p.o. box 2000, fi-33521 tampere, finland abstract: the deer ked (lipoptena cervi) is a haematophagous ectoparasite of moose (alces alces) and other cervids that commonly bites humans in finland. since the 1970s there has been an increasing number of finns who suffer from long-lasting and recurrent dermatitis associated with deer ked bites. forestry workers, hunters, berry and mushroom pickers, and other people who work in or visit forests during late summer and early autumn are especially vulnerable to incidental deer ked infestation and dermatitis. interestingly, negative effects of deer keds on human activities have not been recently reported in countries other than finland. our work indicates that dermatitis caused by deer keds consists of a few to 20-50 red papules which occur mostly on the scalp, neck, and upper back. the papules usually appear 6-24 h after the bites and size varies from a few mm to 1-2 cm. they can persist several weeks and in some people up to 1 year. the rapid range expansion of the deer ked in 1970-1990s seems related to the concurrent increase in moose population density in finland. it is possible that range expansion of the deer ked will be promoted by high densities of semi-domesticated reindeer (rangifer tarandus tarandus) in northern finland. as a result, we predict an increase in the distribution of deer keds and the number of people with deer ked dermatitis requiring medical treatment in finland. alces vol. 45: 73-79 (2009) key words: alces alces, deer ked, deer ked dermatitis, dermatitis, lipoptena cervi, moose. the moose (alces alces) is the most important game species in fennoscandia (lavsund et al. 2003), and its high numbers have been appreciated by recreational hunters. however, the increasing density of moose has also caused frustration among forest-owners and other stakeholders because of the negative impact on forestry and traffic safety (aarnio and härkönen 2007). in addition, there has been growing concern about the impact of high densities of moose on the occurrence of dermatitis caused by the deer ked (lipoptena cervi) in finland (reunala et al. 2008). the deer ked is a haematophagous ectoparasite of moose and other cervids and was first documented in the southeastern region of finland in 1960 (hackman et al. 1983). at present, it is common in southern and central finland and its range is gradually expanding northward. in 2007 the first sightings were made close to 66° n within the southern part of the reindeer (rangifer tarandus tarandus) herding area (kaunisto et al. 2009). the distribution of l. cervi also includes central europe, southeastern norway, southern sweden, some parts of siberia, northern china, and algeria in northern africa (maa 1969); it is an introduced species in the northeastern united states (bequaert 1942). moose are the main host of the deer ked in finland (hackman et al. 1983), but it also parasitizes wild forest reindeer (r. t. fennicus), 4present address: hunters central organization, fantsintie 13-14, fi-00890 helsinki, finland deer ked dermatitis in finland – härkönen et al. alces vol. 45, 2009 74 semi-domesticated reindeer, and white-tailed deer (odocoileus virginianus) (kaunisto et al. 2009). in central europe deer keds also use red deer (cervus elaphus), roe deer (capreolus capreolus), and fallow deer (dama dama) as hosts (haarløv 1964). a related species, l. mazamae, parasitizes white-tailed deer in north america and brocket deer (mazama spp.) in central and south america (bequaert 1942, 1957). females produce one third-instar larva at a time and the larva immediately pupates (hackman et al. 1983). ked pupae drop off their hosts onto the snow or ground, for example, at bedding sites and trails of hosts. the number of pupae produced in the life of a female l. cervi is unknown. in finland, the winged adults emerge from pupae in late julyoctober and seek a new host by flying short distances. upon finding a suitable host, adult keds shed their wings and commence to suck blood recurrently (haarløv 1964). in addition, they also bite humans and other hosts but do not reproduce on them (ivanov 1975, reunala et al. 1980, 2008). surprisingly, a ked bite is barely noticeable to humans as it pierces the skin only about 1 mm. the blood meal taken from humans is typically small (hermosilla et al. 2006). forestry workers, hunters, berry and mushroom pickers, and others who work in or visit forests where moose occur are especially vulnerable to incidental deer ked infestation (reunala et al. 2008). deer ked attacks are mainly annoying and an inconvenience in having to remove dozens of keds from hair and clothes. however, a recent finnish case study (liukkonen et al. 2007) suggests that infestation of deer keds could reduce the recreational value of the hunting experience, especially in western finland. in addition, 55% of the 1,400 citizens replying to a nationwide questionnaire concerning attitudes about moose management in finland identified deer keds as an important or very important reason for reducing the moose population (petäjistö et al. 2005). their occurrence was the third most important reason for controlling moose numbers after road accidents and forest damage. to our knowledge, there are no recent reports from other countries relative to negative effects on human activities (ivanov 1975, alekseev 1985). since the 1970s there have been an increasing number of people in finland who, following deer ked bites, suffer from longlasting and recurrent dermatitis (rantanen et al. 1982, reunala et al. 1980, 2008). in an extreme case, an occupational ige-mediated allergic condition with symptoms in the nose and eyes resulted (laukkanen et al. 2005). in this paper we provide a brief overview of typical symptoms of dermatitis associated with deer ked bites, and evaluate the future potential of this condition relative to range expansion by deer keds and population densities of potential host species in finland. methods patients with dermatitis caused by deer keds were diagnosed and studied at the dermatological out-patient clinics at tampere and helsinki university hospitals. the examinations yielded skin biopsies from the bite lesions and skin and blood tests for ige antibody mediated allergies. the distributions and population estimates for moose, whitetailed deer, roe deer, wild forest reindeer, semi-domesticated reindeer, and fallow deer were evaluated using relevant literature. in addition, harvest statistics of huntable species were collected from the database of the hunters central organization. we assumed that the annual number of harvested animals can be used as an index of population trends (mysterud et al. 2002, grøtan et al. 2005). results and discussion our patients usually developed symptoms a few years after their first contact with l. cervi. however, sensitization to the bites with accompanied symptoms is highly variable and alces vol. 45, 2009 härkönen et al. – deer ked dermatitis in finland 75 can appear in the first season of bites or up to 30 years afterward (rantanen et al. 1982). deer ked dermatitis consists of a few to 2050 red papules occurring mostly on the scalp, neck, and upper back. they usually appear 6-24 h after the bites and their size varies from a few millimeters to 1-2 cm. papules are accompanied by intense itching since they are easily scratched and often become secondarily infected by staphylococcal bacteria (rantanen et al. 1982). according to hackman et al. (1983), some victims show a local wheal and flare reaction within a few minutes of being bitten. often the papules can persist several weeks and in some people for a year (reunala et al. 2008). the histological finding of a papule associated with a recent bite by l. cervi is a marked dermal accumulation of lymphocytes and eosinophils. older lesions can resemble a malignant tumor such as skin lymphoma (reunala et al. 2008). immunohistological findings reveal complement deposits in blood vessel walls suggesting that, in addition to cell mediated mechanisms, complement activation also seems to be involved in the pathogenesis of bite lesions (rantanen et al. 1982). skin tests were performed with wholebody extract of deer keds to confirm the allergic sensitization. the patients with deer ked dermatitis showed positive immediate and delayed skin test reactions in contrast to non-reactive control subjects (rantanen et al. 1982). the involvement of immunoglobulin e (ige) in the pathogenesis of reactions to bites of l. cervi was confirmed by laukkanen et al. (2005). they also identified the igebinding allergenic protein by immunoblotting. this was not further characterized but is obviously a saliva protein, possibly similar to those described earlier from mosquito saliva which frequently sensitizes people (brummerkorvenkontio et al. 1997). after deer keds were present for about 20 years in eastern finland, one third of regularly exposed forest workers became sensitized to the bites (hackman et al. 1983). new bites often cause flare-ups of dermatitis in sensitized victims in the same or subsequent years, and this allergic sensitization persists for years (rantanen et al. 1982, reunala et al. 2008). due to the rapid spread of the deer ked in finland, increasing numbers of finns are now exposed annually and many of these subjects will become sensitized in the future (see also reunala et al. 2008). unfortunately there is no official register for the number of patients having deer ked dermatitis either at pirkanmaa hospital district or any other hospital districts in finland. however, it is well-known that patients with deer ked dermatitis occur throughout finland (reunala et al. 2008). the exact number of finns presenting dermatitis of deer ked origin is not available at present, but is estimated to be several thousands. unfortunately, there are no effective repellents against deer keds (ivanov 1975, alekseev 1985) and the available medical treatment for deer ked dermatitis only provides symptomatic relief. treatments include antihistamine tablets taken orally to relieve pruritus and corticosteroid creams that are applied to the papules (karppinen et al. 2002). it has been suggested that l. cervi might transmit harmful infectious agents such as bartonella spp. (dehio et al. 2004, halos et al. 2004, reeves et al. 2006) which are intracellular, small, gram-negative bacteria transmitted by blood-sucking arthropods; they are considered as emerging pathogens in humans and animals (chang et al. 2001). however, the potential risk that l. cervi could serve as a vector for the transmission of micro-organisms from one host to another is relatively low since the deer ked sheds its wings after finding a potential host, thus making it difficult to change hosts (reunala et al. 2008). nevertheless, it is reported that deer keds may transfer directly from another white-tailed deer to newborn fawns (samuel and trainer 1972). the density of moose has increased in deer ked dermatitis in finland – härkönen et al. alces vol. 45, 2009 76 harvest 0 10000 20000 30000 40000 50000 60000 70000 80000 90000 19 33 19 37 19 41 19 45 19 49 19 53 19 57 19 61 19 65 19 69 19 73 19 77 19 81 19 85 19 89 19 93 19 97 20 01 20 05 fe lle d moose white-tailed deer harvest 0 50 100 150 200 250 19 75 19 77 19 79 19 81 19 83 19 85 19 87 19 89 19 91 19 93 19 95 19 97 19 99 20 01 20 03 20 05 20 07 fe lle d wild forest reindeer fallow deer harvest 0 500 1000 1500 2000 2500 3000 3500 4000 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 20 04 20 05 20 06 20 07 fe lle d roe deer fig. 1. harvest of moose in 1933-2007, white-tailed deer in 1958-2007, wild forest reindeer 1996-2007, fallow deer in 1975-2007, and roe deer in 1991-2007 in finland. cervid species hunting is licensebased and hunters must report felled animals no later than after the end of the hunting season. roe deer hunting was released 2005 and after that time the harvest reports have been on voluntarily basis (i.e., harvest is assumed to be underestimated after 2005). harvest statistics were collected from the database of the hunters central organization. alces vol. 45, 2009 härkönen et al. – deer ked dermatitis in finland 77 finland since the 1970s (torvelainen 2007). the post-harvest moose population was at its highest in 2001 when it was estimated to be 139,000 (4.6 moose/10 km2 land area); moose harvest data indicate a similar trend (fig. 1). the white-tailed deer population is more dense in southwestern finland, whereas the roe deer population is more widely, but sparsely distributed throughout southern and central finland (tiainen and rintala 2008). the post-harvest population of white-tailed deer was estimated as 30,000 in the winter of 2007-2008 and has been increasing steadily (fig. 1); white-tailed deer are also expanding their distribution. the population of roe deer (15,000 in winter 2004-2005) and other cervids are relatively stable (hunters central organization, unpubl. data). the population of wild forest reindeer was approximately 2,000 in 2007 and is distributed in two subpopulations, one in central finland and the other in eastern finland (bisi and härkönen 2007). fallow deer are few in number (~500) and have a restricted distribution in southern finland and are only marginal hosts for deer keds; relatively few wild forest reindeer and fallow deer are harvested (fig. 1). reindeer husbandry occurs north of 66° n in finland, with approximately 200,000 animals in winter herds (torvelainen 2007). we conclude that moose are currently the most important host species for l. cervi as moose populations are relatively dense throughout finland (see pusenius et al. 2008). balashov (1996) reported that fluctuations in abundances of deer ked in northwestern russia were connected to the densities of local moose populations. similarly, we suggest that the rapid range expansion of l. cervi in the 1970-1990s in finland (see reunala et al. 2008) coincided well with an increase of the moose population. to our knowledge, this hypothesis of occurrence of l. cervi in relation to the spatial and temporal moose densities in finland has not been tested scientifically. it is also possible that the range expansion of l. cervi, especially in northern finland, will be promoted by high densities of semi-domesticated reindeer. simultaneously, prevailing indications of climate change may also promote the range expansion and life history of l. cervi (l. härkönen et al., university of oulu, unpubl. data). in light of the above, we suggest that the number of people with dermatitis caused by l. cervi and requiring medical treatment will continue to increase in finland. acknowledgements we thank hunters central organization for technical help during harvest statistics collection. we are also grateful to ed addison and to 2 anonymous referees for their valuable comments on the manuscript. references aarnio, j., and s. härkönen. 2007. hirvestä hyötyjä ja kustannuksia (pros and cons of moose). metsätieteen aikakauskirja 2: 101-106. (in finnish). alekseev, e. a. 1985. initial experience with individual human protection from attack by the deer louse fly lipoptena cervi. medicinskaa parazitologia (mosk.) 6: 56-57. balashov, y. s. 1996. the fluctuations of abundance of the deer louse-fly lipoptena cervi (hippoboscidae) in forests of the north-west russia. parazitologiia 30: 182-184. bequaert, j. c. 1942. a monograph for the melophaginae, or ked flies, of sheep, goats, deer and antelopes (diptera, hippoboscidae). entomologist americana 22: 1-220. _____. 1957. the hippoboscidae or louse flies (diptera) of mammals and birds. ii. taxonomy, evolution and revision of american genera and species. entomologist americana 36: 417-611. bisi, j., and s. härkönen. 2007. status of the wild forest reindeer population. in: deer ked dermatitis in finland – härkönen et al. alces vol. 45, 2009 78 management plan for the wild forest reindeer population in finland. publications of ministry of agriculture and forestry 9b: 21-26. brummer-korvenkontio, h., t. palosuo, g. françois, and t. reunala. 1997. characterization of aedes communis, aedes aegypti and anopheles stephensi mosquito saliva antigens by immunoblotting. international archives of allergy and immunology 112: 169-174. chang, c. c., b. b. chomel, r. w. kasten, v. romano, and n. tietze. 2001. molecular evidence of bartonella spp. in questing adult ixodes pacificus ticks in california. journal of clinical microbiology 39: 1221-1226. dehio, c., u. sauder, and r. hiestand. 2004. isolation of bartonella schoenbuchensis from lipoptena cervi, a blood-sucking arthropod causing deer ked dermatitis. journal of clinical microbiology 42: 5320-5323. grøtan, v., b.-e. saether, s. engen, e. j. solberg, j. d. c. linnell, r. andersen, h. brøseth, and e. lund. 2005. climate causes large-scale spatial synchrony in population fluctuations of a temperate herbivore. ecology 86: 1472-1482. haarløv, n. 1964. life cycle and distribution pattern of lipoptena cervi (l.) (dipt., hippobosc.) on danish deer. oikos 15: 93-129. hackman, w., t. rantanen, and p. vuojolahti. 1983. immigration of lipoptena cervi (diptera, hippoboscidae) in finland, with notes on its biology and medical significance. notulae entomologicae 63: 53-59. halos, l., t. jamal, r. maillard, b. girard, j. guillot, b. chomel, m. vayssiertaussat, and h.-j. boulouis. 2004. role of hippoboscidae flies as potential vectors of bartonella spp. infecting wild and domestic ruminants. applied and environmental microbiology 70: 6302-6305. hermosilla, c., n. pantchev, r. bachmann, and c. bauer. 2006. lipoptena cervi (deer ked) in two naturally infested dogs. the veterinary record 159: 286-287. ivanov, v. i. 1975. anthropophilia of deer blood sucker lipoptena cervi l. (diptera, hibboboscidae). medicinskaa parazitologia (mosk.) 44: 491-495. kaunisto, s., r. kortet, l. härkönen, s. härkönen, h. ylönen, and s. laaksonen. 2009. new bedding site examina-2009. new bedding site examination-based method to analyse deer ked (lipoptena cervi) infection in cervids. parasitology research 104: 919-925. karppinen, a., h. kautiainen, l. petman, p. burri, and t. reunala. 2002. comparison of cetirizine, ebastine and loratadine in the treatment of immediate mosquito-bite allergy. allergy 57: 534-537. laukkanen, a., p. ruoppi, and s. mäkinenkiljunen. 2005. deer ked-induced occupational allergic rhinoconjunctivitis. annals of allergy, asthma and immunology 94: 604-608. lavsund, s., t. nygrén, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. liukkonen, t., j. bisi, h. hallila, and o. joensuu. 2007. feedback from hunters on state-owned lands. metsähallituksen luonnonsuojelujulkaisuja. sarja b 84. (in finnish with english summary). maa, t. c. 1969. studies in hippoboscidae (diptera) part 2. pacific insects monograph 20: 1-312. mysterud, a., r. langvatn, n. g. yoccoz, and n. c. stenseth. 2002. large-scale habitat variability, delayed density effects and red deer populations in norway. journal of animal ecology 71: 569-580. petäjistö, l., j. aarnio, p. horne, t. koskela, and a. selby. 2005. kansalaismielipide hirvikannasta ja sen säätelystä (public attitudes toward the size and regulation of the moose population). metsäntutalces vol. 45, 2009 härkönen et al. – deer ked dermatitis in finland 79 kimuslaitoksen tiedonantoja 945. (in finnish). pusenius, j., m. pesonen, r. tykkyläinen, m. wallén, and a. huittinen. 2008. hirvikannan koko ja vasatuotto 2006 (moose population size and calf production in 2006). pages 7-14 in m. wikman, editor. riistakannat 2007: riistaseurantojen tulokset (monitoring game abundance in finland, 2007). selvityksiä 5/2008. (in finnish). rantanen, t., t. reunala, p. vuojolahti, and w. hackman. 1982. persistent pruritic papules from deer ked bites. acta dermato-venereologica 62: 307-311. reeves, w. k., m. p. nelder, k. d. cobb, and g. a. dasch. 2006. bartonella spp. in deer keds, lipoptena mazamae (diptera: hippoboscidae), from georgia and south carolina, usa. journal of wildlife diseases 42: 391-396. reunala, t., m. laine, m. vornanen, and s. härkönen. 2008. hirvikärpäsihottuma – maanlaajuinen riesa (deer ked dermatitis – a country wide nuisance). duodecim 124: 1607-1613. (in finnish). _____, t. rantanen, p. vuojolahti, and w. hackman. 1980. deer ked (lipoptena cervi l.) causes chronic dermatitis in man. duodecim 96: 897-902. (in finnish with english summary). samuel, w. m., and d. o. trainer. 1972. lipoptena mazamae rondani, 1878 (diptera: hippoboscidae) on white-tailed deer in southern texas. journal of medical entomology 9: 104-106. tiainen, j., and j. rintala. 2008. kulttuuriympäristön riista talvella 2007 (game abundance in agricultural areas in winter 2007). pages 22-25 in m. wikman, editor. riistakannat 2007: riistaseurantojen tulokset (monitoring game abundance in finland, 2007). selvityksiä 5/2008. (in finnish). torvelainen, j. 2007. multiple-use forestry. pages 201-218 in a. peltola, editor. finnish statistical yearbook of forestry 2007. agriculture, forestry and fishery 2007. finnish forest research institute. alces18_301.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces18_veditorialcommitte.pdf alces vol. 18, 1982 alces17_viattendancelist.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_126.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_1.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_64.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces18_xiiworkshopaerialsurveytech.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces16_35.pdf alces vol. 16, 1980 variation in fine-scale movements of moose in the upper koyukuk river drainage, northcentral alaska kyle joly1, timothy craig2,4, mathew s. sorum1, jennifer s. mcmillan3, and michael a. spindler2 1national park service, gates of the arctic national park and preserve, 4175 geist road, fairbanks, alaska 99709; 2us fish and wildlife service, kanuti national wildlife refuge, 101 12th avenue, fairbanks, alaska 99701; 3bureau of land management, central yukon field office, 1150 university avenue, fairbanks, alaska 99709; 4retired abstract: fine-scale movements form the foundation of local habitat selection by animals. in northern interior alaska, the dalton highway corridor management area and other parts of game management unit 24 are accessible to moose hunters from the dalton highway. concern that these areas may be a population sink for moose (alces alces) inhabiting the gates of the arctic national park and preserve and the kanuti national wildlife refuge prompted this study of movements. we found that migratory bulls and cows traveled about the same distance over the course of a year as non-migratory moose. although counterintuitive, this may reflect the selective foraging behavior of a low density (∼0.1 moose/km2) moose population in habitat with abundant forage. maximum movement rates by bulls occurred at the onset of rut at the end of the hunting season. this spike in movement may have given local residents the impression that local moose were migratory and vulnerable to hunting from non-residents. movement rates were lowest in winter for both bulls and cows, and declined with increasing winter severity, but not temperature specifically. reduced movement rates by cows during the calving season were not readily evident and annual fidelity to calving sites was minimal. alces vol. 51: 97–105 (2015) key words: alces alces, moose, movement, seasonality, strategy, winter severity. although moose (alces alces) migrate long distances in northern interior alaska (mauer 1998), prior to this study little was known about migratory patterns in the upper koyukuk river drainage where land and moose management is complicated by a mosaic of lands administered by the state of alaska, national park service, us fish and wildlife service, bureau of land management, and private entities (fig. 1). public wildlife advisory groups were concerned that moose harvested within and around the dalton highway corridor management area (dhcma) were influencing (lowering) moose density in the gates of the arctic national park and preserve (gaar) and kanuti national wildlife refuge (knwr). however, migration of moose between these conservation units and the dhcma is limited and likely not a management concern (joly et al. 2015). analysis of fine-scale movements by moose is useful to better understand their local ecology and behavior that is integral for implementing informed management strategies. just as moose exhibit variation in large-scale migratory movements (mauer 1998, joly et al. 2015), fine-scale movement patterns vary individually, by gender, and in response to local physiographic variables. movement in other boreal regions occurs relative to an individual's need for cover, forage, and reproduction (leblond et al. 2010). improved comprehension of factors 97 influencing fine-scale movements in a heterogeneous landscape is critical to understanding moose behavior and their distribution across the landscape. in this study, we examined fine-scale movement data collected as part of a larger project that assessed moose movements between gaar and knwr, and the dhcma. our goals were to better understand the fine-scale movements of moose in the upper koyukuk river drainage relative to migratory status, sex, season, physiography, temperature, and winter severity. we hypothesized that migratory moose traveling to and from wintering areas would travel farther annually than non-migratory moose, and that harsher winter conditions would reduce movement rate. methods study area the study was in the upper koyukuk river drainage that encompasses the southern slopes of the central brooks range, including the southeastern portion of gaar, all of knwr, and other state, federal (including portions of the dhcma), and native lands (fig. 1). moose density in the upper koyukuk is very low (∼0.1 moose/km2; lawler et al. 2006), and the physiography fig. 1. the upper koyukuk river study area (white polygon) in northcentral alaska which encompassed moose locations derived from gps telemetry data from 2008–2013. 98 upper koyukuk moose movements – joly et al. alces vol. 51, 2015 and habitat types are diverse. in the north, rugged mountains (up to 2000 m in elevation) divided by narrow river valleys dominate the landscape. habitats range from alpine tundra to shrubs, boreal forest, and muskegs with declining elevation. alders (alnus spp.), willows (salix spp.), and dwarf birch (betula glandulosa) dominate shrub habitats. black spruce (picea mariana) is the most prevalent tree species, with white spruce (picea glauca), poplar (populus balsmifera), and numerous shrub species common in riparian areas. birch stands (betula papyrifera) occur on south-facing slopes at moderate elevations and are common in recently burned areas. the landscape becomes progressively flatter (elevations typically <500 m) to the south with more muskegs, streams, and lakes interspersed within boreal forest and broad riparian zones. the regional climate is strongly continental, with long, extremely cold (<−45° c) winters, and brief hot (>30° c) summers. snow depth exceeds 90 cm many winters, with >60 cm in most. moose relocation data and gis analyses adult moose were darted using a mixture of carfentanil citrate and xylazine from robison r-44 helicopters, and instrumented with a gps radio-collar. moose captured north and east of bettles, alaska (fig. 1) were designated as ‘northern moose’ and those in and around knwr as ‘southern moose’. radio-collars deployed in march 2008 collected 1 gps location/day, and thereafter all collected 3 locations/day. about half of the radio-collars were instrumented with temperature sensors. movements of northern and southern moose were contrasted due to the substantial differences in terrain and habitat. migratory status of individual moose was ascertained from net-squared displacement analyses (bunnefeld et al. 2011) as described by joly et al. (2015). average movement rates and distance traveled were calculated based on successive locations; results are reported as mean rates and distances with their associated standard error (se). differences in mean movement rates by sex, season, and migratory status were not statistically assessed because of the varied annual sample size and differential location rates. we analyzed the data on a weekly basis and across 6 designated seasons: spring, 26 march – 27 may; calving, 28 may – 23 june; summer, 24 june – 26 august; hunting, 27 august – 23 september; fall/ rut, 24 september – 25 november; winter, 26 november – 25 march. we did not determine exact calving sites in this remote landscape due to budgetary restrictions preventing survey flights. instead, we assessed an individual cow’s annual fidelity to a calving area by calculating the distance between locations nearest to the expected calving date (i.e., june 1). for example, for each cow we measured the distance between its location on 1 june 2010 with those on 1 june 2008, 2009, and 2011, assuming 2010 was the most central location. movement was related to terrain ruggedness (sappington et al. 2007) and temperature using regression analysis; p < 0.05 was the critical significance level. likewise, we assessed whether moose exhibited elevational migrations to presumably take advantage of potential temperature inversions on extremely cold days (<−20° c). we compared dem-assigned elevations from each location to the temperature at the nearest weather station in bettles, alaska (, accessed july 2014). winter severity was classified in 3 categories from the total number of days with snow and depth of snow as recorded in bettles: mild (<135 days with ≥30 cm snow and <7 days with ≥60 cm snow), moderate (>170 days with >30 cm snow, >50 days with >60 cm, or <14 days with >90 cm snow), or severe (>170 days with ≥30 cm of snow, alces vol. 51, 2015 joly et al. – upper koyukuk moose movements 99 http://www.ncdc.noaa.gov/ http://www.ncdc.noaa.gov/ >100 days with ≥60 cm, or >30 days with ≥90 cm snow). results a total of 37 adult moose (26 cows, 11 bulls) were captured and instrumented with gps radio-collars. eighteen northern cows were marked: 5 in march 2008, 9 in october–november 2008, 2 in november 2009, and 2 in april 2011. eight southern cows and 11 northern bulls were marked in april 2011. the 37 gps units collected a total of 71,675 locations. cows with 1 location/day had ∼10% lower total annual movement (492.2 ± 71.4 km) than cows with 3 locations/day (544.5 ± 34.6 km); this was expected given the more intensive sampling regime that captures tortuous movements (joly 2005). cows (534.6 ± 30.7 km) and bulls (523.6 ± 43.0 km) traveled similar distances over the course of a year. southern cow moose (616.3 ± 75.9 km) walked ∼17% further than northern cow moose in a year (493.7 ± 21.6 km). non-migratory (n = 5; 528.5 ± 68.0 km) and migratory (n = 9; 525.9 ± 50.7 km) cows had similar annual movement distances. likewise, annual movement distances were similar between non-migratory (n = 4; 543.4 ± 78.8 km) and migratory bulls (n = 4; 528.2 ± 78.8 km). bull and cow movement rates were not substantially different in fall, winter, and spring and both moved the least during winter (fig. 2). movement rates of bulls were ∼45% higher than cows during the calving and hunting seasons; conversely, cows moved more than bulls during summer (fig. 2). movement rates of migratory cows were ∼20% less than non-migratory cows in summer (87 ± 10 versus 107 ± 12 m/h) and ∼25% less in fall (50 ± 9 versus 65 ± 10 m/h). southern cow moose had ∼40–65% higher movement rates in spring, fall, and winter (77 ± 6, 69 ± 8, 60 ± 7 m/h, respectively) than northern cow moose (56 ± 4, 49 ± 5, 36 ± 5 m/h, respectively). weekly movement rates were most pronounced by both sexes during calving, continued through summer by cows, and fig. 2. mean seasonal movement rates (meters per hour) and standard errors (se) of gps radio-collared bull (light bars) and cow (dark bars) moose in the upper koyukuk river drainages, northcentral, alaska, 2008–2013. 100 upper koyukuk moose movements – joly et al. alces vol. 51, 2015 increased again by bulls during the rut (fig. 3). locations of individual cows during calving were, on average, ∼8,400 m horizontal distance from previous calving sites (n = 21 cows with 59 potential calving events, range = 227–43,654 m). only 3 (8%) potential calving sites were <1 km from the next nearest location for all potential calving events. winter movement rates of cows were ∼20% longer in mild (39 ± 2 m/h) versus moderate and severe winters (32 ± 2 and 33 ± 3 m/h, respectively). although gps radio-collars were not deployed on bulls during the only severe winter (2008–2009), their movement rates were ∼10% higher in mild (40 ± 5 m/h) than moderate winters (36 ± 5 m/h). we found no evidence that moose utilized higher elevations for thermal advantage during temperature inversions in extreme cold weather. movement rates were significantly, and negatively associated with fine-scale (180 m) terrain ruggedness for 50% of moose (n = 32); conversely, 2 southern cow moose had significant positive associations. cows used the highest elevations during fall (576 ± 2 m) and the lowest in spring (413 ± 2 m). bulls were also at lowest elevations during spring (334 ± 4 m), but at highest elevations in summer (528 ± 4 m). use of low elevations in spring may reflect earlier green-up. the relationship between movement rate and temperature was strongest in spring when 80% of moose had significant positive associations between movement rates and collar temperature (n = 15); no moose had a negative association. an obvious pattern was not observed in other seasons; >33% of movement rates were negatively associated with temperature (i.e., greater movement rates at colder temperatures) and 13% had positive associations. lastly, we found that moose repeatedly crossed the dalton highway, the trans-alaska pipeline system, and its associated maintenance road. discussion fine-scale movements of moose provide critical behavioral and ecological information and are the foundation of large-scale movements such as migration. migration can reduce predation and the energy required to avoid predation, and hence increase productivity (avgar et al. 2013, white et al. 2014); however, migration can impose costs as well. for example, increased exposure to predation can occur along the course of the migration route or at its terminus (middleton et al. 2013), and longer migrations require higher energy output (fancy and white 0 50 100 150 200 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49 51 cows bulls)ruoh/ m(tne mevo m calving winter winter spring summer hunting fall/rut fig. 3. weekly movement rates (m/h) of gps radio-collared bull (light bars) and cow (dark bars) moose in the upper koyukuk river drainages, northcentral, alaska, 2008–2013. week 1 is 1–7 january, and subsequent seasons are delineated accordingly. alces vol. 51, 2015 joly et al. – upper koyukuk moose movements 101 1987). counterintuitively, we found that non-migratory moose moved similar distances in a year as migratory moose, despite having smaller home ranges and not travelling between summer and winter ranges (joly et al. 2015). although high habitat quality typically results in smaller home ranges and lower movement rates (hundertmark 1997, dussault et al. 2005), it is not clear why migratory moose would not travel farther than non-migratory moose. we offer 3 possible explanations related to foraging, terrain ruggedness, and predation. moose are highly selective browsers (risenhoover 1987, hundertmark 1997) and browse utilization rates in our study area are among the lowest in alaska (paragi et al. 2008). where forage quantity and quality are high relative to moose density, such as in our study area, only a small proportion of forage biomass is consumed (hundertmark 1997). under these conditions, moose may move rapidly between stands of readily available, small diameter forage with high digestibility (vivas and saether 1987). these moose adopt an intake maximization strategy rather than an energy conservation strategy, which would not be surprising during productive seasons. in contrast, dussault et al. (2005) found lower movement rates in high quality habitats in eastern canada. terrain ruggedness may influence moose movement as the majority of northern moose inhabiting much more mountainous, rugged terrain had lower movement rates. in the less rugged southern portion there was no clear association between movement rates and ruggedness. moose may move less in rugged terrain simply because of the increased energetic expense and difficulty to do so. moreover, the actual overland distance travelled by moose may be underestimated for northern moose since movement calculations typically do not account for vertical movement required to navigate rough terrain; vertical movement within elevated terrain results in shorter calculated distances than travel in flat terrain (dettki and ericsson 2008). lastly, higher predation pressure could cause increased movement for our moose with small home ranges (see ballard et al. 1980). the dalton highway provides access and facilitates hunting and trapping within the dhcma, and combined with disturbance from human activities in the area, is generally thought to have reduced the abundance of wolves (canis lupus) and bears (ursus arctos and u. americanus) in the dhcma. unfortunately, only limited information is available about predator populations in this region. nonetheless, in gaar where hunting and trapping regulations are restrictive, cows moved less than cows in knwr where limited, but more hunting and trapping occurs – weakening the predator abundance explanation. bulls and cows traveled similar distances over the course of a year and their movement rates were similar in the fall/rut and winter seasons (fig. 2). movement rates for both bulls and cows were lowest during winter as documented elsewhere (fig. 2 and 3; hundertmark 1997, dussault et al. 2005). winter movement rates of bulls and cows were higher during mild winters and more restricted in harsher winters as deep snow impacts movement, distribution, and home range size (van ballenberghe 1977, miquelle et al. 1992, ball et al. 2001). our limited data suggest that bulls may be less sensitive to winter severity than cows, possibly because their larger size facilitates movement through deep snow and they use larger foraging areas. we did not find that lower winter temperatures reduced movement rates or led to elevational migrations. indeed, we found slightly higher movement rates at extreme low temperatures and that moose mostly used lower elevations during winter. bull and cow movement rates varied during other seasons; for example, bull 102 upper koyukuk moose movements – joly et al. alces vol. 51, 2015 movement rates were substantially higher during the hunting season (fig. 2). this abrupt and marked movement likely related to the concern of public hunting advisory groups that moose might be migrating out of the conservation units (gaar and knwr) into areas where hunting regulations were more liberal. related analyses (joly et al. 2015) indicate that such concerns are mostly unwarranted. bulls and cows also had different movement rates during calving. it is intuitive that cow movement would be less than that of bulls when cows tend calves. our data suggest that cow movement rates do not drop substantially during calving, rather, that bull movement rates increase measurably during early summer. in fact, cow movement rates were higher during calving than earlier in spring; however, we could not parse individual cows by their reproductive status. we found that analyzing weekly movement rates was insightful in that these movements identify temporal patterns that are otherwise masked in seasonal timeframes. for instance, increased movement rates of bulls between mid-september and early october corresponded to the hunting season and onset of rut, but were masked by seasonal averages (fig. 3). further, the weekly analysis revealed a spike in cow movement prior to peak calving, which may be an anti-predatory behavior. cows often make substantial shifts (>5 km) just prior to parturition (bowyer et al. 1999, testa et al. 2000, mcgraw et al. 2014), and successive calving locations were located an average >8 km apart: <10% of potential calving sites were located within 1 km of previous potential calving locations and none within 225 m. these data are consistent with other studies (e.g., chekchak et al. 1998, testa et al. 2000, mcgraw et al. 2014) that found little annual fidelity to calving sites. this study is the first of its kind in this region and was an offshoot of a larger investigation about migratory movements. our analysis of fine-scale movements surprisingly revealed that 1) migratory and nonmigratory moose traveled similar distances in a year, 2) terrain ruggedness was related to movement distance, and 3) that bulls increased movement more than cows in spring. we also found that bulls did not seek higher elevation during extreme cold, bulls more than doubled their movement during the rut, and that cows had little annual fidelity to calving sites. further study of fine-scale movements should contribute meaningfully to improved understanding of behavioral adaptations and micro-habitat selection by moose. acknowledgements this project was promoted by public subsistence and wildlife advisory groups. funding was provided by the national park service, us fish and wildlife service, alaska department of fish and game, and the bureau of land management. we thank pilots t. cambier, m. spindler, m. webb, c. cebulski, and a. greenblatt for making this project possible and safe. j. burch, j. caikoski, t. hollis, j. lawler, n. pamperin, t. paragi, c. roberts, l. saperstein, g. stout, and many others provided critical assistance with project implementation. we thank g. stout for his contributions to project management and insights into the ecology of moose within the region. n. bywater, s. miller, r. sarwas, and a. quist provided database and gis expertise. e. addison and anonymous reviewers provided suggestions to improve a previous version of this manuscript. all moose captures adhered to state of alaska animal care and use committee (acuc) guidelines (#07–11). alces vol. 51, 2015 joly et al. – upper koyukuk moose movements 103 literature cited avgar, t., g. street, and j. m. fryxell. 2013. on the adaptive benefits of mammal migration. canadian journal of zoology 91: 481–490. ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by largeungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7: 39–47. ballard, w. b., c. l. gardner, and s. d. miller. 1980. influence of predators on summer movements of moose in southcentral alaska. proceedings of the north american moose conference and workshop 16: 338–359. bowyer, r. t., v. van ballenberghe, j. g. kie, and j. a. k. maier. 1999. birth-site selection by alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80: 1070–1083. bunnefeld, n., l. borger, b. van moorter, c. m. rolandsen, h. dettki, e. j. solberg, and g. ericsson. 2011. a model-driven approach to quantify migration patterns: individual, regional and yearly differences. journal of animal ecology 80: 466–476. chekchak, t., r. courtois, j.-p. ouellet, l. breton, and s. st.-onge. 1998. characteristics of moose (alces alces) calving sites. canadian journal of zoology 76: 1663–1670. dettki, h., and g. ericsson. 2008. screening radiolocation datasets for movement strategies with time series segmentation. journal of wildlife management 72: 535–542. dussault, c., r. courtois, j.-p. ouellet, and i. girard. 2005. space use of moose in relation to food availability. canadian journal of zoology 83: 1431–1437. fancy, s. g., and r. g. white. 1987. energy expenditures for locomotion by barrenground caribou. canadian journal of zoology 65: 122–128. hundertmark, k. j. 1997. home range, dispersal and migration. pages 303–335 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, dc, usa. joly, k. 2005. the effects of sampling regime on the analysis of movements of overwintering female caribou in eastcentral alaska. rangifer 25: 67–74. ———, t. craig, m. s. sorum, j. s. mcmillan, and m. a. spindler. 2015. moose movement patterns in the upper koyukuk river drainage, northcentral alaska. alces 51: 87–96. lawler, j. p., l. saperstein, t. craig, and g. stout. 2006. aerial moose survey in upper game management unit 24, alaska, fall 2004, including state land, and lands administered by the bureau of land management, gates of the arctic national park and preserve, and kanuti national wildlife refuge. national park service technical report nps/ar/nr/ tr-2006-55, fairbanks, alaska, usa. leblond, m., c. dussault, and j.-p. ouellet. 2010. what drives fine-scale movements of large herbivores? a case study using moose. ecography 33: 1102–1112. mauer, f. j. 1998. moose migration: northeastern alaska to northwestern yukon territory, canada. alces 34: 75–81. mcgraw, a. m., j. terry, and r. moen. 2014. pre-parturition movement patterns and birth site characteristics of moose in northeast minnesota. alces 50: 93–103. middleton, a. d., m. j. kauffman, d. e. mcwhirter, j. g. cook, r. c. cook, a. a. nelson, m. d. jimenez, and r. w. klaver. 2013. animal migration amid shifting patterns of phenology and predation: lessons from a yellowstone elk herd. ecology 94:1245–1256. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation 104 upper koyukuk moose movements – joly et al. alces vol. 51, 2015 in alaskan moose. wildlife monographs 122: 1–57. paragi, t. f., c. t. seaton, and k. a. kellie. 2008. identifying and evaluating techniques for wildlife habitat management in interior alaska: moose range assessment. alaska department of fish and game, division of wildlife conservation. final research technical report. grants w-33-4, 5, 6 & 7. project 5.10. juneau, alaska, usa. risenhoover, k. l. 1987. winter foraging strategies of moose in subarctic and boreal forest habitats. ph.d. thesis, michigan technological university, houghton, michigan, usa. sappington, j. j., k. m. longshore, and d. b. thompson. 2007. quantifying landscape ruggedness for animal habitat analysis: a case study using bighorn sheep in the mojave desert. journal of wildlife management 71: 1419–1426. testa, j. w., e. f. lee, and g. r. lee. 2000. movements of female moose in relation to birth and death of calves. alces 36: 155–162. van ballenberghe, v. 1977. migratory behavior of moose in southcentral alaska. transactions of the 13th international congress of game biologists 13: 103–109. vivas, h. j., and b.-e. saether. 1987. interactions between a generalist herbivore, the moose alces alces, and its food resources: an experimental study of winter foraging behavior in relation to browse availability. journal of animal ecology 56: 509–520. white, k. s., n. l. barten, s. crouse, and j. crouse. 2014. benefits of migration in relation to nutritional condition and predation risk in a partially migratory moose population. ecology 95: 225–237. alces vol. 51, 2015 joly et al. – upper koyukuk moose movements 105 variation in fine-cale movements of moose in the upper koyukuk river drainage, northcentral alaska methods study area moose relocation data and gis analyses results discussion acknowledgements literature cited alces16_152.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces17_193.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces16_275.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_360.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces20_79.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces15_iveditorialcommittee.pdf alces vol. 15, 1979 alces16_82.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_527.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_444.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces14_21.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces15_280.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces15_388.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces15_169.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces14_109.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 shoot growth responses at supplementary feeding stations for moose in norway karen marie mathisen1, amandine rémy2, and christina skarpe1 1hedmark university college, department of forestry and wildlife management, anne evenstadsvei 84, 2480 koppang, norway; 2agrocampus-ouest, 65 rue de saint-brieuc cs 84215, 35042 rennes cedex, france abstract: moose browsing pressure in the vicinity of supplementary winter feeding stations eventually declines over time. it is believed that continual winter browsing over multiple years causes locally reduced shoot growth and forage availability for moose (alces alces). we tested this hypothesis by comparing the size of annual shoots of scots pine (pinus sylvestris), downy birch (betula pubescens), and norway spruce (picea abies) along a distance gradient from supplementary feeding stations. contrary to our hypothesis, we found that shoot size was larger at feeding stations than at distances out to 1500 m. this increase in shoot size was probably not related directly to browsing, but to higher nutrient and light availability associated with moose activity at feeding stations. increased use of norway spruce, yet reduced browsing overall at feeding stations, probably reflects the overall decline in abundance of preferred scots pine and downy birch in a local environment substantially altered by an artificially and abnormally high density of moose. alces vol. 51: 123–133 (2015) key words: accumulated browsing, alces alces, betula pubescens, moose, picea abies, pinus sylvestris, plant response, shoot biomass, supplementary feeding understanding plant-herbivore interactions is important for management of moose (alces alces) populations and habitat. plant morphology and architecture influence forage selection of moose (shipley 2007), which in turn can affect plant growth patterns (danell et al. 2003, pastor and de jager 2013). in general, plants respond to herbivory through compensatory growth, thereby reducing the impact of herbivory on growth and fitness (mcnaughton 1983). however, plant species differ in their responses to herbivory; for example, after browsing, deciduous plants may show stronger compensatory growth than conifers because conifers have predetermined shoot growth, and different storage sites for nutrients and distribution of meristems (millard et al. 2001). herbivores may affect forage plants such that the quantity and accessibility of their food source is altered (christie et al. 2014). browsing by moose over time can reduce biomass production in scots pine (pinus sylvestris), downy birch (betula pubescens), and silver birch (b. pendula) depending on the intensity of browsing and the produc‐ tivity of the environment (persson et al. 2005). moderate browsing may increase biomass production in birch, but high intensity browsing over an extended period generally leads to reduced biomass production in both birch and scots pine (persson et al. 2007). shoot size of birch can increase in response to winter-browsing by moose, but the number of shoots and shoot biomass production are reduced (danell et al. 1985, danell karen marie mathisen, hedmark university college, faculty of applied ecology and agricultural sciences, department of forestry and wildlife management, anne evenstadsvei 84, 2480 koppang, norway, karen.mathisen@hihm.no 123 mailto:karen.mathisen@hihm.no et al. 1997). shoot size in scots pine can increase at low browsing intensity and decline at high browsing intensity (edenius 1993, edenius et al. 1993). although short term browsing may have positive effects on shoot biomass, accumulated browsing over multiple years, or at high intensity, reduces shoot biomass. shoot size is an important selection criterion of moose (belovsky 1981). according to the plant vigor hypothesis, herbivores prefer large shoots from vigorous, fast-growing plant modules (price 1991); therefore, moose may avoid smaller shoots from slow growing trees. selection of stems can be predicted by the tradeoff between fast harvesting and high quantity (large twigs), and quick digestion and high quality (small twigs) (shipley 2010), but moose select larger bite diameters when forage availability declines (shipley and spalinger 1995). further, secondary chemical compounds influence shoot selection during winter; for example, moose select tea-leaved willow (salix phylicifolia) stems with lower concentration of phenolics (stolter et al. 2013). supplementary or diversionary winter feeding of moose is a common management practice in fennoscandia (gundersen et al. 2004, sahlsten et al. 2010) that affects the local environment at feeding sites and modifies plant-herbivore interactions (doenier et al. 1997, smith 2001, putman and staines 2004, cooper et al. 2006, rajsky et al. 2008). it also affects spatial distribution, activity, and movement of moose around feeding stations. to some degree, it can be described by central-place foraging theoryhigh moose density at feeding stations and an inverse relationship between moose den‐ sity and distance from the feeding station (van beest et al. 2010a, 2010b, mathisen et al. 2014). the high browsing impact at feeding stations often leads to local food depletion (van beest et al. 2010a), and high nutrient input from dung and urine may alter vegetation composition and structure, with cascading effects on birds and mammals (mathisen and skarpe 2011, mathisen et al. 2012, pedersen et al. 2014). browsing on scots pine and downy birch increases initially (gundersen et al. 2004) but eventually declines at annual feeding stations (van beest et al. 2010a), coincident with increased browsing of norway spruce (picea abies), a less preferred forage species (shipley et al. 1998, månsson et al. 2007b). van beest et al. (2010a) suggested that accumulated browsing over time reduces shoot production in downy birch and scots pine in the vicinity of feeding stations, perhaps explaining the eventual reduced browsing pressure at feeding stations. the goal of this study was to investigate the hypothesis that trees at moose feeding stations produce small shoots due to long-term browsing impact, compared with trees further from feeding stations. study area the study occurred in the stor-elvdal municipality in hedmark county in southeast norway (∼61 °n, 11 °e). the vegetation was primarily boreal forest (moen et al. 1999) below commercial timberline (700 m). it consisted of managed stands of pure or mixed scots pine, norway spruce, downy birch, and silver birch, interspersed with grey alder (alnus incana), rowan (sorbus aucuparia), aspen (populus tremula), and willows (salix spp.). the field layer vegetation was dominated by dwarf shrubs such as vaccinium spp.. the 30-year average summer (may–september) and winter (october–april) temperatures at the valley bottom were 11.2 and −4.8 °c, respectively. the average 30-year annual precipitation and snow depth (october– april) were 766 mm and 31.3 cm, respec‐ tively (nmi 2014). moose are the dominant large herbivore in the area with a winter population density between 1.1–3.4 moose per km2 (gundersen 124 shoot size at moose feeding stations – mathisen et al. alces vol. 51, 2015 et al. 2004, storaas et al. 2005, milner et al. 2012). in winter the population concentrates in the lower valleys, leading to browsing damage in young scots pine stands and vehicular collisions on the main road and railway. landowners have winter-fed moose with grass silage since 1990 (initially to divert moose away from the main road and railway) and this supplementary food now provides ∼60% of the population’s winter forage (gundersen et al. 2004, van beest et al. 2010a, milner et al. 2012), and is provided ad libitum at fixed sites throughout the winter (november–march). the amount has increased from a few hundred kg in 1990 to ∼200 tons in 1998, and almost 2000 tons in 2010. feeding stations now number >100 in stor-elvdal alone, and the radius around feeding stations with heavy browsing impact and local browse depletion has expanded from 0.2 km in 1998 to 1 km in 2008 (van beest et al. 2010a). methods moose density can be 50 x higher at feeding stations than at the landscape level during winter, causing high browsing impact on vegetation and a high deposition rate of moose dung and urine. with increasing distance from feeding stations, browsing impact and dung deposition declines rapidly following a negative exponential function (mathisen et al. 2014). we measured shoot morphology along a gradient at 3 distance categories from feeding stations using a study design employed in previous surveys (mathisen et al. 2012, pedersen et al. 2014): feeding station (fs), 0–30 m; intermediate (int), 150–400 m; far (far), 900–1500 m. to minimize variation among plots, we focused on young, mixed coniferbirch forest in the norwegian forestry cutting classes 2 (trees up to 8 m) and 3 (trees above 8 m but not mature for felling), with a field layer dominated by bilberry (vac‐ cinium myrtillus) or cowberry (v. vitis-idaea) (moen et al. 1999). these criteria yielded a balanced design of 11 plots in each of the 3 distance categories (fs, int, far). since the plots were not designated by random allocation, the study design was not strictly experimental. rather, it was quasiexperimental that includes highly structured observational studies, with some nonrandom treatments unlikely to be fraught with confounding factors (shadish et al. 2002). at each site, samples of current annual shoots from scots pine, downy birch, and norway spruce were clipped after the growing season (3 september – 2 october 2013). these tree species were the most common and present at all 3 distances categories, whilst other species were less abundant and not present at all distances. trees were selected for clipping by starting with the closest tree to plot center, and moving sequentially outwards until we clipped a minimum of 10 lateral shoots per species. if the closest tree had no twigs at browsing height, we moved to the next closest tree, and proceeded sequentially outwards. the maximum 3 lateral shoots were clipped from the same tree. shoots were sampled at 3 height intervals (0.5–1, 1.1– 1.5, and 1.51–2.0 m) reflecting the height of trees available and avoiding shoots <50 cm height which are normally covered by snow. the closest shoot to the plot center at each height interval was clipped. top shoots were sampled if they were <3 m (maximum moose browsing height). for each tree, an index of accumulated browsing was assigned: 0 = no previous browsing; 1 = previous browsing present, but architecture of the tree had not changed; 2 = previous browsing present and architecture of the tree had changed; 3 = previous browsing present and architecture of tree had strongly changed (skarpe et al. 2000). stem diameter was measured (nearest 0.1 mm) at alces vol. 51, 2015 mathisen et al. – shoot size at moose feeding stations 125 the base of the shoot, and length was measured (nearest 0.1 cm) from the base of the shoot to the base of the terminal bud. stem samples were then collected, oven dried at 100 °c to constant weight, and weighed (nearest 0.01g). the relationship between distance to feeding station and shoot diameter, length, and dry weight of each species were analyzed separately using linear mixed models in the package nlme in r 3.1.0 (r development core team 2014). although these are essentially 3 different measurements to describe the size of a shoot and are not independent variables, we chose to analyse them as such to investigate if they responded differently to distance to feeding station. distance to feeding station (fs/int/far), accumulated browsing index, shoot type (top/lateral), and height interval were included as explanatory variables. in addi‐ tion, the interaction between distance to feeding station and an accumulated browsing index was included for scots pine and downy birch to test if the effect of feeding station differed from the effect of previous browsing; norway spruce was not included because it was only browsed at feeding stations. specific tree identification and specific tree within site identification were included as random intercepts to account for dependency within shoots sampled from the same tree or site. shoot length and dry weight were log-transformed to fulfill assumptions of normal distribution and homogeneity of variance. the effect of explanatory variables was assessed by comparison of nested models by dropping one explanatory variable at a time following zuur et al. (2009). the like‐ lihood ratio between models including/ excluding each variable was evaluated, and variables showing a significant relationship (p < 0.05) with shoot morphology were included in the model with estimates presented for these variables. results we measured 1253 shoots from 580 trees. shoot length and dry weight of all 3 species varied with distance to feeding stations (table 1, fig. 1). shoots tended to be larger at feeding stations compared to int and far plots, whilst shoot size was more similar between int and far plots (fig. 1). scots pine had the strongest response to feeding stations with shoot length, diameter, and dry weight significantly greater at feeding stations than at int and far plots; dry weight increased 140%, length 102%, and diameter 42% at feeding stations compared with far plots. for downy birch and norway spruce, shoot dry weight and length increased from feeding stations to far plots 49% and 56%, and 34% and 67%, respectively. shoot diameter of downy birch and norway spruce did not vary significantly with distance to feeding stations (table 1). shoot size was generally larger for top shoots than lateral shoots in all 3 species (fig. 1), and increased with height (table 2). accumulated browsing of downy birch declined slightly with increasing distance from feeding stations, whereas, it was high at all distance categories in scots pine (fig. 2). many scots pine were unbrowsed at feeding stations but <50 cm high, shorter than the average snow depth in the area and protected from browsing. for norway spruce, accumulated browsing was high at feeding stations, but browsing levels were minimal at further distances (fig. 2). we found no effect of the interaction between accumulated browsing and distance to feeding stations on shoot morphology of downy birch and scots pine (table 1). shoot length of scots pine was the only response variable that differed significantly with accumulated browsing; it declined as the accumulated browsing index increased (table 1, fig. 3). the difference in shoot length of scots pine between feeding stations and far plots was greater (102%) than the relationship 126 shoot size at moose feeding stations – mathisen et al. alces vol. 51, 2015 between accumulated browsing and shoot length. shoot length with an accumulated browsing index of 3 was 26% shorter than where the index was 0 (fig. 3). discussion moose browsing on natural vegetation declined over time at the feeding stations in our study area. our starting hypothesis was that this decline was related to intensive accumulated browsing at feeding stations which causes production of smaller shoots avoided/less preferred by moose. we found some weak support for this hypothesis in scots pine, but overall and contrary to our hypothesis, shoot size in all 3 tree species was larger at feeding stations. the increase in shoot length and dry mass was predominantly at feeding stations as the increase in shoot size did not extend to the farther distance plots. the increase could reflect a compensatory response to previous browsing (mcnaughton 1983) as several studies indicate that trees browsed by moose produce larger, but fewer shoots (danell et al. 1985, bergström and danell 1987, danell and bergström 1989). a strong gradient between accumulated browsing data and distance to feeding stations was evident only in norway spruce, although it was only browsed measurably at feeding stations. browsing on downy birch and scots pine was high at all distance categories as the accumulated browsing index was 2–3 on >70% of trees. previous studies at the table 1. significance of distance to supplementary feeding stations (fs.dist) for accumulated browsing (acc.br) and their interaction, height above ground, and shoot type (top/lateral) on current annual shoot morphology analysed by linear mixed models, stor-elvdal, norway. likelihood ratio (l), p, and df are presented for each variable. species response variable interaction fs. dist*acc.br (df = 6) distance to fs (df = 2) accumulated browsing (df = 3) height (df = 2) top / lateral (df = 1) scots pine diameter l = 2.89 l = 14.54 l = 0.74 l = 13.68 l = 134.4 p = 0.823 p < 0.001 p = 0.864 p = 0.001 p < 0.001 length l = 6.74 l = 15.02 l = 7.95 l = 5.10 l = 110.10 p = 0.348 p < 0.001 p = 0.047 p = 0.078 p = 0.078 dry weight l = 1.37 l = 15.06 l = 1.70 l = 11.81 l = 136.76 p = 0.968 p < 0.001 p = 0.637 p = 0.003 p < 0.001 downy birch diameter l = 4.05 l = 3.45 l = 1.71 l = 7.53 l = 16.31 p = 0.671 p = 0.178 p = 0.634 p = 0.110 p < 0.001 length l = 11.46 l = 10.88 l = 2.28 l = 2.42 l = 19.56 p = 0.075 p = 0.004 p = 0.516 p = 0.660 p < 0.001 dry weight l = 8.71 l = 7.97 l = 2.78 l = 11.91 l = 24.56 p = 0.190 p = 0.019 p = 0.427 p = 0.018 p < 0.001 norway spruce diameter – l = 1.48 l = 1.42 l = 48.82 l = 230.90 p = 0.477 p = 0.702 p = <0.001 p < 0.001 length – l = 17.53 l = 2.04 l = 30.11 l = 24.28 p < 0.001 p = 0.563 p = <0.001 p < 0.001 dry weight – l = 7.30 l = 0.70 l = 60.04 l = 115.52 p = 0.026 p = 0.874 p = <0.001 p < 0.001 alces vol. 51, 2015 mathisen et al. – shoot size at moose feeding stations 127 feeding stations noted that a distance gradient in browsing pressure occurs from preferred species such as downy birch and scots pine to less preferred species such as norway spruce (van beest et al. 2010a, mathisen et al. 2014). further, accumulated browsing had no effect on shoot size in norway spruce and downy birch, but was related negatively 0 2 4 6 8 10 12 14 fs int far sh oo t l en gt h (c m ) scots pine lateral top 0 0.5 1 1.5 2 2.5 3 3.5 4 fs int far sh oo t d ry w ei gh t (g ) scots pine lateral top 0 5 10 15 20 fs int far sh oo t l en gt h (c m ) downy birch 0 0.1 0.2 0.3 0.4 0.5 fs int far sh oo t d ry w ei gh t (g ) downy birch 0 2 4 6 8 10 12 14 fs int far sh oo t l en gt h (c m ) distance from supplementary feeding sta�on for moose norway spruce 0 0.2 0.4 0.6 0.8 1 1.2 1.4 fs int far sh oo t d ry w ei gh t (g ) distance from supplementary feeding sta�ons for moose norway spruce fig. 1. effects of distance from supplementary feeding stations for moose on shoot length and dry weight (mean ± 2 se) of current annual shoots of common tree species in storelvdal, norway (fs = feeding station, int = intermediate distance [150–400 m], and far [900–1500 m]). 128 shoot size at moose feeding stations – mathisen et al. alces vol. 51, 2015 with shoot length in scots pine. therefore, it seems unlikely that the increase in shoot size at feeding stations (compared to farther distances) is explained entirely by browsing, at least for scots pine and downy birch. an alternative explanation may be that the increase in shoot size at feeding stations is related to fertilization by an abnormally high input of moose dung and urine. for example, fertilization increases shoot length and twig biomass in downy birch (månsson et al. 2009). in stor-elvdal up to 2000 tons of silage is converted to moose feces and urine annually, and most is deposited within 50 m of feeding stations (mathisen et al. 2014). although local scots pines have access to abundant nutrients because their canopies are limited from heavy browsing, these trees presumably invest in regrowth of a few large shoots. increased shoot growth could also be an indirect result of heavy browsing at feeding stations that promotes more open canopy and sunlight (persson et al. 2005, mathisen et al. 2010). previous studies found table 2. relationship between shoot dry weight (mean ± 2 se) and height above ground for lateral current annual shoots at moose feeding stations in stor-elvdal, norway. shoot mass (g) species height (m) mean 2 se downy birch 0.5–1 0.14 0.02 1–1.5 0.15 0.03 1.5–2 0.18 0.05 norway spruce 0.5–1 0.23 0.04 1–1.5 0.29 0.05 1.5–2 0.27 0.05 scots pine 0.5–1 0.98 0.13 1–1.5 1.35 0.63 1.5–2 1.82 1.37 0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% fs int far fs int far fs int far betula pubescens pinus sylvestris picea abies fr eq ue nc y of tr ee s ab 3 ab 2 ab 1 ab 0 fig. 2. frequency of trees with different accumulated browsing index (ab 0 = no previous browsing; ab 1 = previous browsing present, but architecture of the tree had not changed; ab 2 = previous browsing present and architecture of the tree had changed; ab 3 = previous browsing present and architecture of tree had strongly changed) sampled at different distances from supplementary feeding stations for moose in stor-elvdal, norway (fs = feeding station, int = intermediate distance [150–400 m], and far [900–1500 m]). alces vol. 51, 2015 mathisen et al. – shoot size at moose feeding stations 129 that species composition in the field layer vegetation at feeding stations changed towards more nitrogen and light-demanding species (mathisen and skarpe 2011, mathisen et al. 2012, pedersen et al. 2014). interestingly, limitation of light can have an equally suppressive effect on plant height as browsing (mclaren 1996). chemical defense and nutrient composition in shoots may also have changed due to the combination of previous browsing and altered environmental conditions at feeding sites. as nutrients are often more limiting than light in boreal forests, trees utilize carbon-based rather than nitrogen-based defenses (bryant et al. 1983). trees at feeding sites had high nutrient availability due to fertilization, yet minimal canopy that may reduce the availability of carbon. consequently, these trees could have a relative surplus of nutrients and shortage of carbon. how the artificial environment at feeding stations affects plant chemistry and forage choice is uncertain; however, because previous browsing, in general, increases palatability of shoots for moose (bergqvist et al. 2003), as does increased nutrient availability (ball et al. 2000, mansson et al. 2009), it seems an unlikely explanation for the reduction in browsing pressure over time. in general, moose select diets to optimize digestible energy intake per day, and are more sensitive to changes in plant morphology than plant chemistry (shipley 2010). seasonal diets and consumption levels, particularly in winter, are largely influenced by environmental conditions that dictate forage quality and quantity. stolter et al. (2013) investigated moose browsing in the study area year-round, and found that the relative importance of plant morphology and plant chemistry varied seasonally. browsing on less preferred tree species such as norway spruce at feeding stations (van beest et al. 2010a) may indicate that supplementary feeding changes browsing preferences; however, we did not directly measure consumption or availability of the 3 browse species, and it is highly unlikely that fundamental forage preferences changed. in fact, over time the peak in browsing pressure on scots pine moved outwards from 12.5 to 500 m from feeding stations (van beest et al. 2010a), and gps-collared moose at feeding stations traveled daily up to 500 m distant (mathisen et al. 2014). because supplementary food measurably elevates the nutritional status of local moose (milner et al. 2012) and represents 60% of their winter forage requirements, it is more likely that fed moose demonstrate less selective browsing than unfed animals, and simply consume the mix of browse available at feeding stations where they consume the majority of their food. 0 2 4 6 8 0 1 2 3 sh oo t l en gt h (c m ) accumulated browsing index fig. 3. effects of accumulated browsing on shoot length (mean ± 2se) in scots pine at feeding stations for moose in stor-elvdal, norway. results are averaged over levels of distance to feeding stations and shoot type. the accumulated browsing index was defined as: 0 = no previous browsing; 1 = previous browsing present, but architecture of the tree had not changed; 2 = previous browsing present and architecture of the tree had changed; 3 = previous browsing present and architecture of tree had strongly changed. 130 shoot size at moose feeding stations – mathisen et al. alces vol. 51, 2015 our results suggest that the temporal decline in browsing pressure at feeding stations was not due to higher availability of smaller shoots, as hypothesized previously (van beest et al. 2010a). an alternative explanation is that the number of shoots per tree and total production of browse biomass decline at feeding stations over time (danell et al. 1997, persson et al. 2007). indeed, the number of shoots per tree on both scots pine and downy birch at feeding stations decreased over time (f. van beest, unpublished data), which might reduce preference at the individual tree level (senft et al. 1987, månsson et al. 2007a), a hierarchal level of browse selection above our measurements and study design. acknowledgements we wish to thank students f. stelter and w. sturris for assistance with the fieldwork. we are grateful for support from hedmark university college, department of forestry and wildlife management, and the alces conference for our poster display. references ball, j. p., k. danell, and p. sunesson. 2000. response of a herbivore community to increased food quality and quantity: an experiment with nitrogen fertilizer in a boreal forest. journal of applied ecology 37: 247–255. belovsky, g. e. 1981. food plant-selection by a generalist herbivore the moose. ecology 62: 1020–1030. bergqvist, g., r. bergström, and l. edenius. 2003. effects of moose (alces alces) rebrowsing on damage development in young stands of scots pine (pinus sylvestris). forest ecology and management 176: 397–403. bergström, r., and k. danell. 1987. effects of simulated winter browsing by moose on morphology and biomass of 2 birch species. journal of ecology 75: 533–544. bryant, j. p., f. s. chapin, and d. r. klein. 1983. carbon/nutrient balance of boreal plants in relation to vertebrate herbivory. oikos 40: 357–368. christie, k. s., r.w. ruess, m. s. lindberg, and c. p. mulder. 2014. herbivores influcence growth, reproduction, and morphology of a widespread arctic willow. plos one 9: 1–9. cooper, s. m., m. k. owens, r. m. cooper, and t. f. ginnett. 2006. effect of supplemental feeding on spatial distribution and browse utilization by white-tailed deer in semi-arid rangeland. journal of arid environments 66: 716–726. danell, k., and r. bergström. 1989. winter browsing by moose on 2 birch species impact on food resources. oikos 55: 11–18. ———, ———, l. edenius, and g. ericsson. 2003. ungulates as drivers of tree population dynamics at module and genet levels. forest ecology and management 181: 67–76. ———, e. haukioja, and k. hussdanell. 1997. morphological and chemical responses of mountain birch leaves and shoots to winter browsing along a gradient of plant productivity. ecoscience 4: 296–303. ———, k. huss-danell, and r. bergström. 1985. interactions between browsing moose and 2 species of birch in sweden. ecology 66: 1867–1878. doenier, p. b., g. d. delgiudice, and m. r. riggs. 1997. effects of winter supplemental feeding on browse consumption by white-tailed deer. wildlife society bulletin 25: 235–243. edenius, l. 1993. browsing by moose on scots pine in relation to plant resource availability. ecology 74: 2261–2269. ———, k. danell, and r. bergström. 1993. impact of herbivory and competition on compensatory growth in woodyplants-winter browsing by moose on scots pine. oikos 66: 286–292. alces vol. 51, 2015 mathisen et al. – shoot size at moose feeding stations 131 gundersen, h., h. p. andreassen, and t. storaas. 2004. supplemental feeding of migratory moose alces alces: forest damage at two spatial scales. wildlife biology 10: 213–223. mathisen, k. m., f. buhtz, k. danell, r. bergström, c. skarpe, o. suominen, and i. l. persson. 2010. moose density and habitat productivity affects reproduction, growth and species composition in field layer vegetation. journal of vegetation science 21: 705–716. ———, j. m. milner, f. m. van beest, and c. skarpe. 2014. long-term effects of supplementary feeding of moose on browsing impact at a landscape scale. forest ecology and management 314: 104–111. ———, s. pedersen, e. b. nilsen, and c. skarpe. 2012. contrasting responses of two passerine bird species to moose browsing. european journal of wildlife research 58: 535–547. ———, and c. skarpe. 2011. cascading effects of moose (alces alces) management on birds. ecological research 26: 563–574. mclaren, b. e. 1996. plant-specific response to herbivory: simulated browsing of suppressed balsam fir on isle royale. ecology 77: 228–235. mcnaughton, s. j. 1983. compensatory plant-growth as a response to herbivory. oikos 40: 329–336. millard, p., a. hester, r. wendler, and g. baillie. 2001. interspecific defoliation responses of trees depend on sites of winter nitrogen storage. functional ecology 15: 535–543. milner, j. m., t. storaas, f. m. van beest, and g. lien. 2012. improving moose forage with benefits for the hunting, forestry and farming sectors. final report nr 1–2012. hedmark university college, elverum, norway. (in norwegian with english summary). moen, a., a. lillethun, and a. odland. 1999. vegetation. norwegian mapping authority, hønefoss, norway. månsson, j., h. andren, a. pehrson, and r. bergström. 2007a. moose browsing and forage availability: a scale-dependent relationship? canadian journal of zoology 85: 372–380. ———, r. bergström, and k. danell. 2009. fertilization-effects on deciduous tree growth and browsing by moose. forest ecology and management 258: 2450–2455. ———, c. kalen, p. kjellander, h. andren, and h. smith. 2007b. quantitative estimates of tree species selectivity by moose (alces alces) in a forest landscape. scandinavian journal of forest research 22: 407–414. (nmi) norwegian meterological institute. 2014. (accessed september 2014). pastor, j., and n. r. de jager. 2013. simulated responses of moose populations to browsing-induced changes in plant architecture and forage production. oikos 122: 575–582. pedersen, s., k. m. mathisen, l. gorini, h. p. andreassen, e. røskaft, and c. skarpe. 2014. small mammal responses to moose supplementary winter feeding. european journal of wildlife research 60: 527–534. persson, i. l., r. bergström, and k. danell. 2007. browse biomass production and regrowth capacity after biomass loss in deciduous and coniferous trees: responses to moose browsing along a productivity gradient. oikos. 116: 1639–1650. ———, k. danell, and r. bergström. 2005. different moose densities and accompanied changes in tree morphology and browse production. ecological applications 15: 1296–1305. price, p. w. 1991. the plant vigor hypothesis and herbivore attack. oikos 62: 244–251. 132 shoot size at moose feeding stations – mathisen et al. alces vol. 51, 2015 http://eklima.met.no http://eklima.met.no putman, r. j., and b. w. staines. 2004. supplementary winter feeding of wild red deer cervus elaphus in europe and north america: justifications, feeding practice and effectiveness. mammal review 34: 285–306. r development core team. 2014. version 3.1.0. r foundation for statistical computing, vienna, austria. rajsky, m., m. vodnansky, p. hell, j. slamecka, r. kropil, and d. rajsky. 2008. influence supplementary feeding on bark browsing by red deer (cervus elaphus) under experimental conditions. european journal of wildlife research 54: 701–708. sahlsten, j., n. bunnefeld, j. månsson, g. ericsson, r. bergström, and h. dettki. 2010. can supplementary feeding be used to redistribute moose alces alces? wildlife biology 16: 85–92. senft, r. l., m. b. coughenour, d. w. bailey, l. r. rittenhouse, o. e. sala, and d. m. swift. 1987. large herbivore foraging and ecological hierarchies. bioscience 37: 789–799. shadish, w. r., t. d. cook, and d. t. campbell. 2002. experimental and quasi-experimental designs for generalised causal inference. houghton mifflin, boston, massachusetts, usa. shipley, l. a. 2007. the influence of bite size on foraging at larger spatial and temporal scales by mammalian herbivores. oikos 116: 1964–1974. ———. 2010. fifty years of food and foraging in moose: lessons in ecology from a model herbivore. alces 46: 1–13. ———, s. blomquist, and k. danell. 1998. diet choices made by free-ranging moose in northern sweden in relation to plant distribution, chemistry, and morphology. canadian journal of zoology 76: 1722–1733. ———, and d. e. spalinger. 1995. influence of size and density of browse patches on intake rates and foraging decisions of young moose and white-tailed deer. oecologia 104: 112–121. skarpe, c., r. bergström, a. l. braten, and k. danell. 2000. browsing in a heterogeneous savanna. ecography 23: 632–640. smith, b. l. 2001. winter feeding of elk in western north america. journal of wildlife management 65: 173–190. stolter, c., j. p. ball, and r. julkunentiitto. 2013. seasonal differences in the relative importance of specific phenolics and twig morphology result in contrasting patterns of foraging by a generalist herbivore. canadian journal of zoology 91: 338–347. storaas, t., k. b. nicolaysen, h. gundersen, and b. zimmermann. 2005. moose – traffic in stor-elvdal 2000–2004 – how to avoid moosevehicle accidents on roads and railway lines. report 1–2005. hedmark university college, elverum, norway. (in norwegian with english summary). van beest, f. m., h. gundersen, k. m. mathisen, j. m. milner, and c. skarpe. 2010a. long-term browsing impact around diversionary feeding stations for moose in southern norway. forest ecology and management 259: 1900–1911. ———, l. e. loe, a. mysterud, and j. m. milner. 2010b. comparative space use and habitat selection of moose around feeding stations. journal of wildlife management 74: 219–227. zuur, a. f., e. n. ieno, n. j. walker, a. a. saveliev, and g. m. smith. 2009. mixed effects models and extensions in ecology with r. springer, new york, new york, usa. alces vol. 51, 2015 mathisen et al. – shoot size at moose feeding stations 133 shoot growth responses at supplementary feeding stations for moose in norway study area methods results discussion acknowledgements references combining photography and a geographic information system to measure winter browse use roy v. rea, jamie d. svendsen, and hugues b. massicotte ecosystem science and management, university of northern british columbia, 3333 university way, prince george, british columbia, canada, v2n 4z9 abstract: browse use surveys such as the twig-length method typically used to assess browsing by ungulates are time-consuming and costly. here, we describe a modification of the twig-length method that utilizes digital photography and a geographic information system (gis) technique to quantify browse shoot removal. linear regression analysis indicated that the cumulative shoot length (cm) and biomass removal (g) estimated with our indirect method was similar to direct measurements on scouler’s willows (salix scouleriana). our results suggest that this indirect browse assessment procedure could reduce field time, presumably increase sample size and efficiency, and create a photographic record of each plant for long-term assessment of moose (alces alces) browsing. alces vol. 52: 67–72 (2016) key words: browse, clipping, forage, gis, range, survey, twig, ungulate, willow measuring winter browse utilization of trees and shrubs by ungulates is performed by ecologists to understand ungulate diet choice and feeding requirements, and to provide important information for sound range and ungulate management programs (jensen and urness 1981). quantifying browse removal by herbivores is also fundamental to understanding the ecology of shrub and tree communities that are consumed (bilyeu et al. 2007). methods for determining browse use include detailed twig counts before and after ungulate browsing and percent shoot removal calculations (stickney 1966, dumont et al. 2000, ball and dahlgren 2002), as well as various techniques to estimate the amount of biomass removal (bobek and bergström 1978, rutherford 1979, persson et al. 2005) including broad browsed/form classification systems (schmutz 1983, luttmerding et al. 1990). many of these procedures, however, are time-consuming and expensive, with certain techniques requiring both fall and spring field visits in addition to mark‐ ing and tracking use of individual twigs (jensen and urness 1981). the twig-length method assesses utilization of shrubs and trees by measuring the amount of plant material removed by browsing livestock and/or wildlife (smith and urness 1962). utilization is determined by measuring current annual growth on browse plants both before and after use by browsers, typically during fall and spring in temperate zones. it is an accurate and unbiased method, but has been criticized as labour intensive and requiring lengthy field time (jensen and scotter 1977) criticisms commonly directed at most techniques that provide robust and accurate estimates of forage/browse use (hyder et al. 2003, rea et al. 2010). digital photography has been used to simplify field counts and to provide estimates of leaf area index (macfarlane et al. 2007), corresponding author: roy v. rea, ecosystem science and management, university of northern british columbia, 3333 university way, prince george, british columbia, canada, v2n 4z9, reav@unbc.ca 67 mailto:reav@unbc.ca canopy closure (guevara-escobar et al. 2005), and fruit yields (zaman et al. 2008). recently, photographic methods have been tested for their accuracy and objectivity in estimating forage use (hyder et al. 2003) and simulated browse removal (rea et al. 2010). in this study, we combined the use of digital photography and geographic information system (gis) technology (a digital measurement technique) to quantify winter browse use by moose (alces alces). specifically, in the laboratory we combined these two approaches as an indirect method to estimate the length and biomass of willow twigs removed by simulated (clipping) moose browsing. we hypothesized that the combination of high resolution photography and digital measurements would provide accurate estimates of shoot removal and a more efficient field method to estimate browse use while maintaining the accuracy of the twig-length method. methods during the fall months of 2010, we removed at the stump, 50 whole saplings (~1.5 – 2.0 m tall) of scouler’s willow (salix scouleriana barratt) from the forested lands surrounding the university of northern british columbia, prince george, british columbia, canada (53.895033 °n, �122.816162 °w). the above-ground biomass of each willow sapling was weighed and photographed in the lab in front of a 10 cm lined grid (fig. 1) using a tri-pod mounted canon 5-d digital camera positioned at 130 cm above the ground (just above the mid-point height of our average sapling) and placed 4 m in front of the grid (fig. 2). the camera was equipped with a wide angle lens (ef 24105mm canon zoom) set at a 50 mm focal length, high resolution, and on automatic focus and exposure. images were framed around the midpoint of the grid so that neither the camera position nor its focal length were adjusted between photographs. after photographing each plant, stems were hand-clipped at an approximate diameter of 4 mm (average bite diameter of local moose; carson et al. 2007) at different intensities. the stem mass removed varied between 3 and 86 g, and cumulative shoot a. b. c. fig. 1. photographs of a ~2.0 m tall willow plant before (a) and after (b) simulated browsing. digitized shoots in (c) are marked in highlighter showing those removed by clipping. 68 browse assessments using gis. – rea et al. alces vol. 52, 2016 length between 90 and 1180 cm per plant. plants were re-weighed and photographed a second time in front of the grid as described above. for later comparison, the total length (cm) of stems removed from each plant was measured directly with a hand ruler (nearest mm). photographic analysis pre-browse and post-browse photographs were imported into arcgis (version 9.3.1, esri 2010, redlands, california, usa) and assessed side-by-side to identify which shoots were removed by clipping. following calibration of each photograph with the cells on the measurement grid, we used arcview’s measurement tool to overlay (see fig. 1c) and measure the length (cm) of each shoot removed by clipping. the cumulative shoot length removed (‘browsed’) was then calculated for each willow sapling. data analysis we used linear regression analysis to determine 1) the relationship between the cumulative shoot removal measured with the simulated browsing photographic/gis technique (indirect) and the direct measurements, and 2) the relationship of biomass removal to cumulative shoot length removal associated with the indirect and direct measurement techniques. all analyses were conducted in statistica 9.0 (statsoft 2009). results there was a strong and significant relationship (indirect shoot length (cm) = 0.991 [direct shoot length] + 2.1455; fig. 3) between shoot length estimated with the photographic/gis (indirect) technique and the direct measurements (f1,48 = 3853.9, p < 0.0001; r2 = 0.988). the relationship between biomass removal and cumulative shoot length was strong and significant for both the indirect estimates (f1,48 = 521.327, p < 0.0001; r 2 = 0.916) and direct measurements (f1,48 = 510.495, p < 0.0001; r2 = 0.914). the predictive relationships were similar producing nearly identical regression lines (fig. 4): indirect shoot length = 14.247 (biomass removed) + 86.811 and direct shoot length = 14.280 (biomass removed) + 87.661. discussion our indirect photographic/gis browse assessment technique described here performed extremely well for estimating winter 0 200 400 600 800 1000 1200 1400 length (cm) hand measurement 0 200 400 600 800 1000 1200 1400 le ng th ( cm ) d ig ita l m ea su re m en t fig. 3. the relationship (indirect shoot length (cm) = 0.991[direct shoot length] + 2.1455; r2 = 0.988) between the handmeasured (direct) and the photographic/ gis (indirect) measurements of cumulative shoot length removal from scouler’s willows (n = 50). 4m 130cm fig. 2. a photo-geometric representation of the protocol used to photograph willow saplings in front of our lined grid. alces vol. 52, 2016 rea et al. – browse assessments using gis. 69 shoot length and biomass removal from willow saplings. we expect digital photography coupled with gis technology to be equally useful for estimating browse shoot removals from other deciduous species, although certain differences may occur due to variable plant architecture (rea et al. 2010). the method was subsequently found equally useful in estimating the amount and position of winter twigs removed from both scouler’s willow and paper birch (betula papyrifera marsh.) by moose in cafeteria style feeding trials (rea et al. 2015). although tested here in a laboratory setting, the design of an easily transportable simple plastic or cloth sheet (see schmutz 1983) with a superimposed grid as a backdrop would provide for the photographic component to be performed in the field for preand post-browse utilization assessments on in situ plants (rea et al. 2010). to test field applicability, a much longer-term study could be designed and executed where plants are photographed at the end of summer and the following spring before leaf flush. the challenge would be to mark plants such that both photographs are taken from the same perspective to ensure an accurate estimate of browse removal. the length of shoot removal is closely correlated with biomass removal (jensen and scotter 1977), and our results were also strongly correlated. the indirect estimates of cumulative shoot length removal so closely approximated the direct measurements that we accurately estimated biomass removal using photography/gis. we did not measure diameter where shoots were clipped (bite diameters), as is often done in browse surveys (portinga and moen 2015). the diameter at the browsed tip, however, is often predicted with regression equations developed from shoot biomass or length or vice versa, with biomass removal more typically predicted from bite diameter (ruyle et al. 1983). instead, we used the length measured with gis to predict browse biomass which, we believe, circumvented the need to consider or calculate diameter at point of clipping/biting. browse assessments in the field that employ more traditional and direct measurement techniques can be costly and time consuming (jensen and scotter 1977). the use of this indirect, digital technique described here could reduce the field time spent at each in situ plant by replacing direct measurements with photography. as such, there is the potential to increase the number of plants assessed in the field within a given time frame, thereby increasing sample size and presumably improving analytical accuracy. digital photographs taken at the time of assessment would additionally provide a permanent record of each plant that would allow for multiple assessments and reduce observer bias. databases containing such information could better describe activity, health, and population status of local moose, as well as ecological impacts on vegetative communities. 0 10 20 30 40 50 60 70 80 90 biomass (g wet wt) removed 0 200 400 600 800 1000 1200 1400 dna h dnalati gi d ) mc( ht gnel m ea su re d hand measured gis measured fig. 4. with respect to simulated browsing, similar relationships between shoot biomass (g wet weight) removed from willow saplings (n = 50) and the cumulative shoot length (cm) as measured with the direct and indirect techniques are shown. the direct relationship was: direct shoot length = 14.280 (biomass removed) + 87.661; r2 = 0.914. the indirect relationship was: indirect shoot length = 14.247 (biomass removed) + 86.811; r2 = 0.916. note: regression lines are superimposed on one another. 70 browse assessments using gis. – rea et al. alces vol. 52, 2016 although our technique required time spent working with the gis to outline/digitize the twigs removed by clipping (fig. 1c), computer algorithms and intelligent vision systems (mccarthy et al. 2010) designed to interpret the differences in plant morphology (e.g., between preand post-browsed plants) could reduce the time required to calculate cumulative shoot length and biomass removal. the creation of an artificial cartesian coordinate grid in the gis for registering each photograph to that grid would provide a more systematic measurement protocol. ensuring the angle and perspective at which the preand post-browsed photographs are taken is critical for proper interpretation of the data. standardized lighting, iso, quality, and depth of field settings on the camera also need to be harmonized between preand post-browse photos. like any browse assessment procedure, results will vary by species relative to plant form and twig growth characteristics. how the method performs with plants of complex architecture remains untested; however, accommodation for different plant forms could be approached with some resourcefulness. for example, browsed plants taller or wider than the grid could be imaged by subsections that are later summed for whole plant assessments. plant and browsing height will likely define the practical limits of this technique. nevertheless, when considering allocated field time, working within seasonal windows (e.g., assessing plants after snow melt, but before leaf flush), or attempting to increase plant numbers and data sample sizes – our technique offers an efficient and quick field method for collecting snapshots of browse use on specific plants that can be examined more closely within a controlled laboratory setting regardless of time and weather constraints. acknowledements we would like to thank s. o’keefe, t. windsor and l. tackaberry for their assistance with this project. we would also like to thank the staff of the enhanced forestry laboratory at unbc for providing us with the lab space and equipment necessary to complete this project. we thank s. emmons and h. butow for assistance with our gis analyses. references ball, j. p., and j. dahlgren. 2002. browsing damage on pine (pinus sylvestris and p. contorta) by a migrating moose (alces alces) population in winter: relation to habitat composition and road barriers. scandinavian journal of forest research 17: 427–435. bilyeu, d. m., d. j. cooper and n. t. hobbs. 2007. assessing impacts of large herbivores on shrubs: tests of scaling factors for utilization rates from shoot-level measurements. journal of applied ecology 44: 168–175. bobek, b., and r. bergström. 1978. a rapid method of browse biomass estimation in a forest habitat. journal of range management 31: 456–458. carson, a. w., r. v. rea, and a. l. fredeen. 2007. extent of stem dieback in trembling aspen (populus tremuloides) as an indicator of time-since simulated browsing. rangeland ecology and management 60: 543–547. dumont, a., m. crete, j. p. ouellet, j. huot, and j. lamoureux. 2000. population dynamics of northern white-tailed deer during mild winters: evidence of regulation by food competition. canadian journal of zoology 78: 764–776. guevara–escobar, a., j. tellez, and e. gonzalez-sosa. 2005. use of digital photography for analysis of canopy closure. agroforestry systems 65: 175–185. doi: 10.1007/10457-005-0504-y. hyder, p. w., e. l. fredrickson, m. d. remmenga, r. e. estell, r. d. pieper, and d. m. anderson. 2003. a digital photographic technique for assessing alces vol. 52, 2016 rea et al. – browse assessments using gis. 71 forage utilization. journal of range management 56: 140–145. jensen, c. h, and g. w. scotter. 1977. a comparison of twig-length and browsedtwig methods of determining browse utilization. journal of range management 30: 64–67. ——, and p. j. urness. 1981. establishing browse utilization from twig diameters. journal of range management 34: 113–116. luttmerding, h. a., d. a. dermarchi, e. c. lea, d. v. meidinger, and t. vold. 1990. describing ecosystems in the field. second edition. moe manual 11. ministry of environment, victoria, british columbia, canada. macfarlane, c., m. hoffman, d. eamus, n. kerp, s. higginson, r. mcmurtrie, and m. adams. 2007. estimation of leaf area index in eucalypt forest using digital photography. agricultural and forest meteorology 143: 176–188. mccarthy, c. l., n. h. hancock, and s. r. raine. 2010. applied machine vision of plants: a review with implications for field deployment in automated farming operations. intelligent service robotics 3: 209–217. persson, i. l., k. danell, and r. bergström. 2005. different moose densities and accompanied changes in tree morphology and browse production. ecological applications 15: 1296–1305. portinga, r. l. w., and r. a. moen. 2015. a novel method of performing moose browse surveys. alces 51: 107–122. rea, r.v., o. hjeljord, and m. gillingham. 2015. factors influencing the use of willow and birch by moose in winter. european journal of wildlife research 61: 231–239. ——, d. p. hodder, j. trelenberg, and t. m. o’brien. 2010. the use of stereoscopic photography to estimate browse use by large ungulates. northwest science 84: 103–108. rutherford, m. c. 1979. plant-based techniques for determining available browse and browse utilization: a review. the botanical review 45: 203–228. ruyle, g. b., j. e. bowns, and a. f. schlundt. 1983. estimating snowberry [symphoricarpos oreophilus] utilization by sheep from twig diameter-weight relations. journal of range management 36: 472–474. schmutz, e. m. 1983. browsed-class method of estimating shrub utilization. rangeland ecology and management 36: 632–637. smith, a. d. and p. j. urness. 1962. analyses of the twig length method of determining utilization of browse. bulletin 62–9. utah division of fish and game, salt lake city, utah. usa. statsoft. 2009. statistica for windows, version 9.0. tulsa, oklahoma, usa. stickney, p. f. 1966. browse utilization based on percentage of twig numbers browsed. journal of wildlife management 30: 204–206. zaman, q. u., a. w. schumann, d. c. percival, and r. j. gordon. 2008. estimation of wild blueberry fruit yield using digital color photography. transactions of the american society of agri‐ cultural and biological engineers 51: 1539–1544. 72 browse assessments using gis. – rea et al. alces vol. 52, 2016 combining photography and a geographic information system to measure winter browse use methods photographic analysis data analysis results discussion acknowledements references alcessupp1_52.pdf alces20_259.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alcessupp1_121.pdf alces14_194.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alcessupp1_91.pdf alcessupp1_11.pdf alces19_preface.pdf alces vol. 19, 1983 alces20_161.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces19_136.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alcessupp1_207.pdf alces19_274.pdf alces vol. 19, 1983 alcessupp1_139.pdf alces18_45.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces19_246.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces17_xvdistinguishedmoosebio.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces18_viattendancelist.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces19_71.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 f:\alces\supp2\pagema~1\rus4s.pdf alces suppl. 2, 2002 badlo and simakov nutrition and nitrogen metabolism 19 seasonal features of nutrition and nitrogen metabolism in moose larisa p. badlo and anatoly f. simakov institute of physiology, komi scientific center, ural division of the russian academy of sciences, 167610, syktyvkar gsp, komi republic, russia abstract: the seasonal concentration of nitrogen (n) and nitrogen metabolism in rumen digesta and blood serum were measured in two rumen-fistulated moose. concentration of n and nitrogen metabolism in the rumen and blood varied seasonally. concentration of protein n and metabolic processes in rumen digesta were highest in spring through summer and lower in autumn; the converse was true in blood serum. these seasonal differences were related to the variation in nutritional content of seasonal forage. alces supplement 2: 19-22 (2002) key words: blood serum, moose, nitrogen, rumen metabolism the moose is a ruminant that feeds primarily on arboreal plants. many investigations of their seasonal feeding habits and digestion of natural forages have been conducted (knorre 1959, kaletskyy 1967, sablina 1973, timofeeva 1974, kochanov et al. 1981). we conducted this study to measure seasonal concentration and metabolism of nitrogen in rumen digesta and blood serum, and to relate these data to the seasonal variation in forage quality and forage use by moose. methods this experiment used 2 rumen-fistulated male moose held at the pechora-ilych reservation. the processes of rumen digestion and metabolism were studied during spring, summer, and autumn. the moose were kept on pasture during the experiment but were confined during sampling. they were maintained on food and water ad libitum. rumen digesta samples were taken once daily. blood samples were taken from the jugular vein before feeding. the chemical compounds of food, samples of rumen digesta, and feces were analyzed according to standard procedures. total nitrogen was determined with the kjeldahl method and protein nitrogen with the barnshtein method; non-protein nitrogen equaled the difference between total nitrogen and protein nitrogen. urea concentration was measured with the biotest (cssr). amino nitrogen in the blood was measured according to the uzbeckova method as modified by chulkova. fermentation activity of alt and ast was measured according to the roitmann and frenkel unified method. results and discussion moose differ from domesticated ruminants because environmental influences dictate seasonal variation in forage availabilty, choice, and intake rate. intake rates of moose are high during spring when buds and leaves appear (knorre 1959). biochemical analyses showed that leaves and young sprouts of aboreal plants are rich in nitrogen and crude protein during spring. for example, leaves of rowan-tree and willow have 3.28% nitrogen (20.5% crude protein), aspen leaves have slightly less nitrogen (3.06% or 19.1% crude protein), and, comparanutrition and nitrogen metabolism badlo and simakov alces suppl. 2, 2002 20 tively, birch leaves have lower nitrogen content (2.25% or 14.7% protein). young sprouts of these plant species have quite different concentrations of protein than their leaves; young sprouts of aspen have higher concentration of protein than those of birch and willow. the ratio of protein to non-protein nitrogen (pn:npn) in young sprouts of birch was 3.0, and those of the willow and aspen were 2.8 and 1.7, respectively. aspen bark is also eaten by moose during spring and has a protein concentration of 11%. although the nutritional elements in bark are relatively low, moose consume it because of its high concentration of tannin. both arboreal plants and grass are consumed by moose during summer, but grass represents only 15% of the diet (knorre 1959). during summer balance experiments, birch, rowan-tree, willow, and blossoming willow-herb (rose-bay) were consumed by moose. the concentration of crude protein in arboreal plants in july (13%) was lower than that in spring, although willow-herb had high concentration of crude protein (18.4%). seasonal variation in intake rate helps explain why moose obtain more crude protein during summer than spring. for example, daily consumption of 11.5 kg of natural forage during spring includes 3.55 kg dry matter and 609.4 g protein; whereas, in summer they may consume 22 kg containing 6.6 kg dry matter and 946 g protein, about 2 times and 1.5 times as much dry matter and protein, respectively. variety in forage consumption declines in autumn versus summer. leaves and sprouts of willow and birch, and pine branches, were consumed by the experimental moose during autumn. our analyses indicated that autumn forage declined in nutrition; concentration of crude protein in willow and birch branches was only 7.2% and 6.6%, respectively. although the concentration of protein was high relative to the pn:npn ratio (table 1), the lower absolute values of pn and the decline in daily intake rate to 8.3 kg yielded only 4.2 kg dry matter and 315 g crude protein daily. because the availability of crude protein in forage varies seasonally, similar variation in the intensity of rumen metabolism is likely. it is known that the forestomach of all ruminants plays a primary role in assimilation of nitrogen-containing substances. protein is degraded to peptides, amino acids, and ammonia by enzymatic activity of microorganisms. the elements formed may be used for synthesis of bacterial protein, be table 1. the concentration of nitrogen-containing substances in the rumen digesta of moose. characteristics spring summer autumn nitrogen: total n, g% 3.61 ± 0.13 3.50 ± 0.50 2.06 ± 0.17 protein n, g% 2.40 ± 0.06 2.40 ± 0.12 1.40 ± 0.13 non-protein n, g% 1.20 ± 0.12 0.99 ± 0.12 0.66 ± 0.18 amino-n, mg% 29.30 ± 1.70 55.09 ± 6.20 37.50 ± 3.70 urea, mg% 53.20 ± 4.90 15.20 ± 3.00 18.20 ± 2.00 ammonia, mg% 13.20 5.30 6.80 ratio (pn:npn): in rumen 2.0 2.4 2.1 in food 1.8 2.2 2.5 alces suppl. 2, 2002 badlo and simakov nutrition and nitrogen metabolism 21 partially absorbed through the rumen wall into blood, or flow to post-rumen sections of the digestive tract. hydrolysis of nonprotein nitrogen substances occurs in the rumen; formed elements may be used by microorganisms for the synthesis of amino acids and microbial protein. simultaneously, nitrogen-containing substances may come to the forestomach from salivary secretions and blood through the rumen wall. thus, the main role of the rumen in the metabolism of nitrogen-containing substances is to change or increase the amino acid compounds. saturation of rumen digesta by nitrogen-containing substances was revealed by our measurements of total protein, nonprotein nitrogen, ammonia, and amino-nitrogen. the highest concentration of total nitrogen (3.61 g %) occurred in the springsummer period. variation in diet reflected a drop to 2.06% in autumn (table 1). the concentration of protein and non-protein nitrogen in spring and summer was 1.5 times higher than that in autumn (table 1), indicating the intensive hydrolytic and synthetic processes in spring and summer. comparing seasonal pn:npn ratios of rumen contents revealed increasing content of protein nitrogen from spring to summer and decreasing content from summer to autumn, reflecting the relative level of seasonal microbial activity in the rumen. during spring and summer, nutrient rich forage causes growth and reproduction of rumen microflora, producing valuable microbial protein in the rumen digesta. during autumn these processes slow, due to reduced protein intake, and protein content in the rumen subsequently declines. the essential changes of amino-nitrogen, ammonia, and urea concentration in the rumen are peculiar to the nitrogen metabolism of moose. the highest concentration of ammonia and urea, and the lowest of amino-nitrogen, occurred during spring; the converse was true during summer. these characteristics had intermediate values during autumn (table 2). thus, metabolism of nitrogen-containing substances in the rumen was dependent upon seasonal factors, as were hydrolytic processes in the rumen. results from balance experiments indicated that the highest nutrient digestibility of protein occurred during spring-summer (70–75%), with measurable decline in autumn (52%). the nitrogen balance in summer was 64 g/d; as a table 2. the concentration of nitrogen substances (mg %) in blood serum of moose. characteristics spring summer autumn nitrogen: total n 931 1001 964 protein n 872 922 964 non-protein n 59 79 30 amino-n 3.3 4.1 4.3 urea 30 20 19 protein: refractometry 6.45 6.80 7.04 act ed/ml 74.8 55.6 24.0 alt ed/ml 57.8 36.1 11.0 creatinine 2.1 3.1 nutrition and nitrogen metabolism badlo and simakov alces suppl. 2, 2002 22 result, moose must intake great amounts of nutritive elements in autumn. this intake is indicated by the highest concentration of protein and amino-nitrogen in blood serum during autumn (table 2). similar changes were observed by leresche et al. (1974). in autumn, protein metabolism in the blood declines, as does non-protein nitrogen, urea, and transaminase activity (table 2). high creatinine concentrations in blood during autumn indicated increased metabolism processes in tissues. references kaletskyy, a. a. 1967. moose food during the winter period and overall annual feeding patterns. pages 216221 in biology and harvest of moose. russia publishing, moscow, russia. (in russian). knorre, e. p. 1959. ecology of moose. proceedings of the pechoro-ilych state reserve 7:5–22. (in russian). kochanov, n. e., g. m ivanova., a. e., weber, and a. f. symakov. 1981. the movements of wild animals (northern caribou and moose). nauka, leningrad, russia. (in russian). leresche, r. e., u. c. seal, p. d. karns, and a. w. franzmann 1974. a review of blood chemistry of moose and other cervidae with emphasis on nutritional assessment. naturaliste canadien 101:263–290. sablina, t. b. 1973. fundamental diet of moose in the natural environment. pages 40-53 in the domestication of moose. nauka, moscow, russia. (in russian). timofeeva, e. k. 1974. moose: ecology, distribution, economic importance. leningrad state publishing house, leningrad, russia. (in russian). alces17_136.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces18_142.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces17_15.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 f:\alces\supp2\pagema~1\rus 18s alces suppl. 2, 2002 lozinov and kuznetsov impact of moose on ash 81 the impact of moose on ash productivity georgy l. lozinov and german v. kuznetsov institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: data characterizing the impact of moose (alces alces) on ash (fraxinus spp.) seedlings in the broad–leaved forests of the tula region are given. resistance of ash to the strong browsing pressure is shown and also the resulting peculiarities of crown structure, including the position of shoots and branches. when ash is isolated from moose, annual accretion of shoot phytomass is 5 times more than in locations where ash is exposed to browsing; in leaves, phytomass is 10–12 times more than browsed ash. the large number of shoots and a great quantity of large leaves are considered as an adaptation of the ash tree to survival under the browsing pressure of moose. alces supplement 2: 81-84 (2002) key words: ash tree, fraxinus, moose, productivity a number of workers have reported on the impact of moose browsing on forest plants, including those of the taiga zone ( k a l e t s k a y a 1 9 5 9 , k o z l o v s k y 1 9 6 0 , dinesman and shmalgauzen 1961, perovsky 1973, timofeyeva 1974, fyodorov 1979, kuznetsov 1980, filonov 1983, smirnov 1987). there are fewer reports of moose influence on the arboreal species in the zone of broad–leaved forests. the purpose of this paper is to estimate the moose impact on the ash tree in the closed broad–leaved forests of the russian plateau. study area the forest of the tula abatis consists mainly of middle–aged and old stands and harvested sites containing dense regeneration (kurnayev 1980). it presents a complex mosaic picture. the widely dispersed harvest sites, which are the principal source of browse, and characteristic features of snow cover, determine moose travel routes and feeding locations. methods to estimate the impact of moose browsing activity on ash seedlings on the harvested sites, constant monitoring of the development of sample trees was carried out from 1979 to 1989. as a control, moose were excluded from some of the seedlings. before the onset and at the end of the period of revegetation, we measured the supply of available forage and the mass of shoot accretion both on the moose–isolated trees and those exposed to browsing. the mass of shoot accretion was determined by measuring the lengths of all shoots available to moose and calculating the dry weight in grams. as a basis for the calculation, the mass of a shoot was determined (kuznetsov 1983). besides the index of annual accretion and removal of the phytomass, we measured the height of the trees, diameter of the trunk, number and the length of shoots of the current year, and the quantity and average length of the leaves. to evaluate the density of shoots, special measurements of shoot length were carried out within a volume of space. the parameters impact of moose on ash – lozinov and kuznetsov alces suppl. 2, 2002 82 of the environment, the characteristics of plants, and the herbaceous layer at the beginning of the investigation were similar in both the control area and the area exposed to browsing by moose. the density of moose per 1,000 hectare averaged 5 individuals and changed very little during the study period. results in the tula abatis, ash is the most important forest tree species, forming both pure middle–aged ash stands and young plantations mixed with oak (quercus sp.), elm (ulnus sp.), linden (tilia sp.), and maple (acer sp.). the ash tree, along with oak, is the main forage of moose during the entire year. at the beginning of the investigation in 1979, the mean annual accretion of ash tree shoots made up 11.0 g of dry matter per tree. however, the small quantity of annual shoot accretion was supplemented with a large accretion of leaf phytomass. the annual accretion of phytomass of leaves on stems protected from moose exceeded the accretion of shoot phytomass of exposed stems as much as 10 times, and as much as 12–13 times in some years. the comparison of mean value of ash accretion in the control areas and on stems exposed to browsing shows that the mean annual accretion in isolation is approximately 5 times more than on the exposed stems. the browsing impact of moose on the ash tree is substantial and uniformly distributed with removal of annual growth reaching 40–60%. however, during the first 10 years of monitoring, no ash tree sampled was noted to be dried up. in our opinion, their relative stability was related to the position of shoots on the stem and the size of shoots. the ash tree shoots are rather sharply divided into short (1–5 cm) and long ones up to 1 m in length. another characteristic is the thickness, which reaches 8–10 mm in short shoots that are very thick at the base while at the top the thickness slowly decreases insignificantly up to 10–12 mm. moose changed the crown structure and position of branches, and plants could not escape the reach of moose. stem height remained within the limits from 1 to 1.6 m. browsing of shoots gave the ash crowns a dense appearance. to get objective data characterizing the differences in density of plant shoots subjected to browsing by moose and those on the control area, total length of shoots was measured in a definite volume (a cube with a side of 15 cm), situated in the tree crown (table 1). the dense growth of ash tree leaves should be considered as a special adaptation to moose browsing. this is supported by the correlation of moose–damaged phytomass of shoots and leaves (table 2). discussion from the data given in table 2, we can conclude that leaves are damaged more often than shoots. thus, large complex pinnate leaves of the ash tree create a definite “screen”, protecting young shoots from damage by moose. this leads to the table 1. the total length of ash tree shoots from browsed and protected trees in a volume of 3,375 cm3. length of ash–tree shoots (cm) browsed trees protected trees living shoots 36.2 ± 2.3 24.4 ± 2.6 dead shoots 7.0 ± 2.0 0.0 alces suppl. 2, 2002 lozinov and kuznetsov impact of moose on ash 83 preservation of some accessible shoots and limits phytomass removal. nevertheless, steady moose browsing pressure does not allow ash trees to grow normally. that is why the height–growth of ash trees located where moose are excluded outstrips that of stems exposed to moose browsing. thus, during the period of monitoring, ash trees protected from browsing reached 12–15 m in height and 15 cm in diameter, and successfully reached the first woody canopy layer. but though such sizes do not allow moose to reach ash tree shoots, this does not protect them against moose damage. ash tree bark is very attractive forage for moose. but under our conditions, where constant moose pressure occurs, not only was the forage base decreased, but planted ash and oak seedlings were killed, leading to the overgrowing of harvested plots with tree species that are not subjected to browsing by moose (linden, hazel [corylus sp.]). in addition to this, during winter, moose frequently used plots occupied by apple plantings, and the decrease of these plots played a definite part in increasing damage to ash and oak. according to our data, during the winter in the apple orchards, moose ate shoots 30 times more and bark 12 times more than in the forest per 100 m of daily movement. thus, during the winter– spring period of 1988–1989, moose began to use the bark of large ash trees, 25–30 cm in diameter, situated within large forest tracts. the bark gnawing often girdles large trees and leads to their death. to study the size, degree, and pattern of damage to tree trunks by moose, 6 plots (10 x 10 m) separated by 50 m were established on the harvested site of oak forest covered with young bushes. we examined 903 tree trunks, including 141 ash (table 3). bark gnawing was noted on ash, oak, elm, and common maple. linden, hazel, and table 2. damage to ash tree leaves and shoots by moose in 1983–1987 (average per year, %). 1983 1984 1985 1986 1987 average shoots 17.3 ± 5.4 21.5 ± 8.3 9.7 ± 4.9 7.4 ± 5.2 7.3 ± 2.5 12.6 leaves –– 23.9 ± 2.9 43.1 ± 10.9 11.4 ± 3.9 6.0 ± 2.1 21.2 table 3. damage to the bark of ash trees by moose. plot number damaged average average average of ash trees area of area of weight trees (%) fresh old of fresh gnawings gnawings gnawings / tree / tree / tree (cm2) (cm2) (g) 1 20 66.1 441.5 254.9 51.0 2 43 58.1 25.7 86.3 17.3 3 26 96.2 526.1 429.0 85.8 4 14 100.0 354.2 244.0 48.8 5 35 80.0 809.6 319.5 63.9 6 3 100.0 0.0 121.0 24.2 impact of moose on ash – lozinov and kuznetsov alces suppl. 2, 2002 84 norway maple were not damaged. however, ash was most often damaged. to determine the area of gnawing, the size and the location of gnawing on the trunk was established, and the age of the damage was determined if possible: fresh (last year) or old. the average diameter of damaged ash trees was insignificantly larger than the undamaged ones. considering that the area of the bark on the tree accessible to moose was situated at the height of 0.5–2.5 m, the consumption of bark by moose reached 1.7% of the total area of available bark of ash trees or 2.0% of the area of the bark of damaged ash trees. the average area of bark removed per tree during 1988–1989 was 1.4–fold higher than for the previous period. this relates to the increased moose impact on the vegetation. thus, the kind and level of damage to ash by moose (the percentage of phytomass of shoots and bark removed compared to other species) is complicated and changes as the forest community develops. references dinesman, l. g., and v. i. shmalgauzen 1961. the role of moose in the formation of the primary forest. biology and hunting of moose. russian publishing, moscow, russia. (in russian). filonov, k.p. 1983. moose and forest industry. moscow, russia. (in russian). fyodorov, f. f. 1979. pine tree injuries by moose and dependence on forest inspection of reforestation efforts. pages 45–57 in questions of forest sport hunting. moscow, russia. (in russian). kaletskaya, m. l. 1959. damage by moose to pine seedlings in the darwin animal preserve. proceedings of the institute of silviculture, russian academy of sciences 13:63–69. (in russian). kozlovsky, a. a. 1960. protection of f o r e s t s f r o m d a m a g e b y m o o s e . vnilm, moscow, russia. (in russian). kurnayev, s. f. 1980. broadleaf shade forests of the russian plains and urals. nauka, moscow, russia. (in russian). kuznetsov, g. v. 1980. the role of ungulates in forest ecology (some insights and perspectives from research). pages 88–100 in herbivores in vegetative communities. nauka, moscow, russia. (in russian). ———. 1983. the impact of moose on forest vegetation in the southern taiga. moip bulletin 88:28–35. (in russian). perovsky, m. d. 1973. composition of moose diet in different geographic areas. biological bulletin of the advancement, productivity and conservation of game and fur animals. works vsxizo 63:38–44. (in russian). smirnov, k. a. 1987. the role of moose in biocenoses of the southern taiga. nauka, moscow, russia. (in russian). timofeeva, e. k. 1974. moose: ecology, distribution, economic importance. leningrad state publishing house, leningrad, russia. (in russian). f:\alces\supp2\pagema~1\rus 22s alces suppl. 2, 2002 petrov et al. – morphological changes of organs 105 morphological changes of organs during ontogeny of moose (alces alces) compared to domesticated ruminants anatoley k. petrov, nicholas e. plesnakov, and j. a. isajenkov ivanovo agricultural institute, ivanovo, russia abstract: the ontogenesis of body mass, the skeleton, organs of the alimentary tract, and the endocrine system of moose (alces alces), cattle, sheep, and goats were studied in the ivanovo region using morphological and ontogenetic methods. we noted that changes in the rates of growth and development of body mass and some organs were connected with their functional characteristics in different periods of development. histological studies showed that decline in growth rate of all the organs were accompanied by an increase in complexity in their structural organization. moose had earlier formation of thyroid and adrenal glands, thymus, ossification centers in the skeleton, and of characteristic structures in the organs of the alimentary tract in comparison with domestic animals. alces supplement 2: 105-107 (2002) key words: alimentary tract, ecology, endocrine system, goat, growth, moose, ontogenesis, sheep, skeleton comparative ontogenesis of body mass, skeleton, alimentary tract, and endocrine system of moose (alces alces) and other ruminants are given in the report. characteristics, similarity, and differences in moose, goat, and cattle development are revealed. materials and methods the material for moose research was taken both from dead animals and those specially shot in the ivanovo region. shooting of moose was conducted systematically during the last 2–3 days of every month for the purpose of age selection, focusing in particular on development of fetuses. when defining the age of moose we considered the season of their collection and estimates of age derived from the tooth wear. six regions of ivanovo district provided the material for cattle, sheep, and goats. sex and mass of all animals were recorded. internal organs were taken from dead animals. the skeleton was cleaned of soft tissue. materials were not allowed to dry because that would have changed their mass. bones and internal organs were weighed. the volume, linear dimensions, and mass of bones and internal organs were measured prior to fixing in alcohol–ether for further investigation. histological preparations were made by encasing samples in paraffin prior to slicing. microscopic sections were stained by haematoxylin–eosin. results and discussion observations over several decades in ivanovo region let us draw the following conclusions. the number of moose in our district is increasing every year. moose are recorded in every region. moose in our region have become accustomed to the presence of humans. sometimes they come to settlements, parks, and even industrial areas. meeting moose in the forest is a common occurrence. in our region adult moose are 180–200 centimeters in height, 180–190 centimeters in length, and have a mass of 400–450 kg. morphological changes of organs – petrov et al. alces suppl. 2, 2002 106 thus the moose in ivanovo district are very massive animals. sexual maturity is achieved when moose are 1.5 years old. the rut occurs mainly in september, but in some cases mating may take place in october. duration of the rut is related to the fatness of moose at that time. according to our observations, pregnancy of moose females lasts 7.5 months. young moose are usually born in may. moose females have few outward signs of pregnancy. they usually bear two calves (86%) and young individuals bear only one. moose twins are usually of different sexes (67%), and seldom are of the same sex (33%). unisexual twins are usually males rather than females. hair covering of young animals, 2.5 months of age, has a bright red color and is abundantly smeared with oils. when young moose are 2.5 months old they begin to shed hair. shedding of hair begins in the area of the ulna, knee–joint, and surface of the abdomen, then it spreads to the sides of the body. shedding ends when young moose are 4 months old and hair covering has the appearance of an adult. in studying age characteristics, we ascertained that moose growth takes place primarily within one period of great intensity and then declines. in fetus development, growth is high during the first half of pregnancy and decreases by the time of birth. in the postnatal period, increased growth takes place from birth until 2 months (sheep), 3 months (cattle), or 6 months (moose). increase of mass during the first 6 months of life is 10–fold in moose compared with a 5–fold increase in cattle, and moose of that age are 41% of adult mass, whereas cattle are 28% of their adult mass. growth subsequently decreases gradually until adult size is reached. it should be mentioned that the period from birth to 6 months of age in moose must be used as much as possible for creation of favorable conditions for the purpose of growth and development. fetus development of moose occurs quicker than that of cattle and is comparable to the development of sheep and goats, having shorter terms. moose and cattle fetuses aged 3 months have different attributes. for example, moose have more height at the shoulder, a longer helix, and a shorter tail. at 5 months, the moose fetus has a body covered with short hair (cattle achieve this only during the eighth month of fetal development). on the basis of comparative studies of body formation in moose ontogenesis, we believe that moose grow quicker after birth than they do during fetal development. skeletal growth takes place with the same regularity as body formation. we noted that the growth of the peripheral skeleton of moose prevails over axial growth during development of the fetus but after birth this pattern reverses. the appearance of centers of ossification occurs in moose in the same pattern as in cattle. the main difference is that centers of ossification in moose occur earlier than in cattle, but the age coincides with sheep and goats. the process of ossification of wrist bones differs in moose and sheep. in moose, ossification occurs from medial to lateral, and in cattle and sheep it occurs from lateral to medial. hence, it appears that a great load in moose falls on medial bones of the wrist rather than on lateral bones, and in cattle and sheep it is the opposite. differences in patterns of ossification of wrist bones thus depend on characteristics of locomotion, which evolved in connection with different life history requirements. summarizing the results of alimentary tract research in ontogenesis of moose, cattle, sheep, and goats, we drew the following conclusions. stomach and intestine grow irregularly. on the whole they grow rapidly in moose and cattle through the fourth month of fetal development (in sheep, 2–3 months). an intensified growth of the alces suppl. 2, 2002 petrov et al. – morphological changes of organs 107 alimentary tract occurs in moose fetuses but that growth takes place after birth in cattle. growth of the stomach and intestines in moose basically ends at 1.5 months of age whereas in cattle it is still in progress. stomach chambers of these animals grow simultaneously but unequally. the reticulum and omasum of cattle, sheep, and goats grow in the same way as the stomach as a whole, with great intensity in the postnatal period, but the abomasum grows more quickly, especially in volume, during fetal development. only the rumen in moose grows more intensively after birth, but the reticulum and omasum grow rapidly during fetal development. intestines as a whole grow more rapidly during fetal development. the rumen and large intestine are better developed in adult moose in comparison with cattle, sheep, and goats, which in turn have greater development of the abomasum, omasum, and small intestine. endocrine system formation occurs earlier in moose than cattle, sheep, and goats. fetuses aged 2–2.5 months showed formation of thyroid, thymus, and adrenal glands. the first signs of their functioning were detected at 3–3.5 months of fetal age. the first signs of thymus involution – an indicator of sexual maturity – were documented in a 1–month–old moose. alces18_xlviimooseharvestworkshop.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces17_78.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_229.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces18_210.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces15_vattendees.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 f:\alces\supp2\pagema~1\rus13s. alces suppl. 2, 2002 kochetkov moose population dynamics 57 factors determining moose population dynamics in the central forest reserve vitaliy v. kochetkov central forest biosphere reserve, 172513, nelidovo, tver region, russia abstract: we determined the main factors that led to a decrease in the moose population of the central biosphere reserve. the role and importance of the factors in this process were defined. the key role of predation by wolf in the moose population decline is emphasized. the predominant factor leading to the decrease in the moose population was wolf predation, which exerted a pronounced effect on the moose population number and on its age and sex composition. alces supplement 2: 57-61 (2002) key words: moose, population dynamics, predation, wolf review of the literature on ungulate– forest interactions shows some conflicting opinions that will require further detailed study. nevertheless, there are data that allow estimation of the role of this group of animals in natural ecosystems. the appraisal of moose–forest interaction is complex and, to a certain degree, contradictory due to differing points of view of authors with respect to the forest as a whole and to moose in particular (filonov 1983). moose damage plant cover and thus cause changes not only in the structure and productivity of brush and woodland vegetation but also in the composition of leaf litter and properties of soils (pastor et al. 1988). the extent of influence by moose on vegetation depends on moose population density (gatikh 1980) and determines the characteristics of changes in natural forests. this is especially important for the ecology of reserves. study area observations were performed in a 1,000 km2 area in the central forest biosphere reserve, including a wildlife protection area and a sporting zone. methods the data described below were obtained during 1975–1985 in the central forest reserve. several points were confirmed by data of the tver state hunting inspectorate. we also used the reserve’s archives. causes of moose deaths were assigned to one of four categories: (1) wolf predation; (2) brown bear predation; (3) illegal human hunting; and (4) drowning. changes in the composition of the moose population by sex and age caused by wolf predation and human harvest in the study area were estimated on the basis of inspection of wolf–killed prey and human harvest. results and discussion during the study period the moose population decreased from 340/1,000 km2 in 1975 to 66–80/1,000 km2 during 1980–1985. the archives and questioning of inhabitants revealed that the moose number in the study area also fluctuated before this period. moose density during 1917–1919 was 180/ 1,000 km2 in this area. at that time the moose density for the adjacent territories was 300/1,000 km2. due to hunting, the moose population declined and during 1931– 1932 there were only 2 moose in the remoose population dynamics kochetkov alces suppl. 2, 2002 58 serve. the population density started to increase again from 44/1,000 km2 in the winter 1940–1941 to 231/1,000 km2 in 1949– 1950 (yurgenson and yurgenson 1951). the highest moose populations were during 1957–1962; however, a decrease occurred, followed by another increase from the late 1960s to the early 1970s. the population number started to decline again in 1977 (the density in the reserve was 320/1,000 km2 and 274/1,000 km2 in the protection area). in 1978 the density was 280/1,000 km2 in the reserve and 230/1,000 km2 in the protection area, and in 1979 the density was down to 120/1,000 km2 in both the reserve and the protection area. thus, the population dynamics in the territory of the reserve are characterized by increases and decreases of moose numbers and by periods of stabilization at both high and low density. lower moose densities from 1919 to 1931 reflected losses mainly due to anthropogenic factors; from 1977 to 1983 other influences should be taken into account. density reduction was observed everywhere in the tver region. in 1976, 1977, 1978, 1984, and 1985 the moose population numbers were 17,616, 18,080, 19,548, 12,000, and 7,000, respectively (the first 3 values were derived from winter migration observations and the others from aircraft– assisted observations). a decrease in population was observed in most parts of the tver region. in the toropetskii district there were 968 moose in 1971 yet only 382 moose in 1983; at the same time the population decreased 2–fold in the selizharovskii district and 3–fold in the lesnoi district. direct observations and tracking analysis in the territory of the reserve revealed that during june through august (n = 441), on average, there was 0.93 calves per cow and 0% twins in 1965, 0.40 and 0% in 1970, 1.00 and 50% in 1973, 1.00 and 47% in 1974, 0.75 and 33% in 1975, 1.50 and 33% in 1977, 0.90 and 33% in 1978, 1.13 and 47% in 1979, 1.00 and 0% in 1980, 1.00 and 0% in 1981, and 0.75 and 33% in 1982. mean values for cow moose embryo counts and twinning rates in the tver region during 1977–1979, based on the results of licenced hunting, were that the embryo number was 1.32 per cow with calf and 0.78 per mature cow, with a 32% occurrence of twins. in the yaroslavl region the mean values of these parameters during the same period were very close to those mentioned above. in the moscow region, the percentage of twins increased from 16% in 1977 up to 43% in 1980, and the embryo number per cow with calf increased from 1.16 to 1.43 (filonov 1983). we can conclude that moose productivity in the reserve did not decline during 1977–1970, and the decrease in population number was therefore due to some other cause. there were no changes in climate that deviated from normal mean values during the long–term observations. human factors (both direct and indirect influences) were not significant. on average, no more than 10–12 moose were harvested per year (only 2.5–3.0% of the population at the beginning of the biological year). the natural increase of the population was 16.3% in 1975 and 16.4% in 1976. hence, anthropogenic factors could not play a crucial role in the decline of the moose population. during the period under study, abiotic factors did not differ greatly from the mean values for many years. no mass moose migrations were observed. if the factors mentioned above cannot explain the population fluctuations, it is reasonable to consider the role of predators. causes of moose deaths (n = 71) were assigned to one of four categories: (1) wolf predation (79%); (2) brown bear predation (19%); (3) illegal human hunting (1%); and (4) drowning (1%). it was assumed that during 1975–1977 in the study area, the bear harvest was 3–5% and wolf harvest alces suppl. 2, 2002 kochetkov moose population dynamics 59 21% of the population. the number of wolf–killed moose was greater than the natural rate of population increase; thus, wolves crucially affected the moose population. this assumption is corroborated by the fact that in spite of the total decline of the moose population, several areas were characterized by an increase in moose numbers, depending on the pressure of wolves on moose. in the vesyegonskii district the moose population number increased from 825 in 1977 to 830 in 1979, 1,050 in 1981, and 1,105 in 1983. this increase was promoted by a low predator population number; from 1976 to 1982 the wolf density was 5.4– 9.3 per 1,000 km2 (the average density in the entire tver region was 16–22 per 1,000 km2) and the wolf–to–moose ratio was in the range of 1:75–1:58. in the kashinskii district the moose number was 256 in 1976 and 700 from 1981 to 1983; the wolf density was 2–7 per 1,000 km2 from 1978 to 1983 (there were no wolves registered in the district during 1976–1977) and the wolf–to– moose ratio varied from 1:64 to 1:50. changes in the composition of the moose population by sex and age caused by wolf predation and human harvest in the study area were estimated on the basis of inspection of wolf–killed prey and human harvest (table 1). wolf predation was disproportionately heavy on males, whereas human harvest included both sexes equally. calves and old moose (older than 10 years) constituted 53% of the wolf kills but only 6% of the harvest by hunters (fisher criterion f = 20.5 calculated according to zaitsev 1984); moose in the age range of 3.5–7.5 years constituted 23 and 51% of the kills, respectively (f = 12.1). using track size measurements (yazan 1961), we classified the moose population in the summer of 1976 as 23% calves (<0.5 years), 11% yearlings (<1.5 years), and 66% adult (>2.5 years). in the pripyat reserve the age composition from 1971 to 1975 was 21.6% calves and 12.6% yearlings (gatikh 1976). in the berezin reserve from 1959 to 1980, these parameters were 14.8% and 8.8%, respectively (kozlo 1983). in june through august 1957–1977, the calf:yearling:adult ratios for moose were (%) 24:12:64 in the leningrad region, 19:8:73 in the novgorod region, and 28:13:59 in the pskov region (vereshchagin and rusakov 1979). according to teplov‘s data (see yurgenson 1964) for 12 regions of russia in 1961, the percentages in each age group ranged from 20 to 31% for calves, 6 to 16% for yearlings, and 57 to 65% for adult animals. during summer, 25% of the moose population were calves and 12% were yearlings, but in winter calves composed only 12.3% and yearlings 9.0% of the population (vereshchagin and rusakov 1979). based on these findings, it appears there is a clear selection for calves in wolf predation, whereas adult moose are selectively killed by hunters. similar results were obtained for other areas. in the darwin reserve wolves killed mainly young animals (61%) with no selection for sex (kaletskaya 1973). estimation of the proportion of kills by age on isle royale showed that 28% were calves, 21% animals 8–15 years of age, 38% animals 10–18 years of age, and 7% animals 20 years old (mech 1970). according to the literature (yazan 1961, vereshchagin and rusakov 1979, filonov 1983), moose 3.5– 7.5 years old have the highest reproductive activity. thus, in the reserve, most moose preyed upon were of lower reproductive activity. there are differences in the age and sex composition of moose killed by pairs and packs (table 1). males are preyed upon by pairs more frequently; no selection by sex was observed for kills by packs (the difference was not statistically significant). kills by pairs appeared to consist of less productive moose; only 13% of all kills by moose population dynamics kochetkov alces suppl. 2, 2002 60 table 1. composition (%) by sex and age of moose killed by wolves (n = 79) and harvested by man (n = 68) in the central forest reserve. killed sex age (years) b y female male <1 <2 2.5 3.5 <5.5 <7.5 <9.5 >10 hunters 51 49 3 10 25 18 11 22 8 3 wolves 40 60 43 4 10 3 16 4 10 10 in packs 48 52 47 6 16 5 21 5 0 0 in pairs 29 71 35 4 4 0 9 4 22 22 pairs were moose in the age range 3.5–7.5 years compared to 31% for packs, although the difference was not statistically significant. the differences in kills by pairs and packs for moose 7.5–9.5 years of age (f = 15.1) and for the animals older than 10 years (f = 15.1) are statistically significant. using the data on age composition of prey (table 1), biomass values of males and females for each group (yazan 1961), and wolf consumption level and population number, it was calculated that on average a wolf harvests 5–6 moose per year and consumes 4.75–5.75 kg of meat per day. according to observations in the study area, the consumption rate was, on average, 1 moose/7 days for a pack of 5 wolves and 1 moose/5 days for a pack of 7 wolves. assuming that the average mass is 171 kg for a female moose and 228 kg for a male moose, one can calculate that in the first case 4.89 kg of meat were available per wolf per day, and in the second case 6.51 kg per wolf per day, the mean value being 5.7 kg/wolf/day. these approximations are very close to the consumption rates calculated above. thus, the moose population decline may be ascribed to both environmental and human influences, but the predominant factor was wolf predation, which exerted a pronounced effect not only on the moose population number but also on its age and sex composition. references filonov, k. p. 1983. moose. lesnaya promyshlennost, moscow, russia. (in russian). gatikh, v. s. 1976. study of the moose population structure in pripyat polesye. pages 84–88 in biological aspects of development, reconstruction and prot e c t i o n o f t h e a n i m a l w o r l d o f byelorussia. (in russian). . 1980. moose influence on brush and woodland vegetation in byelorussian polesye. pages 137–139 in ungulates of ussr fauna. nauka, moscow, russia. (in russian). kaletskaya, m. l. 1973. wolf and its role as predator in the darwin reserve. proceedings of darwin state reserve 11:41–58. (in russian). kozlo, p.g. 1983. ecological/morphological analysis of moose population. science and technology. minsk, belarus. (in russian). mech, l. d. 1970. the wolf: the ecology and behavior of an endangered species. the natural history press, garden city, new york, usa. pastor, j., r. naiman, b. dewey, and p. mcinnes. 1988. moose, microbes, and the boreal forest. bioscience 38:770– 777. vereshchagin, n. k., and o. s. rusakov. 1979. ungulates from the north–western part of the ussr. nauka, leninalces suppl. 2, 2002 kochetkov moose population dynamics 61 grad, russia. (in russian). yazan, y. p. 1961. biological features and u t i l i z a t i o n o f m i g r a t i n g m o o s e populations in the pechora taiga. proceedings of pechora–ilychskii state reserve 9:114–201. (in russian). yurgenson i. a., and p.b. yurgenson. 1951. ecological review of mammals in central forest state reserve and its environs, (1931–1950). unpublished manuscript. (in russian). yurgenson, p. b. 1964. moose population structure and composition in forest hunting zones. pages 13–34 in biology and harvest of moose. (in russian). zaitsev, g. n. 1984. mathematical statistics in experimental botany. nauka, moscow, russia. (in russian). alces16_374.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 population genetic structure of moose (alces alces) of south-central alaska robert e. wilson1,2, thomas j. mcdonough3, perry s. barboza1, sandra l. talbot4, and sean d. farley5 1university of alaska fairbanks, institute of arctic biology, fairbanks, alaska 99775; 3alaska department of fish and game, 3298 douglas st., homer, alaska 99603; 4u. s. geological survey, alaska science center, 4210 university drive, anchorage, alaska 99508; 5alaska department of fish and game, 333 raspberry road, anchorage, alaska 99518 abstract: the location of a population can influence its genetic structure and diversity by impacting the degree of isolation and connectivity to other populations. populations at range margins are often thought to have less genetic variation and increased genetic structure, and a reduction in genetic diversity can have negative impacts on the health of a population. we explored the genetic diversity and connectivity between 3 peripheral populations of moose (alces alces) with differing potential for connectivity to other areas within interior alaska. populations on the kenai peninsula and from the anchorage region were found to be significantly differentiated (fst = 0.071, p < 0.0001) with lower levels of genetic diversity observed within the kenai population. bayesian analyses employing assignment methodologies uncovered little evidence of contemporary gene flow between anchorage and kenai, suggesting regional isolation. although gene flow outside the peninsula is restricted, high levels of gene flow were detected within the kenai that is explained by male-biased dispersal. furthermore, gene flow estimates differed across time scales on the kenai peninsula which may have been influenced by demographic fluctuations correlated, at least in part, with habitat change. alces vol. 51: 71–86 (2015) key words: alaska, genetic diversity, gene flow, moose, population genetic structure the pattern of geographical variation in genetic diversity and divergence is dictated by the interaction of genetic drift, gene flow, and natural selection (eckert et al. 2008), and these evolutionary processes can be influenced by the location of a population within the species’ geographic range (briggs 1996, wisely et al. 2004, howes and lougheed 2008). at the local and regional scales, the relative location of a population can strongly impact patterns of dispersal and degree of isolation influenced by both historical and contemporary events (vucetich and waite 2003, eckert et al. 2008), ultimately determining the level of genetic structure and diversity. genetic diversity is lowest at the range margins and highest at the center of a species distribution (yamashita and polis 1995, schwartz et al. 2003, eckert et al. 2008, howes and lougheed 2008). marginal populations are more likely to be isolated, occur in patchy habitats, and may reflect recent colonization. peripheral populations are less likely to receive immigrants whereas the core populations typically occupy prime habitat and experience greater levels of gene flow (hoffmann and blows 1994, brown et al. 1995, wisely et al. 2004, miller et al. 2010, schrey et al. 2011). evolutionary theory suggests that the reduction of genetic diversity within peripheral populations can impede adaptation to 2present address: u. s. geological survey, alaska science center, 4210 university drive, anchorage, alaska 99508 71 changing environmental conditions (bradshaw 1991, hoffmann and parsons 1991, hoffmann and blows 1994, blows and hoffmann 2005). such adaptation is largely determined by the availability of additive genetic variation in heritable traits with fitness consequences. several studies have shown that even small changes in genetic variation can have large effects on population fitness (frankham 1995, reed and frankham 2003) including juvenile survival (coulson et al. 1999, mainguy et al. 2009, silva et al. 2009), antler growth (von hardenberg et al. 2007), and parasite resistance (coltman et al. 1999) within ungulates. however, adaptability is also affected by effective dispersal which can have either positive or negative effects on the population depending on the rate of gene flow and the strength of selection acting on the local population (garcía-romos and kirkpatrick 1997, akerman and bürger 2014, bourne et al. 2014, frankham 2015). thus, examining conditions (habitat, genetic diversity, gene flow rates, and life history) under which peripheral populations exist can aid in understanding the processes that maintain geographical ranges, predicting the consequences of climate change (parmesan and yohe 2003, root et al. 2003, hampe and petit 2005), and conserving populations at range margins (howes and lougheed 2008). the kenai peninsula is a peripheral region situated in south-central alaska that was separated from the mainland by a narrow (16 km wide) isthmus at the end of the last ice age. due to its diverse landscape, biodiversity in this region is unusually high at this latitude (morton et al. 2009), and the moose (alces alces) is one of the most recognizable and socio-economically important species. moose populations on the kenai peninsula are characterized by fluctuations in population size, peaking after the occurrence of forest fires that promote optimal forage habitat (oldemeyer et al. 1977). while moose populations on the kenai have fluctuated between 5,000–8,000 animals over the past several decades (t. j. mcdonough, unpublished data), these fluctuations have not been uniform across moose management units on the peninsula. while population size in game management unit (gmu) 15c in southwest kenai has increased, numbers in gmu 15a (northwest) have declined drastically, ∼40% in the last 20 years as quality forage has diminished since the last major fire in 1969. relative isolation from neighboring regions with a strong history of fluctuations in population size might lead to reduced genetic variability on the kenai peninsula which could ultimately be detrimental to the long-term health of moose in this region. using microsatellite loci, we compared levels of genetic variation and gene flow in 2 areas within the kenai peninsula and the anchorage area. these 3 areas are situated on the periphery of overall moose distribution in alaska but differ in levels of potential connectivity to the core area of interior alaska. first, we investigated the connectivity between gmus on the kenai peninsula that have been affected by a long history of land alteration and demographic changes. second, we predicted that 2 sites within the disjunct kenai peninsula region, where opportunities for genetic exchange may be more limiting than anchorage, would exhibit relatively lower genetic diversity. methods sample collection a total of 163 moose were sampled from 3 populations in south-central alaska (fig. 1). ear-plugs and blood were taken from 33 collared female moose in 2008–2010 and 2012 from the city of anchorage and adjacent eagle river (called anchorage hereafter). in addition, muscle tissue was taken from 32 hunter-killed moose (16 female, 15 male, and 1 unknown) during the winter of 72 population genetic structure – wilson et al. alces vol. 51, 2015 2011–2012. in spring 2012, blood was taken from radio-collared female moose from gmu 15a (n = 49; 3,367 km2) and gmu 15c (n = 49; 3,030 km2) on the kenai peninsula, the borders of which are approximately 20 km apart. anchorage samples are archived at the molecular ecology laboratory, u.s. geological survey, anchorage, alaska, and kenai peninsula samples at the alaska department of fish and game, homer, alaska. all animal capturing and genetic sampling were conducted under division of wildlife conservation acuc approval (# 2012–2007, 2013–2021, and 90–05) and under the university of alaska fairbanks iacuc approval (# 14885 and 182744). molecular techniques genomic dnawas extracted from blood and tissue samples using a “salting out” procedure described by medrano et al. (1990), with modifications described in sonsthagen et al. (2004). genomic dna concentrations were quantified using fluorometry and diluted to 50 ng ml–1 working solutions. individuals were initially screened at 17 microsatellite loci. thirteen autosomal loci were found to be polymorphic of which 9 with dinucleotide repeat motifs were selected for further analysis that were polymorphic in all populations: bl42, bm888, bm203, bm2830 (bishop et al. 1994), nvhrt21, nvhrt22 (røed and midthjell 1998), rt1, rt5, and rt30 (wilson et al. 1997). polymerase chain reaction (pcr) amplification and electrophoresis followed protocols described in roffler et al. (2012). ten percent of the samples were amplified and genotyped in duplicate for the 9 microsatellite loci for quality control. analysis of genetic diversity and population genetic subdivision we calculated allelic richness, inbreeding coefficient (fis), observed and expected heterozygosities, and tested for deviations from hardy-weinberg equilibrium (hwe) and linkage disequilibrium (ld) for each microsatellite locus and population in fstat ver. 2.9.3 (goudet 1995). the degree of genetic subdivision among moose populations was assessed by calculating overall and pairwise fst and rst, adjusting for multiple comparisons using bonferroni correction (α = 0.05) in arlequin v3.5.1.3 (excoffier and lischer 2010). because the upper possible fst value for a set of microsatellite loci is usually <1.0 (hedrick 2005), we used recodedata, version 1.0 (meirmans 2006) to calculate the uppermost limit of fst for our data set. we also used the bayesian-clustering program stucture 2.2.3 (pritchard et al. 2000) to determine the level of population structure in the autosomal microsatellite data set. we performed 2 sets of analyses to look at structure within south central alaska: 1) between anchorage and kenai peninsula and 2) within the kenai peninsula (gmu 15a and 15c). structure assigns individuals to populations maximizing hardy-weinberg equilibrium and minimizing linkage disequilibrium. the analysis was conducted for 1– 10 populations (k) using an admixture fig. 1. sampling areas for three moose populations in south-central alaska: anchorage, game management unit (gmu) 15a (northwest kenai peninsula), and gmu15c (southwest kenai peninsula). alces vol. 51, 2015 wilson et al. – population genetic structure 73 model with 100,000 burn-in iterations and 1,000,000 markov chain monte carlo (mcmc) iterations without providing a priori information on the geographic origin of the individuals; the analyses were repeated 10 times for each k to ensure consistency across runs. we used the ▵k method of evanno et al. (2005) and evaluation of the estimate of the posterior probability of the data given k, ln p(d), to determine the most likely number of groups at the uppermost level of population structure. for the kenai peninsula analysis we used the locprior which is able to detect weak signals of population structure in datasets not detectable under standard models (hubisz et al. 2009). we determined if location was informative by the value of r, which parameterizes the amount of information contained by the location of the samples. values of r > 1 indicates either there is no population structure or that structure is independent of locality. gene flow we estimated gene flow between moose populations using 2 methodologies: migrate v3.2.16 (beerli and felsenstein 1999, 2001) and bayesass 3.0 (wilson and rannala 2003). these programs differ in the underlying model used to estimate gene flow. migrate uses a steady-state twoisland coalescent model of population differentiation which incorporates parameters scaled to the mutation rate (µ), the effective population size parameter θ (4neµ), and the migration rate m (m/µ) between populations. bayesass uses an assignment methodology which does not incorporate genealogy or assume that populations are in hardy-weinberg equilibrium (wilson and rannala 2003). thus, estimates of migration rate can be interpreted differently and at different temporal scales. bayesass reflects gene flow over the last several generations (referred to as contemporary gene flow hereafter) whereas migrate gene flow estimates are averaged over the past n generations, where n equals the number of generations the populations have been at mutation-drift equilibrium (beerli and felsenstein 1999, 2001). it is generally agreed that microsatellite mutation rates are several orders of magnitude higher than mutation rates of dna sequences (mitochondrial or nuclear; schlötterer 2000, ellegren 2004). thus, microsatellite markers can reflect recent (within the last 10,000 years) and almost contemporaneous events, but increases in homoplasy associated with microsatellites reduce their ability to capture older demographic events (hartl and clark 2007, hughes 2010). therefore, migrate analyses are referred to as estimating recent gene flow. migrate was run with a full migration model; θ (4neµ, composite measure of effective population size and mutation rate) and all pairwise migration parameters were estimated individually from the data. gene flow was estimated using maximum likelihood search parameters; 10 short chains (5000 trees used out of 1,500,000 sampled), 10 long chains (15,000 trees used out 5,250,000 sampled), and 5 static heated chains (1, 1.33, 2.0, 4.0, and 1,000,000; swapping interval = 1). full models were run 10 times to ensure the convergence of parameter estimates. bayesass was initially run with the default delta values for allelic frequency (p), migration rate (m), and inbreeding (f). subsequent runs incorporated different delta values to ensure that acceptance rate for proposed changes was between 20–40% for each parameter to maximize log likelihood values and ensure the most accurate estimates (wilson and rannala 2003). final delta values used were ▵p = 0.5 (27% acceptance rate), ▵m = 0.2 (27%), and ▵f = 0.85 (31%). we performed 10 independent runs (10 million iterations, 1 million burn-in, 74 population genetic structure – wilson et al. alces vol. 51, 2015 and sampling frequency of 1000) and 2 additional longer runs (50 million iterations, 5 million burn-in) with different random seeds to ensure convergence and consistency across runs. convergence was also assessed by examining the trace file in program tracer v1.5 to ensure proper mixing of parameters (rambaut and drummond 2007). population demography to estimate the effective population size (ne) for each gmu on the kenai peninsula and anchorage area, we used the approximate bayesian computation method (beaumont et al. 2002) implemented in the program onesamp 1.2 (tallmon et al. 2008). we used a lower prior of 100 for all populations and a maximum prior that reflected the current census size (1,000 for anchorage, 2,000 for gmu 15a, and 3,000 for gmu 15c). similar values were obtained for larger maximum possible effective population sizes. lastly, we used bottleneck which compares the number of alleles and gene diversity at polymorphic loci under the infinite allele model (iam; maruyama and fuerst 1985), stepwise mutation model (smm; ohta and kimura 1973), and twophase model of mutation (tpm; di rienzo et al. 1994; parameters: 79% ssm, variance 9; piry et al. 1999, garza and williamson 2001). one thousand simulations were performed for each population and parameters were changed among 5 runs to evaluate the robustness of results. significance was assessed using a wilcoxon sign-rank test which determines if the average of standardized differences between observed and expected heterozygosities is significantly different from zero (cornuet and luikart 1996). significant heterozygote deficiency relative to the number of alleles indicates recent population growth, whereas heterozygote excess relative to the number of alleles indicates a recent population bottleneck (cornuet and luikart 1996). bottleneck compares heterozygote deficiency and excess relative to genetic diversity, not to hardyweinberg equilibrium expectation (cornuet and luikart 1996). results genetic diversity and population subdivision multilocus genotypes were collected from 163 individuals and each individual had a unique genotype. the number of alleles per locus observed ranged from 3.4– 4.7 per population with an overall estimate of 5.1 (tables 1, 2). the observed heterozygosity ranged from 43–55% with an overall mean heterozygosity of 49%. the kenai peninsula exhibited a 19% lower allelic richness (20% in gmu 15a and 25% in gmu 15c) compared to the anchorage area, and 3x more private alleles were observed in the anchorage region (table 1). in addition, the observed (ho) and expected (he) heterozygosity was significantly lower (all p-values < 0.0001) in the kenai peninsula (ho by 18%, he by 16% expected), in gmu 15c (15%, 14%), and in gmu 15a (22%, 20%). on average, individuals on the kenai showed a greater level of homozygosity (kenai: 4.94 loci [sd = 1.46] vs. anchorage: 3.98 loci [sd = 1.51]; t = 4.02, p < 0.0001). the inbreeding coefficient (fis) did not differ significantly from zero in any population (table 1). all loci and populations were in hwe and linkage equilibrium. significant genetic structure was observed at the 9 microsatellite loci between anchorage and the two gmus on the kenai peninsula (table 3). no significant difference was found within the kenai peninsula. the upper limit of the fst for our microsatellite data set was 0.499. therefore, the overall fst of 0.071 accounted for 14.2% of the maximum possible level of genetic structure and 19% for the pairwise alces vol. 51, 2015 wilson et al. – population genetic structure 75 comparisons between anchorage and kenai peninsula gmus. structure uncovered genetic partitioning within south central moose populations, supporting a two-population model (δk = 188.3, average ln p(d) = −2758.7). most individuals from anchorage were assigned to one genetic cluster (87.7%), whereas individuals from kenai gmu 15a and 15c were assigned to a second cluster with high probability, 93.6 and 92.6%, respectively (fig. 2). seven anchorage individuals were assigned to the anchorage cluster with <60% certainty, conversely, only a single kenai individual was assigned to the kenai cluster with <60% certainty. genetic partitioning was not observed within kenai peninsula, as including capture location (lociprior) was not informative (r > 9). gene flow restricted gene flow over the past several generations was observed under the bayesass model between anchorage and kenai peninsula, with 96.8% (93.3–100%) of the anchorage population comprised of a non-migrant origin (fig. 3). within the kenai peninsula, there was a signal of a northern direction of contemporary gene flow from gmu 15c into 15a (proportion of individuals with migrant origin: 27.8% in 15a vs. 6.9% in 15c); although the 95% confidence intervals do overlap (fig. 3). asymmetrical recent gene flow as estimated by migrate was observed among sampled populations. the directionality of gene flow was from kenai peninsula into anchorage (fig. 3). the number of migrants per generation (nem) ranged from 2.56 (gmu 15a; 1.97–3.29) and 2.78 (gmu 15c; 2.20–3.48) into anchorage and 0.99 (0.74–1.32) and 1.08 (0.78–1.47) into the kenai gmu 15a and 15c, respectively. within kenai there was a signal of asymmetrical gene flow from gmu 15a into 15c (3.3 migrants/generation; fig. 3). population demography the estimated effective size using the bayesian computation method for the anchorage region was 74.3 (95% ci: 67.6– 83.0). gmus 15a and 15c on the kenai peninsula had lower estimated effective sizes with non-overlapping confidence intervals with anchorage (table 1). the bottleneck analysis showed no evidence of significant table 1. estimates of genetic diversity of the moose sampled from three locales in south-central alaska, including: average number of alleles, allelic richness (ar), observed and expected heterozygosities (ho/he), inbreeding coefficient (fis), effective population size (ne) estimated in onesamp and sample size (n) calculated from nine microsatellite loci. allelic richness is based on smallest sample size of 65 for anchorage and overall kenai. within kenai peninsula (gmu 15a and gmu 15c) based on sample size of 49. kenai peninsula anchorage gmu 15a gmu 15c overall kenai no. alleles 4.67 3.67 3.44 4.00 no. private alleles 10 1 2 3 ar 4.59 3.67 3.44 3.78 ho (sd) / 0.55 (0.02)/ 0.43 (0.02)/ 0.47 (0.02)/ 0.45 (0.02)/ he (sd) 0.56 (0.06) 0.45 (0.07) 0.48 (0.05) 0.47 (0.06) fis 0.007 0.056 0.031 0.043 ne 74.3 (67.6–83.0) 47.9 (43.2–56.8) 36.8 (33.3–43.9) 145.5 (123.3–255.9) n 65 49 49 98 76 population genetic structure – wilson et al. alces vol. 51, 2015 heterozygosity excess or deficit under the smm or tpm. however, there was evidence of a recent population decline (heterozygote excess) based on the infinite allele model (iam) for kenai gmu 15c. discussion climatic and glaciation history has played a major role in shaping the evolutionary history of many taxa in south-central alaska. it was not until approximately 7,000 years before present that the kenai peninsula became distinct and relatively isolated from the mainland by a 16 km wide mountainous isthmus (pielou 1991, muhs et al. 2001). this isolation has fostered genetically or morphologically distinct populations for a variety of taxa (e.g., wolverine table 2. estimates of observed and expected heterozygosity, number of alleles per locus for nine autosomal nuclear microsatellite loci assayed in three moose populations in south-central alaska. all loci were in hardy-weinberg equilibrium. sample size is in parentheses; ho = heterozygosity observed, he = heterozygosity expected, and na = number of alleles. kenai peninsula locus anchorage (65) gmu 15a (49) gmu 15c (49) overall kenai (98) all populations (165) nvhrt22 ho/he 0.69/0.76 0.49/0.54 0.57/0.53 0.53/0.53 0.60/0.68 na 6 5 4 6 6 nvhrt21 ho/he 0.49/0.50 0.55/0.46 0.39/0.45 0.47/0.45 0.48/0.48 na 5 3 2 3 5 rt1 ho/he 0.49/0.46 0.27/0.29 0.39/0.38 0.31/0.34 0.38/0.40 na 2 2 2 2 2 rt5 ho/he 0.54/0.52 0.18/0.21 0.25/0.32 0.21/0.26 0.34/0.40 na 4 3 3 3 4 rt30 ho/he 0.55/0.58 0.69/0.67 0.74/0.72 0.71/0.70 0.65/0.66 na 5 4 4 4 5 bm203 ho/he 0.20/0.20 0.37/0.41 0.51/0.50 0.44/0.46 0.34/0.38 na 5 3 4 4 6 bm2830 ho/he 0.46/0.49 0.37/0.43 0.41/0.41 0.39/0.42 0.42/0.45 na 2 2 2 2 2 bm888 ho/he 0.63/0.65 0.22/0.26 0.20/0.27 0.21/0.27 0.38/0.46 na 4 4 3 4 4 bl42 ho/he 0.91/0.84 0.71/0.81 0.80/0.75 0.76/0.79 0.82/0.83 na 9 6 7 8 12 overall loci ho/he 0.55/0.56 0.43/0.45 0.47/0.48 0.45/0.47 0.49/0.53 na 4.67 3.67 3.44 4.00 5.11 table 3. pairwise and overall values of fst and rst calculated from nine microsatellite loci. significant values after bonferroni correction (p < 0.0001) are marked with an asterisk. fst rst anchorage – kenai gmu 15a 0.094* 0.014 – kenai gmu 15c 0.092* 0.028 kenai gmu 15a – kenai gmu 15c 0.001 0.000 overall 0.071* 0.016 alces vol. 51, 2015 wilson et al. – population genetic structure 77 [gulo gulo], tomasik and cook 2005; american marten [ursus americanus], robinson et al. 2007; song sparrow [melospiza melodia], patten and pruett 2009). the moose populations residing on the kenai are no exception. using a multi-locus approach, we observed that moose on the kenai were genetically distinct from those in the 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 pr o b ab ili ty o f a ss ig n m en t anchorage kenai gmu 15a kenai gmu 15c fig. 2. structure analysis showing posterior probability of assignment of individuals to each (k = 2) genetic cluster. white bar represents the estimated probability of assignment to cluster one and grey bar is the estimated probability of assignment to cluster two. 0.99 (0.74-1.32) 2.56 (1.97-3.29) 3.29 (2.63-4.07) 1.18 (0.89-1.54) a b 2.78 (2.20-3.48) 1.08 (0.78-1.47) 0.02 (0.00-0.05) 0.01 (0.00-0.03) 0.28 (0.13-0.43) 0.07 (0.00-0.25) 0.01 (0.00-0.02) 0.01 (0.00-0.04) fig. 3. estimates of (a) recent (number of migrants per generation, nem) estimated in migrate and (b) contemporary (proportion of individuals with migrant origin, m) calculated in bayesass for moose populations in south-central alaska as calculated from nine microsatellite loci with relative magnitude indicated by width of arrow; 95% confidence intervals are in parentheses. 78 population genetic structure – wilson et al. alces vol. 51, 2015 mainland anchorage population and exhibited significantly lower levels of genetic diversity at microsatellite loci. loss of genetic diversity between peninsula and mainland populations residing in areas with barriers that limit dispersal (e.g., peninsulas and islands) across the landscape are expected to have lower genetic variation (gaines et al. 1997). our results were consistent with gaines et al. (1997) prediction: moose occupying the kenai peninsula had significantly reduced genetic diversity (∼18%) compared to the nearest mainland population in anchorage. a reduction of genetic variability has also been reported for other alaskan moose populations (hundertmark 2009, schmidt et al. 2009) as well as other mammals on the kenai peninsula (e.g., canada lynx [lynx canadensis], schwartz et al. 2003). the loss of genetic variation in peripheral populations may be due to numerous factors such as limited number of connections to other populations or smaller population size (schwartz et al. 2003). cook inlet waters, mountains, and a highway and railways may represent formidable dispersal barriers for moose between these regions. although kenai peninsula and anchorage are in close geographic proximity (straight line distance over land is ∼105 km), the costs of dispersal over the rugged terrain and highways or swimming across the inlet are likely high. in agreement with limited effective dispersal, we found restricted contemporary gene flow between kenai peninsula and mainland anchorage populations with confidence intervals suggesting there has been limited genetic exchange over the past several generations. telemetry studies of the sampled females in this study showed that individuals remained in the same general area throughout the year (farley et al. 2012, t. j. mcdonough, unpublished data), further suggesting a low likelihood of long-distance dispersal between these two regions. however, connectivity could be mediated through a contact zone north of the isthmus located at portage valley that is used by black bears (ursus americanus) (robinson et al. 2007). the isthmus is not an absolute/strong barrier as movements of radio-collared moose occur across the isthmus; this movement was restricted within intermountain valleys that spanned both sides of the isthmus (t. lohuis, alaska department of fish and game, unpublished data). furthermore, structure analysis estimated a low probability assignment to a genetic cluster for ∼12% of the individuals in anchorage, suggestive of genetic exchange that has occurred during or after population divergence, with higher rate going into the anchorage area based on the migrate analysis. this northward direction of gene flow is also found in other peninsular populations (schmidt et al. 2009) and may reflect post-colonization gene flow rates. further study of moose in areas between anchorage and kenai peninsula might identify if a contact zone exists for moose at the isthmus as seen in other mammals, or if these regions are truly isolated as indicated by the contemporary gene flow analysis. relationships within the peninsula unlike the potential strong barriers to dispersal between the peninsula and mainland populations, there are relatively few natural barriers to movement in the western part of the peninsula, and gene flow estimates suggest that there is ongoing genetic exchange. the directionality of gene flow across the western kenai peninsula has not remained constant over time. differences in directionality across time scales may be attributed to the fluctuating nature of moose population dynamics that is correlated at least in part with habitat change, in particular alces vol. 51, 2015 wilson et al. – population genetic structure 79 in gmu 15a where population size fluctuates with major fire events (oldemeyer et al. 1977, schwartz and franzmann 1989, loranger et al. 1991). the habitat in gmu 15a has changed drastically over the last century after major fires in 1947 and 1969 transformed previously low quality habitat to ideal foraging habitat, which subsequently declined to the current condition (oldemeyer et al. 1977, schwartz and franzmann 1989). if periodic population increase has been sufficiently frequent throughout the history of moose in this area, and dispersal is influenced by population density and habitat quality, we might expect the directionality of gene flow to change over time with more moose dispersing from areas of high productivity into areas of lower density or less preferred habitat as competition for resources increases. indeed, contemporary gene flow estimated in bayesass indicates gene flow from a higher density area (gmu 15c) with better quality habitat into an area characterized by poor habitat conditions and lower density area (15a). moose populations on the kenai peninsula have also fluctuated in size partially due to human activities (land development and forest fires), with changes in habitat potentially affecting fertility and survival of young (klein 1970, franzmann and arneson 1973, schwartz and franzmann 1989, testa and adams 1989). while moose populations initially respond positively to wildfires through the emergence of optimal habitat, populations eventually decline as the habitat changes to late succession (non-optimal forage) vegetation. during the 20 years following the last major fire in gmu 15a (1969), the population has declined by approximately 40%. current and previous assessment of calf survival from this area has identified low calf survival (franzmann et al. 1980, t. j. mcdonough, unpublished data). such a drastic decline in population size coupled with low productivity can negatively impact genetic diversity of a population; this may partially explain the significantly low genetic diversity on the kenai peninsula. a reduction in genetic diversity can lower viability and fecundity (falconer 1981, ralls et al. 1983, frankham 1995, crnokrak and roff 1999), and at the extreme can lead to inbreeding depression; decreased viability and fecundity occur currently on the kenai peninsula (franzmann and schwartz 1985, adf&g 2013, unpublished data). whether lower reproductive rates (twinning rates and calf survival) on the kenai peninsula are correlated solely with genetic variability or are influenced in addition, or solely by environmental factors, is an area for future investigation. conservation implications although the effects of inbreeding depression can diminish over time (lynch 1977), a general loss of genetic diversity can be detrimental over evolutionary time as it may lower the ability of populations to respond to environmental stressors such as novel predators, parasites, or climatic conditions (lacy 1987, quattro and vrijenhoek 1989, leberg 1993). following the recommendations of frankham et al. (2014), all 3 populations fall below the minimum effective population size of 1,000 required to maintain long-term viability. in addition, the gmus on the kenai peninsula, when considered separately, have an effective population size lower than both recent (> 100; frankham et al. 2014) and earlier (> 50; franklin 1980, soulé 1980) recommendations to avoid inbreeding depression. indeed, the kenai peninsula does have a higher inbreeding coefficient (although not significantly different from zero) and higher levels of homozygosity. however, when considering the kenai peninsula as a single population, the effective population exceeds 100 but remains below the threshold for long-term viability. 80 population genetic structure – wilson et al. alces vol. 51, 2015 neutral loci are commonly used to infer evolutionary history of populations and make inferences about overall variation (see howes and lougheed 2008), but it is still unclear whether the trends in putatively neutral loci are reflective of quantitative-trait variation found in genes for physiological, morphological, or life history traits that are likely important for the adaptive potential of populations (merilä and crnokrak 2001, reed and frankham 2001, eckert et al. 2008). therefore, a conclusion that reduced genetic diversity observed at neutral microsatellite markers reflects reduction of diversity in the genome overall is premature. our results showing significant population structure and limited connectivity to outside populations for the kenai peninsula provide a working hypothesis for the potential effects on genetic diversity, which can be tested by assaying both selectively neutral and functional diversity. such studies can provide greater resolution on the processes responsible for the distribution of genetic diversity among moose populations within southcentral alaska. acknowledgements funding was provided by joint base elmendorf-richardson, under the guidance of h. griese, c. garner, d. battle, and r. graham, and u. s. geological survey. extensive field assistance was provided by the military conservation agent program on elmendorf air base. j. crouse, j. selinger, and pilots j. decreft, t. levanger, and j. fieldman, and the kenai national wildlife refuge provided field assistance on the kenai peninsula. z. grauvogel, k. sage, and s. sonsthagen provided laboratory and manuscript advice, while k. hundertmark provided advice on marker selection and g. roffler and e. solomon provided help with maps. any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the u.s. government. references akerman, a., and r. bürger. 2014. the consequences of gene flow for local adaptation and differentiation: a two-locus deme model. journal of mathematical biology 68: 1135–1198. beaumont, m. a., w. zhang, and d. j. balding. 2002. approximate bayesian computation in population genetics. genetics 162: 2025–2035. beerli, p., and j. felsenstein. 1999. maximum-likelihood estimation of migration rates and effective population numbers in two populations using a coalescent approach. genetics 152: 763–773. ———, and ———. 2001. maximum likelihood estimation of a migration matrix and effective population sizes in n subpopulations by using a coalescent approach. proceedings of the national academy of sciences usa 98: 4563–4568. bishop, m. d., s. m. kappes, j. w. keele, r. t. stone, s. l. f. sunden, g. a. hawkins, s. s. toldo, r. fries, m. d. grosz, j. yoo, and c. w. beattie. 1994. a genetic linkage map for cattle. genetics 136: 619–639. blows, m. w., and a. a. hoffmann. 2005. a reassessment of genetic limits to evolutionary change. ecology 86: 1371–1384. bourne, e. c., g. bocedi, j. m. j. travis, r. j. pakeman, r. w. brooker, and k. schiffers. 2014. between migration load and evolutionary rescue: dispersal, adaptation and the response of spatially structured populations to environmental change. proceedings of royal society b 281: 20132795. bradshaw, a. d. 1991. the croonian lecture, 1991: genostasis and the limits to evolution. philosophical transactions of the royal society b: biological sciences 333: 289–305. alces vol. 51, 2015 wilson et al. – population genetic structure 81 briggs, j. c. 1996. biogeography and punctuated equilibrium. biogeographica 72: 151–156. brown, j. h., d. w. mehlman, and g. c. stevens. 1995. spatial variation in abundance. ecology 76: 2028–2043. coltman, d. w., j. g. pilkington, j. a. smith, and j. m. pemberton. 1999. parasite-mediated selection against inbred soay sheep in a free-living, island population. evolution 53: 1259–1267. coulson, t., s. albon, j. slate, and j. pemberton. 1999. microsatellite loci reveal sex-dependent responses to inbreeding and outbreeding in red deer calves. evolution 53: 1951–1960. cornuet, j. m., and g. luikart. 1996. description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. genetics 144: 2001–2014. crnokrak, p., and d. a. roff. 1999. inbreeding depression in the wild. heredity 83: 260–270. di rienzo, a., a. c. peterson, j. c. garza, a. m. valdes, m. slatkin, and n. b. freimer. 1994. mutational processes of simple-sequence repeat loci in human populations. proceedings of the national academy of sciences usa 91: 3166–3170. eckert,c. g., k. e. samis, and s.c. lougheed. 2008. genetic variation across species’ geographical ranges: the central-marginal hypothesis and beyond. molecular ecology 17: 1170–1188. ellegren, h. 2004. microsatellites: simple sequences with complex evolution. nature reviews genetics 5: 435–445. evanno, g., s. regnaut, and j. goudet. 2005. detecting the number of clusters of individuals using the software structure: a simulation study. molecular ecology 14: 2611–2620. excoffier, l., and h. e. l. lischer. 2010. arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under linux and windows. molecular ecology resources 10: 564–567. falconer, d. s. 1981. an introduction to quantitative genetics. longman, london, england. farley, s., p. barboza, h. griese, and c. gardner. 2012. characterization of moose movement patterns, movement of black bears in relation to anthropogenic food sources, and wolf distribution and movement on jber lands, of elmendorf afb and fort richardson ap. alaska department of fish and game report, juneau, alaska, usa. franklin, i. r. 1980. evolutionary change in small populations. pages 135–149 in m. e. soulé and b. a. wilcox, editors. conservation biology: an evolutionaryecological perspective. sinauer associates, sunderland, massachusetts, usa. frankham, r. 1995. inbreeding and conservation: a threshold effect. conservation biology 9: 792–799. ———. 2015. genetic rescues of small inbred populations: meta-analysis reveals large and consistent benefits of gene flow. molecular ecology doi: http:// 10.1111/mec.13139. ———, c. j. a. bradshaw, and b. w. brook. 2014. genetics in conservation management: revised recommendations for the 50/500 rules, red list criteria and population viability analyses. biological conservation 170: 56–63. franzmann, a.w., and p. d. arneson. 1973. moose research center studies. alaska department of fish and game report, soldotna, alaska, usa. ———, and c. c. schwartz. 1985. moose twinning rates: a possible population condition assessment. journal of wildlife management 49: 394–396. ———, ———, and r. o. peterson. 1980. moose calf mortality in summer on the kenai peninsula, alaska. journal of wildlife management 44: 764–768. gaines, m. s., j. e. diffendorfer, r. h. tamarin, and t. s. whittam. 1997. the 82 population genetic structure – wilson et al. alces vol. 51, 2015 effects of habitat fragmentation on the genetic structure of small mammal populations. journal of heredity 88: 294–304. garcía-ramos, g., and m. kirkpatrick. 1997. genetic models of adaptation and gene flow in peripheral populations. evolution 51: 21–28. garza, j. c., and e. g. williamson. 2001. detection of reduction in population size using data from microsatellite loci. molecular ecology 10: 305–318. goudet, j. 1995. fstat (version 1.2): a computer program to calculate f-statistics. journal of heredity 86: 485–486. hampe, a., and r. j. petit. 2005. conserving biodiversity under climate change: the rear edge matters. ecology letters 8: 461–467. hartl, d. l., and a. g. clark. 2007. principle of population genetics, 4th edition. sinauer associates, sunderland, massachusetts, usa. hedrick, p. w. 2005. a standardized genetic differentiation measure. evolution 59: 1633–1638. hoffmann, a. a., and m. w. blows. 1994. species borders: ecological and evolutionary perspectives. trends in ecology and evolution 9: 223–227. ———, and p. a. parsons. 1991. evolutionary genetics and environmental stress. oxford university press, oxford, england. howes, b. j., and s. c. lougheed. 2008. genetic diversity across the range of temperate lizard. journal of biogeography 35: 1269–1278. hubisz, m. j., d. falush, m. stephens, and j. k. pritchard. 2009. inferring weak population structure with the assistance of sample group information. molecular ecology resources 9: 1322–1332. hughes, a. l. 2010. reduced microsatellite heterozygosity in island endemics supports the role of long-term effective population size in avian microsatellite diversity. genetica 138: 1271–1276. hundertmark, k. 2009. reduced genetic diversity in two introduced and isolated moose populations in alaska. alces 45: 137–142. klein, d. r. 1970. food selection by north american deer and their response to over-utilization of preferred plant species. pages 25–46 in a. watson, editor. animal populations in relation to their food sources. british ecological society symposium 10. blackwell, oxford, england. lacy, r. c. 1987. loss of genetic diversity from managed populations: interacting effects of drift, mutation, immigration, selection, and population subdivision. conservation biology 1: 143–158. leberg, p. 1993. strategies for population reintroduction: effects of genetic variability on population growth and size. conservation biology 7: 194–199. loranger, a. j., t. n. bailey, and w. w. larned. 1991. effects of forest succession after fire in moose wintering habitats on the kenai peninsula, alaska. alces 27: 100–109. lynch, c. b. 1977. inbreeding effects upon animals derived from a wild population of mus musculus. evolution 31: 526–537. mainguy, j., s. d. côté, and d. w. coltman. 2009. multilocus heterozygosity, parental relatedness and individual fitness components in a wild mountain goat, oreamnus americanus population. molecular ecology 18: 2297–2306. maruyama, t., and p. a. fuerst. 1985. population bottlenecks and non-equilibrium models in population genetics. ii. number of alleles in a small population that was formed by a recent bottleneck. genetics 111: 675–689. medrano, j. f., e. aasen, and l. sharrow. 1990. dna extraction from nucleated red blood cells. biotechniques 8: 43. meirmans, p. g. 2006. using the amova framework to estimate a standardized alces vol. 51, 2015 wilson et al. – population genetic structure 83 genetic differentiation measure. evolution 60: 2399–2402. merilä, j., and p. crnokrak. 2001. comparison of genetic differentiation at marker loci and quantitative traits. journal of evolutionary biology 14: 892–903. miller, m. j., e. bermingham, j. klicka, p. escalante, and k. winker. 2010. neotropical birds show a humped distribution of within-population genetic diversity along a latitudinal transect. ecological letters 13: 576–586. morton, j. m., m. bower, e. berg, d. magness, and t. eskelin. 2009. long term ecological monitoring program on the kenai national wildlife refuge, alaska: an fia adjunct inventory. pages 1–17 in w. mcwilliams, g. moisen, and r. czapiewski, compilers. proceedings of the forest inventory and analysis (fia)symposium 2008. rmrsp-56cd. usda forest service, rocky mountain research station, fort collins, colorado, usa. muhs, d., t. a. ager, and j. e. beget. 2001. vegetation and paleoclimate of the last interglacial period, central alaska. quaternary science reviews 20: 41–61. ohta, t., and m. kimura. 1973. a model of mutation appropriate to estimate the number of electrophoretically detectable alleles in a finite population. genetic research 22: 201–204. oldemeyer, j. l., a. w. franzmann, a. l. brundage, p. d. arneson, and a. flynn. 1977. browse quality and the kenai moose population. journal of wildlife management 41: 533–542. parmesan, c., and g. yohe. 2003. a globally coherent fingerprint of climate change impacts across natural systems. nature 421: 37–42. patten, m. a., and c. l. pruett. 2009. the song sparrow, melospiza melodia, as a ring species: patterns of geographic variation, a revision of subspecies, and implications for speciation. systematics and biodiversity 7: 33–62. pielou, e. c. 1991. after the ice age: the return of life to glaciated north america. university of chicago press, chicago, illinois, usa. piry, s., g. luikart, and j. m. cornuet. 1999. bottleneck: a computer program for detecting recent reductions in the effective population size using allele frequency data. journal of heredity 90: 502–503. pritchard, j. k., m. stephens, and p. donnelly. 2000. inference of population structure using multilocus genotype data. genetics 155: 945–959. quattro, j. m., and r. c. vrijenhoek. 1989. fitness differences among remnant populations of the sonoran topminnow, poeciliopsis occidentalis. science 245: 976–978. ralls, k., j. ballou, and r. l. brownell jr. 1983. genetic diversity in california sea otters: theoretical considerations and management implications. biological conservation 25: 209–232. rambaut, a., and a. j. drummond. 2007. tracer v1.4. < http://beast.bio.ed.ac.uk/ tracer> (accessed august 2013). reed, d. h., and r. frankham. 2001. how closely correlated are molecular and quantitative measures of genetic variation? a meta-analysis. evolution 55: 1095–1103. ———, and ———. 2003. correlation between fitness and genetic diversity. conservation biology 17: 230–237. robinson, s. j., l. p. waits, and i. d. martin. 2007. evaluating population structure of black bears on the kenai peninsula using mitochondrial and nuclear dna analyses. journal of mammalogy 88: 1288–1299. røed, k. h., and l. midthjell. 1998. microsatellites in reindeer, rangifer tarandus, and their use in other cervids. molecular ecology 7: 1773–1776. roffler, g. h., l. g. adams, s. l. talbot, g. k. sage, and b. w. dale. 2012. range overlap and individual movements 84 population genetic structure – wilson et al. alces vol. 51, 2015 http://beast.bio.ed.ac.uk/tracer http://beast.bio.ed.ac.uk/tracer during breeding season influence genetic relationships of caribou herds in southcentral alaska. journal of mammalogy 95: 1318–1330. root, t. l., j. t. price, k. r. hall, s. h. schneider, c. rosenzweig, and j. a. pounds. 2003. fingerprints of global warming on wild animals and plants. nature 421: 57–60. schlötterer, c. 2000. evolutionary dynamics of microsatellite dna. chromosoma 109: 365–371. schmidt, j. i., k. j. hundertmark, r. t. bowyer, and k. g. mccracken. 2009. population structure and genetic diversity of moose in alaska. journal of heredity 100: 170–180. schrey, a. w., m. grispo, m. awad, m. b. cook, e. d. mccoy, h. r. mushinsky, t. albaryrak, s. bensch, t. burke, l. k. butler, r. dor, h. b. fokidis, h. jensen, t. imboma, m. m. kesslerrios, a. marzal, i. r. k. stewart, h. westerdahl, d. f. westneat, p. zehtindjiev, and l. b. martin. 2011. broad-scale latitudinal patterns of genetic diversity among native european and introduced house sparrow (passer domesticus) populations. molecular ecology 20: 1133–1143. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose and forest succession, some management considerations on the kenai peninsula, alaska. alces 25: 1–10. schwartz, m. k., l. s. mills, y. ortega, l. f. ruggiero, and f. w. allendorf. 2003. landscape location effects genetic variation of canada lynx (lynx canadensis). molecular ecology 12: 1807–1816. silva, a. d., j.-m. gaillard, n. g. yoccoz, a. j. m. hewison, m. galan, t. coulson, d. allaine, l. vial, d. delorme, g. van laere, f. klein, and g. luikart. 2009. heterozygosity-fitness correlations revealed by neutral and candidate gene markers in roe deer from a long-term study. evolution 63: 403–417. sonsthagen, s. a., s. l. talbot, and c. m. white. 2004. gene flow and genetic characterization of northern goshawks breeding in utah. condor 106: 826–836. soulé, m. e. 1980. thresholds for survival: maintaining fitness and evolutionary potential. pages 151–169 in m. e. soulé and b. a. wilcox, editors. conservation biology: an evolutionary-ecological perspective. sinauer associates, sunderland, massachusetts, usa. tallmon, d. a., a. koyuk, g. h. luikart, and m. a. beaumont. 2008. onesamp: a program to estimate effective population size using approximate bayesian computation. molecular ecology resources 8: 299–301. testa, j. w., and g. p. adams. 1989. body condition and adjustments to reproductive effort in female moose (alces alces). journal of mammalogy 79: 1345–1354. tomasik, e., and j. a. cook. 2005. mitochondrial phylogeography and conservation genetics of wolverine (gulo gulo) of northwestern north america. journal of mammalogy 86: 386–396. von hardenberg, a., b. bassano, m. festabianchet, g. luikart, p. lanfranchi, and d. coltman. 2007. age-dependent genetic effects on a secondary sexual trait in male alpine ibex, capra ibex. molecular ecology 16: 1969–1980. vucetich, j. a., and t. a. waite. 2003. spatial patterns of demography and genetic processes across the species’ range: null hypotheses for landscape conservation genetics. conservation genetics 4: 639–645. wilson, g. a., and b. rannala. 2003. bayesian inference of recent migration rates using multilocus genotypes. genetics 163: 1177–1191. ———, c. strobeck, l. wu, and j. w. coffin. 1997. characterization of microsatellite loci in caribou rangifer tarandus, and their use in other artiodactyls. molecular ecology 6: 697–699. alces vol. 51, 2015 wilson et al. – population genetic structure 85 wisely, s. m., s. w. buskirk, g. a. russell, k. b. aubry, and w. j. zielinksi. 2004. genetic diversity and structure of the fisher (martes pennanti) in a peninsular and peripheral metapopulation. journal of mammaology 85: 640–648. yamashita, t., and g. a. polis. 1995. a test of the central-marginal model using sand scorpion populations (paruroctonus mesaensis, vaejovidae). journal of arachnology 23: 60–64. 86 population genetic structure – wilson et al. alces vol. 51, 2015 population genetic structure of moose (alces alces) of south-entral alaska methods sample collection molecular techniques analysis of genetic diversity and population genetic subdivision gene flow population demography results genetic diversity and population subdivision gene flow population demography discussion loss of genetic diversity between peninsula and mainland relationships within the peninsula conservation implications acknowledgements references alces16_171.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_106.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces14_prefaceiii.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces16_463.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_289.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_editorialcommittee.pdf alces vol. 16, 1980 alces14_32.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alcessupp1_65.pdf alces14_126.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alcessupp1_101.pdf alcessupp1_127.pdf alcessupp1_16.pdf alces16_549.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces15_303.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces20_283.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces19_148.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces19_3.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alcessupp1_156.pdf alces14_209.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces20_187.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alcessupp1_213abstracts.pdf using aerial survey observations to identify winter habitat use of moose in northern maine haley a. andreozzi1, peter j. pekins1, and lee e. kantar2 1department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa; 2maine department of inland fisheries and wildlife, bangor, maine 04401, usa abstract: winter habitat use by moose (alces alces) is typically comprised of regenerating forest and softwood cover in the northeastern united states, and globally, high winter densities are of concern relative to forest damage. habitat variables associated with winter locations of moose collected during aerial surveys in maine in 2011 and 2012 were compared to available habitat at multiple landscape scales. mixed forest was the most used land cover type at both the location and 5 ha scales (35.1% and 31.3%, respectively). although regenerating forest habitat was used only in proportion to availability, the proximity to recent clearcuts, light partial cuts, and heavy partial cuts was an important predictor of moose location. the used proportion of coarse habitat variables (i.e., mature and regenerating forest) were similar to those available in each aerial survey block, indicating that heterogeneous and productive moose habitat is widely available across the commercial forest landscape of northern maine. moose locations derived from aerial surveys can provide insight about spatial distribution and habitat use across the landscape, identify local density in areas where forest regeneration is of concern, and monitor population responses to commercial forest management practices. alces vol. 52: 41–53 (2016) key words: aerial surveys, alces alces, winter, habitat use, moose, maine moose (alces alces) exhibit patterns of habitat use that indicate generalist behavior, but often have seasonal preference for specific habitat variables. peek (1997) considered moose “selective generalists” due to their selective use of certain habitats when seasonally advantageous. habitat selection in all seasons is primarily driven by food abundance and quality (vivas and saether 1987), and access to adequate thermal cover (karns 1997, dussault and ouellet 2004). in northern new england, commercial timber harvesting typically provides heterogeneous forests with stands of varying age that provide high quality forage and cover for moose (leptich and gilbert 1989, scarpitti et al. 2005). although moose are reasonably mobile in typical winter conditions in northern new england, habitat use is influenced by weather, snow depth, forage availability, and cover. moose minimize energy expenditure and reduce home range in winter (peek 1997, renecker and schwartz 1997b), indirect evidence of the importance of winter habitat rela‐ tive to individual and population productivity. at the fine scale, cut/regeneration habitat is used more than other habitat types during winter in new hampshire, presumably because of high forage availability and preference (scarpitti 2006). the most significant landscape characteristic influencing winter locations in northeast vermont was proximity to forest openings/timber cuts that presumably 41 provide important seasonal browse (millette et al. 2014). areas where moose concen‐ trate habitually in high seasonal density are often associated with forest damage globally (heikkila et al. 2003). for example, high winter densities of moose were associated with heavy browsing, limited growth, and regeneration of birch (betula spp.) in newfoundland (bergerud and manuel 1968), and low regeneration in specific cutover sites adjacent to traditional wintering areas in new hampshire (bergeron et al. 2011). it was predicted that such sites were shifting from hardwood to coniferous dominance in both northeast vermont (andreozzi et al. 2014) and northern new hampshire (bergeron et al. 2011). winter aerial surveys conducted by the maine department of inland fisheries and wildlife (mdifw) in 2011 and 2012 measured moose abundance in specific northern wildlife management districts (wmd) with presumed high moose density (kantar and cumberland 2013), and additional surveys determined sex-age composition. each observation had an associated gps location providing the ability to identify and assess habitat use by moose during the survey period. while habitat use patterns are generally known for moose throughout their range, and specifically in new hampshire (miller 1989, scarpitti et al. 2005, scarpitti 2006) and maine (leptich and gilbert 1989, thompson et al. 1995), it is important to con‐ tinually examine how these patterns are ex‐ pressed on a local scale and respond to habitat (forest) change (peek 1997). identifying the seasonal habitat use of moose should pro‐ vide information on the relative proximity and dispersion of forage and cover resources (hundertmark 1997) and provide regional insight about the relationship between forest harvest practices and moose populations. this gis analysis was conducted to measure habitat and landscape characteristics associated with locations of moose observed during winter aerial surveys in northern maine. millette et al. (2014), who originally measured moose abundance, used a similar approach to identify winter habitat use relationships in northern vermont. the primary objectives were to identify the habitat type associated with locations, determine if locations were random relative to habitat availability, and identify land cover characteristics related to locations. study area the study area encompassed those wmds flown in each survey year: wmds 2, 3 and 6 were flown in winters 2010 and 2011, and wmds 1, 2, 3, 4, 5, 8, 11 and 19 in winters 2011 and 2012 (fig. 1). the survey area totaled ~32,950 km2 and included aroostook county and northern portions of adjacent franklin, hancock, penobscot, piscataquis, somerset, and washington counties which are dominated by commercial forests comprised primarily of spruce (picea spp.), balsam fir (abies balsamea), northern white cedar (thuja occidentalis), and white pine (pinus strobus), with mixed hardwoods of aspen (populus spp.), birch (betula spp.), beech (fagus grandifolia), and maple (acer spp.) (kantar and cumberland 2013). the forest composition in each wmd was described by 7 forest habitat variables (maine office of geographic information system 2004), and each survey block was representative of the proportional availability of the 7 habitat variables within a wmd. most blocks were dominated by uncut (>50% combined) and cut (>20%, various treatments) forest; recent cuts and regenerating habitat were available in all survey blocks (table 1; l. kantar, unpublished data). methods the wmds with highest moose den‐ sity (based on hunter sighting rates and highest harvest rates and permit allocations) were prioritized for the aerial surveys except 42 winter habitat use – andreozzi et al. alces vol. 52, 2016 wmd 11 which was surveyed to evaluate the reliability of the survey technique at lower density. survey blocks were 15 × 24 km rectangles selected by assessing the propor‐ tion of habitat variables within each survey block, and prioritizing the block that was most representative of the overall habitat in the wmd. the double-count survey occurred when moose mobility was unrestricted (snow depths <61 cm), ambient temperature was relatively cold (<−12 °c), and no obvious group‐ing was evident (kantar and cumberland 2013); the same conditions were met during the composition surveys. moose locations (n = 481; ≥1 moose/location) were acquired during abundance surveys in 2011 fig. 1. maine wildlife management districts (shaded) used for double-count aerial surveys and sexage composition surveys during winters 2011 and 2012, northern maine, usa. alces vol. 52, 2016 andreozzi et al. – winter habitat use 43 (28 january – 1 february) and 2012 (13 december 2011 – 8 february 2012), and composition surveys in 2012 (13 december 2011 – 3 february 2012) (table 2). the gps coordinates were collected at each observation location; the number of moose at each location ranged from 1 (n = 215, 45%) to 16 (n = 1), with groups >5 restricted to composition surveys. habitat use all gps locations were defined as used locations and mapped in arcgis (esri 2010) to identify habitat characteristics in a useavailability analysis. an equal number of random points were generated using the “generate random points” tool (esri 2011) to represent available locations within each flight survey block. because most moose were moving (disturbed) from the helicopter, and to evaluate a reasonable spatial scale of habitat use, a circular buffer (4.9 ha) was placed around each used and available location as a conservative estimate of diurnal habitat use; this buffer also accounted for any gps error. the ~5 ha scale is representative of a circular polygon with a radius of 125 m that was the average distance moved by moose in a 2-hour period in northwest wyoming (becker 2008). land cover types were identified using the maine landcover dataset 2004 (melcd: table 1. forest composition (%) of survey blocks within wildlife management districts (wmd) flown in double-count and age-sex composition aerial moose surveys during winters 2011 and 2012, northern maine, usa. wmd mixed forest (%) deciduous forest (%) coniferous forest (%) partial cuts (%) recent cuts/ regenerating forest/ scrub-shrub (%) wetland (%) crops/ grassland (%) 1 21.7 10.4 35.9 21.4 6.7 4.0 0.0 2 40.7 17.8 14.1 12.9 9.9 4.5 0.1 3 30.5 18.1 23.9 3.3 10.3 7.1 6.9 4* 17.4 16.3 21.7 18.0 17.6 8.9 0.1 4* 15.5 25.8 14.6 20.3 16.9 6.7 0.1 5 42.2 4.9 23.3 15.6 7.8 6.0 0.1 6 33.0 12.0 20.0 4.7 4.0 10.0 15.6 8 15.2 21.0 29.7 20.6 10.9 2.6 0.1 11 42.8 7.0 20.0 15.4 4.1 9.5 1.2 19 30.4 6.8 34.0 11.3 7.6 9.6 0.4 table 2. survey dates and number of locations collected during aerial double-count and composition count surveys by wmd during winters 2011 and 2012, northern maine, usa. date wmd locations (n) double-count 28-jan-11 2 33 31-jan-11 3 24 1-feb-11 6 13 13-dec-11 2 27 8-jan-12 5 11 9-jan-12 4 40 11-jan-12 1 26 22-jan-12 19 21 26-jan-12 8 17 2-feb-12 4 31 8-feb-12 11 4 composition count 13-dec-11 2 66 28-dec-11 3 63 22-dec-11 4 55 3-feb-12 8 50 44 winter habitat use – andreozzi et al. alces vol. 52, 2016 maine office of geographic information system 2004), and applied to used and available units at all spatial scales. cover types that were not utilized or did not occur in the study area or flight paths were not used in analysis. relevant cover types were aggregated into 7 habitat variables previously used in the selection of survey blocks for the aerial surveys: 1) mixed forest, 2) deciduous forest, 3) coniferous forest, 4) partial cuts, 5) recent clearcuts/regenerating forest/scrub-shrub, 6) wetlands, and 7) crops/grasslands (kantar and cumberland 2013). additionally, recent clearcuts, partial cuts, regenerating forest, and scrub-shrub were analyzed as a combined variable to reflect an overall regenerating land cover class providing typical winter browse. a separate habitat variable was created combining recent clearcuts, heavy partial cuts, and light partial cuts to evaluate the proximity of used and available units to forest cuts in general. used and available units were also analyzed for proximity to mature conifer, using the coniferous forest land cover classification from melcd. national elevation data (ned) from the u.s. geological survey was used to assess elevation, slope, and aspect. statistical analysis general linear mixed model (glmm) analysis was performed using jmp software (sas institute, cary, north carolina, usa) to identify individual habitat variables that differed between used and available units at all landscape scales. these individual hypothesis tests were used to inform variable selection for use in eventual model selec‐ tion under an information-theoretic approach (anderson et al. 2001). land cover classes, elevation, slope, aspect, proximity to cuts, and proximity to mature conifer were treated as fixed-effects for individual analyses; wmd was treated as a random effect in all analyses to remove variation due to habitat differences within wmds. habitat variables with significant difference (p<0.05) between used and available units were used as inputs for model selection. model selection was performed with a mixed effects logistic regression model using r statistical software (r development core team 2013) using the lme4 package (bates et al. 2012); this analysis was used to identify those combinations of habitat vari‐ ables that most influence moose presence. significant habitat variables from the individual glmm analyses were treated as fixed effects; wmd was treated as a random effect in all models. model comparisons were made using the akaike information criterion (aicc) scores; top competing models were those with δaicc <2.0 and the best fitting model was determined by identifying the model with the lowest aicc score and highest akaike weight (burham and anderson 2002). model parameter coefficients were averaged for top competing models (i.e., δaicc <2.0) using the mumin package (barton 2013) in r. significance values were not reported for model parameter coefficients as they are considered inappropriate when using the information-theoretic approach (anderson et al. 2001). results are presented throughout as x̄ ± se. results habitat composition of used units was dominated (~95%) by 5 habitat variables that were similar (0–4% different) at the location and 5 ha scale. the primary composition at the location scale was 35.1% mixed forest, 19.1% deciduous forest, 14.5% coniferous forest, 15.2% partial cuts, and 11.8% recent cuts/regenerating forest/scrub-shrub. similarly, habitat composition at the 5 ha scale was 31.3% mixed forest, 20.7% deciduous forest, 17.56% partial cuts, 14.5% coniferous forest, and 11.4% recent cuts/regenerating forest/scrub-shrub. significant differences were found between used and available units at locations alces vol. 52, 2016 andreozzi et al. – winter habitat use 45 and the 5 ha scale in the individual glmm analyses of habitat variables. used locations included more deciduous forest (7.3%, f = 8.92, p = 0.003) than at available locations; conversely, wetlands (3.3%, f = 5.64, p = 0.018), crops/grassland (2.1%, f = 7.64, p = 0.006), and coniferous forest (3.7%, f = 3.11, p > 0.05) were less common at used than available locations (fig. 2). at the 5 ha scale, used areas included more deciduous forest (6.3%, f = 8.18, p = 0.004) and partial cuts (4.1%, f = 4.50, p = 0.034) than in available areas; coniferous forest (3.7%, f = 4.58, p = 0.033), wetlands (2.2%, f = 4.56, p = 0.033), and crops/grassland (1.9%, f = 9.85, p = 0.002; fig. 2) were used less than available. there was no detectable difference (p > 0.05) between used and available units in the combined regener‐ ating habitat variable (recent clearcuts, partial cuts, regenerating forest, and scrubshrub) at either scale. similarly, there was no detectable difference in mixed forests that represented the largest proportion of used and available units at both scales (28.8–35.1%; fig. 2). used locations were in closer proximity to cuts (299.4 ± 66.8 m) than available locations (410.4 ± 66.8 m, p < 0.0001); similarly, at the 5 ha scale (p < 0.0001; fig. 3) used units were closer to cuts (215.1 ± 62.2 m) than available units (319.1 ± 62.2 m). there was no detectable difference (p > 0.05) in proximity to mature conifer between used and available units at either scale. elevation was higher at used (291.6 ± 39.4 m) than available locations (280.1 ± 39.4 m; p = 0.012), and likewise at the 5 ha scale (291.3 ± 39.6 m vs. 280.1 ± 39.6 m, p = 0.014). directional aspect was fig. 2. proportion (%) of cover types within used (u) and available (a) units for locations and the 5 ha landscape scale during winters 2011 and 2012 in northern maine, usa. units are starred (*) that are different (p < 0.05) within each cover type at the location and the 5 ha scale. 46 winter habitat use – andreozzi et al. alces vol. 52, 2016 not different (p > 0.05) at either scale, with the exception of northeast-facing slopes used less than available at locations (4.0%, f = 3.86, p = 0.049). flat aspects accounted for <3% of available units and had no data points at used units; this aspect class was removed from the analysis. there was no detectable difference in slope (p > 0.05) at either scale. the habitat parameters used in the logistic regression mixed effects models were deciduous forest, coniferous forest, wetlands, distance to cuts, and elevation. the model that best explained (lowest aicc score) moose presence included deciduous forest, distance to cut, and wetlands at both the location and 5 ha scales (table 3). specifically, locations were most influenced by a higher proportion of deciduous forest (β = 0.516, se = 0.179), shorter distance to cuts (β = −0.269, se = 0.072), and smaller proportion of wetlands (β = −0.596, se = 0.318); likewise, at the 5 ha scale used areas were most influenced by deciduous forest (β = 0.180, se = 0.066), distance to cuts (β = −0.197, se = 0.054), and wetlands (β = −0.107, se = 0.069; table 3). models with δaicc <2.0 also included smaller proportions of coniferous forest and elevation at both the location (β = −0.209, se = 0.184 and β = −0.029, se = 0.071, respectively) and 5 ha scales (β = −0.086, se = 0.069 and β = −0.044, se = 0.072, respectively; table 3). discussion the collection of accurate moose locations during winter aerial surveys in maine resulted in a robust dataset that, while time-specific, was efficient, relatively cheap compared to long-term collaring efforts, and repeatable. the number (n = 481) of locations (i.e., moose) and 5 ha areas analyzed in this study over a ~2 month time period was reasonable when compared to traditional studies. thompson et al. (1995) assessed winter habitat use of cow (n = 10) and bull (n = 4) moose in maine with a seasonal mean of 5.8 and 5.4 observations, respectively. in new hampshire, scarpitti (2006) evaluated seasonal habitat use of cow moose using 42 and 54 core areas (2.6–3.7 km2) in early and late winter, respectively. this modeling exercise indicated that proximity to regenerating forests in the form of recent clearcuts, light partial cuts, or heavy partial cuts is an important predictor of the location of moose during winter in northern maine (table 3, fig. 3). in previous research, 87% of winter observations in maine were in areas that had been logged within 10–30 years (thompson et al. 1995). similarly, cut/regeneration habitat was used more than expected in early winter and dictated habitat use at the fine scale in new hampshire (scarpitti 2006), regenerating stands were used more than available in early winter in massachusetts (wattles 2011), and moose locations were influenced by the relative distance to forest openings associated with timber harvest in vermont (millette et al. 2014). unlike in summer when high quality forage is available fig. 3. mean distance (m) to cuts of used and available units at locations and the 5 ha landscape scale during winters 2011 and 2012, northern maine, usa. distance was less (p < 0.05) at both used scales; bars reflect the se. alces vol. 52, 2016 andreozzi et al. – winter habitat use 47 in more habitat types (scarpitti 2006), regenerating forests are preferentially used in winter because concentrated, abundant browse allows moose to forage efficiently (belovsky 1981). while the distance to cut was shorter for used than available units at both the location and 5 ha scales (fig. 3), the combination of habitat variables reflecting regenerating forest habitat (recent clearcuts, partial cuts, regenerating forest, and scrub-shrub) was used in proportion to availability at both scales. it is possible that partial cuts have a shorter distance to edge that provides both browse and cover in closer proximity, and therefore are more influential in moose use. for example, moose in ontario showed preference for edge provided by strips (100– 200 m) of uncut timber over locations within table 3. the total number of parameters (k), log likelihood statistic (loglik), aicc score, delta aicc, and model weight for top competing location and 5 ha landscape scale models (i.e., delta aicc scores <2), and the estimates and standard error (se) for the model-averaged coefficients. locations model selection based on aicc k loglik aicc delta weight deciduous + distance to cut + wetlands 5 −652.41 1314.87 0 0.40 coniferous + deciduous + distance to cut + wetlands 6 −651.76 1315.61 0.73 0.28 deciduous + distance to cut 4 −654.26 1316.55 1.68 0.17 deciduous + distance to cut + elevation + wetlands 6 −652.32 1316.73 1.86 0.16 model-averaged coefficients estimate se (intercept) −0.060 0.081 deciduous 0.516 0.179 distance to cut −0.269 0.072 wetlands −0.610 0.318 coniferous −0.209 0.184 elevation −0.029 0.071 5 ha landscape scale model selection based on aicc k loglik aicc delta weight deciduous + distance to cut + wetlands 5 −653.79 1317.65 0 0.24 deciduous + distance to cut 4 −654.96 1317.96 0.31 0.21 coniferous + deciduous + distance to cut + wetlands 6 −652.96 1318.00 0.36 0.20 coniferous + deciduous + distance to cut 5 −654.27 1318.60 0.95 0.15 deciduous + distance to cut + elevation + wetlands 6 −653.64 1319.36 1.72 0.10 coniferous + deciduous + distance to cut + elevation + wetlands 7 −652.73 1319.57 1.92 0.09 model-averaged coefficients estimate se (intercept) −0.032 0.066 deciduous 0.180 0.070 distance to cut −0.197 0.054 wetlands −0.107 0.069 coniferous −0.086 0.069 elevation −0.044 0.072 48 winter habitat use – andreozzi et al. alces vol. 52, 2016 clearcuts without edge (mastenbrook and cumming 1989). however, caution should be taken in examining narrowly-defined habitat variables in use:availability analysis. variables don’t necessarily describe behavioral recognition or choice, and importance could reflect high/low availability and not absolute use. for example, moose may seem to be specialists under certain variables (i.e., distance to cut) and generalists under others, particularly as they are combined or made coarser (i.e., regenerating/foraging habitat). additionally, high availability of a habitat can mask the importance of its use; for example, despite being used in proportion to its availability, mixed forest was the most used land cover type at both the location and 5 ha scales (35.1% and 31.3%, respectively; fig. 2). deciduous forests were preferentially used and were important in predicting locations (fig. 2, table 2). winter habitat use in maine is primarily influenced by food availability until snow depth becomes restrictive, and moose are commonly located where sufficient hardwood browse is available (morris 1999). moose feed mostly on deciduous vege‐ tation (renecker and schwartz 1997a) and seek out the highest biomass of dormant shrubs and palatable forage during the period of time after the rut and into winter (peek 1997). while not included in the top competing model at either landscape scale, locations were associated with a smaller proportion of coniferous forest (table 3). while forage is likely more accessible and nutritious in deciduous, mixed, and regenerating forests during early winter, cover provided by coniferous forest is probably an important habitat variable when snow depth impedes movement or as thermal cover in later winter/early spring as ambient temperature rises, conditions avoided in this study. moose in new brunswick showed preference for more open and deciduous forest types in early winter and preference for dense conifer stands in late winter (telfer 1970), and radio-collared moose in central massachusetts showed increasing selection for conifer stands as winter progressed (wattles 2011). abundance of food resources, not availability of cover, is likely the most important factor in predicting habitat use in early winter in maine, but a heterogeneous forest that provides both forage and shelter probably increases in use as winter progresses. elevation, while not included in the best fitting model, was higher in used than available units throughout the study area. the slightly higher elevation (~11 m) may reflect avoidance of wetlands in winter as used locations had a smaller proportion of wetlands than available habitat (table 3). wetland habitats at lower elevations may be important predictors of moose locations from late spring through autumn when insects, thermoregulation, and aquatic forage influence habitat use, but play no role in winter habitat selection (peek et al. 1976, peek 1997). previous research in maine found that moose moved from lowland (<305 m) into mid-elevation areas (367–427 m) in early winter, and occurred at slightly higher elevations later in winter (thompson et al. 1995). however, the ~11 m difference in elevation that we measured is probably biologically insignificant. trends in used habitat variables were similar at locations and the 5 ha scale; specifically, the majority of used units were found in mature (mixed, deciduous, and coniferous) and regenerating forest (recent clearcuts, partial cuts, regenerating forest, and scrub-shrub, table 4). the used proportion of these coarser habitat variables (i.e., mature and regenerating forest) were similar to those defined for each survey block, and ultimately the respective wmd (table 4). northern maine is considered high quality moose habitat due to commercial timber harvesting that produces stands of varying age and size providing adequate forage and cover throug‐ hout the region, and this heterogeneous alces vol. 52, 2016 andreozzi et al. – winter habitat use 49 habitat is key to the current high regional population (mdifw 2012). because moose browsing can substantially alter plant communities and affect the structure and dynamics of forest ecosystems (mcinnes et al. 1992, renecker and schwartz 1997a), there are important implications for forest management since moose prefer forage in clearcut and early successional habitat (westworth et al. 1989, scarpitti et al. 2005). browse consumption is strongly determined by its spatial distribution (vivas and saether 1987) and forage availability is an important factor in moose foraging behavior, irrespec‐ tive of scale (dussault et al. 2005, månsson et al. 2007). integrated management of an abundant moose population with commer‐ cial forestry in northern maine requires balancing moose density with their potential post-harvest influence on forest regenera‐ tion and stand composition (bergeron et al. 2011, andreozzi et al. 2014). extensive use of cutover areas by female moose in maine is indicative of how forest harvesting practices create beneficial interspersion of food and cover (leptich and gilbert 1989). the best moose habitat in maine is associated with commercially harvested forest (morris 1999), and >25% of the study area was classified as some form of cut habitat. however, there are economic, political, and social issues associated with forest harvest practices and mandated changes could influence the relative abundance of moose in northern maine (morris 1999). concern about the effects of heavy clearcutting in the 1970s and 1980s, particularly in response to a substantial spruce budworm (choristoneura fumiferana) outbreak (griffith and alerich 1996), resulted in the maine legislature passing the maine forest practices act in 1989 (maine forest service 1999). this act limited the size of clearcuts (<250 acres) and led to a dramatic shift from clearcutting to partial harvests beginning in the early 1990s; for example, ~93% of the 444,339 acres harvest in maine was defined as partial harvest in 2011 (maine forest service 2011). this harvest practice will presumably produce abundant and patchily distributed browse and cover in closer proximity than created by larger clearcuts. the relatively high moose density estimates in much of the study area (2.0–4.0 moose/km2; kantar and cumberland 2013) may reflect such habitat change. table 4. a comparison of the mean proportion (%) ± se of coarse cover types (i.e., mature and regenerating forest) within used units at locations and within the 5 ha landscape scale and the proportions within survey blocks during winters 2011 and 2012, northern maine, usa. proportion mature forest (%) wmd location (used) 5 ha (used) survey block 1 73.1 ± 8.9 70.7 ± 6.8 68.0 2 73.0 ± 4.0 75.8 ± 3.1 72.6 3 80.5 ± 4.3 77.5 ± 3.1 72.4 4 57.8 ± 5.2 57.8 ± 4.2 55.4 4 54.8 ± 9.1 50.0 ± 7.3 55.9 5 90.9 ± 9.1 70.1 ± 11.5 70.4 6 53.8 ± 14.3 62.6 ± 11.6 65.0 8 52.8 ± 5.9 52.4 ± 4.5 65.9 11 100.0 ± 0.0* 80.2 ± 7.7 69.8 19 76.2 ± 9.5 73.9 ± 7.9 71.2 proportion regenerating forest (%) wmd location (used) 5 ha (used) survey block 1 26.9 ± 8.9 25.7 ± 6.9 28.1 2 24.6 ± 3.9 22.3 ± 3.0 22.8 3 17.2 ± 4.1 17.6 ± 3.0 13.6 4 41.1 ± 5.2 40.9 ± 4.2 35.6 4 35.5 ± 8.7 42.9 ± 7.5 37.2 5 9.1 ± 9.1 23.2 ± 11.0 23.4 6 15.4 ± 10.4 9.6 ± 6.1 8.7 8 40.3 ± 5.8 40.5 ± 4.4 31.5 11 0.0 ± 0.0 18.1 ± 8.0 19.5 19 19.0 ± 8.8 20.9 ± 6.9 18.9 *small sample size (n=4) likely influenced proportions. 50 winter habitat use – andreozzi et al. alces vol. 52, 2016 analyses and modeling with data from single locations of individuals, rather than continuous locations from radio-collared animals, pose some concern and limitation for application across spatial and temporal scales. however, our habitat use information was analogous with past regional studies, and importantly, was labor efficient and provided added value to annual surveys. subsequent surveys in the same wmds should identify temporal changes in moose abundance and distribution important in developing management strategy. expansion and continuation of such analyses should also prove useful in examining the spatial distribution of moose across the landscape, the concentration of moose in habitat vulnerable to browsing damage, and long-term temporal relationships between moose population responses and timber harvesting practices in northern maine. acknowledgements we are grateful to the mdifw for providing the data used in this analysis. we thank the mdifw flight crew for collecting the data, especially k. marden for helping to organize and transfer the data. we are also thankful to m. ducey for providing knowledge and guidance during data analysis. references anderson, d. r., w. a. link, d. h. johnson, and k. p. burnham. 2001. suggestions for presenting the results of data analyses. the journal of wildlife management 65: 373–378. andreozzi, h. a., p. j. pekins, and m. l. langlais. 2014. impact of moose browsing on forest regeneration in northeast vermont. alces 50: 67–79. barton, k. 2013. mumin: multi-model inference. r package version 1.9.0 http://cran. r-project.org/package=mumin (accessed july 2013). bates, d. m., and m. maechler. 2012. lme4: linear mixed-effects models using s4 classes . r package version 0.999999-0 http://cran.r-project.org/package=lme4 (accessed july 2013). becker, s. a. 2008. habitat selection, condition, and survival of shiras moose in northwest wyoming. m.s. thesis. university of wyoming, laramie, wyoming, usa. belovsky, g. e. 1981. food plant selection by a generalist herbivore: the moose. ecology 62: 1020–1030. bergeron, d. h., p. j. pekins, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. the journal of wildlife management 32: 729–746. burham, k., and d. anderson. 2002. model selection and multivariate inference: a practical information–theoretical approach. springer-verlag, new york, new york, usa. dussault, c., r. courtois, j. p. ouellet, and i. girard. 2005. space use of moose in relation to food availability. canadian journal of zoology 83: 1431–1437. –––, c., and j. ouellet. 2004. behavioural responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. esri. 2011. arcgis desktop: release 10. environmental systems research institute. redlands, california, usa. griffith, d. m., and c. l. alerich. 1996. forest statistics for maine, 1995. usda forest service northeastern forest experiment station resource bulletin ne-135. radnor, pennsylvania, usa. heikkila, r., p. hokkanen, m. kooiman, n. ayguney, and c. bassoulet. 2003. the impact of moose browsing on tree species composition in finland. alces 39: 203–213. alces vol. 52, 2016 andreozzi et al. – winter habitat use 51 http://cran.r-project.org/package=mumin http://cran.r-project.org/package=mumin http://cran.r-project.org/package=lme4 hundertmark, k. j. 1997. home range, dispersal and migration. pages 303–336 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. jmp. 2012. version 10. sas institute inc., cary, north carolina, usa. kantar, l. e., and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose abundance in maine. alces 49: 31–39. karns, p. d. 1997. population distribution, density and trends. pages 125–140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. the journal of wildlife management 53: 880–885. maine forest service. 1999. forest regeneration and clearcutting standards: msf rules, chapter 20. augusta, maine, usa. –––. 2011. silvicultural activities report: including annual report on clearcutting and precommercial activities. department of agriculture, conservation and forestry, forest policy and management division, augusta, maine, usa. maine office of geographic information system. 2004. maine land cover dataset. augusta, maine, usa. månsson, j., h. andren, å. pehrson, and r. bergström. 2007. moose browsing and forage availability: a scale-dependent relationship? canadian journal of zoology 85: 372–380. mastenbrook, b., and h. cumming. 1989. use of residual strips of timber by moose within cutovers in northwestern ontario. alces 25: 146–155. mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73: 2059–2075. mdifw (maine department of inland fisheries and wildlife). 2012. maine’s moose population estimated at 76,000 after new survey. maine department of inland fisheries and wildlife, augusta, maine, usa. miller, b. k. 1989. seasonal movement patterns and habitat use of moose in northern new hampshire. m.s. thesis. university of new hampshire, durham, new hampshire, usa. millette, t. l., e. marcano, and d. laflower. 2014. winter distribution of moose at landscape scale in northeastern vermont: a gis analysis. alces 50: 17–26. morris, k. i. 1999. moose assessment. maine department of inland fisheries and wildlife, augusta, maine, usa. peek, j. m. 1997. habitat relationships. pages 351–375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. –––, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3–65. r development core team. 2013. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. http://www. r-project.org/ (accessed july 2013). renecker, l. a., and c. c. schwartz. 1997a. food habits and feeding behavior. pages 403–439 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. –––, and –––. 1997b. nutrition and energetics. pages 403–439 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american 52 winter habitat use – andreozzi et al. alces vol. 52, 2016 http://www.r-project.org/ http://www.r-project.org/ moose. smithsonian institute press, washington, d.c., usa. scarpitti, d. 2006. seasonal home range, habitat use, and neonatal habitat characteristics of cow moose in northern new hampshire. m.s. thesis. university of new hampshire, durham, new hampshire, usa. –––, c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. telfer, e. s. 1970. winter habitat selection by moose and white-tailed deer. journal of wildlife management 34: 553–559. thompson, m. e., j. r. gilbert, g. j. matula, jr., and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31: 223–245. vivas, h. j., and b.-e. saether. 1987. interactions between a generalist herbivore, the moose (alces alces), and its food resources: an experimental study of winter foraging behaviour in relation to browse availability. journal of animal ecology 56: 509–520. wattles, d. w. 2011. status, movements, and habitat use of moose in massachusetts. m.s. thesis. university of massachusetts amherst, amherst, massachusetts, usa. westworth, d., l. brusnyk, j. roberts, and h. veldhuzien. 1989. winter habitat use by moose in the vicinity of an open pit copper mine in north-central british columbia. alces 25: 156–166. alces vol. 52, 2016 andreozzi et al. – winter habitat use 53 using aerial survey observations to identify winter habitat use of moose in northern maine study area methods habitat use statistical analysis results discussion acknowledgements references alces18_1.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 minimizing mortality of moose neonates from capture-induced abandonment william j. severud1, glenn d. delgiudice1,2, and tyler r. obermoller1,2 1department of fisheries, wildlife, and conservation biology, university of minnesota, 2003 upper buford circle, suite 135, saint paul, minnesota, usa 55108; 2forest wildlife populations and research group, minnesota department of natural resources, 5463-c west broadway avenue, forest lake, minnesota, usa 55025 abstract: neonatal moose (alces alces) may be prone to maternal abandonment induced by capture activities. we observed unexpectedly high levels of abandonment during the first year of our study of calf survival and cause-specific mortality in northeastern minnesota. in response, we crafted a capture-induced abandonment contingency plan to reduce calf deaths caused by such abandonment. locations and movements of dams relative to calves were used to gauge whether abandonment was occurring and to trigger retrieval of live calves. the minnesota zoo and a private facility accepted abandoned calves in viable condition. as undesirable as it is to remove calves from the population and landscape, we found it preferable to leaving them to succumb to starvation, hypothermia, or predation. we believe variations of this plan may be used in other study areas to mitigate neonate mortality due to capture-induced abandonment. alces vol. 52: 73–83 (2016) key words: abandonment, alces alces, calves, capture-induced abandonment, gps collars, humaninduced abandonment, moose neonates capture-induced abandonment of ungulate neonates is a little-understood phenomenon in which mothers permanently reject offspring ostensibly in response to the disturbance of capture, marking of the neonate, or some combination of factors (livezey 1990, swenson et al. 1999). it is defined as “the permanent separation of mother and young causing death of the young,” occurring ≤1 day after marking (livezy 1990:193). the estimated rate of capture-induced abandonment of moose (alces alces) neonates varies widely from 4.6 to 41.7% (ballard et al. 1981, keech et al. 2011, patterson et al. 2013). in our recent study of moose calf survival and cause-specific mortality in northeastern minnesota, global positioning system (gps) collars fit to neonates of gps-collared dams revealed that both exhibited complex behaviors before ultimate abandonment over ≥ 48 h post-capture (delgiudice et al. 2015). this supports livezey’s (1990) contention that studies using very high frequency (vhf) collars or infrequent direct observation may not always recognize capture-induced abandonment. minimizing undue stress and mortality of study subjects during capture is an animal welfare issue, and we are bound by the ethics of our profession to ameliorate adverse effects of capture as much as possible (sikes et al. 2011). because the northwestern minnesota moose population declined from 4,000 to <100 animals from the mid-1980s to 2007 (murray et al. 2006, lenarz et al. 2009), and the northeastern population by ~55% from 2006 to 2016 (delgiudice 2016), public interest in the current moose research has been particularly keen (khouri 2012, marcotty 2013, associated press 2015). 73 because losses associated with capture occurred in the current study, we responded with measures in an attempt to minimize such mortality (butler et al. 2013, carstensen et al. 2014, 2015; delgiudice et al. 2014, 2015; severud et al. 2015). due to an unexpected and unacceptable level of capture-induced abandonment early in the calf study (2013), our objective was to develop and employ a formal contingency plan for recovering affected neonates. this plan relied on our increased understanding of dam and calf movement behavior indicative of captured-induced abandonment, made possible from our hourly monitoring of gpscollared dams and neonates during the 2013 calving season. during our 2014 capture operations, we implemented this plan and successfully mitigated mortality of abandoned, newly collared neonates. we further refined our plan for use in 2015, but those revisions remain untested due to the implementation of executive order 15-10 (28 april 2015) through which the governor of minnesota prohibited additional capture and collaring of moose in the state. study area we conducted this study in a 6,068-km2 area of northeastern minnesota, usa, located between 47° 00′ n and 47° 56′ n, 89° 57′ w and 92° 17′ w. the area was characterized as northern superior uplands (minnesota department of natural resources [mndnr] 2015) and was interspersed with lakes, wetlands, logging roads, and low-density human settlements. stands of northern white cedar (thuja occidentalis), black spruce (picea mariana), and tamarack (larix laricina) predominated in the lowlands, and balsam fir (abies balsamea), jack (pinus banksiana), eastern white (p. strobus), and red pine (p. resinosa) were prevalent on the uplands, where mixed stands of trembling aspen (populus tremuloides) and white birch (betula papyrifera) also occurred. open areas included deciduous shrub and sedge (carex spp.) meadows (mndnr 2015). white-tailed deer (odocoileus virginianus) populations occurred at pre-fawning density of ≤4 deer/ km2 (grund 2014). predators of moose in the area included gray wolves (canis lupus; 3 wolves/100 km2, erb and sampson 2013) and black bears (ursus americanus; 23 bears/100 km2, garshelis and noyce 2011). moose harvests last occurred in 2012 (delgiudice 2012). mean daily minimum temperature at ely in may 1991–2012 was 5 °c, and mean maximum was 18 °c. the mean minimum on 1 may was 1 °c and rose to 7 °c by the end of the month. mean daily maximum was 14 to 20 °c (https://weatherspark.com/ averages/30172/5/ely-minnesota-unitedstates). the mean daily minimum temperature at grand marais in may 1997–2012 was 5 °c, and the mean maximum was 15 °c. the mean minimum on 1 may was 1 °c and increased to 6 °c by the end of the month. mean daily maximum was 12 to 17 °c (https:// weatherspark.com/averages/29912/5/grandmarais-minnesota-united-states). methods neonate capture, handling, and abandonment considerations 2013. — we computer-monitored preparturient gps-collared female moose beginning 1 may 2013 to identify a calving movement followed by localization (severud et al. 2015). we allowed ≥36 hours before attempting to capture and handle calves of localized dams. a capture crew (quicksilver air, inc., fairbanks, alaska, usa) located the calf(ves) via helicopter, fit a gps collar (vectronic aerospace gmbh, berlin, germany), drew a blood sample, and measured several morphometrics and rectal temperature; search time for the dam via helicopter was not recorded. if twins were observed, both calves were handled and released together. at this age, neonates likely 74 reducing calf mortality due to abandonment – severud et al. alces vol. 52, 2016 https://weatherspark.com/averages/30172/5/ely-minnesota-united-states https://weatherspark.com/averages/30172/5/ely-minnesota-united-states https://weatherspark.com/averages/30172/5/ely-minnesota-united-states https://weatherspark.com/averages/29912/5/grand-marais-minnesota-united-states https://weatherspark.com/averages/29912/5/grand-marais-minnesota-united-states https://weatherspark.com/averages/29912/5/grand-marais-minnesota-united-states had received colostrum and milk from their dams. suckling and brown adipose tissue stores are necessary for thermoregulation and maintaining body temperature (schoonderwoerd et al. 1986). captive moose neonates typically nurse an average 8 times/day (±1.5) for 130 sec/bout and consume 375 g milk/bout (reese and robbins 1994). assuming that nursing occurred regularly over a 24-h cycle, this translates to 1 nursing bout every 3 h (2.2–4.8). we estimated the number of feedings missed, as it relates to potential abandonment, and the number of hours without food based on the dam’s time away from the calf, assuming the calf last nursed 3 h prior to capture and will nurse immediately upon reuniting. all captures and handling methods were approved by the university of minnesota’s institutional animal care and use committee (iacuc; protocol 1302-30328a) and were consistent with guidelines recommended by the american society of mammalogists (sikes et al. 2011). additional details can be found in severud et al. (2015). based on other studies using similar calf-handling methods (keech et al. 2011, patterson et al. 2013), we anticipated a low rate of capture-induced abandonment and did not have a formal contingency plan in place. prior to capture operations, however, our staff veterinarian contacted several canadian zoos that agreed to accept abandoned female calves, and the minnesota zoo agreed to accept 2 female calves. since no location would accept males, we initially considered euthanizing abandoned males as a more humane alternative to allowing them to succumb to exposure, starvation, or predation. 2014. — beginning 1 may 2014, we again remotely monitored cows for calving activity. we modified our capture and handling methods in 2014 in response to high levels of abandonment in 2013 (delgiudice et al. 2014). based on our analyses of capture-induced abandonment data from 2013 (delgiudice et al. 2015), we initially retained all 2013 handling methods, but conducted all captures with a ground crew of 3–4 people (5 once) without the use of helicopter assistance. search times were not recorded, but captures generally involved handlers briskly approaching the calving site coordinates with little time spent searching for calves. we only approached 1 dam/ day and waited until she was reunited with her calf(ves) before approaching another eligible dam the following day. we used the mean distance (256 m) of non-abandoning dams from their calves (2013) during the first 48 h post-capture as a threshold distance to indicate that mothers and calves had reunited (delgiudice et al. 2015). based on our findings from 2013 (see results and discussion, and delgiudice et al. 2015), we developed a formal captureinduced abandonment contingency plan to retrieve rejected calves that were alive and in “good condition” (fig. 1). in cooperation with the minnesota zoo, we contacted several zoos in the usa that agreed to accept and maintain calves; the minnesota zoo agreed to function as a staging area and eventually a receiving site for multiple calves. additionally, private captive facilities were interested in accepting abandoned calves (males and females). because dams may repeat calf abandonment in multiple years (patterson et al. 2013), cows abandoning in 2013 were not approached in 2014. results general capture results in 2013, the total time that handlers were on the ground (from drop from the helicopter to pick up) was not consistently recorded, but averaged 30.5 min (range = 18–56 min) for 10 captures. the average handling time was 9.1 min (range = 3–18 min; see severud et al. 2015 for additional results), and neonates averaged 2.2 days old (see delgiudice alces vol. 52, 2016 severud et al. – reducing calf mortality due to abandonment 75 et al. 2015 and severud et al. 2015 for additional results). the average rectal temperature of 43 neonates at capture was 38.7 °c. we observed no difference between neonates abandoned versus those not, and viewed this as evidence that calves had likely nursed. in 2014, average handling time was 7.5 min (range = 5.0 – 10.4, n = 8) and calf age was 1.7 days (range = 0.5 – 2.7, n = 12). when dams exhibited similar or higher levels of abandonment than in 2013, we further modified our methods by reducing our capture team to 2 people and limited handling to fitting a gps-collar and sex determination. mean handling time was reduced to 0.7 min (range = 0.2 – 2.2, n = 13) and calf age was 2.5 days (range = 1.7 – 4.1, n = 13). fig. 1. the original protocol used in handling and collaring neonatal moose calves in northeastern minnesota, may 2014. ‘warm’ indicates that the daily minimum temperature is <9 °c below the mean minimum temperature in may. ‘cold’ indicates that the daily minimum temperature is ≥9 °c below the mean minimum temperature in may. ‘wet’ indicates any measureable precipitation. 76 reducing calf mortality due to abandonment – severud et al. alces vol. 52, 2016 mortality from capture-induced abandonment 2013. — due to changing concerns over chronic wasting disease, recovered moose calves ultimately could not be transported to canadian zoos that had previously agreed to accept them. however, because abandonment by moose and other ungulates was so poorly understood at the time, we were unable to apply strict guidelines or thresholds (e.g., distance between dam and neonate, time away) for recognizing capture-induced abandonment. this also made us reticent to revisit dam-calf pairs after collaring, concerned that further disturbance might induce abandonment. in general, dams did not defend their calves when handlers approached to capture neonates (delgiudice et al. 2015). we attempted to recover the first of 9 abandoned calves of 49 captured and handled. this calf was female from a mixed set of twins that were 65 h old at capture. they were captured on 8 may, and we retrieved the abandoned calf on 10 may at ~1000 hr. the dam made one return visit without the male twin (9 may at 0435 hr) before she ultimately abandoned the female. the female calf was hypothermic upon retrieval and our staff veterinarian administered 50–70 cc of lactated ringer’s solu‐ tion subcutaneously and attempted to slowly warm the calf. a thermometer was not immediately available upon retrieval, but after ~3 h of warming, its rectal temperature was 33.6 °c, and the calf was warmed to 37.8 °c in the next 2 h; however, it died shortly thereafter, en route to the zoo. the second abandoned calf was male and euthanized. subsequently, due to the uncertainty of interpreting dam behavior relative to abandonment, a decision was made to leave all apparently abandoned calves in the field hoping that dams might reunite with them. in 2013, 7 dams abandoned at least 1 calf, and made 0–2 return trips to their calf (ves) within the 48-h post-capture period before ultimately abandoning (delgiudice et al. 2015). of the 24 non-abandoning dams, 19 returned directly to their calves after the capture process, and stayed with the calf until the calf died of natural cause or was abandoned naturally (delgiudice et al. 2015). the other 5 dams made 1–3 short return trips before reuniting with their calves until we removed the collars in february 2014. the non-abandoning dams stayed away an average of 4.7 h and most returntrips occurred within 1–12 h post-capture. one dam (12605) abandoned a calf, but we successfully re‐united them and the other twin via helicopter; another (12559) seemed to be moving between her twins (13080 and 13097) before all rejoined. dam 12488 abandoned her calf, but this was assigned “abandonment of unknown cause” since the pair was together for 15 h post-capture, after which the dam left without returning. the general pattern exhibited by 19 of the 24 non-abandoning dams was to initially flee at capture, but upon return, remain with the calf (delgiudice et al. 2015). abandoned calves died an average of 56 h post-capture (delgiudice et al. 2015) and 32 h after the last visit by their dam (“visit” defined as located ≤ 256 m from calf; delgiudice et al. 2015). overall, nonabandoning dams began returning to calves at 13–18 h post-capture. we documented one collared calf that died following natural abandonment the dam and calf had reunited for 80 h (severud et al. 2015). 2014. — two mortalities following aban‐ donment were a singleton male and a female from a set of twins. the singleton’s dam made 3 return trips to the vicinity (within an average 256 m), but we resisted retrieving this calf to not further disturb the dam. when it became apparent she was not returning, we initiated a retrieval response but the calf was dead; it died 68.5 h post-capture, and 25 h after the third return visit. alces vol. 52, 2016 severud et al. – reducing calf mortality due to abandonment 77 the second mortality resulted from aban‐ donment of a set of female twins. because the dam’s collar was not transmitting locations, we were unsure of her location postcapture; upon transmission it was clear that she had abandoned the twins and we initiated retrieval. this dam made 2 return trips within an average 86 m from the twins prior to their recovery when we found a dead and a viable calf. the dead calf died 98 h postcapture and 63 h after the dam last returned within 256 m. in 2014, 6 of 19 dams abandoned 9 of 25 calves (3 singletons and 3 sets of twins; delgiudice et al. 2014); however, follow‐ ing modification of our capture protocol, the abandonment rate fell from 5 of 9 dams abandoning 7 of 12 calves, to 1 of 10 abandoning 2 of 13 calves. of the 9 abandoned calves, 7 were retrieved in viable condition and brought into captivity. they nursed vigorously on formula (milk matrix 30/52, zoo‐ logic, hampshire, illinois, usa), although we limited consumption to about 1, 12-oz bottle every ~4 h to prevent diarrhea from overfeeding (schwartz 1992). rectal temperatures were all within normal range. in a few cases where a skin test indicated dehydration, we administered saline subcutaneously. the 7 calves were retrieved 50.9 ± 11.7 h post-capture and all survived in captivity up to 18 months of age (at publication). one collared calf (from a set of twins) died following natural abandonment and had an umbilicus infection. it was initially abandoned but we reunited it with its dam, after which she made several trips between this calf and its healthy twin before abandoning (severud et al. 2014). as in 2013, dams in 2014 did not tend to vigorously defend their calves when handlers approached neonates. in 2013 and 2014, dams (n = 35) not abandoning their calves were within 256 m of their calves for 19.1 (95% confidence interval [ci] 17.19–21.1) and 41.5 hourly fixes (95% ci 38.5–44.6) within 24 and 48 h post-capture, respectively. in stark contrast, abandoning dams (n = 13) were within 256 m of their calves for only 3.8 (95% ci 2.1–5.4) and 6.2 hourly fixes (95% ci 3.5–8.8) within 24 and 48 h post-capture, respectively (delgiudice et al., unpublished data). discussion the establishment and refinement of an abandonment contingency plan enabled our team to retrieve viable calves that had ostensibly been abandoned by their dams due to handling or collaring. when we initiated collaring in 2013, we did not know what movement patterns, distances between dams and their calves, or times apart could readily be used to identify abandonment. natural abandonment of neonatal ungulates has been anecdotally reported in the literature, but the level and causes of naturally occurring abandonment of moose neonates are unknown. based on our assumptions of behavior and movement, we observed a single case each in 2013 and 2014. our plan was intended to be flexible and adaptable to circumstances in the field, such as location of the dam, number of calves (twins or singleton), return-visits made by the dam, timing and duration of return-visits, current proximity of dam and calf, sign of compromised condition of the calf at capture (e.g., small size, injury, deformity or anomaly), fresh predator sign in the vicinity, or adult collar malfunctioning (failure to transmit gps locations, vhf can be used to check for gross proximity). many of these instances would have triggered a quicker response to retrieve calves. nine calves died following captureinduced abandonment in 2013, but in the process we gained an understanding of this behavior. because of the population’s rather rapid decline over the previous 7 years, we considered it important to initially attempt to obtain as much data at capture as reasonably possible. in 2014 we rescued viable calves 78 reducing calf mortality due to abandonment – severud et al. alces vol. 52, 2016 using our abandonment contingency plan and further refined our capture methods to mini‐ mize capture-induced abandonment. considerations for revising the 2015 protocol. — based on results from 2014 (delgiudice et al. 2014), we further refined our contingency plan (fig. 2) with addition of thresholds that would trigger cessation of captures and consideration to reunite abandoned calves with their dams, as done with one calf in 2013 and 2014. the thresholds for discontinuing capture operations were: 1) 3 capture-related mortalities (direct, e.g., trampled by dam; indirect, e.g., related to captured-induced abandonment); 2) availability at zoos for calves recovered in viable condition was exhausted, and not to exceed 6 capture-induced abandonments; 3) the proportion of capture-induced abandonments that would discontinue capture operations would have changed with the total number of calves collared (i.e., more liberal proportions when the sample size was smaller, more restrictive proportions as sample size increased). we would continue operations until we reached 6 calves collared. when the sample size accumulated to 7–30 calves, we would proceed with caution if 15–20% of calves had been abandoned, and ceased operations if >20% were abandoned. when >30 calves had been captured, these levels would have changed to 5–10% and >10% of calves abandoned. capture would discontinue only as the result of a new abandonment, not as a result of changing thresholds (e.g., when capturing the seventh or 31st calf). we also considered developing a plan to reunite individual abandoned calves with their dams. however, we considered this a risky option; many factors (e.g., current location of the dam, presence of a twin) were weighed before launching the attempts in 2013 and 2014. a formal plan was not created for 2015 due to the governor’s executive order. members of the media and the public were very concerned with the fates of individual animals, and capture-related mortality was essentially unacceptable at any level. as researchers studying a declining moose population, we were concerned with population-level processes such as overall birth and death rates, and understanding these demographic parameters was key to discovering the underlying mechanism of the population decline. an online petition that maintained focus on the deaths of individual animals convinced minnesota’s governor to issue an executive order on 28 april 2015 which barred placement of any additional collars on moose by the state. thus, we were not able to implement our revised protocol in 2015, but keep it in reserve should this executive order be rescinded. maternal rejection of offspring is an abnormal, little-understood behavior, although not uncommon across species (e.g., beale and smith 1970, livezey 1990, linnell et al. 2000). the manifestation of this behavior is a significant risk that researchers assume when designing a study that includes handling neonates. without both dams and calves wearing gps-collars, we would not have observed the patterns which allowed us to develop an effective plan to recover abandoned calves in viable condition (delgiudice et al. 2015, unpublished data). vhf telemetry and direct observational studies may underestimate capture-induced abandonment because we identified some dams that reunited temporarily with their calf(ves), but eventually abandoned them. mortalities may then be attributed erroneously to birth defects, predation, disease, or malnutrition (livezey 1990), biasing estimates of neonatal survival and cause-specific mortality (gilbert et al. 2014); albeit, natural abandonment is not well understood either. our study was annually reviewed by the university of minnesota’s iacuc, and our abandonment contingency plan represents a alces vol. 52, 2016 severud et al. – reducing calf mortality due to abandonment 79 fig. 2. a modified protocol developed from field experience in 2014 for handling and collaring neonatal moose calves in northeastern minnesota. ‘warm’ indicates that the daily minimum temperature is <18 °c below the mean minimum temperature in may. ‘cold’ indicates that the daily minimum temperature is ≥18 °c below the mean minimum temperature in may. ‘wet’ indicates any measureable precipitation. 80 reducing calf mortality due to abandonment – severud et al. alces vol. 52, 2016 refinement and improvement of our initial methods towards the advancement of animal welfare during field research (sikes et al. 2011). capture-induced mortality has significant cost with respect to animal welfare, budgets, time, personnel, data lost, and public relations. our protocol proved effective in the recovery of abandoned neonates, and these neonates contribute to educational, zoo, and captive facility goals, including diversifying the gene pool of captive moose. we believe this protocol and the information therein will be useful in future studies to help recognize and mitigate losses associated with capture, handling, and marking neonatal moose. acknowledgements the following individuals helped us try to make sense of the abandonment issue over many restless nights: k. foshay, b. betterly, t. enright, j. lodel, a. jones, l. ross, m. haas, m. carstensen, e. butler, e. hildebrand, d. pauly, m. dexter, d. plattner, b. patterson, r. moen, a. mcgraw, j. forester, m. schrage, m. keech, quicksilver air, vectronic aerospace, m. larson, l. cornicelli, j. rasmussen, t. kreeger, j. crouse, t. wolf, s. moore, minnesota zoo, j. weickert, and other zoos that agreed to take calves (columbus zoo, milwaukee county zoo, green bay [new] zoo). wjs received financial support provided by the albert w. franzmann and distinguished colleagues memorial award. funding was provided by the mndnr section of wildlife’s wildlife populations and research unit, the wildlife restora‐ tion (pittman–robertson) program, and the minnesota deer hunters association. the university of minnesota department of fisheries, wildlife, and conservation biology provided technical and other support. we appreciate the constructive feedback provided by e. bergman and 2 anonymous reviewers. references associated press. 2015. researchers to collar minnesota moose again to study decline. washington times. 20 april 2015. ballard, w. b., t. h. spraker, and k. p. taylor. 1981. causes of neonatal moose calf mortality in south central alaska. journal of wildlife management 45: 335–342. beale, d. m., and a. d. smith. 1970. forage use, water consumption, and productivity of pronghorn antelope in western utah. journal of wildlife management 34: 570–582. butler,e.,m.carstensen,e.hildebrand, and d. pauly. 2013. determining causes of death in minnesota’s declining moose population: a progress report. pages 97–105 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2012. minnesota department of natural resources, st. paul, minnesota, usa. carstensen, m., e. hildebrand, d. pauly, r. g. wright, and m. dexter. 2014. determining cause-specific mortality in minnesota’s northeast moose population. pages 133–143 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2013. minnesota department of natural resources, st. paul, minnesota, usa. ———, ———, d. plattner, m. dexter, c. janelle, and r. g. wright. 2015. determining cause-specific mortality of adult moose in northeast minnesota. pages 161–171 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2014. minnesota department of natural resources, st. paul, minnesota, usa. delgiudice, g. d. 2012. 2012 minnesota moose harvest. minnesota department alces vol. 52, 2016 severud et al. – reducing calf mortality due to abandonment 81 of natural resources, st. paul, minnesota, usa. ———. 2016. 2016 aerial moose survey. minnesota department of natural re‐ sources, st. paul, minnesota, usa. ———, w. j. severud, t. r. obermoller, k. j. foshay, and r. g. wright. 2014. determining an effective approach for capturing newborn moose calves and minimizing capture-related abandonment in northeastern minnesota. pages 25–39 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2013. minnesota department of natural resources, st. paul, minnesota, usa. ———, ———, ———, r. g. wright, t. a. enright, and v. st-louis. 2015. monitoring movement behavior en‐ hances recognition and understanding of capture-induced abandonment of moose neonates. journal of mammalogy 96: 1005–1016. erb, j., and b. a. sampson. 2013. distribution and abundance of wolves in minnesota, 2012–2013.minnesotadepartmentofnatural resources, st. paul, minnesota, usa. garshelis, d., and k. noyce. 2011. status of minnesota black bears, 2010. final report to bear committee, minnesota department of natural resources, st. paul, minnesota, usa. gilbert, s. l., m. s. lindberg, k. j. hundertmark, and d. k. person. 2014. dead before detection: address‐ ing the effects of left truncation on sur‐ vival estimation and ecological inference for neonates. methods in ecology and evolution 5: 992–1001. grund, m. 2014. monitoring population trends of white-tailed deer in minnesota 2014. status of wildlife populations. minnesota department of natural re‐ sources, st. paul, minnesota, usa. keech, m. a., m. s. lindberg, r. d. boertje, p. valkenburg, b. d. taras, t. a. boudreau, and k. b. beckmen. 2011. effects of predator treatments, individual traits, and environment on moose survival in alaska. journal of wildlife management 75: 1361–1380. khouri, a. 2012. minnesota is missing its moose. los angeles times. 22 december 2012. lenarz, m. s., j. fieberg, m.w. schrage, and a. j. edwards. 2009. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1012–1023. linnell, j. d. c., j. e. swensen, r. andersen, and b. barnes. 2000. how vulnerable are denning bears to disturbance? wildlife society bulletin 28: 400–413. livezey, k. b. 1990. toward the reduction of marking-induced abandonment of newborn ungulates. wildlife society bulletin 18: 193–203. marcotty, j. 2013. gps will help biologists address decline of moose. star tribune. 5 january 2013. minnesota department of natural re‐ sources (mndnr). 2015. ecological classification system. minnesota department of natural resources, st. paul, minnesota, usa. (accessed may 2015). murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166. patterson, b. r., j. f. benson, k. r. middel, k. j. mills, a. silver, and m. e. obbard. 2013. moose calf mortality in central ontario, canada. journal of wildlife management 77: 832–841. reese, e. o., and c. t. robbins. 1994. characteristics of moose lactation and neonatal growth. canadian journal of zoology 72: 953–957. 82 reducing calf mortality due to abandonment – severud et al. alces vol. 52, 2016 http://www.dnr.state.mn.us/ecs/index.html http://www.dnr.state.mn.us/ecs/index.html schoonderwoerd, m., c. e. doige, g. a. wobeser, and j. m. naylor. 1986. protein energy malnutrition and fat mobilization in neonatal calves. the canadian veterinary journal 27: 365–371. schwartz, c. c. 1992. techniques of moose husbandry in north america. alces supplement 1: 177–192. severud, w. j., g. d. delgiudice, t. r. obermoller, t. a. enright, r. g. wright, and j. d. forester. 2015. using gps collars to determine parturition and cause-specific mortality of moose calves. wildlife society bulletin 39: 616–625. ———, ———, ———, k. j. foshay, and r. g. wright. 2014. using gps collars to determine moose calving and causespecific mortality of calves in northeastern minnesota: progress report on second field season. pages 40–56 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2013. minnesota department of natural resources, st. paul, minnesota, usa. sikes, r. s., w. l. gannon, and the animal care and use committee of the american society of mammalogists. 2011. guidelines of the american society of mammalogists for the use of wild mammals in research. journal of mammalogy 92: 235–253. swenson, j. e., k. wallin, g. ericsson, g. cederlund, and f. sandegren. 1999. effects of ear-tagging with radiotransmitters on survival of moose calves. journal of wildlife management 63: 354–358. alces vol. 52, 2016 severud et al. – reducing calf mortality due to abandonment 83 minimizing mortality of moose neonates from �capture-nduced abandonment study area methods neonate capture, handling, and abandonment considerations results general capture results mortality from capture-nduced abandonment discussion title_bkm_9 acknowledgements references alces19_83.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces17_xixbusinessmeeting.pdf alces vol. 17, 1981 alces_a_160182_o 99..112 status and trends of moose populations and hunting opportunity in the western united states m. steven nadeau1, nicholas j. decesare2, douglas g. brimeyer3, eric j. bergman4, richard b. harris5, kent r. hersey6, kari k. huebner7, patrick e. matthews8, and timothy p. thomas9 1idaho department of fish and game, 600 s. walnut, boise, idaho 83709, usa; 2montana fish, wildlife and parks, 3201 spurgin road, missoula, montana 59804, usa; 3wyoming game and fish department, box 67, jackson, wyoming 83001, usa; 4colorado parks and wildlife, 317 w. prospect avenue, fort collins, colorado 80526, usa; 5washington department of fish and wildlife, 600 capital way north, olympia, washington 98504, usa; 6utah division of wildlife resources, box 146301, salt lake city, utah 84114, usa; 7nevada department of wildlife, 60 youth center road, elko, nevada, 89801; 8oregon department of fish and wildlife, 65495 alder slope road, enterprise, oregon 97828, usa; 9wyoming game and fish department, box 6249, sheridan, wyoming 82801, usa. abstract: we review the state of knowledge of moose (alces alces shirasi) in the western us with respect to the species’ range, population monitoring and management, vegetative associations, licensed hunting opportunity and hunter harvest success, and hypothesized limiting factors. most moose monitoring programs in this region rely on a mixture of aerial surveys of various formats and hunter harvest statistics. however, given the many challenges of funding and collecting rigorous aerial survey data for small and widespread moose populations, biologists in many western states are currently exploring other potential avenues for future population monitoring. in 2015, a total of 2,263 hunting permits were offered among 6 states, with 1,811 moose harvested and an average success rate per permit-holder of 80%. the spatial distribution of permits across the region shows an uneven gradient of hunting opportunity, with some local concentrations of opportunity appearing consistent across state boundaries. on average, hunting opportunity has decreased across 56% of the western us, remained stable across 17%, and increased across 27% during 2005–2015. generally, declines in hunting opportunity for moose are evident across large portions (62–89%) of the “stronghold” states where moose have been hunted for the longest period of time (e.g., idaho, montana, utah, and wyoming). in contrast, increases in opportunity appear more common at peripheries of the range where populations have expanded, including most of colorado, northeastern washington, southern idaho, and eastern montana. there are many factors of potential importance to moose in this region, including parasites, predators, climate, forage quality, forage quantity, and humans. state wildlife agencies are currently conducting a variety of research focused on population vital rates, the development of monitoring techniques, forage quality, trace mineral levels, and evaluation of relative impacts among potential limiting factors. alces vol. 53: 99–112 (2017) key words: alces alces shirasi, colorado, hunter harvest, idaho, montana, nevada, oregon, population trends, range, shiras moose, utah, washington, wyoming the occupied range of moose (alces alces) extends southward into the western united states along the rocky mountains and the western cordillera ecoregion (cec 1997; fig. 1). here, we review the state of knowledge of moose in the western us with respect to range, population monitoring and management, vegetative associations, licen‐ sed hunting opportunity and hunter harvest success, and hypothesized limiting factors. 99 in particular, we use hunter opportunity and harvest data to address spatio-temporal trends in regional moose populations. we focus specifically on trends in hunting opportunity over the past decade (2005–2015) because they facilitate a simple assessment of local trends with comparable data across the 8-state region. range and habitat moose in the rocky mountains of the usa and southern canada were described as a distinct subspecies during the early part of the 20th century (a. a. shirasi; nelson 1914, peterson 1952), though historical accounts suggest they were rare throughout the us rocky mountains until the mid-1800s (karns 2007). after periods of population expansion and subsequent declines due to overharvest during the late 1800s, it is largely believed that moose populations increased to new highs during the early to mid-1900s within portions of idaho, montana, utah, and wyoming (brimeyer and thomas 2004, fig. 1. predicted range of moose in western usa circa 2015, based on compilation of state distribution data characterizing known occupancy by resident moose and predicted occupancy through species distribution modeling of beauvais et al. (2013). 100 status of moose in western us – nadeau et al. alces vol. 53, 2017 toweill and vecellio 2004, wolfe et al. 2010, decesare et al. 2014). during the latter half of the 20th century, moose also naturally colonized eastern portions of washington and oregon, and translocations were used to introduce moose to colorado and unoccupied portions of wyoming and idaho, as well as to augment populations in utah (kufeld 1994, olterman et al. 1994, base et al. 2006, wolfe et al. 2010, matthews 2012). sightings of moose occurred in nevada as early as the late 1980s, but have increased in recent years, including multiple verified sightings of cows with calves in 2016. while numbers in outlying states such as colorado and washington appear to be stable or increasing, declines have been noted recently in previous stronghold portions of idaho, montana, utah, and wyoming (harris et al. 2015, monteith et al. 2015, decesare et al. 2016). this geographic area is primarily occupied by the shiras subspecies of moose (a. a. shirasi) which exists throughout the rocky mountains from colorado and utah northward to southern alberta and british columbia. recent range expansion of moose into eastern montana has come with some uncertainty regarding subspecies identity. it is unclear whether animals east of the rocky mountain chain in the eastern portions of montana, with neighboring moose populations in saskatchewan and north dakota, would be better described as a. a. shirasi or as belonging to the northwestern subspecies a. a. andersoni. habitats occupied by moose vary throughout the western us with respect to gradients in abiotic conditions (e.g., elevation, temperature, and precipitation), vege‐ tative associations, and large mammal predator-prey communities. we summarized moose management units across each state within which licensed moose hunting is offered, in terms of mean values of elevation, annual precipitation (cm), and minimum january and maximum july temperatures (°c) in a gis using a digital elevation model and 30-year climate averages during 1981–2010 (prism climate group 2016). though moose management units in colorado and utah contain the southern-most introduced and naturally occurring moose populations in the world, the relatively high elevations of occupied habitats in these regions result in climates similar to neighboring states further north (table 1). average elevation ranged from 911 m in washington to 2712 m in colorado, average annual precipitation ranged from 62–84 cm annually, and minimum january and maximum july temperatures averaged �14 to �6 °c and 24 to 27 °c, respectively, among states during 1981–2010 (table 1). table 1. descriptive statistics characterizing means per moose management unit in elevation, precipitation, minimum january temperature, maximum july temperature, summarized across units within each western state where licensed moose harvest is allowed, 1981–2010. elevation (m) precipitation (cm) minimum january temperature (°c) maximum july temperature (°c) state mean range mean range mean range mean range colorado 2712 (2075–3304) 63 (35–111) �13 (�16 to �9) 24 (20–30) idaho 1656 (804–2353) 84 (24–150) �9 (�14 to �4) 26 (23–31) montana 1849 (814–3071) 71 (32–119) �11 (�15 to �6) 24 (18–29) utah 2226 (1896–2715) 64 (48–83) �11 (�14 to �8) 26 (23–28) washington 911 (723–1169) 66 (46–101) �6 (�8 to �5) 27 (24–28) wyoming 2387 (1945–2999) 62 (20–123) �14 (�18 to �11) 24 (19–29) alces vol. 53, 2017 nadeau et al. – status of moose in western us 101 studies of vegetation associations inhabited by moose in the western us include a preponderance of evidence for selection of willow (salix spp.) communities and plants for space use and food habits, particularly during winter (mcmillan 1953, knowlton 1960, dorn 1970, wilson 1971, pierce and peek 1984, van dyke et al. 1995, kufeld and bowden 1996, dungan and wright 2005, baigas et al. 2010, vartanian 2011, burkholder et al. 2017). the importance of willow has been documented primarily in relatively colder and drier portions of the range (e.g., in portions of colorado, utah, wyoming, southeast idaho, and southwest montana), though exceptions to the importance of willow exist in local populations in these areas. for example, in certain portions of western colorado and utah, moose have also colonized upland shrub communities including oakbrush (quercus spp.), serviceberry (amelanchier spp.), and mountain mahogany (cercocarpus spp). in portions of southwest montana moose occupy forested stands of aspen (populus spp.) and douglas-fir (pseudo‐ tsuga menziesii) and feed on a range of other shrubs and saplings, including but not limited to serviceberry, huckleberry (vaccinium spp.), red-osier dogwood (cornus sericea), and subalpine fir (abies lasiocarpa) (stevens 1970). in southeast idaho moose forage primarily on bitterbrush (purshia tridentata), willow, serviceberry, chokecherry (prunus virginiana), and aspen (populus tremuloides) (ritchie 1978). west of the continental divide where occupied portions of the range include more maritime-influenced climates, moose associate with various forest communities where willow-dominated lowlands are less prevalent. in wetter climates such as in northwest montana, moose show strong associations with conifer forests, including but not limited to regenerating vegetation following timber harvest (matchett 1985, langley 1993). forage species in these areas have included red-osier dogwood, serviceberry, and menziesia (menziesia ferruginea) (matchett 1985). in north-central idaho moose select conifer forest including old growth and mature mixed-age stands of grand fir (abies grandis) and subalpine fir (pierce and peek 1984) where they forage primarily on pacific yew (taxus brevifolia), sitka alder (alnus viridis), and menziesia (pierce 1984). moose in washington occupy habitats typically characterized by dense conifer forest, including mixed stands of western red cedar (thuja plicata) and western hemlock (tsuga heterophylla) and consume diets of willow, ceanothus spp., and other shrubs and forbs during summer, and conifers such as cedar and hemlock during winter (j. goerz, university of montana, pers. comm.). population monitoring moose are widespread across broad regions of the western usa but densities are low relative to populations further north in canada, alaska, and the northeastern usa. recent population estimates (2014) in our study area jurisdictions were 2,400 in colorado, 10,000 in idaho, 4,000 in montana, 20 in nevada, 70 in oregon, 2,625 in utah, 3,200 in washington, and 4,650 in wyoming (timmermann and rodgers 2017; nevada estimate unpublished). these abundance estimates are not necessarily comparable as they were each derived with different methods and generally, with uncertainty. resources to monitor moose are relatively sparse when compared with those devoted to more abundant elk (cervus canadensis), deer (odocoileus spp.), and pronghorn (antilocapra americana) populations. thus, population monitoring of moose in many areas is met with challenges of limited data and low statistical power (harris et al. 2015, decesare et al. 2016). to date, most monitoring programs rely primarily on a mixture of aerial surveys of various formats and hunter harvest statistics. 102 status of moose in western us – nadeau et al. alces vol. 53, 2017 colorado parks and wildlife biologists use spreadsheet population models that are based on estimates of survival and recruitment largely obtained opportunistically during surveys for other species (white and lubow 2002). in idaho biologists have similarly collected survey data incidental to elk surveys for several decades, but have not developed a statewide trend index or population estimation technique. montana fish, wildlife and parks biologists conduct minimum count aerial surveys in a subset of moose hunting units, and rely primarily on hunter harvest statistics elsewhere for assessment of trends (decesare et al. 2016). moose have more recently colonized portions of oregon and nevada where state agency biologists primarily monitor them via public sighting reports and opportunistic ground and aerial survey detections. the utah division of wildlife resources conducts aerial surveys specifically targeting moose every 3 years within hunted units, in addition to collecting incidental observations during elk surveys. washington department of fish and wildlife biologists conduct annual helicopter surveys to construct an annual index that is useful to track general trends and recruitment rates within most of the core moose range in northeastern washington; however, despite incorporating covariates affecting detection probabilities, these data remain imprecise (harris et al. 2015). wyoming fish and game biologists conduct annual moosespecific aerial surveys in the more abundant herd units and use these data in a modified spreadsheet simulation model (similar to that used in colorado); incidental observations also occur in surveys of other big game populations (monteith et al. 2015). given the many challenges of funding and collecting rigorous aerial survey data for small and widespread moose populations, many western state agencies are currently exploring other potential avenues for future population monitoring. these include exploration of forward-looking infrared (flir) technology in comparison to aerial surveys in northern idaho (sensu storm et al. 2011), patch occupancy modeling of hunter sightings data in montana (sensu rich et al. 2013), and development of a smartphone app for collecting public sightings of moose in washington (sensu teacher et al. 2013). hunting opportunity, success rates, and trends licensed hunting of moose began in the late 19th century in idaho, montana, and wyoming, but these seasons were subsequently closed in 1897–1899 due to concerns of overharvest. hunting seasons were (re)instated within 6 rocky mountain states in the following chronological order: 1912 (wyoming), 1945 (montana), 1946 (idaho), 1958 (utah), 1977 (washington), and 1985 (colorado). the number of harvested moose has declined in the past decade in 4 of 6 western states (idaho, montana, utah, and wyoming) that allow hunting (fig. 2). conversely, harvest numbers have continued to increase in colorado and washington through added opportunities in traditional and newly opened hunting units (fig. 2). to date, moose hunting has not been initiated in nevada or oregon, though a bull moose was incidentally harvested in nevada in the 1950s. in 2015, a total of 2,263 hunting permits were offered by the 6 states, resulting in 1,811 harvested moose and an average harvest success rate of 80% (table 2). since 1990, the highest single-year state harvest was 1,215 moose in wyoming in 2001 (fig. 2). in contrast, the highest harvest in 2015 was 666 moose in idaho. proportionate harvest of antlerless (cows and calves) moose ran‐ ged from 0 to 17% in states with generally declining opportunity, compared to 35% and 49% antlerless harvest in washington and colorado, respectively, where hunter opportunity continued to increase through 2015 alces vol. 53, 2017 nadeau et al. – status of moose in western us 103 (table 2). for the states focal to this summary, moose hunting is managed through lottery permit systems that produce high harvest success rates averaging 75-95% annually (table 2). there is a tendency for somewhat lower rates of success on antlerless permits versus antlered-only or either-sex permits (table 2). spatial distribution of permits across the region shows an uneven gradient of hunting opportunity (i.e., number of moose hunting permits) across the states (fig. 3a). although moose occupy areas beyond hunting units alone, multi-state concentrations of moose hunting opportunity appear in approximately 4 portions of the range: 1) the northern rockies and columbia mountains region of washington, northern idaho, and northwest montana; 2) the middle rockies region of southwest montana, southeast idaho, and western wyoming; 3) the southern rockies region of north-central colorado; and 4) the wasatch and uinta ranges in north-central utah (fig. 3a). we summarized trends in hunter op‐ portunity across the 11-year period of 2005–2015 for all hunting management units or districts (hereafter “units”) across the entire region. given differences in popula‐ tion monitoring techniques across states, hunter opportunity data provide the most standardized means of assessing and fig. 2. moose harvest trends in the western usa from 2000 to 2015. 104 status of moose in western us – nadeau et al. alces vol. 53, 2017 table 2. numbers of moose hunting permits issued and moose harvested, percentage of antlerless moose (cows and calves) in the harvest, and success rates of hunters holding antlered (including either-sex) and antlerless permitsamong states in the western usa during the 2015 hunting season. state permits harvest % antlerless in harvest % hunter success, antlered1,2 % hunter success, antlerless1 colorado 313 233 48.9% 85.6% 65.5% idaho 873 666 17.4% 77.6% 73.9% montana 362 268 13.6% 73.8% 76.1% nevada 0 0 – – – oregon 0 0 – – – utah 143 137 0% 95.8% – washington 168 142 35.2% 91.6% 72.1% wyoming 411 365 16.4% 90.8% 80.0% 1hunter success measured as harvested moose per permit allocated, not accounting for number of permit-holders that actually hunted. 2hunter success for antlered moose includes the combined success rates of both antlered-only and either-sex permit-holders, given either-sex permit-holders harvest predominately antlered moose (e.g., 94% of hunters in washington). fig. 3. regional patterns in a) the availability of moose hunting permits in 2015 and b) changes in permit availability per hunting unit between 2005 and 2015 across the western usa. note some hunting units were merged together for one or both years to facilitate comparisons of equal areas among years. alces vol. 53, 2017 nadeau et al. – status of moose in western us 105 comparing trends among states. in the absence of consistent sampling-based surveys, we treat hunting opportunity as an index to population status, assuming that changes in opportunity within units reflect relative trends in abundance. in order to allow comparison of equal areas across years, we merged some units together for either the 2005 or 2015 seasons to accommodate changes in unit boundaries or regulations among years. in total, we assessed changes in opportunity across 273 units within 6 states from 2005 to 2015 (table 3). overall, moose hunting opportunity declined by 23% across the entire region, from 2,970 permits in 2005 to 2,279 permits in 2015, but trends varied among local units. on average, hunting opportunity decreased across 56% of the units, remained stable across 17%, and increased across 27% (table 3). visual display of trends per hunting unit shows a diversity of dynamics, from areas that were closed to hunting to areas newly opened to hunting (fig. 3b). generally, declines in hunting opportunity for moose are evident across much of their occupied area in the western us, including large portions (62–89%) of “stronghold” states where moose have been abundant enough to support hunting for the longest period of time (i.e., idaho, montana, utah, and wyoming). we note that this trend was evident prior to 2005 in wyoming where loss of 527 permits occurred between 2000 and 2005. in contrast, increases in hunting opportunity were more common at peripheries of the range where populations expanded, including most of colorado, central utah, southern idaho, northeastern washington, and eastern montana. the increases in eastern montana mirror those in similar prairie habitats of neighboring jurisdictions in southeast alberta, western north dakota, and southwest saskatchewan where moose populations have generally increased in recent years (laforge et al. 2016). increases in hunting opportunity in colorado were partially in response to continued introductions of moose into new areas including the grand mesa national forest (2005–2007) and the white river national forest (2009–2010). not coincidentally, the few units in colorado where hunting opportunity declined were also those serving as source populations for translocations. the wasatch unit has the most recently established population and offers the most hunting opportunity in utah. table 3. trends in moose hunting opportunity (i.e., number of moose hunting permits) from 2005–2015, summarized per hunting unit across 6 western states. trends in moose permits per unit during 2005–2015 state nunits median unit area (km 2) decreased stable increased colorado 42 1127 5% 5% 90% idaho 92 834 62% 21% 17% montana 85 1490 67% 20% 13% utah 9 2431 89% 0% 11% washington 7 2468 14% 0% 86% wyoming 38 1810 66% 24% 11% area-weighted average1 56% 17% 27% 1area-weighted averages were estimated by weighting states according to the product of the number of hunting units and median area per unit for each state. 106 status of moose in western us – nadeau et al. alces vol. 53, 2017 potential limiting factors there are many factors of potential importance to moose population dynamics across the western usa, and data concerning their presence, prevalence, or effects on moose vary across populations. multiple parasites and diseases including the arterial worm (elaeophora schneideri), winter tick (dermacentor albipictus), giant liver fluke (fascioloides magna), chronic wasting disease, hydatid worm (echinococcus granulosus), and other tapeworms (taenia spp.) have been documented in the region from examination of hunter-killed, live-captured, or opportunistically collected specimens of moose or other ungulate species (e.g., worley et al. 1972, samuel et al. 1991, dunkel et al. 1996, pessier et al. 1998, henningsen et al. 2012, levan et al. 2013). large predator communities may have up to 4 species black bears (ursus americanus), grizzly bears (ursus arctos), cougars (puma concolor), and wolves (canis lupus) that vary in density and potential to affect moose population dynamics within individual states (sensu griffin et al. 2011, brodie et al. 2013). presence of wolves, grizzly bears, and chronic wasting disease vary most across states (table 4). the potential importance of nutritional limitations has been documented in certain western moose populations. for example, ruprecht et al. (2016) reported low fat levels in moose in utah relative to northern populations, as well as lower pregnancy and twinning rates than in other north american populations at lower latitudes. both are suggestive of nutritional limitation, although nutritional condition is affected by factors other than forage including disease, parasites, and combined influences related to climate change. climate change may have particularly pronounced effects on populations at the periphery of a species’ range (hampe and petit 2005). although climate change has been linked to expansion of moose in alaska through habitat change (tape et al. 2016), and some southern populations remain stable or increasing (murray et al. 2012), we share the speculative concern of lenarz et al. (2010) over the long-term viability of moose populations along their southern range edge. current climate projections suggest that temperature will increase during all seasons in the rocky mountain region, and precipitation may increase in northern portions (rocca et al. 2014). warmer temperatures may possibly induce heat stressrelated impacts in free-ranging moose as measured in captive moose (renecker and hudson 1986, mccann et al. 2013). indirect effects of warmer climate may also impact moose, as mediated by changes in parasitehost communities or plant communities and/or phenology (rempel 2011, monteith et al. 2015). interestingly, we find that the southernmost global populations of moose found in colorado appear to be the most stable in the western usa as evidenced by population growth and increased hunting opportunity (fig. 3b). however, latitude alone may not sufficiently characterize variation in climate across this region, as portions of colorado are similar in temperature regime to more northern areas (table 1), and increasing populations such as those in colorado or washington are also relatively young and expanding into unoccupied habitats. it is pos‐ sible that the dynamics of these populations are still in accordance with the earlier stages of the eruptive cycle of introduced ungulate populations identified by caughley (1970). the complexities of time since establishment and density dependence may confound comparisons among populations with respect to climateor habitat-related conditions; however, direct physiological impacts of heat stress should manifest regardless. in short, multiple factors other than latitude alone are influential on the shortand long-term alces vol. 53, 2017 nadeau et al. – status of moose in western us 107 t ab le 4 .d o cu m en te d p re se n ce ,a b se n ce o r u n ce rt ai n ty re g ar d in g p ar as it es an d p re d at o rs o f p o te n ti al im p o rt an ce to m o o se d y n am ic s ac ro ss st at es o f th e w es te rn u s a , ci rc a 2 0 1 5 . p ar as it es an d d is ea se p re d at o rs s ta te e la eo p h o ra sc h n ei d er i d er m a ce n to r a lb ip ic tu s e ch in o co cc u s g ra n u lo su s c h ro n ic w as ti n g d is ea se f a sc io lo id es m a g n a p a re la p h o st ro n g yl u s te n u is b la ck b ea r g ri zz ly b ea r c o u g ar w o lf c o lo ra d o + + + + + + – – + + – + + – id ah o + + + + + + – + + – + + + + + + + m o n ta n a + + + + + + – + + – + + + + + + + + n ev ad a + + – – – – + – + + – o re g o n + + + + – – + + + + u ta h + + + + + + – + + + + w as h in g to n + + + – – + + + + + + + w y o m in g + + + + + + – + + + + + + + + + + = d o cu m en te d as co m m o n ly p re se n t in m o o se o r ar ea s o cc u p ie d b y m o o se . + = d o cu m en te d as p re se n t am o n g u n g u la te p o p u la ti o n s, b u t ra re in m o o se o r ar ea s o cc u p ie d b y m o o se . – = n o t d o cu m en te d an d p re se n ce se en as u n li k el y in m o o se o r ar ea s o cc u p ie d b y m o o se . [b la n k ] = n o t d o cu m en te d an d p re se n ce u n k n o w n in m o o se o r ar ea s o cc u p ie d b y m o o se . 108 status of moose in western us – nadeau et al. alces vol. 53, 2017 dynamics of moose populations in north america. state-regulated hunter harvest, tribal harvest, and illegal harvest of moose are expected to play some role in regulating moose populations across the western usa. hunting permits are allocated considering local objectives throughout this region, and range from conservative permit numbers to minimize impacts of hunting, to liberal numbers of permits to reduce populations and their impacts on humans and the environment (table 2; decesare et al. 2014). tribal harvest of moose is permitted in most western states through treaty rights, but the availability of specific data varies by jurisdiction. in montana, tribal harvest was estimated to increase the total annual harvest by 7–16% during 1986–2012. illegal harvest may also have measurable impact on moose; for example, the result of multiple studies in idaho suggested that 31–50% of known mortality was associated with illegal harvest (ritchie 1978, pierce et al. 1985, toweill and vecellio 2004). future research state wildlife agencies are conducting research in conjunction with universities and coordinating research among states to leverage resources across jurisdictions. research objec‐ tives include work focused on population vital rates (e.g., adult female survival, fecun‐ dity, and calf survival), movements and spatial ecology, resource selection, nutritional ecology and forage monitoring, baseline disease and mineral monitoring, the development of monitoring techniques, and identification of limiting factors. research objectives regarding limiting factors include the assessment of relative impacts of predation, parasites, climate change (direct effects of heat stress and indirect effects on moose foraging behavior and parasite loads), and habitat changes (e.g., decline of early seral forests) on moose vital rates. acknowledgements we thank m. atamian, d. base, j. goerz, h. ferguson, s. hansen, a. holland, m. lloyd, j. newby, j. oyster, k. podruzny, a. prince, t. smucker, and p. wolff for their contributions to the collection and compilation of moose monitoring data that went into this manuscript. j. gude, j. maskey, m. mitchell, and j. smith contributed to meetings and discussions that led to the development of this manuscript. we thank a. apa, k. logan, the associate editor and 1 anonymous reviewer for helpful edits and feedback. references baigas, p., r. a. olson, r. m. nielson, s. n. miller, and f. g. lindzey. 2010. modeling seasonal distribution and spatial range capacity approximations of moose in southeastern wyoming. alces 46: 89–112. base, d. l., s. zender, and d. martorello. 2006. history, status, and hunter harvest of moose in washington state. alces 42: 111–114. beauvais, g., m. andersen, d. keinath, j. aycrigg, and j. lonneker. 2013. predicted vertebrate species habitat distributions and species richness. pages 58–110 in j. aycrigg, m. andersen, g. beauvais, m. croft, a. davidson, l. duarte, j. kagan, d. keinath, s. lennartz, j. lonneker, t. miewald, and j. ohmann, editors. ecoregional gap analysis of the northwestern united states: northwest gap analysis project. draft report. universityofidaho,moscow,idaho,usa. brimeyer, d. g., and t. p. thomas. 2004. history of moose management in wyoming and recent trends in jackson hole. alces 40: 133–144. brodie, j., h. johnson, m. mitchell, p. zager, k. proffitt, m. hebblewhite, m. kauffman, b. johnson, j. bissonette, c. bishop, j. gude, j. herbert, k. hersey, m. hurley, p. m. lukacs, alces vol. 53, 2017 nadeau et al. – status of moose in western us 109 s. mccorquodale, e. mcintire, j. nowak, h. sawyer, d. smith, and p. j. white. 2013. relative influence of human harvest, carnivores, and weather on adult female elk survival across western north america. journal of applied ecology 50: 295–305. burkholder, b. o., n. j. decesare, r. a. garrott, and s. j. boccadori. 2017. heterogeneity and power to detect trends in moose browsing of willow communities. alces 53: 23–39. caughley, g. 1970. eruption of ungulate populations, with emphasis on himalayan thar in new zealand. ecology 51: 53–72. commission for environmental cooperation (cec). 1997. ecological regions of north america – toward a common perspective. cec, montreal, quebec, canada. decesare, n. j., j. r. newby, v. j. boccadori, t. chilton-radandt, t. thier, d. waltee, k. podruzny, and j. a. gude. 2016. calibrating minimum counts and catch per unit effort as indices of moose population trend. wildlife society bulletin 40: 537–547. ———, t. d. smucker, r. a. garrott, and j. a. gude. 2014. moose status and management in montana. alces 50: 35–51. dorn, r. d. 1970. moose and cattle food habits in southwest montana. journal of wildlife management 34: 559–564. dungan, j. d., and r. g. wright. 2005. summer diet composition of moose in rocky mountain national park, colorado. alces 41: 139–146. dunkel, a. m., m. c. rognlie, g. rob johnson, and s. e. knapp. 1996. distribution of potential intermediate hosts for fasciola hepatica and fascioloides magna in montana, usa. veterinary parasitology 62: 63–70. griffin, k. a., m. hebblewhite, h. s. robinson, p. zager, s. m. barbermeyer, d. christianson, s. creel, n. c. harris, m. a. hurley, d. h. jackson, b. k. johnson, w. l. myers, j. d. raithel, m. schlegel, b. l. smith, c. white, and p. j. white. 2011. neonatal mortality of elk driven by climate, predator phenology and predator community composition. journal of animal ecology 80: 1246–1257. harris, r., m. atamian, h. ferguson, and i. keren. 2015. estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty. alces 51: 57–69. hampe, a., and r. j. petit. 2005. conserving biodiversity under climate change: the rear edge matters. ecology letters 8: 461–467. henningsen, j. c., a. l. williams, c. m. tate, s. a. kilpatrick, and w. d. walter. 2012. distribution and prevalence of elaeophora schneideri in moose in wyoming. alces 48: 35–44. karns, p. d. 2007. population distribution, density, and trends. pages 125–140 in a. w. franzmann andc.c.schwartz, editors. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. knowlton, f. f. 1960. food habits, movements and populations of moose in the gravelly mountains, montana. journal of wildlife management 24: 162–170. kufeld, r. c. 1994. status and management of moose in colorado. alces 30: 41–44. ———, and d. c. bowden. 1996. movements and habitat selection of shiras moose (alces alces shirasi) in colorado. alces 32: 85–99. laforge, m. p., n. l. michel, a. l. wheeler, and r. k. brook. 2016. habitat selection by female moose in the canadian prairie ecozone. journal of wildlife management 80: 1059–1068. langley, m. a. 1993. habitat selection, mortality and population monitoring of shiras moose in the north fork of the flathead river valley, montana. m.s. 110 status of moose in western us – nadeau et al. alces vol. 53, 2017 thesis, university of montana, missoula, montana, usa. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. levan, i. k., k. a. fox, and m. w. miller. 2013. high elaeophorsis prevalence among harvested colorado moose. journal of wildlife diseases 49: 666–669. matchett, m. r. 1985. habitat selection by moose in the yaak river drainage, northwestern montana. alces 21: 161–190. matthews, p. e. 2012. history and status of moose in oregon. alces 48: 63–66. mccann, n. p., r. a. moen, and t. r. harris. 2013. warm-season heat stress in moose (alces alces). canadian journal of zoology 91: 893–898. mcmillan, j. f. 1953. some feeding habits of moose in yellowstone park. ecology 34: 102–110. monteith, k. l., r. w. klaver, k. r. hersey, a. a. holland, t. p. thomas, and m. j. kauffman. 2015. effects of climate and plant phenology on recruitment of moose at the southern extent of their range. oecologia 178: 1137–1148. murray,d.l.,k.f.hussey,l.a.finnegan, s. j. lowe, g. n. price, j. benson, k. m. loveless, k. r. middel, k. mills, d. potter, a. silver, m.-j. fortin, b. r. patterson, and p. j. wilson. 2012. assessment of the status and viability of a population of moose (alces alces) at its southern range limit in ontario. canadian journal of zoology 90: 422–434. nelson, e. w. 1914. description of a new subspecies of moose from wyoming. proceedings of the biological society of washington 27: 71–74. olterman, j. h., d. w. kenvin, and r. c. kufeld. 1994. moose transplant to southwestern colorado. alces 30: 1–8. pessier, a. p., v. t. hamilton, w. j. foreyt, s. parish, and t. l. mcelwain. 1998. probable elaeophorosis in a moose (alces alces) from eastern washington state. journal of veterinary diagnostic investigation 10: 82–84. peterson, r. l. 1952. a review of the living representatives of the genus alces. contributions of the royal ontario museum of zoology and palaentology 34. royal ontario museum, toronto, ontario, canada. pierce, j. d. 1984. shiras moose forage selection in relation to browse availability in north-central idaho. canadian journal of zoology 62: 2404–2409. ———, and j. m. peek. 1984. moose habitat use and selection patterns in northcentral idaho. journal of wildlife management 48: 1335–1343. ———, b. w. ritchie, and l. kuck. 1985. an examination of unregulated harvest of shiras moose in idaho. alces 21: 231–252. prism climate group. 2016. prism gridded climate data. oregon state university, corvallis, oregon, usa. (accessed october 2016). rempel, r. s. 2011. effects of climate change on moose populations: exploring the response horizon through biometric and systems models. ecological modelling 222: 3355–3365. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. rich, l. n., e. m. glenn, m. s. mitchell, j. a. gude, k. podruzny, c. a. sime, k. laudon, d. e. ausband, and j. d. nichols. 2013. estimating occupancy and predicting numbers of gray wolf packs in montana using hunter surveys. journal of wildlife management 77: 1280–1289. ritchie, b. w. 1978. ecology of moose in fremont county, idaho. wildlife bulletin no. 7. idaho department of fish and game, boise, idaho, usa. rocca, m. e., p. m. brown, l. h. macdonald, and c. m. carrico. 2014. alces vol. 53, 2017 nadeau et al. – status of moose in western us 111 http://prism.oregonstate.edu http://prism.oregonstate.edu climate change impacts on fire regimes and key ecosystem services in rocky mountain forests. forest ecology and management 327: 290–305. ruprecht, j. s., k. r. hersey, k. hafen, k. l. monteith, n. j. decesare, m. j. kauffman, and d. r. macnulty. 2016. reproduction in moose at their southern range limit. journal of mammalogy 97: 1355–1365. samuel, w. m., d. a. welch, and b. l. smith. 1991. ectoparasites from elk (cervus elaphus nelsoni) from wyoming. journal of wildlife diseases 27: 446–451. stevens, d. r. 1970. winter ecology of moose in the gallatin mountains, montana. journal of wildlife management 34: 37–46. storm, d. j., m. d. samuel, t. r. van deelen, k. d. malcolm, r. e. rolley, n. a. frost, d. p. bates, and b. j. richards. 2011. comparison of visualbased helicopter and fixed-wing forwardlooking infrared surveys for counting white-tailed deer odocoileus virginianus. wildlife biology 17: 431–440. tape, k. d., d. d. gustine, r. w. ruess, l. g. adams, and j. a. clark. 2016. range expansion of moose in arctic alaska linked to warming and increased shrub habitat. plos one 11:e0152636. teacher, a. g. f., d. j. griffiths, d. j. hodgson, and r. inger. 2013. smartphones in ecology and evolution: a guide for the app-rehensive. ecology and evolution 3: 5268–5278. timmermann, h. r., and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. toweill, d. e., and g. vecellio. 2004. shiras moose in idaho: status and management. alces 40: 33–43. van dyke, f., b. l. probert, and g. m. van beek. 1995. seasonal habitat use characteristics of moose in south-central montana. alces 31: 15–26. vartanian, j. m. 2011. habitat condition and the nutritional quality of seasonal forage and diets: demographic implications for a declining moose population in northwest wyoming, usa. m.s. thesis, university of wyoming, laramie, wyoming, usa. white, g. c., andb. c.lubow.2002.fitting population models to multiple sources of observed data. journal of wildlife management 66: 300–309. wilson, d. e. 1971. carrying capacity of the key browse species for moose on the north slopes of the uinta mountains, utah. m.s. thesis, utah state university, logan, utah, usa. wolfe, m. l., k. r. hersey, and d. c. stoner. 2010. a history of moose management in utah. alces 46: 37–52. worley, d. e., c. k. anderson, and k. r. greer. 1972. elaeophorosis in moose from montana. journal of wildlife diseases 8: 242–244. 112 status of moose in western us – nadeau et al. alces vol. 53, 2017 status and trends of moose populations and hunting opportunity in the western united states range and habitat population monitoring hunting opportunity, success rates, and trends potential limiting factors future research acknowledgements references alces15_54.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces17_147.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_30.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 analysis of age, body weight and antler spread of bull moose harvested in maine, 1980-2009 haley a. andreozzi1, peter j. pekins1, and lee e. kantar2 1department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa; 2maine department of inland fisheries and wildlife, bangor, maine 04401, usa abstract: age, field-dressed body weight, and antler spread data collected from 11,566 harvested moose (alces alces) were analyzed to assess whether temporal change has occurred in the physical characteristics of bull moose from 1980–2009 in maine. the annual proportion and antler spread of trophy bulls (spread ≥ 137 cm; n = 851) were also analyzed. there was no evidence of a measurable decline in the body weight or antler spread of adult bull moose (≥1.5 years old), similar to findings in vermont and new hampshire in a recent >20 year temporal analysis. there was a slight increase in physical characteristics of yearlings that contrasted with the trend in new hampshire and vermont where it is speculated that parasitism by winter ticks (dermacentor albipictus) reduces growth rate and recruitment by yearlings. the proportion of trophy bulls in the harvest declined proportionally ∼26% (9.3 to 6.9%) as harvest increased >2x from 1980–1987 to 2005–2009; however, the mean spread of trophy bulls declined by only 2% (p = 0.002). additionally, there were no differences (p > 0.05) in the proportion of harvested bulls within each age class between 1980–1987 and 2005– 2009, and the relatively stable proportion of mature bulls (>5 years old) in the harvest across time periods (30–44%) does not suggest selective harvest of older, trophy bulls. in the face of the declining regional population, continued monitoring of harvested moose is warranted to best manage the largest and longest harvested population in the northeastern united states. alces vol. 51: 45–55 (2015) key words: alces alces, bull moose, body weight, antler spread, physical characteristics, trophy, harvest measurement of physical characteristics of harvested moose (alces alces) provides an opportunity to assess temporal trends and relative condition of a moose population. it is usually assumed that a direct relationship exists between habitat quality and physical condition. age-specific body weight of male and female moose should reflect health and production (schwartz and hundertmark 1993). antler measurements are used similarly because of the correlation between antler size and nutritional condition (bubenik 1997). adams and pekins (1995) concluded that yearling moose are useful to estimate overall herd health because their potential growth rate reflects variance in body weight and onset of ovulation. antler morphology in cervids is determined by nutrition and genetics, and antler growth and size are strongly influenced by forage availability, quantity, and quality (schmidt et al. 2007). age also influences the size and formation of antlers as larger, older males invest less in body growth and allocate more resources toward antler growth, symmetry, and size (stewart et al. 2000, bowyer et al. 2001). as body size and age are strongly correlated with antler size and mating success (clutton-brock 1982), dominant males have the ability to limit the 45 mating opportunities of younger males (van ballenberghe and miquelle 1996). hunting may influence ungulate populations by altering age and social structure, sex-ratio, and population dynamics (milner et al. 2006). mortality patterns in harvested populations commonly deviate from those in non-harvested populations, often with an increase in the mortality of prime-aged males (ginsberg and milner-gulland 1994, milner et al. 2006). selective harvest is often applied as a management technique throughout north america to protect adult cow moose and maximize productivity (timmermann 1987), often causing higher harvest of adult bulls. high harvest of older bull moose has the potential to impact normal age structure, and reduce average body size and antler spread in a population over time (solberg et al. 2000); younger, smaller males are eventually predominant in the harvest (schmidt et al. 2007). although hunting for older, large antlered moose can be a local economic stimulant and management tool (monteith et al. 2013), an increasing focus on and popularity of trophy hunting further concentrates harvest on prime bulls (mccullough 1982, timmermann and buss 1997). possible effects of trophy hunting include genetic selection for smaller antlers as well as negative demographic consequences due to other fitness-related genetic traits of trophy males; however, few studies have explored such implications (festa-bianchet and lee 2009). since the initiation of modern moose hunting in 1980, the maine legislature set the moose hunting seasons and harvest levels. the overall goals, developed during the 1985 planning process, were to maintain the moose population at the 1985 level, increase harvest, and maintain viewing opportunities; permits were either sex prior to 1999. since 2001, the maine department of inland fisheries and wildlife (mdifw) has set the moose hunting seasons and harvest levels under a moose management system that describes the decision process and actions necessary to meet population goals and objectives set by a public working group (morris 2002). desired levels of hunting opportunity, viewing opportunity, and road safety are assessed to categorize each wildlife management district (wmd) into either a recreation, road safety, or compromise management area. addressing population goals in a wmd includes determining age structure of harvested animals, age and sex composition from sightings by deer and moose hunters and more recently from helicopter surveys (kantar and cumberland 2013). among other measures, both the proportion of bulls and the percentage of mature bulls (≥5 years old) in each wmd are examined annually and harvest quotas are adjusted to achieve desired levels, a marked shift in management strategy because bull composition was not a prior criteria. a recent >20 year analysis (1988–2009) of physical parameters of harvested moose in new england indicated that body weight and ovulation rate of yearling cows in new hampshire and vermont have decreased; conversely, body weight of yearling cows increased in maine. further, body weight and most antler measurements of harvested bulls in new hampshire and vermont have also declined (bergeron et al. 2013). given this temporal decline in physical characteristics of bulls, there is reason to investigate baseline and trend data in bull moose harvested in maine given the >30 year history of modern moose hunting in maine where harvest has increased from 636 in 1980 to 2,582 in 2011, with higher permit allocations likely to continue (mdifw 2011). importantly, age, antler spread, and body weight of harvested bulls have been measured since 1980. this study provides a temporal assessment of these physical characteristics to identify trends in the relative growth and 46 maine harvest analysis – andreozzi et al. alces vol. 51, 2015 condition of bulls harvested in maine from 1980–2009. the objectives were to assess trends in body weight and antler spread within age classes, the relative proportion of age classes, and the proportion and physical characteristics of trophy bulls in the harvest. study area northern maine is located at the extreme northeast corner of the united states, above 44° 38’ n. it is bordered by quebec and new brunswick to the north, new hampshire to the west, and the atlantic ocean to the south and east. maine is 90% forested and commercial timber harvesting is common throughout the northern portion of the state (hoving et al. 2004). the sub-boreal acadian forest has a mixture of spruce (picea spp.) and balsam fir (abies balsamea) stands and northern hardwood forests; common species include beech (fagus grandi‐ folia), maple (acer spp.), hemlock (tsuga canadensis), birch (betula spp.), spruce, and balsam fir (hoving et al. 2004). harvest data were analyzed for 12 wildlife management districts (wmd; 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, and 19) in a 45,793 km2 area, roughly the northern half of maine (fig. 1). this area contains a high proportion of suitable moose habitat in the form of active commercial forestlands, has had relatively consistent harvest over the study period (1980–2009; l. kantar, pers. comm.), and these wmds represent the core of maine’s moose population (mdifw 2013). fig. 1. locations of maine wildlife management districts (shaded) from which data from harvested bull moose were used to assess temporal trends in physical characteristics, 1980–2009. alces vol. 51, 2015 andreozzi et al. – maine harvest analysis 47 methods biological data collected at moose check stations in 1980–2009 were used to assess temporal trends in the physical characteristics of bull moose. specific measurements were field-dressed body weight, antler spread, and age. field-dressed body weight was defined as the entire carcass weight minus the heart, liver, lungs, and rumenreticulum and was measured on certified scales at registration stations. antler spread was equal to the greatest width (cm) on a plane perpendicular to the skull (l. kantar, pers. comm.). age was determined from cementum annuli counts on cross-sectioned canines (sergeant and pimlott 1959) performed by mdifw biologists. trophy bulls were defined as those with spreads ≥137 cm (54 in) which is similar to the minimum entry for canada moose in the boone and crockett club trophy record-book (boddington 2011). incomplete records of physical characteristics were excluded in order to allow for analysis of the proportional relationships between physical parameters. data were broken into 4 time periods (1980–1987, 1988–1998, 1999–2004, and 2005–2009) to maintain similarity with recent assessments of regional harvest data (adams and pekins 1995, musante et al. 2010, bergeron et al. 2013). data were also analyzed by individual year for some tests; data were unavailable for 1981 (no harvest) and 1985 (data not age-specific). analysis of variance (anova) was used to test for age-specific differences in physical parameters between years and time periods including body weight-age relationships, antler spread-age relationships, age class distribution, and relative condition of the population over time. age classes were 1.5, 2.5, 3.5, 4.5, 5.5, and ≥6.5 years. tukey’s test was used to make pairwise comparisons; significance for all tests was assigned a priori at α = 0.05. results a total of 11,566 harvested moose were included in the data analysis. the number of records per age class ranged from 1169 (5.5 years) to 2860 (≥6.5 years), with sample size increasing in subsequent time periods: 1619 and 1625 in 1980–1987 and 1988– 1998, and 3789 and 4533 in 1999–2004 and 2005–2009, respectively. overall, there was an upward trend in mean body weight of harvested bulls over the 30-year period. between 1980–1987 and 2005–2009, a 4–10% increase in mean body weight occurred in the youngest 4 age classes (1.5–4.5 years old, p ≤ 0.024); minimal change (1–2%, p > 0.05) occurred in the older classes in the same periods (table 1). the current (2005–2009) mean body weight was higher than the 30 year mean in all age classes, with the exception of the 5.5 year age class (table 1). there was no significant difference (p > 0.05) in mean body weight among any time periods in the ≥6.5 year age class. the maximum mean weights occurred in the 1999–2004 time period for the 2.5–5.5 year age classes, and were significantly higher than in other time periods for 2.5 (p ≤ 0.002) and 3.5 year old bulls (p ≤ 0.005) (table 1). maximum mean weight of yearlings (225 kg) occurred in the 2005– 2009 time period (p ≤ 0.02). the 1.5–4.5 year old classes had an overall significant increase (4.0–8.3%, p ≤ 0.014) in mean antler spread between 1980–1987 and 2005–2009, with some variation in the intermediary periods; bulls ≥5.5 years had minimal change (<3.6%, p > 0.05) (table 1). yearlings were the only age class in which the current (2005–2009) mean spread (60 ± 15.9 cm) exceeded the 30 year mean (58 ± 13.6 cm); this age class had the most substantial increase between 1980–1987 and 2005–2009 (8.3%, p = 0.013). though no significant difference (p > 0.05) existed between 1980–1987 and 2005–2009 in the ≥6.5 year age class, spread significantly 48 maine harvest analysis – andreozzi et al. alces vol. 51, 2015 declined 5% (p < 0.000) between 1988–1998 and 2005–2009. the maximum spread occurred in 1999–2004 for 2.5–5.5 year olds and the mean spread was significantly higher than in other time periods for 2.5 (p ≤ 0.003) and 3.5 year olds (p < 0.000) (table 1). there were no significant differences (p > 0.05) in the proportion of harvested bulls within each age class between 1980– 1987 and 2005–2009; some variation occurred within the intermediary periods for each age class (fig. 2). the proportion of yearlings in the harvest declined significantly (64%, p = 0.0003) between 1988–1998 and 2005–2009; conversely, an increase occurred in the 4.5 year age class (28.5%, p = 0.027). a total of 851 harvested trophy bulls (spread ≥ 137 cm) were included in the analysis. the sample size for time periods varied but increased overall, with 151 and 145 in 1980–1987 and 1988–1998, and 238 and 317 in 1999–2004 and 2005–2009, respectively. the mean antler spread of trophy bulls declined 2% (p = 0.002) from 145.7 ± 6.9 cm to 143.3 ± 6.6 cm between 1980–1987 and 2005–2009. there were no significant differences (p > 0.05) in the mean body weight of harvested trophy bulls between 1980–1987 and 2005–2009. the current (2005–2009) mean body weight was higher than the 30 year mean for trophy bulls (395 ± 40.6 kg; table 1). the proportion of trophy bulls declined (∼26%, p = 0.128) from 9.3% in 1980– 1987 to 6.9% in 2005–2009 as the absolute number increased (∼2x) from 151 to 317 animals from 1980–1987 to 2005–2009 (fig. 3, table 1). there was a significant negative relationship between the annual proportion of trophy bulls and year (r2 = 0.14, n = 28, p = 0.03). the mean age of trophy bulls was between 7 and 8.5 years in all time periods with 85–93% ≥5 years old. across all time periods, 5.5–12.5 year olds accounted for 86–92% of all trophy animals (fig. 4). table 1. mean (± sd) field-dressed body weight (kg) and antler spread (cm) of bull moose harvested in select wildlife management districts in maine by time period and age class, 1980–2009. mean body weight and antler spread of harvested trophy bulls are also presented by time period. sample sizes are in parentheses. within age classes, time periods with a letter in common were not significantly different. age 1980–1987 1988–1998 1999–2004 2005–2009 30 year mean body weight (kg) 1.5 217 ± 29 b (196) 214 ± 26 b (410) 218 ± 29 b (573) 225 ± 36 a (420) 219 ± 30 2.5 253 ± 44 d (269) 268 ± 38 c (264) 285 ± 35 a (1035) 279 ± 35 b (896) 278 ± 38 3.5 302 ± 45 c (219) 298 ± 41 c (245) 323 ± 37 a (657) 316 ± 37 b (805) 314 ± 40 4.5 329 ± 48 c (226) 339 ± 44 bc (193) 351 ± 43 a (433) 346 ± 40 ab (696) 344 ± 43 5.5 353 ± 50 b (174) 360 ± 41 ab (152) 366 ± 42 a (337) 360 ± 40 ab (506) 361 ± 43 ≥6.5 374 ± 50 a (535) 372 ± 46 a (361) 374 ± 45 a (754) 374 ± 41 a (1210) 373 ± 44 trophy 397 ± 47 a (151) 392 ± 39 a (145) 396 ± 41a (238) 396 ± 40 a (317) 395 ± 41 antler spread (cm) 1.5 56 ± 13 bc (196) 55 ± 12 c (410) 59 ± 13 ab (573) 60 ± 16 a (420) 58 ± 14 2.5 73 ± 17 c (269) 79 ± 15 b (264) 82 ± 14 a (1035) 79 ± 13 b (896) 80 ± 14 3.5 89 ± 18 c (219) 92 ± 17 bc (245) 97 ± 15 a (657) 92 ± 15 b (805) 94 ± 16 4.5 99 ± 21 c (226) 107 ± 19 ab (193) 109 ± 19 a (433) 104 ± 16 b (696) 105 ± 18 5.5 109 ± 22 b (174) 118 ± 17 a (152) 120 ± 17 a (337) 113 ± 17 b (506) 115 ± 18 ≥6.5 123 ± 19 bc (535) 128 ± 20 a (361) 125 ± 19 b (754) 122 ± 19 c (1210) 124 ± 19 trophy 146 ± 7 ab (151) 146 ± 7 a (145) 144 ± 7 bc (238) 143 ± 7 c (317) 145 ± 7 alces vol. 51, 2015 andreozzi et al. – maine harvest analysis 49 discussion there was no statistical evidence of a measurable change in the physical parameters of bull moose harvested in northern maine from 1980–2009. a minimal upward trend occurred in mean body weight during the 30-year time period as the 2005–2009 mean body weight exceeded the 30-year mean in all age classes (table 1). similarly, a slight overall increase occurred in the 4.5 5.5 6.5 0% 20% 40% 60% 80% 100% h ar ve st ed b ul l m oo se (% ) time period ≥2.5 3.51.5 fig. 2. proportional age structure of bull moose harvested in maine (wmds 1–11 and 19) by time periods, 1980–2009. 0 1 2 3 4 5 6 7 8 9 10 0 300 600 900 1200 1500 1800 2100 2400 2700 1980-1987 1988-1998 1999-2004 2005-2009 p ro po rt io n of h ar ve st ed b ul ls (% ) m oo se h ar ve st ed time period mean harvest proportion of trophy bulls fig. 3. average annual moose harvest (mdifw 2011) and proportion (%) of harvested bull moose considered trophy bulls (spread ≥137 cm) by time period in maine, 1980–2009. 50 maine harvest analysis – andreozzi et al. alces vol. 51, 2015 mean spread of the 4 youngest age classes across the 30-year time period, with some variability but no clear trend in bulls ≥5.5 years old. the lack of declining trends in adult physical characteristics is similar to that measured in nearby vermont and new hampshire and presumably indicates adequate habitat quality (bergeron et al. 2013). however, unlike in vermont and new hampshire, where declines occurred in both body weight and productivity measures in the yearling age class, the physical characteristics of maine yearlings increased slightly indicating variability within the northeastern united states. the downward trend in the proportional harvest within the yearling age class between 1988–1998 and 2005–2009 could indicate a reduction in the proportion of yearlings in the population possibly due to lower recruitment (fig. 2). however, this decline was not coupled with reduced physical parameters that are indicative of a decline in relative health and nutritional status; both body weight and spread increased in yearlings during the 30-year period. numerous factors can influence physical parameters of moose including habitat quality, weather, and disease and parasites. in nearby new hampshire, parasitism by winter ticks (dermacentor albipictus) is considered a primary negative influence on survival and growth of calves and subsequent productivity of yearlings (musante et al. 2010, bergeron and pekins 2014). declining trends in yearling body weight and antler spread in new hampshire and vermont bulls from 1988–2009 (fig. 5) are suggestive of such impact (bergeron et al. 2013). importantly, the core moose habitat in new hampshire and vermont is not considered poor or inadequate based on forest regeneration surveys in both (bergeron et al. 2011, andreozzi et al. 2014), and that commercial forests dominate both areas as in maine. the lack of measurable decline in physical characteristics of adult bulls and slight increase in physical characteristics of yearling bulls in maine from 1980–2009 suggests that parasitism by winter ticks could be less problematic in maine. the majority of the maine study area lies above 0.05 0.10 0.15 0.20 0.25 p ro po rt io n of h ar ve st ed t ro ph y b ul ls age class fig. 4. proportion (%) of harvested trophy bulls (spread ≥137cm) within each age class in maine (wmds 1–11 and 19), 1980–2009. alces vol. 51, 2015 andreozzi et al. – maine harvest analysis 51 44° 38’ n extending as far north as 47° 28’ n, an area further north than the entirety of new hampshire and vermont, both below 45° 18’ n. because abundance of winter ticks and their annual impact are largely determined by length of winter and snow cover (samuel and welch 1991), the core of maine’s moose population may be less influenced by this parasite. the 2% decline in antler spread of trophy bulls is probably not biologically significant, and unlikely to be harvest related as small variation in antler size is often explained by annual weather influences, or variation in population density and uneven sex ratios (solberg and saether 1994). additionally, the relatively stable proportion (30–44%) of bulls >5 years old in the harvest across time periods does not indicate excessive selective harvest pressure towards older, trophy bulls (fig. 2). the majority of trophy bulls (86–92%) are between 5.5 and 12.5 years old in all time periods, with an average age between 7 and 8.5 years (fig. 4). in alaska, spread was maximum in prime age bulls (7–11 years) and declined with senescence at ∼12 years (bowyer et al. 2001). the high proportion of trophy bulls >5 years old and the declining proportion at age 12 in maine suggests that the proportion of trophy bulls in each age class is likely not influenced by harvest pressure, but reflects normal antler growth and maturation, and senescence. most studies with empirical evidence of the effects of trophy hunting on growth of horn-like structures occurs outside of the moose literature; for example, targeted hunting on bighorn trophy rams (ovis canadensis) over a 30-year period resulted in smaller-horned and lighter rams, and fewer trophy animals (coltman et al. 2003). hundertmark et al. (1998) simulated selective harvest for bull moose based on antler size (>127 cm spread) and showed a significant decrease in the frequency of favorable antler alleles; however, empirical evidence of the genetic impact of trophy hunting is rare and such changes are assumed to be undetectable for many generations (harris et al. 2002). age distribution can shift toward younger age classes as harvest intensity increases (jenks et al. 2002). therefore, selective harvest that targets older, larger males can result fig. 5. mean field-dressed body weight (kg) and mean antler spread (cm) of harvested yearling bull moose in maine (wmds 1–11 and 19), vermont, and new hampshire (1988– 2009; bergeron et al. 2013). 52 maine harvest analysis – andreozzi et al. alces vol. 51, 2015 in increased breeding by younger bulls and alter age structure of the population by reducing mean bull age and size over time (mccullough 1982). mdifw determines bull composition by analyzing the age of harvested animals, sightings by deer and moose hunters, and more recently aerial surveys (kantar and cumberland 2013). harvest levels and permit types (i.e., sexspecific) are adjusted annually to maintain desired bull composition levels and limit over-harvest of prime age and mature bulls. for example, in wmds 1–10 and 19, the goal is to maintain 17% mature (≥5 years old) bulls, whereas in wmd 11 it is to maintain a ratio of 60 bulls:100 cows (morris 2002). despite fourfold higher harvest after 30 years of moose hunting in maine (mdifw 2011; fig. 3), the study population has maintained consistent age structure. specifically, there has been no measurable decline in the proportion of harvested bulls ≥6.5 years that would indicate an overall younger age structure due to selective harvest of larger, trophy males (fig. 2). maine’s current moose population estimate is >70,000 moose, and mean annual harvest has increased from 816 in 1980– 1987 to 2239 in 2005–2009 (mdifw 2011, 2012, fig. 3). current harvest is only about 3% of the current population estimate, but will probably increase as hunting interest and moose conflicts increase. while this study indicates that physical characteristics of bull moose in maine have not changed appreciably after 30 years of harvest, understanding the potential and realized influences of harvest on age structure and physical parameters of moose populations is fundamental to proper management. similar harvest analyses have indicated recent declines in body weight, antler measurements, and reproductive rate in moose in nearby vermont and new hampshire (bergeron et al. 2013). these productivity measurements have been collected in maine since 2010 in combination with potvin double-count aerial surveys and age-sex composition flights. integration of these techniques with harvest data will provide the essential data necessary for managing moose under the 3 primary management goals in maine. continued monitoring of physical parameters of harvested moose is warranted to monitor the relative condition and best manage the largest and longest harvested moose population in the northeastern united states. acknowledgements we are grateful to the mdifw for providing the data used in this analysis, and to the mdifw biologists and check station workers who collected data for >30 years. we are also thankful to d. bergeron for providing comparative data. reviewer comments were extremely helpful in improving this paper. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53–59. andreozzi, h. a., p. j. pekins, and m. l. langlais. 2014. impact of moose browsing on forest regeneration in northeast vermont. alces 50: 67–79. bergeron, d. h., and p. j. pekins. 2014. evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire. alces 50: 1–15. ———,———, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. ———,———, and k. rines 2013. temporal assessment of physical characteristics and reproductive status of moose in new hampshire. alces 49: 41–50. boddington, c. 2011. field evaluation for boone and crockett score: alaska, yukon, and canada moose. boone and crockett club, montana, usa. alces vol. 51, 2015 andreozzi et al. – maine harvest analysis 53 bowyer, r. t., k. m. stewart, j. g. kie, and w. c. gasaway. 2001. fluctuating asymmetry in antlers of alaskan moose: size matters. journal of mammalogy 82: 814–824. bubenik, a. b. 1997. behavior. pages 173222 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. clutton-brock, t. 1982. the functions of antlers. behaviour 79: 2–4. coltman, d. w., p. o’donoghue, j. t. jorgenson, j. t. hogg, c. strobeck, and m. festa-bianchet. 2003. undesirable evolutionary consequences of trophy hunting. nature 426: 655–658. festa-bianchet, m., and r. lee. 2009. guns, sheep and genes: when and why trophy hunting may be a selective pressure. pages 94–107 in b. dickson, j. hutton, and w. m. adams, editors. recreational hunting, conservation and rural livelihoods: science and practice. blackwell, west sussex, united kingdom. ginsberg, j. r., and e. milner-gulland. 1994. sex-biased harvesting and population dynamics in ungulates: implications for conservation and sustainable use. conservation biology 8: 157–166. harris, r. b., w. a. wall, and f. w. allendorf. 2002. genetic consequences of hunting: what do we know and what should we do? wildlife society bulletin 30: 634–643. hoving, c. l., d. j. harrison, w. b. krohn, w. j. jakubas, and m. a. mccollough. 2004. canada lynx habitat and forest succession in northern maine, usa. wildlife biology 10: 285–294. hundertmark, k. j., h. thelen, and r. t. bowyer. 1998. effects of population density and selective harvest on anter phenotype in simulated moose populations. alces 34: 375–383. jenks, j. a., w. p. smith, and c. s. deperno. 2002. maximum sustained yield harvest versus trophy management. the journal of wildlife management 66: 528–535. kantar, l. e., and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose abundance in maine. alces 49: 31–39. mccullough, d. r. 1982. antler characteristics of george reserve white-tailed deer. the journal of wildlife management 46: 821–826. maine department of inland fishries and wildlife (mdifw). 2011. history of recreational moose hunting in maine, 1980-2011. maine department of inland fisheries and wildlife, augusta, maine, usa. ———. 2012. maine’s moose population estimated at 76,000 after new survey. maine department of inland fisheries and wildlife, augusta, maine, usa. ———. 2013. report to the joint standing committee on inland fisheries and wildlife: proposed actions for moose management in regards to the number of permits issued, the length and timing of the annual moose hunting season. maine department of inland fisheries and wildlife, augusta, maine, usa. milner, j. m., e. b. nilsen, and h. p. andreassen. 2006. demographic side effects of selective hunting in ungulates and carnivores. conservation biology 21: 36–47. monteith, k. l., r. a. long, v. c. bleich, j. r. heffelfinger, p. r. krausman, and r. t. bowyer. 2013. effects of harvest, culture, and climate on trends in size of horn-like structures in trophy ungulates. wildlife monographs 183: 1–28. morris, k. i. 2002. moose management system. maine department of inland fisheries and wildlife, augusta, maine, usa. musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. 54 maine harvest analysis – andreozzi et al. alces vol. 51, 2015 samuel, w. m., and d. a. welch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27: 169–182. schmidt, j. i., j. m. ver hoef, and r. t. bowyer. 2007. antler size of alaskan moose: effects of population density, hunter harvest and use of guides. wildlife biology 13: 53–65. schwartz, c. c., and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. the journal of wildlife management 57: 454–468. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. the journal of wildlife management 23: 315–321. solberg, e. j., a. loison, b.-e. sæther, and o. strand. 2000. age-specific harvest mortality in a norwegian moose (alces alces) population. wildlife biology 6: 41–52. ———, and saether, b.-e. 1994. male traits as life-history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammalogy 75: 1069–1079. stewart, k. m., r. t. bowyer, j. g. kie, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36: 77–84. timmermann, h. 1987. moose harvest strategies in north america. swedish wildlife research supplement 1: 565–580. ———, and m. buss. 1997. population and harvest management. pages 303-336 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. van ballenberghe, v., and d. miquelle. 1996. rutting behavior of moose in central alaska. alces 32: 109–130. alces vol. 51, 2015 andreozzi et al. – maine harvest analysis 55 analysis of age, body weight and antler spread of bull moose harvested in maine, 1980-,1,6,0,0,0pt,0pt,0pt,0pt study area methods results discussion acknowledgements references alces18_lxiiiminutesmeeting.pdf alces vol. 18, 1982 alces18_235.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces18_54.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces18_168.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 44_front_cover v2.pdf instructions for contributors to alces sentence, in which case it is spelled out. italics should only use the name-and-year system to cite published literature. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table (i.e., the space bar). illustrations page. identify each illustration by printing the author’s if necessary, also indicate the orientation of the illustration on . letters and the same the manuscript. photographs must be of high contrast and typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original gerald alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented shop, but data originality, ideas, analyses, interpretation, accuracy, explained in a covering letter and invoice sent to authors with galley proofs. authors should alces”, alces, vol. 34 (1): 1998 updates are posted on the alces please provide an electronic copy of the manushould maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. justify all text sions should be handled similarly. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlethe author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (< next line. following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. . footnotes use only in tables and at the bottom of the differs from the address at the time of the study. names of domesticated animals or cultivated plants. use système international d’unités (si) units and symbols. use alces15_1.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 45, 2009 west – moose conservation in wildlife refuges 59 moose conservation in the national wildlife refuge system, usa robin l. west u.s. fish and wildlife service, kenai national wildlife refuge, p.o. box 2139, soldotna, ak 99669, usa abstract: the national wildlife refuge system in the united states includes about 150 million acres of lands and waters within 550 refuges managed for conservation. a variety of laws, regulations, and management polices help ensure these areas will be preserved for future generations. in a web-based survey, 35 refuges reported having established populations of moose (alces alces) within their boundaries with nearly 40 million acres of moose habitat, 99% in alaska. the 4 recognized subspecies of moose in north america were represented on refuges found in 12 states. approximately 39,000 moose were reported inhabiting refuges in the usa; about 38,000 in alaska. only 9 refuges used management practices specifically to benefit moose, primarily prescribed or wildland fire. moose populations on refuges varied greatly and refuge managers reported numerous concerns including climate change, illegal harvest, habitat loss or degradation, parasites, disturbance, moose-vehicle collisions, predators, and both recreational and subsistence hunting. future management implications of these issues are discussed. alces vol. 45: 59-65 (2009) key words: alces alces, climate change, management, moose, national policy, survey, wildlife refuges. the national wildlife refuge system (nwrs) was created in 1903 when president theodore roosevelt set aside the first refuge at pelican island, florida. today 550 refuges, at least one in each of the 50 states, encompass approximately 150 million acres of lands and waters and are managed for conservation of fish, wildlife, and plants as part of the nwrs in the u.s. department of interior’s fish and wildlife service (u.s. fish and wildlife service 2009). the mission of the nwrs, formalized with the passage of “organic legislation” in 1997 that amended the national wildlife refuge administration act of 1966 is: “… to administer a national network of lands and waters for the conservation, management, and where appropriate, restoration of the fish, wildlife, and plant resources and their habitats within the united states for the benefit of present and future generations of americans.” national wildlife refuges (nwr) support a diverse variety of wildlife species including the 4 recognized north american subspecies of moose as described by bubenik (1997). alaska has the majority of the acreage, the most moose habitat, and the most moose within the nwrs. kenai national wildlife refuge, nearly 2 million acres of boreal forest located in south central alaska, was established as the kenai national moose range by executive order in 1941 and was managed specifically to conserve and protect moose until the refuge’s purposes were expanded in 1980. predator management and enforcement against poaching dominated early activities at the kenai national moose range, but by the 1960s management efforts became more focused on habitat conservation and treatments. methods and objectives a web-based survey was employed early moose conservation in wildlife refuges west alces vol. 45, 2009 60 in 2008 to gather information to 1) better understand the role of the nwrs in moose conservation, and 2) identify the most important issues or constraints facing management of moose on refuges. the survey was developed at surveymonkey.com and included 21 questions including basic questions about the refuge (i.e., size, purpose, date established) and moose-specific questions regarding the abundance, habitat, harvest, and management of moose. refuge biologists and managers were notified of the request through regional biologists in 5 of the 8 administrative regions of the u.s. fish and wildlife service that were within the range of moose. reminders were provided to help ensure near complete responses, and follow-up contacts were made where clarification was needed. results thirty-nine refuges from 12 states responded to the survey. one refuge, kodiak nwr in alaska, reported that moose had been introduced but were no longer present. three refuges, the charles m. russell in montana, rachel carson in maine, and rydell in minnesota reported only incidental sightings of moose, but that healthy populations occurred in nearby areas. the remainder (35 refuges) reported moose as occupying refuge lands on a regular basis (table 1, fig. 1). the total combined area of refuges reporting the presence of moose was 72,024,112 acres; the estimated moose habitat was 39,599,769 acres (55%). these areas were not based on quantifiable data but were estimates that generally eliminated unsuitable moose habitat such as glaciers and alpine tundra. the vast majority (approximately 99%) of the total acreage and suitable moose habitat was on refuges in alaska. population estimates (n = 33; 4 refuges in alaska reported as 2 since fig. 1. location of national wildlife refuges in the united states reporting the presence of moose, 2007. alces vol. 45, 2009 west – moose conservation in wildlife refuges 61 refuge name state refuge size (acres) percent moose habitat population estimate population status annual harvest estimate alaska peninsula ak 3,563,489 & 81% 2,500 ? 51-100 becharof ak 1,200,060 arctic ak 19,286,322 26% 1,000 ? 51-100 innoko ak 3,850,481 100% 3,700 10-50 izembek ak 311,076 10% 101 + <10 kanuti ak 1,430,160 52% 588 ? 10-50 kenai ak 1,912,425 89% 3,481 301-400 koyukuk ak 3,550,160 & 100% 15,000 0 301-400 nowitna ak 1,560,000 selawik ak 2,150,162 100% 2,100 0 51-100 tetlin ak 700,059 96% 1,272 + 51-100 togiak ak 4,101,178 30% 1,600 0 51-100 yukon delta ak 19,162,297 20% 4,700 + 301-400 yukon flats ak 8,633,385 100% 2,500 0 201-300 arapaho co 23,271 19% 20 0 0 bear lake id 18,086 22% 5 0 0 camas id 10,578 60% 8 0 grays lake id 20,125 50% 12 0 0 kootenai id 2,774 100% 12 ? <10 moosehorn me 28,874 100% 20 + <10 sunkhaze meadows me 11,217 87% 25 ? 0 seney mi 93,245 95% 50 ? 0 agassiz mn 61,501 73% 33 ? 0 glacial ridge mn 2,360 50% 7 ? 0 lost trail mt 8,834 38% 5 + 0 red rock lakes mt 68,810 84% 93 + <10 des lacs nd 19,547 51% 10 0 0 j. clark salyer nd 59,376 42% 40 0 0 lostwood nd 26,904 100% 8 0 0 upper souris nd 32,302 69% 4 + 0 lake umbagog nh 13,173 50% 100 0 <10 silvio o. conte vt 26,574 100% 85 10-50 little pend oreille wa 42,594 100% 10 + <10 turnbull wa 17,935 64% 10 0 0 national elk wy 24,778 20% 20 0 table 1. summary data from 39 national wildlife refuges in the united states reporting the presence of moose, 2007. the population status was described as: 0 = stable, + = increasing, = decreasing, and ? = unknown. moose conservation in wildlife refuges west alces vol. 45, 2009 62 they are managed as complexes) were derived from expert opinion (5), incidental observation (11), refuge-specific aerial surveys (12), or state agency aerial surveys (5). population status was reported as increasing (8), decreasing (5), stable (12), and unknown (8). this assessment was made by expert opinion (9), incidental observation (12), or statistical trend analysis (12). hunting occurred on 20 refuges. the estimated annual harvest on refuges that allowed hunting ranged from <10 (6), 10-50 (3), 51-100 (5), 201-300 (1) and 301-400 (3); only alaskan refuges reported harvests >10 moose. the harvest at izembek nwr in alaska was <10 moose with an increasing population. most moose reported at izembek were on an adjacent unit of the alaska peninsula nwr but were managed by izembek. all alaskan refuges that allow hunting have both recreational and subsistence hunting, but harvest estimates were not differentiated by type (table 1). only 9 of 35 refuges reported specific management actions to benefit moose such as prescribed fire (8), re-vegetation (2), willow cutting (1), rest area from grazing (1), and wildland fire use (2). wildland fire use is the practice of allowing naturally ignited fire to burn for resource benefits and differs from prescribed fire that is a management-ignited fire for resource benefit. new terminology being used to describe various strategies to suppress all or part of a wildfire, or permit portions to burn, equate all fire management decisions other than prescribed fire as “appropriate management response” so future habitat treatment by fire management decisions may be more difficult to track. some refuges historically used crushing or chaining to set back forest succession to benefit moose, but these techniques are not currently employed. follow-up conversations with biologists and managers revealed that fire is generally considered cheaper, more ecologically acceptable, and more effective than mechanical treatments. kenai nwr reported that nearly 60,000 acres received vegetation treatment for habitat improvement in 1960-2008 (21,697 acres of mechanical, 4,863 acres of prescribed fire, and 29,638 acres of wildland fire use) with wildland fire use accounting for nearly all acreage treated in the past 5 years. the current assessment of moose habitat on refuges (n = 33) included improving (7), stable (12), declining (5), and unknown (9) conditions. these assessments were largely reported as qualitative (84%) not quantitative. the most important issues or management concerns about moose on refuges were climate change (13), habitat degradation (12), illegal harvest (11), subsistence hunting (10), recreational hunting (8), parasites (5), habitat loss (3), disturbance (2), and moose-vehicle collisions (2). fifteen refuges reported “other” that included practical and political issues involving predators, coordination and education with rural users, drought, vegetation management, practical fire management programs, and reliable population surveys. all issues were identified by at least one manager in both alaska and the lower 48 states except subsistence hunting was identified as a management concern only in alaska; parasites were identified as a concern by lower 48 refuge managers only. discussion refuge managers are charged with achieving specific refuge purposes and the mission of the nwrs. management of the nwrs has evolved from the beginning of the 20th century when refuges were viewed as inviolate sanctuaries, and little or any public use was allowed – to post-world war ii when refuges were managed increasingly for multiple uses – to the current era (post-1997) when refuges are managed primarily for wildlife. human uses are allowed only when such uses are compatible with (i.e., do not materially interfere with or detract from) refuge purposes and the nwrs mission. additionally, in 1997 alces vol. 45, 2009 west – moose conservation in wildlife refuges 63 congress mandated that wildlife-dependent recreational uses (i.e., hunting, fishing, wildlife viewing, photography, and outdoor education and interpretation) were appropriate uses of the nwrs and should be permitted if compatible. it is this mandate that may help ensure opportunity for moose hunting well into the future, though this applies mostly to alaska where the majority of moose and moose habitat occur in the nwrs. kenai nwr has a specific purpose to provide for wildlife-oriented recreation including hunting. the other 15 alaskan refuges have a specific purpose to provide continued opportunities for subsistence hunting and fishing, but also have the general mandate to permit hunting and other wildlife-dependent recreational uses whenever compatible with other purposes. while subsistence hunting in alaska is administered differently under federal law than state managed recreational hunting, ample opportunity exists for both user groups. this dual management program began in 1990 and has resulted in frequent philosophical debate and legal challenge, but no significant conservation concerns have developed to date. the survey indicated no identifiable trend in the status of moose regionally or by state. informal discussions with refuge managers suggest that site-specific habitat variables probably drive moose numbers more than any other factor; however, there were a few exceptions, such as concern over the role of parasites in the population decline of moose in agassiz nwr in minnesota. habitat treatment on refuges is guided by a number of factors including the legal purposes for establishing the area, and other legal mandates, policies, and economics. wilderness designations by congress are relatively new protective layers applied to areas in certain refuges, as well as portions of some national parks, national forests, and bureau of land management lands. wilderness designations provide legal protection from development such as logging, mining, oil and gas extraction, and road building, but also limit the intensive management options of managers. the legal guidelines for wilderness management require natural processes to dominate, but active management may be used to restore or help facilitate natural processes, prevent loss of species, or be implemented in case of emergency. when active management is to be undertaken, or where mechanization is necessary to access the area or complete the proposed work in designated wilderness, federal policies require that the minimum tool practical be employed to successfully complete the task. wilderness designations may prevent some managers from undertaking active moose management practices, but the long-term additional protection given to these areas should ultimately benefit moose and other wildlife and outweigh any detriment from lost management flexibility. refuge management emphasis has also changed to include broader purposes and attention to wildlife diversity from earlier years when certain refuges were established as game ranges such as hart mountain national antelope refuge in oregon, national bison range in montana, national elk refuge in wyoming, and kenai national moose range in alaska. this is especially true in alaska where the majority of moose and moose habitat occurs within the nwrs. in 1980 congress passed the alaska national interest lands conservation act (anilca) which expanded the 7 existing refuges and created 9 new ones, establishing approximately 77 million acres in the nwrs (about 50%). the primary management purpose established by anilca for all alaska refuges was: “to conserve fish and wildlife populations and habitats in their natural diversity …”. anilca also emphasized specific species for which the areas were primarily known. moose were specifically mentioned in 8 refuges: kenai, alaska peninsula, innoko, kanuti, moose conservation in wildlife refuges west alces vol. 45, 2009 64 koyukuk, nowitna, tetlin, and yukon flats. however, the stated emphasis was clearly not exclusive and does not justify management activities benefiting highlighted species while clearly harming other species. the overall goal of managing the largely pristine alaskan refuges is to preserve natural diversity and natural processes. this has provided some unique challenges, but has largely been realized in the 29 years since anilca was enacted. the long-term prognosis is less certain given the increasing issues associated with climate change. this is no simple phenomena but rather a threat of ecosystem level change within decades rather than centuries. increased prevalence of wildfire, drying of lakes and wetlands, elevated levels of forest insect outbreaks, rising tree lines, and melting of glaciers and permafrost are some of the potential effects of climate warming (wiles et al. 1995, klein et al. 2005, berg and anderson 2006, berg et al. 2006, dial et al. 2007, wiles et al. 2008). perhaps most notable in alaska is the predicted shift from a largely spruce (picea spp.)-dominated forest to a deciduous forest because of a projected increase in the fire cycle (chapin et al. 2003, rupp and mann 2005). such a shift should substantially favor moose, but could drastically reduce suitability of large areas to caribou (rupp et al. 2006). while warmer climates (and increased prevalence of fire) may benefit moose, other factors may have the opposite effect such as the emergence of parasitic infections (kutz et al. 2004). refuge managers charged with maintaining natural diversity need to have meaningful philosophical discussions to accompany data gathering, economic analyses, and management planning actions. first, there must be a common understanding of what is “natural diversity” if it is to be a management goal, followed by a decision of whether climate change is natural or anthropomorphic. if anthropomorphic, this could logically justify actions to prevent, reverse, or restore losses reasonably linked to climate change; however, practical consideration of social, technological, and economic issues should be addressed in long-term landscape management. because of the potential magnitude of ecological change, managers may have little choice other than documenting habitat changes and tracking plant and animal diversity, particularly in large remote refuges. it seems evident that if significant climate change is realized in the short-term, species composition in ecological communities will change and moose populations, habitat, and range will likely shift. acknowledgements i am grateful to all the nwrs staff from the 39 refuges that provided information for this summary, and wish to offer special thanks to dr. eric taylor and dr. joel reynolds who helped devise the survey instrument. references berg, e. e. and r. s. anderson. 2006. fire history of white and lutz spruce forests on the kenai peninsula, alaska, over the last two millennia as determined from soil charcoal. forest ecology management 227: 275-283. _____., j. d. henry, c. l. fastie, a. d. de volder, and s. m. matsuoka. 2006. spruce beetle outbreaks on the kenai peninsula, alaska, and kluane national park and reserve, yukon territory: relationship to summer temperatures and regional differences in disturbance regimes. forest ecology management 227: 219-232. bubenik, a. b. 1997. evolution, taxonomy and morphology. pages 77-123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. chapin, f. s., t. s. rupp, a. m. starfield, l. dewilde, e. s. zaveleta, n. fresco, j. henkleman, and a. d. mcguire. 2003. alces vol. 45, 2009 west – moose conservation in wildlife refuges 65 planning for resilience: modeling change in human-fire interactions in the alaskan boreal forest. frontiers in ecology and the environment 1: 255-261. dial, r. j., e. e. berg, k. timm, a. mcmahon, and j. geck. 2007. changes in the alpine forest-tundra ecotone commensurate with recent warming in southcentral alaska: evidence from orthophotos and field plots. journal of geophysical research biogeosciences 112: g04015. klein, e., e. e. berg, and r. dial. 2005. wetland drying and succession across the kenai peninsula lowlands, south-central alaska. canadian journal of forestry research 35: 1931-1941. kutz, s. j., e. p. hoberg, j. nagy, l. polley, and b. elkin. 2004. “emerging” parasitic infections in arctic ungulates. integrative and comparative biology 44: 109-118. rupp, t. s., and d. h. mann. 2005. development of a computer model for management of fuels, human-fire interactions, and wildland fires in the boreal forest of alaska. final report to joint fire science program governing board. _____, m. olson, l. g. adams, b. w. dale, k. joly, j. henkleman, w. b. collins, and a. m. starfiled. 2006. simulating the influences of various fire regimes on caribou winter habitat. ecological society of america 16: 1730-1743. u. s. fish and wildlife service. 2009. annual report of lands under control of the u.s. fish & wildlife service as of september 30, 2008. in press. wiles, g. c., p. e. calkin, and a. post. 1995. glacier fluctuations in the kenai fords, alaska, u.s.a: an evaluation of controls on iceberg-calving glaciers. arctic and alpine research 27: 234-245. _____, d. j. barclay, p. e. calkins, and t. v. lowell. 2008. century to millennialscale temperature variations for the last two thousand years indicated from glacial geologic records of southern alaska. global and planetary change 60: 115125. alces17_95.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces16_392.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces17_257.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 untitled-1 alces suppl. 2, 2002 balciaskas modeling of moose hunting 23 modeling of moose hunting: protection of cows with twins linas p. balciaskas institute of ecology, lithuanian academy of sciences, 232021, akademijos 2, vilnius, lithuania abstract: i developed a simulation model to evaluate moose hunting strategies. the model incorporated age and sex-specific schedules for natural (nonhunting) mortality and fecundity. i evaluated 2 strategies for harvesting cow moose. in the first, cow moose were harvested irrespective of whether they were accompanied by calves. in the second, hunters were not allowed to kill cows accompanied by twin calves. simulation results indicated that 500 calves/1000 cows could be saved under the second harvest strategy. alces supplement 2: 23-26 (2002) key words: hunting, moose, numerical simulation, twins moose are one of the most important game species in the ussr. in the european portion of the ussr, the goal of most game managers is to increase moose numbers. traditional hunters in these regions do not understand or accept the concept of a selective harvest. for this reason, simulation modeling was used to evaluate the effect of different harvest strategies. in the first strategy, cow moose were harvested regardless of whether they were accompanied by calves. in the second, hunters were not allowed to kill cows accompanied by twin calves. based on a comparison of the number of calves and embryos in cow moose harvested in the kostroma region (baskin, unpublished data), i assumed that individual cows consistently produce either single calves or twins. in contrast, moose density is too high in all parts of the lithuanian republic. the model was used to help determine the best strategy to reduce moose density while retaining the sex and age structure of the population. in particular, a strategy was needed that would preserve bulls in the 5.5 – 8.5 year age classes for trophy hunting, maximize meat production, and shooting opportunities. these needs can be addressed using an optimizing model (lopatin and rosolovsky 1990). methods simulations were run on a dvk-3 microcomputer with os rt11sj and the model code was written in pascal. the model tallied moose numbers after each of a sequence of discrete events: natural mortality, hunting mortality, and calving. the postcalving population was then subjected to that sequence in the next iteration of the model. through tabulations, we tracked the sex and age structure of the population, proportions of cows with 0, 1, and 2 calves, and carcass weights in each sex and age category. in modeling the harvest in lithuania, a harvest quota was set, then the antlerless portion (including calves) was set, and finally, the age distribution in each category was identified. simulations ignored the potential effects of weather and nutrition. output from the simulations was displayed on the screen, printed out, or saved as a separate file. for evaluating the twin protection strategies discussed above, i modeled only the modeling of moose hunting balciaskas alces suppl. 2, 2002 24 table 1. initial set of parameters for moose population model. age percent percent percent percent percent carcass class females in females not females with females mortality mass (kg) (years) population pregnant one calf with twins max min 0.5 20.0 100.0 0.0 0.0 25.0 72.0 65.8 1.5 13.0 75.0 25.0 0.0 10.0 128.8 120.6 2.5 10.0 70.0 25.0 5.0 5.0 153.7 143.0 3.5 10.0 40.0 50.0 10.0 3.0 168.7 158.6 4.5 8.0 20.0 50.0 30.0 3.0 187.6 166.3 5.5 7.0 15.0 50.0 35.0 3.0 204.3 167.5 6.5 6.0 20.0 40.0 40.0 3.0 210.6 170.0 7.5 4.0 20.0 45.0 35.0 3.0 211.9 172.0 8.5 3.0 30.0 40.0 30.0 3.0 218.7 177.0 9.5 3.0 35.0 45.0 20.0 3.0 220.0 178.0 10.5 3.5 40.0 40.0 20.0 3.0 220.0 182.0 11.5 3.0 42.0 30.0 28.0 3.0 224.7 185.0 12.5 3.0 45.0 35.0 20.0 3.0 228.0 192.0 13.5 3.0 50.0 35.0 20.0 3.0 230.0 206.0 14.5 2.5 60.0 30.0 10.0 10.0 220.0 200.0 15.5 1.0 85.0 10.0 5.0 100.0 210.0 195.0 female portion of the population. i assumed that natural mortality, pregnancy rates, and fecundity rates remained constant throughout the simulations. furthermore, i assumed that hunting maintains a stable population of mature cows and only the proportion of cows with single or twin calves varies. the probability for a cow to have a single calf was set as p1 (the probability of twins for the same cow was p2 = 1-p1), and the probability that that cow would have a single calf again next year was p3 (thus, the probability that a cow would twin in consecutive years was p4 = 1-p3). as indicated earlier, if p2 is increased, p4 should increase as well because it is assumed that the ability to bear twins is inherited and individual to each cow. biological parameters of the model no single source of data was available for all 16 age classes used in the model, and the initial set of population parameters used in the model were a composite from several areas and authors. several sources of data were used: baleisis (1973, 1977) and the society of hunters and fisherman (baleisis and butautas 1987, 1988) [lithuania]; filinov (1983) and kozlo (1983) [various regions of the ussr]; and sylven et al. (1979) [sweden]. the starting population contained 1,000 moose divided among 16 age classes (table 1). natural mortality was set at 3%, harvest mortality at 20%, and i assumed that equal numbers of cows and bulls were harvested. the sex ratio of newborn calves was 108 male calves/100 female calves. alces suppl. 2, 2002 balciaskas modeling of moose hunting 25 the probability p1 was modeled as varying from 0.4 to 0.8 and p3 varied from 0.3 to 0.7. precise values are unknown. based on data from the kostroma region, we can assume the probability of twins recurring is 1.7, so p1 equals 0.3. however, more data must be analyzed to determine a precise probability level of twins recurring. we modeled female harvest ratios of 5%, 10%, 15%, and 20%. results and management implications the value of protecting cows with twins was determined by comparing the number of calves born during 10 years of the first and second harvest strategies. this difference was expressed as a percentage of (a) the number of calves born under the first strategy and (b) as a percentage of the initial number of cows (table 2). when the harvest rate ranged from 10 to 15% of the mature cows, the protection of twins resulted in 3-5% more calves, compared to a strategy of no protection. over a 10-year period, this could result in an increase of 300-500 individuals. with higher harvest rates, even more calves are recruited; however, the population may become saturated with twin-bearing cows. the harvest was limited to yearling cows, the only age class that would produce a single calf. simulations with the 1987 harvest data from lithuania (table 3) indicate that moose density would remain stable for the first 5 years and that there would be a constant output of meat (table 4). overharvest of cows in the 6.5 age-class, however, results in a female-biased harvest in the sixth year. moose numbers will decrease insignificantly but provide a higher yield of meat. simulations with the 1988 harvest data indicate that moose numbers will decline 4% in the first 5 years and provide 25 tons of meat annually/1,000 moose. overharvest of the population will occur later. the harvest quota can be increased up table 2. results of protecting cows with twins. percent of cows harvested 5 10 15 20 p1 p3 a b a b a b a b 0.8 0.7 0.33 3.6 0.666 7.2 0.98 10.8 1.32 14.4 0.6 0.65 7.2 1.29 14.4 1.94 21.6 2.58 28.8 0.5 0.95 10.8 1.91 21.6 2.86 32.4 3.81 43.2 0.4 1.25 14.4 2.50 28.8 3.75 43.2 5.00 57.6 0.3 1.54 18.0 3.08 36.0 4.62 54.0 6.15 72.0 0.7 0.6 0.45 5.4 0.90 10.8 1.35 16.2 1.81 21.6 0.5 0.88 10.8 1.77 21.6 2.65 32.4 3.53 43.2 0.4 1.30 16.2 2.59 32.4 3.89 48.6 5.18 64.8 0.3 1.69 21.6 3.38 43.2 5.07 64.8 6.76 86.4 0.6 0.5 0.56 7.2 1.11 14.4 1.67 21.6 2.22 28.8 0.4 1.08 14.4 2.16 28.8 3.24 43.2 4.32 57.6 0.3 1.58 21.6 3.16 43.2 4.74 64.8 6.32 86.4 0.5 0.4 0.64 9.0 1.29 18.0 1.94 27.0 2.58 36.0 0.3 1.25 18.0 2.50 36.0 3.75 54.0 5.00 72.0 modeling of moose hunting balciaskas alces suppl. 2, 2002 26 table 4. results of moose population modeling according to the harvest data from lithuania. according to 1987 data according to 1988 data year number number harvest meat number number harvest meat alive born yield1 alive born yield1 max min max min max min max min 1 517 506 135 121 192 25.3 513 508 131 120 192 25.7 2 519 506 129 117 195 25.5 513 502 126 114 194 25.8 3 517 503 126 116 195 25.5 509 498 123 114 193 25.8 4 511 492 119 110 194 25.4 502 488 117 109 191 25.6 5 502 477 114 103 191 25.2 485 472 105 100 188 25.2 1x 1000 kg. to 10% or more with an increased harvest of yearlings and a reduced harvest of cows. it should be stressed that for successful moose population management using the numerical simulation models, precise knowledge of the sex and age structure of the population is needed. gathering these data is a high priority for lithuanian game managers. references baleisis, p. r., and v. butautas. 1987. ungulate harvest during the 1986–1987 hunting season, data from antler inspections. the society of hunters and fishermen, vilnius, lithuania. , and . 1988. ungulate harvest during the 1987–1988 hunting season, data from antler inspections. the society of hunters and fishermen, vilnius, lithuania. baleisis, r. 1973. biology and forestation: significance of moose in lithuania. dissertation paper no. 29. (in russian). . 1977. moose. moskslas, vilnius, lithuania. filonov, k.p. 1983. moose and forest industry. moscow, russia. (in russian). kozlo, p.g. 1983. ecological/morphological analysis of moose population. science and technology. minsk, belarus. (in russian). lopatin, v., and s. rosolovsky. 1990. mathematical analysis of the efficient exploitation of moose populations. 5th congress vto:105–106. (in russian). sylven, s., m. aspers, j.-å. eriksson, and m. wilhelmson. 1979. regulated harvesting of the moose population — a simulation study. report 33. swedish university of agricultural sciences, uppsala, sweden. table 3. moose harvest data from lithuania (% of quota). year age class 0.5 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5 10.5 11.5 12.5 13.5 14.5 15.5 1987 32.7 11.5 14.4 13.4 6.8 8.7 4.0 2.1 2.4 1.2 1.7 0.5 0.4 0.1 0.1 – 1988 36.7 11.5 10.3 10.4 6.7 7.8 4.5 3.8 2.9 2.2 2.0 0.5 0.6 0.2 0.2 0.1 f:\alces\supp2\pagema~1\rus 19s alces suppl. 2, 2002 mikhyeva and gaross – moose in latvia 85 moose in latvia and intensive game management practices ruta v. mikhyeva and vitauts g. gaross research and production association “silava,” riga, latvia abstract: historical population trends of moose in latvia and current information on moose population size, sex and age ratios, annual increment rates, and mortality factors are presented. the authors review moose antler quality, interspecific competition, food habits, and discuss forest damage by moose. a management framework for regulating moose harvests in accordance with carrying capacity, under conditions of intensive forestry, is outlined. alces supplement 2: 85-88 (2002) key words: alces alces, cervus elaphus, ecology, intensive forestry, interspecific competition, latvia, moose, population composition, red deer little information is available in the published literature on moose (alces alces) in latvia. the purpose of this paper is to present some general background information on moose ecology within this region and discuss the role of moose management within latvia’s intensive forestry program. past and present moose population status moose have been common over the land area of present–day latvia since the end of the glacial era. data on moose populations for the last 5 centuries are scanty. indirect evidence, however, indicates that moose were highly valued and populations were large enough to supply people with meat and hides. moose numbers decreased sharply by the end of the 18th century, and 100 years later it was assumed there were no more than 1,000– 2,000 moose. more reliable data indicated there were only 85 moose in 1923 and about 1,000 in 1940. the post–world–war ii period was distinguished by a marked increase in moose all over latvia. the highest number, according to official information, was recorded in 1973 (21,830). however, more reliable methods showed these estimates were incorrect and, in most cases, underestimated the actual size of the moose population. the official numbers represent, at best, only rough estimates of population size. follow– up investigations, using more accurate methods, estimated the number of moose in 1975 at approximately 45,000 or 22 moose/1,000 ha of forest land. on some forestry enterprises and forest ranges, this figure reached 40 moose/1,000 ha. the total harvest between 1954 and 1988 was 111,829 moose. the moose population started declining after 1975 (fig. 1). a reliable estimate of the 1989 spring population was 16,000–17,000, or 6 moose/1,000 ha of forest. in a number of localities, the population density ranged from 1 to 5 moose/1,000 ha of forest. the adult sex ratio of moose between 1935 and 1937 varied from 1 male/female to 1 male/1.7 females. similar sex ratios were recorded in 1963, when there was practically no harvest and the impact of predators was insignificant. these sex ratios are believed typical for latvia. sex ratios in 1975, 1978, and 1989 were 1 male/female, 1 male/0.9 female, and 1 male/1.3 females, moose in latvia – mikhyeva and gaross alces suppl. 2, 2002 88 bark chewing results in the spruce stem becoming infected by fungal diseases, and die–backs occur within 5–15 years. intensive moose–forestry management forestry remains one of the cornerstones of latvia’s national economy, yielding timber worth 240 million rubles (1 us$ ≈ 29 russian rubles) annually, in addition to other forest products. yet moose management runs counter to the forest yield–management practices on which modern forestry should be based. one of the principal reasons why hunting quotas for moose were raised during the 1970s, and the harvest increased (fig. 1), was to reduce moose– related forest damage. unfortunately, because of indecisive and conservative attitudes to the problem in question, as well as inaccuracies in the population estimates, this action was delayed for 5–8 years. this resulted in the forest sector suffering tremendous losses, which will require at least 60–70 years to repair. despite the decline in the moose population after 1975, intensive harvesting continued until 1989. biologists and land managers believe this was another mistake because moose harvest should have been reduced. detailed analyses of the population data and the occurrence of forest crop damage show the so–called “silviculturally optimum” moose population density for the forest of latvia in general, or individual forestry enterprises (covering 30–50,000 ha of forest), to be invalid. the estimate of 5–10 moose/ 1,000 ha of forest, as an index of carrying capacity under conditions of intensive forestry, is considered a very rough estimate. all the factors affecting moose abundance such as the number of predators (mainly wolves), food availability, weather, climate, and man’s activities should be known when estimating carrying capacity. these factors are extremely dynamic and are considered relatively stable only on smaller areas (around 10,000 ha for latvia). in order to harmonize management of the moose–forest system, the following data should be accumulated annually: the occurrence of forest crop damage; moose population size, sex ratio, and annual increment rates; and the number of moose dying from harvest, predation, accidents, diseases, and other factors. the general trend in population density should be maintained upward with the following principle kept in mind: the moose population should be large enough to utilize the annual increment produced by its natural forage, without harming forestry interests. in practice, management proceeds by allowing the population to grow when there is no visible damage to the forest crops; i.e., the kill is further reduced from the previous year and the harvest is set less than or equal to the annual increment. if crop damage is increasing, hunting quotas are raised. moose densities are thus reduced when required to avoid overutilization of natural forage and to reduce forest crop damage. in summary, a prolonged period of intensive hunting on moose in latvia, based upon the hunting techniques used (mainly enclosures), has had no long–term adverse impacts on the moose population. the population estimate for 1989 indicates that the cutbacks currently practiced should be halted and management should change to a policy of population control by implementing the principles discussed above. in this respect, methods that stimulate the growth of natural moose forage are of greatest importance. only an integrated approach will resolve the problem between moose and intensive forestry and increase the overall productivity of the forest biogenocenosis. f:\alces\supp2\pagema~1\rus14s. alces suppl. 2, 2002 kuvshinov – central nervous system of moose 63 histopathological changes in the central nervous system of moose (alces alces) in the ivanovo region of russia vadim l. kuvshinov ivanovo agricultural institute, ivanovo, russia abstract: this work describes histopathological changes in the central nervous system of moose (alces alces) of different ages from various regions of the ivanovo district. moose are affected in certain parts of the world by a variety of infectious agents, such as anthrax, rinderpest, necrobacillosis, and foot and mouth disease, which is contracted by coming into contact with reindeer. all of these agents are capable of causing serious disease. our present work attempted to detect less conspicuous forms of disease in moose that might reflect disturbances or degradation of ecological systems comprising their habitat. we particularly noted lesions in the central system of moose that were characterized by nonpurulent meningo–encephalitis, edema, perivasculitis in extracellular spaces, and focal ischaemic necrosis because of thrombosis of small vessels. such changes presumably resulted from neuro–dynamical, vascular, physical–chemical, and fluid disturbances and were pronounced in the central nervous system of moose inhabiting regions with unfavorable environmental conditions. we observed the most serious changes in regions where chemical weed–killers, pesticides, and mineral fertilizers were irrationally applied. alces supplement 2: 63-64 (2002) key words: biosphere, encephalitis, extracellular edema, herbicides, insecticides, moose, perivasculitis, pollution, tolerance results were obtained from a study of histological changes observed in the central nervous system of moose in relation to the use of agricultural chemicals in different regions of ivanovo district. the brains of different–aged moose were obtained from animals shot by hunters; females predominated over males. routine tissue samples taken for histological investigation included the following parts of the brain: medulla, pons, cerebellum, quadrigeminal bodies, epithalamus, thalamus, hypothalamus, some parts of the cerebral cortex, and the vascular plexus of the lateral ventricles. methods the tissues were fixed in 10% formalin, imbedded in paraffin and thin sections were prepared and stained with haematoxylin– eosin. frozen sections were prepared with specific histochemical stains, including the spilmayer method for myelin, the naut– zeindlow method for nerve fiber degeneration, alexandrovskaya’s modification of the orthega and miyagava method for microglial cells, and the gold–sublimate method of ramon and kahan for glial astrocytes. results and discussion changes observed in the medulla differed according to weight and location. the soft cerebral membrane was thickened by edema and infiltrating lymphoid cells. peri– and endovasculitis were visible in the white matter in pyramidal tracts and consisted mostly of macrophages and plasma cells. in the areas of the thin fasciculi (goll), the wedge–shaped fasciculi (burdah), and the main ventral fasciculi, there was less perivasculitis but some small glial nodules, and central nervous system of moose – kuvshinov alces suppl. 2, 2002 64 demyelinated nerve fibers; mononuclear cells were present in the spinal canal. atypical changes were observed in the varolii pons. instead of a nidus cell reaction of the mononuclear type, there were diffuse accumulations of analogous cells. they were located near the soft cerebral membrane, chiefly in the regions of the posterior long fasciculi, the nuclei of the abducent vestibular and trigeminus nerves, under the ependyma, in the area of the cochlear nerve nuclei, the gover’s fasciculi, and in the base of the cerebellum. here and there, mononuclear infiltrates extended from the sub– apsidimal zone into the more profound parts of the white and gray matter. infiltrations of the white and gray matter by diffuse and focal cell accumulations were characteristic of some moose. such accumulations consisted of lymphatic–type cells, mostly glial cells and macrophages. proliferating macroglial cells and some hypotrophic astrocytes were located primarily along vessels. some animals had a lymphocytic perivasculitis in the epiphysis cerebri. this histological investigation of the central nervous system of moose living in unfavorable ecological regions allowed us to conclude that pollution of the environment can disturb internal homeostasis; in particular, it provokes irreversible changes in the brain. alces14_editorialcommitteevi.pdf alces vol. 14, 1978 alces20_3.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces16_attendancelist.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_203.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces14_56.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alcessupp1_105.pdf alces16_314.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alcessupp1_132.pdf alces14_141.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 seasonal variation of nutritional hormones in captive female moose cory j. stantorf1,2, c. loren buck1,3, duane h. keisler , william b. collins , and donald e. spalinger1 1university of alaska anchorage, anchorage, alaska 99508, usa; university of missouri, columbia, mo 65211, usa; alaska department of fish and game, palmer, alaska 99645, usa abstract: the health status of animals may be inferred from the patterns of hormonal concentrations and other chemical characteristics in blood samples. baseline endocrine data representing the nutritional and reproductive condition of moose are currently unknown. in this study, we examined the seasonal patterns of 3 nutritional hormones (leptin, ghrelin, insulin-like growth factor-1) in 3 captive, non-pregnant female moose (alces alces) fed a maintenance diet from november to august. plasma concentrations for leptin, ghrelin, and igf-1 averaged 1.36 ± 0.81 ng/ml, 0.229 ± 0.110 ng/ml, and 114.0 ± 30.5 ng/ml, respectively; only ghrelin displayed a seasonal change. plasma ghrelin concentration was significantly elevated (p < 0.001) during winter months suggesting it may be sensitive to seasonal changes and indicative of nutritional status. alces vol. 53: 53–64 (2017) key words: alces alces, ghrelin, hormones, hpg axis, igf-1, leptin, moose, reproduction, season northern latitudes are characterized by extreme seasonal differences in temperature, photoperiod, forage availability, and forage quality (chapin et al. 1980, risenhoover 1989). as a consequence, large herbivores such as moose (alces alces) may experience seasonal nutritional constraints that can impact their health and fecundity (cook et al. 2001, tollefson et al. 2010). in south-central alaska, pregnancy and calving rates of moose are positively associated with autumn body condition (testa and adams 1998). in northern alaska, heavier caribou (rangifer tarandus) in autumn were more likely to have a successful pregnancy (cameron et al. 1993), suggesting that a threshold body condition must be attained to trigger reproduction (thomas 1982). however, this evidence is largely circumstantial and fails to provide a mechanistic connection between body condition and reproduction. our study focused on 3 hormones leptin, ghrelin, and insulinlike growth factor-1 (igf-1) that relay satiety information to the central nervous system (cns) and influence reproduction. leptin, an adipose-derived hormone (zhang et al. 1994) and product of the ob gene, is secreted by white adipose tissue in direct proportion to the adiposity of an animal (delavaud et al. 2002, geary et al. 2003) and influences appetite, energy expenditure, and reproductive function (zieba et al. 2008, friedman 2010). correlations between plasma leptin concentration and adiposity are established in various animal species including cattle (bos taurus; block et al. 2001) and sheep (ovis aries; blache et al. 2000). receptors for leptin exist at each tier of the present address: alaska department of fish and game, anchorage, alaska 99518, usa. present address: northern arizona university, flagstaff, arizona 86011, usa. 53 2 3 4 5 4 5 hypothalamus-pituitary-gonadal (hpg) axis, suggesting a mechanism of action on these target tissues by leptin (brann et al. 2002). ghrelin is a gut-derived hormone synthesized and secreted by the abomasal and ruminal tissues in ungulates (hayashida et al. 2001, geary et al. 2003). it is the natural ligand of the growth hormone secretagogue (a substance eliciting release of another substance) receptor, and elicits the release of growth hormone (gh) from the pituitary (shintani et al. 2001). ghrelin is a potent neuroendocrine integrator, influencing a myriad of endocrine and non-endocrine functions (barreiro and tena-sempere 2004, fernandezfernandez et al. 2006). circulating ghrelin levels are elevated above baseline concentrations prior to feeding, and return rapidly to baseline after re-alimentation in sheep (sugino et al. 2002) and rats (rattus norvegicus; toshinai et al. 2001). high ghrelin levels stimulate appetite, decrease energy expenditure, and hinder reproduction in several species (tena-sempere 2005, budak et al. 2006). igf-1, similar to ghrelin and leptin, is a peptide hormone that is influenced by nutrition (daftary and gore 2005). in mammals, igf-1 is primarily produced by the liver (gluckman et al. 1991) and is synthesized and released in response to the presence of gh (ketelslegers et al. 1995). in steers on a low plane of nutrition, igf-1 concentrations are reduced and thought to help conserve energy during times of negative energy balance (breier et al. 1986). it exerts positive effects on bone growth, protein synthesis, and somatic growth, and likely plays a role in regulating reproduction given that gonadotropin releasing hormone (gnrh) neurons in the brain express both igf-1 and igf-1 receptors (suttie and webster 1995, daftary and gore 2005). receptors for leptin, ghrelin, and igf-1 are widely distributed throughout the body, including the hypothalamus, pituitary, and ovaries, indicating possible direct and indirect actions at all levels of the hpg axis (gluckman et al. 1991, hodgkinson et al. 1991, brann et al. 2002, tena-sempere 2005, zhang et al. 2008). studies on sheep, rats, and cattle suggest that leptin and ghrelin are necessary for normal secretion patterns of gnrh (hileman et al. 2000, zieba et al. 2003) and pulsatile rhythms of luteinizing hormone (lh; hileman et al. 2000, nagatani et al. 2000). moreover, evidence exists that leptin, ghrelin, and igf-1 help regulate embryo implantation (kawamura et al. 2003) and timing of puberty (ahima et al. 1997, chehab at al. 1997). collectively, leptin, ghrelin, and igf-1 appear to communicate nutritional status to the cns, which in turn, influences reproductive function (tenasempere 2005, budak et al. 2006). understanding the role of nutritional hormones in regulating appetite, energy expenditure, and reproduction could provide an important tool for assessing nutritional and reproductive status of wild herbivores. assessment of this potential role requires accurate assays and knowledge of baseline seasonal concentrations of these hormones. to date, leptin and ghrelin have not been characterized in moose. our objectives were to: 1) adapt and validate assays for leptin, ghrelin, and igf-1 in moose, 2) provide baseline measures for leptin, ghrelin and igf-1, and 3) examine the influence of season on leptin, ghrelin, and igf-1 in moose on a maintenance diet. methods this study was conducted at the university of alaska fairbanks (uaf) matanuska experiment farm (mef) near palmer, alaska (61º 33´ 57 n, 149º 15´ 05 w). the captive moose were maintained throughout the study in fenced 4-ha paddocks with access to a lake for water. design and species our study was designed to provide baseline seasonal profiles of the 3 nutritional 54 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 hormones in captive, tractable female moose (5 to 9 years of age) over a 10-month period beginning in november 2009. all animals were hand-raised orphans collected from the matanuska-susitna borough and municipality of anchorage in south-central alaska, and were conditioned to the experimental protocol such that moving the animals and drawing blood required neither sedation nor restraint. they were fed a nutritionally balanced pelleted ration (alaska mill & feed supply, anchorage, alaska) at a maintenance level of 1.25% of body weight. they were provided ad libitum access to ensiled brome (bromus inermis) hay and pasture, and had access to native forages in the paddock including paper birch (betula papyrifera), balsam poplar (populus balsamifera), quaking aspen (populus tremuloides), scouler willow (salix scouleriana), rose (rosa acicularis), and low bush cranberry (vaccinium oxycoccos). availability of native forage was limited due to previous overbrowsing by captive moose. blood sampling and processing the moose were moved once monthly from their paddocks to individual 4 x 20 m stalls where blood samples were collected immediately prior to the morning feeding, after which they were returned to the paddock. blood (6 cc) was drawn in a few seconds from the jugular vein through a 21 gauge needle attached to a 10 cc syringe; this procedure was not stressful to the animals. blood samples were evacuated into ethylenediaminetetraacetic acid (edta) treated blood collection tubes (7 ml, 12 mg edta). the blood was mixed by inversion and centrifuged at 4000-x g for 10 min to separate the plasma and cellular components. the plasma fraction was transferred with a long-stemmed pasteur pipette into 5 ml cryovials (fischer scientific, usa) that were immediately transferred to a cooler con‐ taining ice packs for same day transport to the laboratory. each sample was then aliquoted into 5, 1.5 ml snap cap vials (fischer scientific, usa) and stored at -80 ºc until assayed. all methods were approved by the university of alaska anchorage (protocol #181596-2) and university of alaska fairbanks (protocol #182744-2) institutional animal care and use committees. hormone assays and validation hormones were assayed using either enzyme-linked immunosorbent (eia) or radioactive immunosorbent assays (ria). all plasma samples were assayed in duplicate, and each assay contained a pooled plasma sample for tests of interand intra-assay variation. ghrelin levels were measured with a commercially available, double-antibody, ria kit (ghrt-89hk, millipore total ghrelin, st. charles, michigan, usa) following the manufacturer’s recommended protocol with a slight modification to account for low sample volumes. the radioactive pellet was counted on a laboratory technologies genesys gamma counter (genii model ltl1010, maple park, illinois, usa). kits were validated using standard tests of parallelism and accuracy on pooled plasma samples. serial dilutions of pooled samples and standards exhibited parallelism for each respective hormone; specifically, ghrelin had a parallelism r2 value of 0.983. tests of accuracy for ghrelin had an r2 = 0.980 with a minimum detectable concentration of 0.093 ng/ml. intraand inter-assay coefficients of variation (cv) for ghrelin were 12.45 and 9.71%, respectively. leptin and igf-1 concentrations were measured in triplicate with a competitive, liquid-liquid phase, double-antibody leptin/ igf-1 ria at the university of missouri. the leptin ria followed the procedure established for cattle (delavaud et al. 2000) and modified for use with rabbit anti-ovine leptin primary anti-serum #7105. minimum alces vol. 53, 2017 stantorf et al. – nutritional hormones in captive moose 55 detectable concentration was 0.1 ng/tube, and interand intra-assay cvs were 5%. the igf-1 ria followed the procedure established previously for cattle (lalman et al. 2000). minimum detectable concentration was 1.5 ng/tube, and interand intra-assay cvs were < 6%. descriptive statistics were computed (spss 17.0, armonk, new york, usa) for each animal and hormone. to test for seasonal differences, winter was defined as november through april, and summer as may through august; these intervals were chosen to fit with astronomical seasons and temporal change in forage quality. seasonal concentrations of hormones were analyzed in r 3.1.3 using a one-way anova; all differences were considered significant at α ≤ 0.05. results plasma leptin concentration averaged 1.36 ± 0.81 ng/ml (table 1) and mean individual concentrations ranged from 0.73–1.99 ng/ml. although no significant seasonal differences were found (f1,28 = 0.0710, p = 0.792), individual leptin concentrations varied widely across time (table 1, fig. 1). plasma ghrelin concentrations averaged 0.229 ± 0.110 ng/ml (table 1) with individual concentrations ranging from 0.081–0.337 ng/ ml. mean ghrelin plasma concentrations were significantly higher in winter than summer (f1,28 = 42.5, p < 0.001; table 1, fig. 2); concentrations declined in all moose in may, and except for 1 animal (ar) in june, remained < 0.200 ng/ml in june-august (fig. 2). plasma concentrations of igf-1 averaged 114.0 ± 30.5 ng/ml (table 1) with mean concentrations of individuals ranging from 84.7–149.0 ng/ml; no seasonal difference was found (f1,28 = 3.32, p = 0.079; fig. 3). discussion leptin leptin concentrations in this study were lower than those reported in domestic cattle and sheep (chilliard et al. 1998), but similar to the lower range (1.20 – 2.63 ng/ml) measured in male iberian red deer (cervus elaphus hispanicus; gaspar-lópez et al. 2009) and reindeer (rangifer tarandus; soppela et al. 2008). in addition, plasma leptin concentrations in non-pregnant red deer (cervus elaphus) hinds (scott 2011) were higher than those of our moose, indicating that captive moose are in the lower range reported for other table 1. mean monthly hormone concentrations with standard errors for leptin, ghrelin, and igf-1 measured in 3 captive female moose at the matanuska experimental farm in palmer, alaska, november to august 2010. leptin (ng/ml) ghrelin (ng/ml) igf-1 (ng/ml) month n �x se �x se �x se nov 3 1.43 0.11 0.257 0.067 100.0 7.0 dec 3 1.12 0.23 0.337 0.025 119.0 10.2 jan 3 1.09 0.28 0.303 0.032 103.0 14.9 feb 3 1.90 0.57 0.271 0.014 84.7 17.1 mar 3 1.58 0.75 0.301 0.031 111.0 23.8 apr 3 1.26 0.42 0.321 0.033 116.0 25.6 may 3 1.42 0.54 0.157 0.065 134.0 21.7 jun 3 1.43 0.81 0.170 0.057 149.0 8.6 jul 3 1.31 0.57 0.100 0.038 129.0 5.5 aug 3 1.10 0.60 0.081 0.029 91.4 14.9 56 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 cervids. although moose typically gain body fat during summer and fall and deplete these reserves over winter (schwartz et al. 1987, franzmann and schwartz 1998), we did not expect the captive moose to exhibit these typical fluctuations in body condition given their high quality diet. the highly variable leptin concentrations observed throughout the year suggest that leptin may not be regulated solely by adiposity level, and as a result, is not an adequate singular measure of fat mass in moose. while not investigated here, other studies indicate that leptin may respond to the overall nutritional status of an animal and its environment, rather than adiposity (daniel et al. 2002), a possible explanation for the absence of seasonal change. in addition to nutritional status, photoperiod might be a seasonal cue capable of modifying leptin concentrations in ruminants and non-ruminants (bocquier et al. 1998, chilliard et al. 2005, soppela et al. 2008). for example, leptin concentration in well-nourished reindeer declined in early winter as animals maintained body weight and feed intake, suggesting that photoperiod, rather than feed intake and quality, plays an important role in controlling leptin concentration (soppela et al. 2008). further, leptin declined in ovariectomized ewes exposed to short days regardless of nutritional status and changes in adipose mass (bocquier et al. 1998). it is reasonable to expect that photoperiod could influence leptin concentrations in our moose given the extreme changes in daylength at northern latitudes. fig. 1. seasonal concentrations of leptin in 3 captive moose fed a maintenance diet at the matanuska experimental farm in palmer, alaska, november to august 2010. each letter combination represents an individual moose. alces vol. 53, 2017 stantorf et al. – nutritional hormones in captive moose 57 ghrelin ghrelin is a gut-derived hormone that is one of many influences on appetite, fattening, and reproduction via input to the hpg axis (gentry et al. 2003, tena-sempere 2008). ghrelin fluctuates in response to change in the short-term nutritional state of animals, with plasma concentrations significantly elevated in food restricted animals (gualillo et al. 2002, bradford and allen 2008). mean ghrelin concentrations were lower in our moose compared to ad libitum fed steers (0.229 vs 0.123 ng/ml) and holstein heifers (wertz-lutz et al. 2006, field et al. 2013), and slightly lower than mean ghrelin concentrations in fasted holstein heifers (field et al. 2013). we expected ghrelin concentration to remain stable because of the consumption of a high quality maintenance diet. conversely, given the relationship between ghrelin and feeding and gut fill, we expect that concentrations would fluctuate seasonally in wild animals. ghrelin concentration in our captive moose varied seasonally, with higher prefeeding concentrations in winter than summer months for all animals, suggesting that in moose, ghrelin may not solely respond to changes in gut fill or appetite, but may be influenced by seasonal changes in energy expenditure and/or other physiological processes. this contrasts with research on rats (toshinai et al. 2001) and angus steers (wertz-lutz et al. 2008) that indicated ghrelin levels in food-deprived or food-restricted animals are significantly higher than in fed animals, presumably due to energy restriction. additionally, the preprandial ghrelin surge fig. 2. seasonal concentrations of ghrelin in 3 captive moose fed a maintenance diet at the matanuska experimental farm in palmer, alaska, november to august 2010. each letter combination represents an individual moose. 58 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 in suffolk rams can be modified by feeding restriction (sugino et al. 2002). a seasonal response would correspond with seasonal changes in forage quality and consumption level, and facilitate elevated ghrelin concentrations during those seasons with poor forage quality and reduced consumption and energy expenditure. we believe ghrelin concentration is reflective of a longer temporal window, not just the immediate period prior to feeding. the ghrelin spike in moose “ar” during june remains inexplicable. insulin-like growth factor-1 igf-1 is essential in stimulating bone growth and protein synthesis (suttie and webster 1995). additionally, igf-1 receptors are located throughout the body including sites on gnrh neurons, suggesting that it can influence reproduction via neuroendocrine pathways (daftary and gore 2005). concentrations in the current study did not change between winter and summer months, and individual peak concentrations were similar to those in muskoxen (ovibos moschatus; adamczewski et al. 1997) and reindeer (suttie and webster 1995). the overall mean concentration was similar to those reported in wild moose sampled from late fall through late winter in south-central and southeast alaska (90.7–135.0 ng/ml; parillo 2010). in contrast, although captive muskoxen had similar mean concentrations as our moose, wild muskoxen had lower concentrations (adamczewski et al. 1997). reindeer had higher peak levels overall (bubenik et al. 1998), with their bottom range similar to our values. fig. 3. seasonal concentrations of igf-1 in 3 captive moose fed a maintenance diet at the matanuska experimental farm in palmer, alaska, november to august 2010. each letter combination represents an individual moose. alces vol. 53, 2017 stantorf et al. – nutritional hormones in captive moose 59 the absence of any seasonal change in igf-1, and the minor/lack of change in body condition supports the idea that igf-1 in moose is influenced, in part, by active and energetically expensive physiological processes as with other species (e.g., the active rebuilding of body stores and growth). other ungulate species exhibit increasing igf-1 concentration in conjunction with growth in young animals, restoration of lean body mass in adults, increase in photoperiod, and the timing of forage quality and quantity (kerr et al. 1991, suttie et al. 1991, adamczewski et al. 1992, ditchkoff et al. 2001). in contrast, parillo (2010) reported higher igf-1 concentrations in winter than fall in wild alaskan moose, although the winter values were based on a single, rather than multiple collections across the entire season. conclusion our results indicate that leptin, ghrelin, and igf-1 can be measured with accuracy and precision in moose using standard protocols with certain modifications. plasma leptin concentrations were highly variable between and within moose and without a seasonal pattern over the 10-month period, suggesting that leptin is not regulated singularly by adiposity, but by other hormonal/physiological mechanisms. in contrast, plasma ghrelin concentrations exhibited a seasonal pattern over the same period, with higher concentrations measured during winter months; however, it is difficult to determine if this was an artifact of sample size or physiological response. lastly, igf-1 concentrations reflected neither monthly nor seasonal changes; given our small animal sample, it could not be determined if this was due to individual variation or a seasonal response. using any of these hormones as definitive indicators of nutritional status or reproductive condition of moose is not appropriate because levels varied considerably over time and among individuals. further investigations are warranted given the complexity of hormonal regulation in the physiological and reproductive processes in moose. acknowledgements this research was supported by the alaska department of fish and game, lgl, the university of alaska anchorage’s chancellor’s award for research, and the center for global climate change and arctic systems research at the university of alaska fairbanks. we thank the buck lab for their help with analyzing samples and assay troubleshooting, dr. d. keisler and his lab for their help with hormone analysis, dr. d. saalfeld and dr. t. lohuis for reviewing and providing constructive feedback on this manuscript, and the uaf agriculture farm for housing our captive moose and use of their laboratory. references adamczewski, j. z., a. gunn, b. laarveld, and p. flood. 1992. seasonal changes in weight, condition, and nutrition of free-ranging and captive muskox females. rangifer 12: 179–183. ———, s. c. tedesco, b. laarveld, and p. f. flood. 1997. seasonal patterns in growth hormone, insulin and insulinlike growth factor-1 in female muskoxen. rangifer 17: 131–134. ahima, r. s., j. dushay, s. n. flier, d. prabakaran, and j. s. flier. 1997. leptin accelerates the onset of puberty in normal female mice. journal of clinical investigation 99: 391–395. barreiro, m. l., and m. tena-sempere. 2004. ghrelin and reproduction: a novel signal linking energy status and fertility? molecular and cellular endocrinology 226: 1–9. blache, d., r. l. tellam, l. m. chagas, m. a. blackberry, p. e. vercoe, and g. b. martin. 2000. level of nutrition affects leptin concentrations in plasma 60 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 and cerebrospinal fluid in sheep. journal of endocrinology 165: 625–637. block, s. s., w. r. butler, r. a. ehrhardt, a. w. bell, m. e. van amburgh, and y. r. boisclair. 2001. decreased concentration of plasma leptin in periparturient dairy cows is caused by negative energy balance. journal of endocrinology 171: 339–348. bocquier, f., m. bonnet, y. faulconnier, m. guerre-millo, p. martin, and y. chilliard. 1998. effects of photoperiod and feeding level on perirenal adipose tissue metabolic activity and leptin synthesis in the ovariectomized ewe. reproduction nutrition development 38: 489–498. bradford, b. j., and m. s. allen. 2008. negative energy balance increases periprandial ghrelin and growth hormone concentrations in lactating dairy cows. domestic animal endocrinology 34: 196–203. brann, d. w., m. f. wade, k. m. dhandapani, v. b. mahesh, and c. d. buchanan. 2002. leptin and reproduction. steroids 67: 95–104. breier, b. h., j. j. bass, j. h. butler, and p. d. gluckman. 1986. the somatotrophic axis in young steers: influence of nutritional status on pulsatile release of growth hormone and circulating concentrations of insulin-like growth factor 1. journal of endocrinology 111: 209–215. bubenik, g. a., d. schams, r. g. white, j. rowell, j. blake, and l. bartos. 1998. seasonal levels of metabolic hormones and substrates in male and female reindeer (rangifer tarandus). comparative biochemistry and physiology part c: pharmacology, toxicology and endocrinology 120: 307–315. budak, e., m. f. sánchez, j. bellver, a. cerveró, c. simón, and a. pellicer. 2006. interactions of the hormones leptin, ghrelin, adiponectin, resistin, and pyy3–36 with the reproductive system. fertility and sterility 85: 1563–1581. cameron, r. d., w. t. smith, s. g. fancy, k. l. gerhart, and r. g. white. 1993. calving success of female caribou in relation to body weight. canadian journal of zoology 71: 480–486. chapin, f. s., d. a. johnson, and j. d. mckendrick. 1980. seasonal movement of nutrients in plants of differing growth form in an alaskan tundra ecosystem: implications for herbivory. journal of ecology 68: 189–209. chehab, f. f., k. mounzih, r. lu, and m. e. lim. 1997. early onset of reproductive function in normal female mice treated with leptin. science 275: 88–90. chilliard, y., f. bocquier, and m. doreau. 1998. digestive and metabolic adaptations of ruminants to undernutrition, and consequences on reproduction. reproduction nutrition development 38: 131–152. ———, c. delavaud, and m. bonnet. 2005. leptin expression in ruminants: nutritional and physiological regulations in relation with energy metabolism. domestic animal endocrinology 29: 3–22. cook, r. c., d. l. murray, j. g. cook, p. zager, and s. l. monfort. 2001. nutritional influences on breeding dynamics in elk. canadian journal of zoology 79: 845–853. daftary, s. s., and a. c. gore. 2005. igf-1 in the brain as a regulator of reproductive neuroendocrine function. experimental biology and medicine 230: 292–306. daniel, j. a., b. k. whitlock, j. a. baker, b. steele, c. d. morrison, d. h. keisler, and j. l. sartin. 2002. effect of body fat mass and nutritional status on 24-hour leptin profiles in ewes. journal of animal science 80: 1083–1089. delavaud, c., f. bocquier, y. chilliard, d. h. keisler, a. gertler, and g. kann. 2000. plasma leptin determination in ruminants: effect of nutritional status and body fatness on plasma leptin concentration assessed by a specific ria in sheep. journal of endocrinology 165: 519–526. alces vol. 53, 2017 stantorf et al. – nutritional hormones in captive moose 61 ———, a. ferlay, y. faulconnier, f. bocquier, g. kann, and y. chilliard. 2002. plasma leptin concentration in adult cattle: effects of breed, adiposity, feeding level, and meal intake. journal of animal science 80: 1317–1328. ditchkoff, s. s., l. j. spicer, r. e. masters, and r. l. lochmiller. 2001. concentrations of insulin-like growth factor-i in adult male white-tailed deer (odocoileus virginianus): associations with serum testosterone, morphometrics and age during and after the breeding season. comparative biochemistry and physiology part a: molecular & integrative physiology 129: 887–895. fernandez-fernandez, r., m. tena-sempere, v. m. navarro, m. l. barreiro, j. m. castellano, e. aguilar, and l. pinilla. 2006. effects of ghrelin upon gonadotropin-releasing hormone and gonadotropin secretion in adult female rats: in vivo and in vitro studies. neuroendocrinology 82: 245–255. field, m. e., s. e. deaver, r. p. rhoads, r. j. collier, and m. l. rhoads. 2013. effects of prolonged nutrient restriction on baseline and periprandial plasma ghrelin concentrations of postpubertal holstein heifers. journal of dairy science 96: 6473–6479. franzmann, a. w., and c. c. schwartz. 1997. ecology and management of the north american moose. first edition. smithsonian institution press, washington, d. c., usa. friedman, j. m. 2010. leptin and the regulation of body weight. hamdan medical journal 3: 131–135. gaspar-lópez, e., j. casabiell, j. a. estevez, t. landete-castillejos, l. f. de la cruz, l. gallego, and a. j. garcía. 2009. seasonal changes in plasma leptin concentration related to antler cycle in iberian red deer stags. journal of comparative physiology b 179: 617–622. geary, t. w., e. l. mcfadin, m. d. macneil, e. e. grings, r. e. short, r. n. funston, and d. h. keisler. 2003. leptin as a predictor of carcass composition in beef cattle. journal of animal science 81: 1–8. gentry, p. c., j. p. willey, and r. j. collier. 2003. ghrelin, a growth hormone secretagogue, is expressed by bovine rumen. journal of animal science 81: 123. gluckman, p. d., r. g. douglas, g. r. ambler, b. h. breier, s. c. hodgkinson, j. b. koea, and j. h. f. shaw. 1991. the endocrine role of insulin like growth factor i. acta pædiatrica 80: 97–105. gualillo, o., j. e. caminos, r. nogueiras, l. m. seoane, e. arvat, e. ghigo, f. f. casanueva, and c. dieguez. 2002. effect of food restriction on ghrelin in normal-cycling female rats and in pregnancy. obesity research 10: 682–687. hayashida, t., k. murakami, k. mogi, m. nishihara, m. nakazato, m. s. mondal, y. horii, m. kojima, k. kangawa, and n. murakami. 2001. ghrelin in domestic animals: distribution in stomach and its possible role. domestic animal endocrinology 21: 17–24. hileman, s. m., d. d. pierroz, and j. s. flier. 2000. leptin, nutrition, and reproduction: timing is everything. journal of clinical endocrinology and metabolism 85: 804–807. hodgkinson, s. c., g. s. g. spencer, j. j. bass, s. r. davis, and p. d. gluckman. 1991. distribution of circulating insulinlike growth factor-i (igf-i) into tissues. endocrinology 129: 2085–2093. kawamura, k., n. sato, j. fukuda, h. kodama, j. kumagai, h. tanikawa, a. nakamura, y. honda, t. sato, and t. tanaka. 2003. ghrelin inhibits the development of mouse preimplantation embryos in vitro. endocrinology 144: 2623–2633. 62 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 kerr, d. e., j. g. manns, b. laarveld, and m. i. fehr. 1991. profiles of serum igf-i concentrations in calves from birth to eighteen months of age and in cows throughout the lactation cycle. canadian journal of animal science 71: 695–705. ketelslegers, j. m., d. maiter, m. maes, l. e. underwood, and j. p. thissen. 1995. nutritional regulation of insulinlike growth factor-i. metabolism 44: 50–57. lalman, d. l., j. e. williams, b. w. hess, m. g. thomas, and d. h. keisler. 2000. effect of dietary energy on milk production and metabolic hormones in thin, primiparous beef heifers. journal of animal science 78: 530–538. nagatani, s., y. zeng, d. h. keisler, d. l. foster, and c. a. jaffe. 2000. leptin regulates pulsatile luteinizing hormone and growth hormone secretion in the sheep. endocrinology 141: 3965–3975. parillo, a. a. 2010. the effects of nutrient availability and season on the somatotropic axis in free-ranging alaskan moose (alces alces). m.s. thesis, university of connecticut, storrs, connecticut, usa. risenhoover, k. l. 1989. composition and quality of moose winter diets in interior alaska. journal of wildlife management 53: 568–577. schwartz, c. c., w. l. regelin, and a. w. franzmann. 1987. seasonal weight dynamics of moose. swedish wildlife research (supplement) 1: 301–310. scott, i. c. 2011. voluntary food intake of pregnant and non-pregnant red deer hinds during the gestating period. ph.d. thesis, lincoln university, lincoln, new zealand. shintani, m., y. ogawa, k. ebihara, m. aizawa-abe, f. miyanaga, k. takaya, t. hayashi, g. inoue, k. hosoda, m. kojima, k. kanagawa, and k. nakao. 2001. ghrelin, an endogenous growth hormone secretagogue, is a novel orexigenic peptide that antagonizes leptin action through the activation of hypothalamic neuropeptide y/y1 receptor pathway. diabetes 50: 227–232. soppela, p., s. saarela, u. heiskari, and m. nieminen. 2008. the effects of wintertime undernutrition on plasma leptin and insulin levels in an arctic ruminant, the reindeer. comparative biochemistry and physiology part b: biochemistry and molecular biology 149: 613–621. sugino, t., y. hasegawa, y. kikkawa, j. yamaura, m. yamagishi, y. kurose, m. kojima, k. kangawa, and y. terashima. 2002. a transient ghrelin surge occurs just before feeding in a scheduled meal-fed sheep. biochemical and biophysical research communications 295: 255–260. suttie, j. m., and j. r. webster. 1995. extreme seasonal growth in arctic deer: comparisons and control mechanisms. american zoologist 35: 215–221. ———, r. g. white, b. h. breier, and p. d. gluckman. 1991. photoperiod associated changes in insulin-like growth factor-i in reindeer. endocrinology 129: 679–682. tena-sempere, m. 2005. exploring the role of ghrelin as novel regulator of gonadal function. growth hormone and igf research 15: 83–88. ———. 2008. ghrelin as a pleotrophic modulator of gonadal function and reproduction. nature clinical practice endocrinology and metabolism 4: 666–674. testa, j. w., and g. p. adams. 1998. body condition and adjustments to reproductive effort in female moose (alces alces). journal of mammology 79: 1345–1354. thomas, d. c. 1982. the relationship between fertility and fat reserves of peary caribou. canadian journal of zoology 60: 597–602. tollefson, t. n., l. a. shipley, w. l. myers, d. h. keisler, and n. dasgupta. 2010. influence of summer and autumn nutrition on body condition and reproduction in lactating mule deer. alces vol. 53, 2017 stantorf et al. – nutritional hormones in captive moose 63 journal of wildlife management 74: 974–986. toshinai, k., m. s. mondal, m. nakazato, y. date, n. murakami, m. kojima, k. kangawa, and s. matsukura. 2001. upregulation of ghrelin expression in the stomach upon fasting, insulin-induced hypoglycemia, and leptin administration. biochemical and biophysical research communications 281: 1220–1225. wertz-lutz, a. e., j. a. daniel, j. a. clapper, a. trenkle, and d. c. beitz. 2008. prolonged, moderate nutrient restriction in beef cattle results in persistently elevated circulating ghrelin concentrations. journal of animal science. 86: 564–575. ———, t. j. knight, r. h. pritchard, j. a. daniel, j. a. clapper, a. j. smart, a. trenkle, and d. c. beitz. 2006. circulating ghrelin concentrations fluctuate relative to nutritional status and influence feeding behavior in cattle. journal of animal science 84: 3285–3300. zhang, w., z. lei, j. su, and s. chen. 2008. expression of ghrelin in the porcine hypothalamo-pituitary-ovary axis during the estrous cycle. animal reproduction science 109: 356–367. zhang, y., r. proenca, m. maffei, m. barone, l. leopold, and j. m. friedman. 1994. positional cloning of the mouse obese gene and its human homologue. nature 372 (6505): 425. zieba, d. a., m. amstalden, s. morton, j. l. gallino, j. f. edwards, p. g. harms, and g. l. williams. 2003. effects of leptin on basal and ghrhstimulated gh secretion from the bovine adenohypophysis are dependent upon nutritional status. journal of endocrinology 178: 83–89. ———, m. szczesna, b. klocek-gorka, and g. l. williams. 2008. leptin as a nutritional signal regulating appetite and reproductive processes in seasonallybreeding ruminants. journal of physiology and pharmacology 59: 7–18. 64 nutritional hormones in captive moose – stantorf et al. alces vol. 53, 2017 seasonal variation of nutritional hormones in captive female moose methods design and species blood sampling and processing hormone assays and validation results discussion leptin ghrelin insulinike growth factor-,1,6,2,0,0pt,0pt,0pt,0pt outline placeholder conclusion acknowledgements references alces16_482.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces16_571.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces19_162.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces14_227.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces20_209.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alcessupp1_162.pdf moose movement patterns in the upper koyukuk river drainage, northcentral alaska kyle joly1, timothy craig2,4, mathew s. sorum1, jennifer s. mcmillan3, and michael a. spindler2 1national park service, gates of the arctic national park and preserve, 4175 geist road, fairbanks, alaska, 99709; 2us fish and wildlife service, kanuti national wildlife refuge, 101 12th avenue, fairbanks, alaska, 99701; 3bureau of land management, central yukon field office, 1150 university avenue, fairbanks, alaska, 99709; 4retired abstract: understanding movement patterns of moose (alces alces) is critical to understanding their ecology and sound management. our study was prompted by concern that the dalton highway corridor management area (dhcma), where the dalton highway facilitates access for non-local hunting, may be a population sink for moose that also reside in more remote and protected areas like gates of the arctic national park and preserve (gaar) and kanuti national wildlife refuge (knwr). we did not detect substantial migrations between dhcma and gaar or knwr. however, we estimated that 14–60% of moose in our study area were migratory depending on sex, location within our study area, and methodology utilized to differentiate migratory behavior. a quarter of the animals displayed mixed-migratory strategies where migration is exhibited by a single individual in some years but not others. the percentage of moose that were migratory in our study population, and the distances they migrated, were lower than reported from studies elsewhere in interior alaska. we hypothesize this may be related to their very low density (∼ 0.1 moose/km2) and/or higher terrain ruggedness in part of the study area. winter severity did not appear to impact migration, but home range sizes were smaller in severe winters. alces vol. 51: 87–96 (2015) key words: alaska, alces alces, conservation, migration, moose, strategy, winter severity moose (alces alces) across their circumpolar range often exhibit migratory patterns (e.g., van ballenberghe 1977, ball et al. 2001, white et al. 2014), but migration is not a ubiquitous trait among all populations (mueller et al. 2011) and they are not widely perceived as migratory by the general public. migration is thought to increase fitness because animals gain access to better quality or quantity of forage, experience a reduction in predation, or possibly reduced exposure to parasites or disease (van ballenberghe 1977, avgar et al. 2013). variability in migration among individuals may be the norm where habitat conditions sustain resident populations but fitness varies spatially (fryxell and holt 2013). movements of moose in the upper koyukuk river drainage (fig. 1) are not well understood. on the lower koyukuk river, an area where moose density is comparatively high (5/km2), 83% of adult moose and 58% of cow-calf pairs were migratory, and movement between summer and winter ranges averaged 42 and 31 km, respectively (osborne and spindler 1993). in the arctic national wildlife refuge and neighboring canada, northeast of our study area and where moose density is much lower (∼0.4/km2; caikoski 2010) than in the lower koyukuk study area, 88% of moose were classified as migratory (mauer 1998). there, the mean maximum migration distance was 123 km (i.e., the mean longest migration 87 distance by individual moose). moose density is even lower in the upper koyukuk section of our study area (∼0.1/km2; lawler et al. 2006). land and wildlife management in our study area falls within a complicated patchwork of authorities and includes lands administered by the state of alaska, national park service, us fish and wildlife service, bureau of land management, and private entities. public wildlife advisory groups interested in hunting opportunities and management in the upper koyukuk river drainage were concerned that harvest was high within and near the dalton highway corridor management area (dhcma) which contains the only road in northern interior alaska (fig. 1). harvest opportunity was open to all alaskan residents in the dhcma, whereas only local residents fig. 1. the upper koyukuk river study area (white polygon) in northcentral alaska, encompassing moose locations derived from gps and vhf telemetry data from 2008–2013. 88 migration in northcentral alaska – joly et al. alces vol. 51, 2015 harvest moose within gates of the arctic national park and preserve (gaar) and much of the kanuti national wildlife refuge (knwr). subsistence hunters in the upper koyukuk were concerned that moose were migrating long distances, as in the eastern brooks range (mauer 1998), from their exclusively subsistence hunting areas into general harvest areas, causing a local decline in density and hunting opportunity. the possibility that hunter effort and success in one portion of the region was negatively affecting another area where distinct user groups hunted was an important management concern. therefore, our goal was to better understand the migratory patterns of moose in the upper koyukuk river drainage, and we were specifically interested to identify if moose moved between lands with different management goals and hunting regimes. methods study area our study area falls within the upper koyukuk river drainage in north-central alaska (fig. 1), and encompassed the southern flanks of the central brooks range, including the southeastern portion of gaar, all of knwr, and other state, federal, private, and native lands. it included portions of the dhcma which contains an all-weather highway and had special hunting restrictions. the terrain and vegetation communities are diverse. rugged mountains to 2000 m in elevation with narrowly-confined glacial river valleys are covered with a mix of alpine, shrub, and boreal forest habitat types in the northern portion. shrub habitats were dominated by alders (alnus spp.), willows (salix spp.), and dwarf birch (betula glandulosa). black spruce (picea mariana) is the most prevalent tree species, with white spruce (picea glauca) and poplar (populus balsmifera) common in riparian areas, and birch stands (betula papyrifera) occur on south-facing slopes and in areas that burned. extensive tracts of tussock (eriophorum spp.) tundra occur in wetter, flatter areas. the landscape becomes progressively flatter (elevations typically <500 m) to the south with more muskegs, streams, and lakes interspersed within boreal forest and broad riparian zones. the regional climate is strongly continental, with long, extremely cold (dropping below -45o c) winters, and brief, but hot (temperatures >30o c) summers. forest fires are common during summer months, and snow depth exceeds 90 cm many winters with 60 cm during most. moose relocation data and gis analyses the project ran between march 2008 and april 2013. adult moose were darted using a mixture of carfentanil citrate and xylazine from robison r-44 helicopters. moose were instrumented with either a gps collar equipped with a vhf beacon or a standard vhf collar. moose captured north and east of bettles, alaska (fig. 1) were designated as ‘northern moose’ and those in and around knwr as ‘southern moose’. gps collars deployed in march 2008 collected 1 gps location/day, thereafter, all collected 3 locations/day. an attempt to relocate moose visually from small aircraft using the vhf beacons occurred ∼monthly from march 2008 to april 2013. these efforts were much more consistent for the southern moose due to budgetary restraints, weather, and logistics. movements of northern and southern moose were compared due to differences in terrain, habitats, and sampling effort. migratory status was assessed using 2 different methods using individual moose as the sample unit for both techniques. first, for both gpsand vhf-collared moose, we calculated distances between summer (i.e., the closest relocation july 1) and winter (i.e., the closest relocation to january 15) locations to assess migratory status and distance moved. moose that consistently alces vol. 51, 2015 joly et al. – migration in northcentral alaska 89 demonstrated separation between summer and winter locations were categorized as migratory (i.e., winter locations were not located sympatrically with summer locations or vice versa), whereas those without consistent separation between ranges were categorized as non-migratory. distance between summer and winter ranges was not a determinate. moose that had separation in some years but not others, or did not have a discernible pattern, were categorized as having a mixed-migratory strategy. distances between winter and summer locations were averaged for moose that had multiple years of data. while somewhat subjective, this methodology mimics other studies (e.g., osborne and spindler 1993, mauer 1998) that were conducted adjacent to ours, thereby allowing more direct comparison. second, net squared displacement (nsd) of gps-collared moose was used to determine migratory status on an annual basis (see bunnefeld et al. 2011). the nsd measures straight line distances between a starting location and all successive relocations within the entire annual movement path. the first position was set to july 1 when moose are in their summer range, and we excluded all individuals sampled <330 days. moose displaying “home range” movement patterns according to bunnefeld’s nomenclature were categorized as nonmigratory, “migration” as migratory, and “mixed migratory” and “dispersal” as mixed migratory (no moose dispersed from the study area). moose that did not consistently have the same movement status among years were considered mixed migratory. the shape of movement patterns, not nsd distance, was the determinate of migratory status (bunnefeld et al. 2011). there were insufficient vhf data to quantify migration patterns with the nsd technique for any individual moose. the 2 methods for determining migratory status are highly disparate and caution is required when comparing results. gps and vhf locations were also used to determine if individual moose utilized conservation units (gaar, knwr), the dhcma exclusively, or multiple units. winter severity was classified from the total number of days with snow and snow depth as recorded in bettles, alaska. the categories were mild (<135 days with ≥30 cm snow or <7 days with ≥60 cm snow), moderate (>170 days with ≥30 cm snow, >50 days with ≥60 cm, or <14 days with ≥90 cm snow), or severe (>170 days with ≥30 cm snow, >100 days with ≥60 cm, or >30 days with ≥90 cm snow). annual and winter home ranges for gpscollared moose were determined by establishing fixed kernels using the 95% utilization distribution and least square cross validation method for smoothing using arcgis (esri, redlands, ca; worton 1989, seaman and powell 1996). there were insufficient data to develop kernels for vhf collars. results moose relocation data in total we captured 120 adult moose (27 bulls and 93 cows); 58 in march 2008, 10 in october 2008, 15 in november 2009, and 37 in april 2011. one presumed capture myopathy (adult cow in 2008) was censured. there were 52 southern moose and 67 northern moose providing ∼265 moose-years of data to analyze allopatric migration and movements among conservation units. aerial telemetry flights yielded 2119 high-quality relocations (positive visual identification). gps collars were deployed on 14 northern cow moose in 2008 (25 moose-years of data), 2 in 2009 (2 moose-years), and 2 in 2011 (2 mooseyears); 8 gps collars were deployed on southern cows (8 moose-years of data) and 11 on northern bulls in 2011 (18 mooseyears). in total, the 37 gps units collected 71,675 locations. 90 migration in northcentral alaska – joly et al. alces vol. 51, 2015 migration allopatry of winter and summer ranges and the migratory status of 86 cows and 21 bulls were ascertained through the use of both vhf and gps data. of the 10 moose with only a single year of data, 2 displayed non-migratory behavior and 8 displayed mixed-migratory behavior. using only gps data for nsd-based models, we determined the migratory status of 20 cows and 11 bulls; 25 had 2–4 years of data (5 nonmigratory, 8 migratory, and 12 mixed) and 6 had only a single year of data (4 non-migratory, 1 migratory, and 1 mixed). non-migratory, migratory and mixedmigratory strategies were exhibited in varying proportions by sex, location, and methodology (fig. 2). 0 10 20 30 40 50 60 non-migratory migratory mixed-migratory all moose allopatry all moose nsd all cows allopatry all cows nsd southern cows allopatry southern cows nsd northern cows allopatry northern cows nsd all bulls allopatry all bulls nsd southern bulls allopatry southern bulls nsd northern bulls allopatry northern bulls nsd pe rc en ta ge fig. 2. percentages of non-migratory, migratory, and mixed-migratory moose in the upper koyukuk river drainage, northcentral alaska, 2008–2013. light bars represent cows and darker bars represent bulls. rectangular shaped bars represent percentages determined by winter and summer range allopatry, and cylindrical bars represent those calculated by net squared displacement (nsd) methods outlined by bunnefeld et al. (2011). percentages associated with southern moose are represented with diagonal striping, and northern moose are represented by horizontal striping; stippled bars represent all moose combined. note: missing bars occur where n = 0. alces vol. 51, 2015 joly et al. – migration in northcentral alaska 91 mean distances moved between successive summer and winter locations were 2.5 times greater for migratory cows (n = 86) than non-migrators (table 1). migratory southern cows (n = 46) tended to move less between summer and winter locations than migratory northern cows (n = 40; table 1). migratory distances (mean per individual) for all cows ranged from 557-88,402 m. bulls (n = 21) had a similar relative pattern with individual mean distance between summer and winter locations ranging from 2,886–52,624 m (table 1). summer ranges of both sexes were located in all cardinal directions relative to winter ranges; however, 59% of winter ranges of migratory moose were located north of summer ranges. net displacement (nd) of gps-collared moose (mean per individual) ranged from 10,970– 94,972 m for cows and 13,060–52,660 m for bulls (means in table 1). only 45% of moose were categorized with the same migratory pattern using the nsd techniques when compared to the winter-summer range allopatry analysis. although we did not detect any movements that led to dispersal or forays outside the study area, we also did not collar younger moose that would be more likely to undertake measurable dispersal movements. home range the mean annual home ranges were 243.5 ± 96.9 km2 for gps-collared cows (n = 21; range = 106.5 – 498.3 km2) and 262.0 ± 67.7 km2 for gps-collared bulls (n = 11; range = 185.7 – 400.9 km2). as expected (joly 2005), home ranges of cows (n = 4; 305.0 ± 47.2 km2) calculated from 1 gps location/day were larger than cows (n = 17; 228.9 ± 22.9 km2) with 3 locations/ day. the mean home range size of nonmigratory (nsd technique) cows was ∼10–15% smaller (n = 5; 228.1 ± 45.2 km2) than that of migratory cows (n = 7; 247.9 ± 38.2 km2) or cows with mixed-migration strategies (n = 8; 261.7 ± 35.7 km2). nonmigratory (n = 4; 260.3 ± 33.4 km2) and migratory (n = 4; 295.9 ± 33.4 km2) bulls conformed to this same pattern; bulls with mixedmigration strategies had home ranges about 25% smaller than other bulls (n = 3; 219.0 ± 38.5 km2). the mean annual home range was smallest for cows found primarily within the knwr (n = 5; 199.6 ± 63.0 km2), table 1. distance (mean ± se) between summer and winter locations (allopatry) and net displacement (nd; sensu bunnefeld et al. 2011) of cow and bull moose for the entire and portions of the upper koyukuk river drainage study area, northcentral alaska, 2008–2013. methodology location/sex non-migratory (m) migratory (m) mixed migratory (m) allopatry all cows 8621 ± 2241 21598 ± 2462 17598 ± 2827 allopatry southern cows 7697 ± 2233 16128 ± 2233 18619 ± 2658 allopatry northern cows 9494 ± 3809 29347 ± 4665 16373 ± 5110 allopatry all bulls 5923 ± 5923 15709 ± 3419 18820 ± 5297 allopatry southern bulls 5015* 906 ± 3261 2886* allopatry northern bulls 6226 ± 7176 17644 ± 4143 22804 ±6215 nd all cows 23584 ± 8410 49046 ± 8410 29084 ± 5947 nd southern cows 16526 ± 7663 15604* 40430 ± 5419 nd northern cows 28290 ± 10197 57406 ± 8831 21520 ± 7210 nd northern bulls 22267 ± 6473 37757 ± 6473 22808 ± 7474 *n = 1, no se reported. 92 migration in northcentral alaska – joly et al. alces vol. 51, 2015 increased in size in the dhcma (n = 11; 227.9 ± 88.5 km2), and was largest for cows (n = 3; 279.1 ± 77.4 km2) in the gaar. the 2 cows that utilized both knwr and the dhcma had the largest annual home ranges (384.9 ± 160.0 km2). all 11 gps bulls were collared in gaar and had larger average annual home ranges than cows in gaar. east of gaar, home ranges often overlapped the dalton highway and trans-alaska pipeline system (taps). for all gps moose, winter home ranges were slightly larger (n = 32; 24.2 ± 13.0 km2) in mild winters than moderate (n = 36; 20.2 ± 8.6 km2) and severe winters (n = 11; 21.0 ± 9.0 km2). using only 3 locations/day data, we similarly found larger winter home ranges during mild than moderate winters, and smallest ranges during severe winters (23.3 ± 12.5 km2, 19.4 ± 7.0 km2, and 16.4 ± 8.1 km2, respectively). cows appeared to be more sensitive than bulls to severe weather; their winter home ranges were 24.2 ± 13.7 km2, 19.0 ± 7.7 km2, and 16.4 ± 8.0 km2 in mild, moderate and severe winters, respectively. no gps bulls were collared during a severe winter, but their winter home range sizes were nearly identical in mild and moderate winter (20.9 ± 9.3 km2, 20.2 ± 8.6 km2, respectively). movement relative to conservation units vhf and gps data (n = 116 moose) were combined to examine movements in and around gaar, knwr, and the dhcma. thirty-two moose stayed exclusively within the confines of knwr, including all 5 bulls originally collared there. an additional 21 used knwr and areas outside the refuge, and of these, only 1 bull and 2 cows (14.3%) were also found within the dhcma; 1 cow primarily used the dhcma (1 location within knwr) and 1 bull primarily used the gaar. two (1 bull and 1 cow) of these 21 moose used gaar extensively and did not enter the dhcma, and the cow consistently migrated from gaar in the winter to calve within knwr. the gaar had 16 moose (10 bulls and 6 cows) that stayed within its borders and 24 that moved in and out of the area; of the 24, 9 bulls and 11 cows (83.3%) moved into the adjacent dhcma (fig. 1). nonetheless, most (∼70%) moose using the gaar and dhcma resided primarily in one or the other unit, with a single or few relocations falling in the other. moose located commonly in both units had home ranges with measurable overlap of both. robust migratory movements between these 2 units were not detected for bulls or cows. discussion moose in the upper koyukuk river drainage exhibited partial migration (25–34% of cows and 36–57% of bulls), similar to other areas in alaska (van ballenberghe 1977, osborne and spindler 1993, mauer 1998, white et al. 2014). however, the proportion of moose that were migratory in the upper koyukuk river (35–38%) was less than in the lower drainage (osborne and spindler 1993) or the eastern brooks range (mauer 1998). similarly, the distances traveled from winter to summer range by our migrators (22 km for cows and 16 km for bulls) were less than those measured in adjacent studies (31–123 km; osborne and spindler 1993, mauer 1998) or elsewhere in alaska (see review in mauer 1998). our nsd calculations revealed longer migration distances than obtained by simply comparing mid-summer to mid-winter locations (range allopatry). although not directly comparable to earlier studies because migration was calculated differently, moose in the upper koyukuk river drainage exhibited migratory behavior less frequently than moose elsewhere in the region, and migrated shorter distances. this may be due to lower moose density, differences in habitat quality and alces vol. 51, 2015 joly et al. – migration in northcentral alaska 93 distribution, and/or physiographic differences between study areas. however, moose migration patterns are more complicated than parsing individuals into migratory or non-migratory categories. in addition to partial migration, we documented that a large segment (26–50% of cows and 24–27% of bulls) exhibited a mixed-migration strategy (i.e., migratory behavior only in certain years), and that the annual strategy varied individually. our estimates are conservative as a portion of our population was classified during a single year only. this phenomenon was documented in southcentral alaska and believed related to snow depth (van ballenberghe 1977). alternatively, adaptive migration may convey fitness benefits especially in highly variable climates (e.g., our study area) and in rapidly changing environments such as the arctic. however, migration strategy does not appear to be unilaterally linked to environmental conditions because not all our moose, or entire populations in other areas (van ballenberghe 1977, ball et al. 2001), respond similarly in a given year. as our analyses reveal and as expected (singh et al. 2012, fryxell and holt 2013), a mix of migratory strategies are displayed by sympatric moose under varied environmental circumstances. the mixed-migration strategies we detected highlight the importance of longterm studies and procuring an adequate sample size. the migratory characteristics of a study population would probably not be categorized accurately by short duration studies because mixed-migratory behavior could be erroneously categorized as either migratory or non-migratory behavior (dettki and ericsson 2008). long-term, more focused studies are required to determine if migratory strategy is a behavioral trait acquired from maternal experience as documented elsewhere (e.g., sweanor and sandegren 1988). if migratory strategy is a learned behavior, it would explain, in part, why multiple strategies exist. our results also reveal that the methodology used to determine migratory status is important because: “migratory behavior is persistent and straightened out movement effected by the animal’s own locomotory exertions or by its active embarkation upon a vehicle. it depends on some temporary inhibition of station keeping responses but promotes their eventual disinhibition and recurrence” (kennedy 1985, dingle and drake 2007). the nsd technique appears to more robustly and objectively parse moose migratory status relative to this definition. however, the nsd requires data beyond identifying if summer and winter locations are allopatric. because the allopatric technique is simpler, requires less data, and has been used in past studies, we employed it to make more direct comparisons to regional studies. because both techniques have positive and negative aspects, identifying objectives and definitions are key in determining which is more appropriate for a specific study. we found home ranges to be larger in the northern part of our study, which was consistent with past studies (see review by hundertmark 1997). highly variable seasonality and climatic conditions in the north where forage productivity is lower, and habitat is patchier due to more rugged and higher terrain, may foster migratory behavior and larger home ranges (ballard et al. 1991, hundertmark 1997). the size of home ranges appeared to be influenced by winter severity as smaller home ranges occurred in harsher winters. our study was promoted by public concern about moose moving between conservation units (gaar and knwr) and areas with different hunting regulations and accessibility, such as the dhcma. peak movement rates of bulls occurred from midseptember through early october (joly et al. 94 migration in northcentral alaska – joly et al. alces vol. 51, 2015 2015) coinciding with much of the hunting season; presumably this timing of bull movement stimulated public concern. exposure of knwr moose to the dhcma was negligible, likely a result of the separation (>10 km apart) of the units in most areas and smaller home ranges of moose in that area. in contrast, gaar and the dhcma shared a common boundary and a large proportion (50%) of the moose collared in gaar used both units. nonetheless, a preponderance of these moose resided primarily in either unit, with a single or a few relocations at the edge of the adjacent unit. moose that spent large portions of time in both units had home ranges that measurably overlapped both units. with a single exception, we did not detect regular or substantial migratory movements between the dhcma and gaar or the more distant knwr. this study was the first of its kind in this region of alaska and these data provide a basis for evaluating moose movements relative to harvest concerns and distribution of the moose resource, specifically in our study area. understanding the spatial ecology of moose improves the understanding of moose ecology, behavior, and demographic and genetic processes. it is also critical to developing a comprehensive management program (hundertmark 1997), especially where harvest allocation, resource damage, and local management issues are influenced by seasonal migration and distribution of moose. acknowledgements this project was promoted by public subsistence and wildlife advisory groups. funding was provided by the national park service, us fish and wildlife service, alaska department of fish and game, and the bureau of land management. we thank pilots t. cambier, m. spindler, m. webb, p. zaczkowski, c. cebulski, l. dillard, p. christian, a. greenblatt, h. bartlett, n. guldager, d. sowards, and s. hamilton for making this project possible and safe. s. backensto, j. burch, j. caikoski, k. degroot, j. dillard, m. flamme, a. morris, c. harwood, t. hollis, e. julianus, h. kristenson, j. lawler, n. pamperin, t. paragi, k. rattenbury, c. roberts, l. saperstein, g. stout, and many others provided critical assistance with project implementation. we thank g. stout for his contributions to project management and insights into the ecology of moose within the region. n. bywater, c. harwood, s. miller, r. sarwas, and a. quist provided database and gis expertise. we appreciate the constructive reviews by e. addison, r. churchwell, and anonymous reviewers. all moose captures adhered to state of alaska animal care and use committee (acuc) guidelines (#07–11). literature cited avgar, t., g. street, and j. m. fryxell. 2013. on the adaptive benefits of mammal migration. canadian journal of zoology 91: 481–490. ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7: 39–47. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114: 1–49. bunnefeld, n., l. borger, b. van moorter, c. m. rolandsen, h. dettki, e. j. solberg, and g. ericsson. 2011. a modeldriven approach to quantify migration patterns: individual, regional and yearly differences. journal of animal ecology 80: 466–476. caikoski, j. 2010. units 25a, 25b, and 25d moose. pages 611–642 in p. harper, editor. moose management report of survey and inventory activities 1 july 2007 – 30 june 2009. alaska department alces vol. 51, 2015 joly et al. – migration in northcentral alaska 95 of fish and game. project 1.0 juneau, alaska, usa. dettki, h., and g. ericsson. 2008. screening radiolocation datasets for movement strategies with time series segmentation. journal of wildlife management 72: 535–542. dingle, h., and v. a. drake. 2007. what is migration? bioscience 57: 113–121. fryxell, j. m., and r. d. holt. 2013. environmental change and the evolution of migration. ecology 94: 1274–1279. hundertmark, k. j. 1997. home range, dispersal and migration. pages 303–335 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, wildlife management institute, washington, dc, usa. joly, k. 2005. the effects of sampling regime on the analysis of movements of overwintering female caribou in east-central alaska. rangifer 25: 67–74. ———, t. craig, m. s. sorum, j. s. mcmillan, and m. a. spindler. 2015. variation in fine-scale movements of moose in the upper koyukuk river drainage, northcentral alaska. alces 51: 97–105. kennedy, j. s. 1985. migration: behavioral and ecological. pages 5–26 in m. a. rankin, editor. migration: mechanisms and adaptive significance. contributions in marine science 27 (suppl.). university of texas, austin, texas, usa. lawler, j. p., l. saperstein, t. craig, and g. stout. 2006. aerial moose survey in upper game management unit 24, alaska, fall 2004, including state land, and lands administered by the bureau of land management, gates of the arctic national park and preserve, and kanuti national wildlife refuge. national park service technical report nps/ar/nr/ tr-2006-55, fairbanks, alaska. 30 pp. mauer, f. j. 1998. moose migration: northeastern alaska to northwestern yukon territory, canada. alces 34: 75–81. mueller, t., k. a. olson, g. dressler, p. leimgruber, t. k. fuller, c. nicolson, a. j. novaro, m. j. bolgeri, d. wattles, s. destefano, j. m. calabrese, and w. f. fagan. 2011. how landscape dynamics link individualto populationlevel movement patterns: a multispecies comparison of ungulate relocation data. global ecology and biogeography 20: 683–694. osborne, t. o., and m. a. spindler. 1993. moose population identification study. three day slough, koyukuk national wildlife refuge, alaska, game management unit 21d. progress report 93-3. us fish and wildlife service, fairbanks, alaska, usa. seaman, d. e., and r. a. powell. 1996. an evaluation of the accuracy of kernel density estimators for home range analysis. ecology 77: 2075–2085. singh, n. j., l. borger, h. dettki, n. bunnefeld, and g. ericsson. 2012. from migration to nomadism: movement variability in a northern ungulate across its latitudinal range. ecological applications 22: 2007–2020. sweanor, p. y., and f. sandegren. 1988. migratory behavior of related moose. holarctic ecology 11: 190–193. van ballenberghe, v. 1977. migratory behavior of moose in southcentral alaska. transactions of the 13th international congress of game biologists 13: 103–109. white, k. s., n. l. barten, s. crouse, and j. crouse. 2014. benefits of migration in relation to nutritional condition and predation risk in a partially migratory moose population. ecology 95: 225–237. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70: 164–168. 96 migration in northcentral alaska – joly et al. alces vol. 51, 2015 moose movement patterns in the upper koyukuk river drainage, northcentral alaska methods study area moose relocation data and gis analyses results moose relocation data migration home range movement relative to conservation units discussion acknowledgements literature cited alces17_165.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces16_37.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces17_44.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces19_98.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces18_329.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces17_iipreface.pdf alces vol. 17, 1981 alces vol. 17, 1981 provincial population and harvest estimates of moose in british columbia gerald w. kuzyk ministry of forests, lands and natural resource operations, p.o. box 9391, victoria, british columbia v8w 9m8, canada abstract: provincial population and harvest estimates of moose in british columbia, canada were assessed over a 28-year period from 1987 to 2014. the population generally remained stable, whereas the licensed hunter harvest declined gradually by about half despite constant hunter effort. the annual population estimate ranged from a low of 157,000 moose in 1994 to a high of 190,000 in 2011, with an overall mean of 172,000 ± 9900 (sd). in 2014, the relative status of hunted populations within 7 wildlife administrative units was 1 increasing, 3 stable, and 3 in decline. the mean annual licensed harvest was 10,038 ± 2137 (sd) moose, and the mean harvest rate was 6 ± 1.3% (sd). in december 2013, british columbia initiated a 5-year (2013–2018) research project to identify factors contributing to the decline of the moose population and licensed harvest. alces vol. 52: 1–11 (2016) key words: alces alces, british columbia, harvest, moose, population periodic updates of moose (alces alces) abundance are necessary to assess management objectives (brown 2011), evaluate sustainable harvest (timmerman and buss 2007), and to provide information to the public. assessing licensed harvest concurrent with population estimates should provide better understanding and explanation of population fluctuations over time. moose population estimates are also used for comparison among jurisdictions to assess patterns of broad-scale population trends. in north america, there is current concern for declining populations in southern parts of moose range (murray et al. 2006, lenarz et al. 2009), whereas populations remain stable in other areas (murray et al. 2012). explanations for population change include human-caused habitat alterations (rempel et al. 1997), climate change (rempel 2011), and a combination of natural and humaninfluenced variables (murray et al. 2006, brown 2011). moose in british columbia are highly valued for food, social, and ceremonial purposes by first nations, for recreational and commercial harvest opportunities by licensed hunters, and for wildlife viewing. specific management objectives for moose harvest are to manage for first nations use, support a sustainable licensed hunter harvest, and provide for diverse hunter opportunities (bc flnro 2015). assessment of abundance and licensed harvest estimates is required to ensure that harvest levels are sustainable (hatter 1999), objective information is available for management decisions, and to provide accurate information on the status of moose to stakeholders and the public (bc flnro 2015). the purpose of this paper is to provide an overview of the population abundance and licensed harvest of moose in british columbia from 1987 to 2014. study area british columbia is an ecologically diverse province (meidinger and pojar 1991) 1 where moose are widely distributed (fig. 1) and occupy a range of landscapes including wet coastal habitats, dry interior forests, cold northern forests, and montane habitats (eastman and ritcey 1987). at the provincial scale, moose co-exist with several ungulate species including bison (bison bison), mule deer (odocoileus hemionus), white-tailed deer (odocoileus virginianus), elk (cervus elaphus), and caribou (rangifer tarandus) (shackleton 1999). the main predators of moose are wolves (canis lupus), grizzly bears (ursus arctos), and black bears (u. americanus), with cougars (puma concolor) important in southern british columbia (spalding and lesowski 1971). bull hunts were mostly open seasons, with antler restrictions or limited entry hunts occurring between 15 august and 30 november. antlerless harvest was largely restricted to limited entry hunts with some general open seasons for calves in select areas. seasons for antlerless moose occurred between 1 october and 10 december (bc moe 2010). fig. 1. distribution and population status (i.e., stable, increasing, decreasing) of moose in 7 wildlife administrative units in british columbia, canada, 2014. 2 populations and harvest of moose in bc – kuzyk alces vol. 52, 2016 hunting seasons were generally available throughout the distribution of moose with the exception of regions 1 and 2 which have few moose (i.e., <130 combined) and national parks (<1% of land area) where licensed hunting is prohibited. methods moose population estimates were produced by regional biologists in 7 wildlife administration units (regions; fig. 1) from 1987 to 2014, and then combined for a provincial total. there were 3–5 year intervals between estimates to provide time to assess potential changes in moose abundance at the provincial scale. minimum and maximum estimates were derived from 2000 to 2014 because of the need to convey uncertainty when comparing estimates between years. these estimates were developed using the best available information from a combination of sources including aerial surveys, big game stock assessments, and expert opinion. a third degree polynomial was used to fit a long-term population trend line to the abundance estimates from 1987 to 2014. the polynomial was preferred to a linear or log-linear trend line because the polynomial was sensitive to fluctuations in population size. in the 7 regions where moose were hunted, the trend (stable, declining, increasing) was determined from the change in abundance estimates and the slope of the trend line from 2011 to 2014. aerial surveys were the most important source of information because they provided data for estimation of population size, density, and composition. all surveys were required to follow provincial standards that are based on defensible scientific methods (risc 2002). stratified random block surveys were used (gasaway et al. 1986) or modified to include habitat-based stratification (heard et al. 2008). a standard sightability correction factor was applied to account for detection probability based on research with radio-marked moose in central british columbia (quayle et al. 2001). aerial surveys were required to conform to standards for accuracy and precision (1-α) and to produce a 90% ci with allowable error (±15–25%). the frequency of stratified random block surveys was based on available funds and prioritization criteria which in‐ cluded time since last survey, first nations concerns, impact to hunter opportunity, population objectives, and if the survey was part of an ongoing monitoring program (bc flnro 2015). aerial composition surveys were also conducted to determine bull:cow and calf:cow ratios; ground-based surveys following provincial standards were used occasionally (risc 1998, d’ eon et al. 2006). big game stock assessments were used to help estimate population size and sustainable harvest levels as outlined in the provincial moose harvest management procedure (bc moe 2010). these assessments helped maximize information from aerial surveys and hunter harvest (griffiths and hatter 2011), and incorporated uncertainty associated with extrapolating area-based survey results to regional population estimates. they helped determine the maximum allowable mortality and accounted for first nations harvest and road/rail mortality where available. population models were one component of big game stock assessments and were occasionally used in the regional population estimates by fitting annual licensed harvest data to periodic survey data (white and lubow 2002). population variables used in the models generally included annual licensed harvest data, posthunt population size and composition, overwinter survival, and recruitment rates (griffiths and hatter 2011). if empirical information was lacking about a population, regional biologists used a broad spectrum of expert opinion including field information gathered from resident hunters and trappers, guide-outfitters, first nations, and other alces vol. 52, 2016 kuzyk – populations and harvest of moose in bc 3 resource professionals. this information was gathered during a variety of forums and locations including formal stakeholder meetings and informal discussions. licensed harvest of moose was monitored annually from 1987 to 2014 with a provincial resident hunter survey, and guide declarations for non-resident hunters. harvest information from first nations was not part of the provincial hunter survey and was largely unknown (bc flnro 2015), with the exception of certain first nations communities that voluntarily provided information. estimates of licensed hunter harvest (resident and non-resident combined), hunter days, and hunter numbers were available, all with 95% confidence intervals (ci). these estimates were produced from mail-out questionnaires sent to a random sample of resident hunters; from 2008 to 2014 an average of 13,003 questionnaires were mailed annually with an average response rate of 61%. licensed harvest rates were calculated from the provincial population estimate for a given year and the average of the 3 nearest harvest estimates; 2014 was an exception when the average of the 2 nearest harvest estimates were used because of delay in the 2015 estimate. combined resident and nonresident hunting license sales from 1989 to 2014 were used to further measure hunter interest. results the mean annual population estimate of moose in british columbia was 172,000 ± 9900 (sd) from 1987 to 2014. annual estimates were relatively stable ranging from a low of 157,000 moose in 1994 to a high of 190,000 in 2011 (fig. 2). the minimum and maximum estimates (i.e, from 2000 to 2014) reflected varied levels of uncertainty (fig. 2). the 2014 estimates varied among the 7 regions with hunted populations: 3 were considered stable (regions 3, 6, and 7b; figs. 1 and 3), 3 were declining (regions 4, 5, and 7a; figs. 1 and 4), and one region was increasing (region 8; figs. 1 and 3). the mean annual licensed harvest from 1987 to 2014 was estimated as 10,038 ± 2137 (sd). total harvest declined gradually by about one-half during this period, yet hunter effort (average days hunted) remained stable (fig. 5). the mean annual licensed harvest rate from 1987 to 2014 was 6 ± 1.3% (sd), ranging two-fold from a high of 8% in 1987 to a low of 4% in 2011. from 1987 to 2014, the mean number of licensed hunters (resident and non-resident combined) was 33,721 ± 4292 (sd) that spent 273,622 ± 32,521 (sd) days of hunter effort (table 1). the mean annual hunting license sales was 39,815 ± 4158 (sd) from 1989 to 2014 and varied minimally from 1993 to 2014 (table 1). discussion the annual moose population in british columbia during 1987–2014 was relatively stable, averaging 172,000. in 2014 hunted populations were stable in 3 regions, decreasing in 3 regions, and increasing in one. although both provincial and regional population estimates had varied levels of uncertainty, they remain important for resource managers to address management objectives (bc flnro 2015), and to inform first nations, stakeholders, and the general public about the status of moose in british columbia. the estimation error was partially responsible for the uncertainty reported in the abundance estimates. the variation in the population estimates may reflect the varied abundance and composition of local and regional predators (ballard and van ballenberghe 2007), human-altered landscape change (rempel et al. 1997) which may enhance forage quality and quantity while facilitating predator and hunter access to moose, and variation in licensed and unlicensed harvest levels (timmerman and 4 populations and harvest of moose in bc – kuzyk alces vol. 52, 2016 buss 2007). other factors such as weather, disease, parasites, and accidents including road and rail mortality also influence local moose abundance. the quality of data used to develop the population estimates could also be improved with increased financial and logistical support that would provide more aerial surveys over a broader geographical area. of most concern to stakeholders were recent (2008–2014) population declines in regions 4, 5, and 7a (fig. 4). in two regions (region 5 and 7a) the moose declines coincided with a mountain pine beetle (dendroctonus ponderosae) epidemic (chan-mcleod 2006) which led to increased salvage logging and associated road building. this type of landscape change can presumably alter the spatial dynamics of moose, predators, and hunters, ultimately influencing moose abundance and harvest rate. although moose should benefit from salvage logging through increased forage production (janz 2006), those benefits are not immediate and may be offset by higher harvest and predation due to easier access afforded by high density of roads and cutblocks (ritchie 2008). to address the recent moose population declines, british columbia initiated a provincially-coordinated research project in 2013 to evaluate the landscape change hypothesis (kuzyk and heard 2014) and to increase science-based information for moose management. to date, unpublished data from this research has provided no evidence that low pregnancy rates, infectious disease, or parasites are influencing the moose population (h. schwantje, bc flnro, personal communication). similarly, preliminary adult survival rates are within the limits of a stable moose population (92 ± 8% in 2013–2014 and 92 ± 5% in 2014–2015; kuzyk et al. 2015). in southeastern british columbia (region 4), declining forage production in older burns and wolf predation are believed limiting to moose population growth (stent 2009, 2012). further, in an attempt to reduce predation of an endangered caribou population, the local moose density was reduced which lowered wolf abundance in a small fig. 2. provincial population estimates of moose and trend line derived from inventories, population modeling, and expert opinion from 1987 to 2014 in british columbia, canada. minimum and maximum ranges in population estimates are presented from 2000 to 2014. alces vol. 52, 2016 kuzyk – populations and harvest of moose in bc 5 portion of the region (~6,375 km2) (serrouya et al. 2011, serrouya 2013). given stakeholder and public concern for declining moose populations, it is important to maintain a balanced, provincial-level assessment and approach that also addresses regions with stable or increasing populations. the large northwestern (region 6) and northeastern (region 7b) regions with stable fig. 4. regional moose population estimates and declining trend lines in regions 4, 5, and 7a as derived from inventories, population modeling, and expert opinion, 1987–2014, british columbia, canada. minimum and maximum ranges in population estimates are presented for 2000–2014. fig. 3. regional moose population estimates and trend lines in regions 3, 6, 7b, and 8 as derived from inventories, population modeling and expert opinion, 1987–2014, british columbia, canada. minimum and maximum ranges in population estimates are presented for 2000–2014. 6 populations and harvest of moose in bc – kuzyk alces vol. 52, 2016 moose populations are more remote than those in the southern half of the province and have not undergone landscape change that presumably facilitates hunter and predator access. these regions also experienced little impact from the mountain pine beetle outbreak compared to the central interior regions (region 5 and 7a). the one stable population in the south was largely affected by the mountain pine beetle and salvage logging, but had lower wolf density compared to northern regions (bc flnro 2014, kuzyk and hatter 2014). the increasing population in the southern region (region 8) overlapped with a recolonizing wolf population (bc flnro 2014). further, this regional estimate was revised in 2013 with a habitat-based model (gyug 2013) that may have amplified the estimated increase in abundance between 2011 and 2014. the average (6%) and range (4–8%) of the provincial licensed harvest rate were mid-range of values reported throughout north america (2–16%; crête 1987). more conservative harvest rates of 5% are recommended for northern systems where predation is believed to limit moose density (e.g., yukon; hayes et al. 2003), and may be appropriate in northern regions of british columbia (hatter 1999). first nations harvest of moose is thought to be broadly distributed province-wide (bc flnro 2015), but because no formal method exists to quantify first nations harvest, the total harvest and rates reported here are underestimated and conservative. for example, local harvest may have been underestimated by up to 40% in ontario by not accounting for first nations harvest (leblanc et al. 2011). harvest information from first nations in british columbia would benefit future management efforts to ensure sustainable harvests for all users including first nations, recreational hunters, and the guide-outfitting industry (bc flnro 2015). an important outcome from this assessment was documentation of the gradual decline in licensed harvest by approximately half over 28 years from 1987 to 2014, despite constant hunter effort, indicating that the kill fig. 5. annual estimates of provincial moose harvest and hunter effort (average days hunted) by licensed hunters, british columbia, 1987–2014. alces vol. 52, 2016 kuzyk – populations and harvest of moose in bc 7 per unit of effort (kills/hunter days) had declined. the disparity between these two trends may be related to difficulties producing accurate provincial population estimates that are driven by wide regional variation. further, changes in the hunting season structure in the early 1990s reduced harvest levels in some regions (hatter 1999), and similarly, a regulatory change allowing shared limited entry hunts in the early 2000s raised hunter effort through increased opportunity to hunt moose, without increasing harvest. finally, although hunters maintained constant hunting effort as harvest declined, lower hunter success often reflects inclement weather and human disturbance that influence moose distribution. given the number, frequency, and variable proportional influence of these factors, kill per unit of effort is probably not a reliable measurement to assess moose table 1. a summary of annual moose license sales and annual estimates of licensed hunters, hunter days, and moose harvest in british columbia, canada, 1987–2014. year licensed hunters licensed hunter days licensed harvest license sales 1987 42,526 338,482 13,463 n/a 1988 42,679 334,246 13,539 n/a 1989 41,979 332,852 14,070 51,520 1990 42,104 334,718 13,457 50,367 1991 39,400 304,852 12,251 46,010 1992 38,973 314,613 11,557 45,289 1993 33,236 252,647 10,025 38,538 1994 31,423 247,039 9944 37,714 1995 31,778 248,281 11,047 38,018 1996 30,923 245,617 9701 35,948 1997 32,085 251,582 10,494 37,243 1998 35,617 276,206 11,438 41,089 1999 29,840 250,287 7459 35,612 2000 31,106 255,569 9182 36,221 2001 30,988 272,771 10,290 36,145 2002 31,829 256,975 10,803 37,010 2003 31,493 238,983 11,309 36,608 2004 27,293 214,743 9571 40,438 2005 31,498 253,619 9980 37,175 2006 32,010 247,409 9939 38,374 2007 31,719 260,126 8000 38,069 2008 31,368 267,654 8730 37,125 2009 32,880 291,920 8074 40,371 2010 32,242 270,781 8836 39,733 2011 32,324 280,931 7660 40,503 2012 32,277 276,699 7576 40,236 2013 32,420 280,133 6890 40,109 2014 30,172 261,677 5773 39,723 mean 33,721 ± 4292 273,622 ± 32,521 10,038 ± 2137 39,815 ± 4158 8 populations and harvest of moose in bc – kuzyk alces vol. 52, 2016 abundance in british columbia (hatter 2001). further research should help identify the relationships among moose abundance, harvest rate, hunter effort, and landscape changes. it is important that regional and provincial moose abundance estimates and harvest data be monitored and evaluated on a regular basis to improve regional, provincial, and rangewide status of moose. acknowledgements i would like to thank b. cadsand, d. heard, s. maciver, s. marshall, c. procter, p. stent, c. thiessen, m. bridger, h. schwantje, and a.walker for their comments and discussions on early drafts of this paper and m. klaczek for producing fig. 1. special thanks to i. hatter who helped interpret regional moose population estimates and provided useful revisions to a later version of this manuscript. i appreciate the useful input from associate editor e. bergman and two anonymous reviewers which improved this manuscript. references ballard, w., and v. van ballenberghe. 2007. predator-prey relationships. pages 247–274 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. british columbia ministry of environment (bc moe). 2010. moose harvest management procedure manual. fish and wildlife branch, victoria, british columbia, canada. british columbia ministry of forests, lands and natural resource operations (bc flnro). 2014. management plan for the grey wolf (canis lupus) in british columbia. british columbia ministry of forests, lands and natural resource operations, victoria, british columbia, canada. –––. 2015. provincial framework for moose management in british columbia. fish and wildlife branch, victoria, british columbia, canada. brown, g. s. 2011. patterns and causes of demographic variation in a harvested moose population: evidence for the effects of climate and density-dependent drivers. journal of animal ecology 80: 1288–1298. doi: 10.1111/j.1365-2656. 2011.01875.x. chan-mcleod, a. c. a. 2006. a review and synthesis of the effects of unsalvaged mountain-pine-beetle-attacked stands on wildlife and implications for forest management. british columbia journal of ecosystems and management 7: 119–132. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research, supplement 1: 553–563. d’eon, r. g., s. f. wilson, and d. hamilton. 2006. ground-based inventory methods for ungulates: snow-track surveys. standards for components of british columbia’s biodiversity no. 33a. resource information standards committee, british columbia ministry of environment, victoria, british columbia, canada. eastman, d., and r. ritcey. 1987. moose habitat relationships and management in british columbia. swedish wildlife research supplement 1: 101–117. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22, institute of arctic biology. griffiths, f., and i. hatter. 2011. population modelling for big game stock assessment: an introductory guide. british columbia ministry of forests, lands and natural resource operations, victoria, british columbia, canada. gyug,l.w.2013.okanagan moose inventory 2012–2013. british columbia ministry alces vol. 52, 2016 kuzyk – populations and harvest of moose in bc 9 of forests, lands and natural resource operations, penticton, british columbia, canada. hatter, i. w. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. –––. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces 37: 71–77. hayes, r. d., r. farnell, r. m. p. ward, j. carey, m. dehn, g. w. kuzyk, a. m. baer, c. l. gardner, and m. o’donoghue. 2003. experimental reduction of wolves in the yukon: ungulate responses and management implications. wildlife monographs 152: 1–35. heard, d. c., a. b. d. walker, j. b. ayotte, and g. s. watts. 2008. using gis to modify a stratified random block survey design for moose. alces 44: 111–116. janz, d. w. 2006. mountain pine beetle epidemic – hunted and trapped species sensitivity analysis. british columbia ministry of the environment, prince george, british columbia, canada. kuzyk, g. w., and i. w. hatter. 2014. using ungulate biomass to estimate abundance of wolves in british columbia. wildlife society bulletin 38: 878–883. doi: 10. 1002/wsb.475. –––, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. british columbia ministry forest, lands and natural resource operations. victoria, british columbia, canada. –––, s. marshall, m. klaczek, and m. gillingham. 2015. determining factors affecting moose population change in british columbia: testing the landscape change hypothesis. progress report, february 2012–july 2015. wildlife working report no. wr-122. british columbia ministry forest, lands and natural resource operations, victoria, british columbia, canada. leblanc, j. e., b. e. mclaren, c. pereira, m. bell, and s. atlookan. 2011. first nations moose hunt in ontario: a community’s perspectives and reflections. alces 47: 163–174. lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–10. doi: 10.2193/2008-265. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. british columbia ministry of forests, special report series number 6. british columbia ministry of forests, victoria, british columbia, canada. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate change influences on a declining moose population. wildlife monographs 166: 1–30. –––, k. f. hussey, l. a. finnegan, s. j. lowe, g. n. price, j. benson, k. m. loveless, k. r. middel, k. mills, d. potter, a. silver, m. j. fortin, b. r. patterson, and p. j. wilson. 2012. assessment of the status and viability of a population of moose (alces alces) at its southern range limit in ontario. canadian journal of zoology 90: 422–434. doi: 10.1139/z2012-002. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–54. rempel, r. 2011. effects of climate change on moose populations: exploring the response horizon through biometric and systems models. ecological modelling 222: 3355–3365. doi: 10.1016/ j.ecolmodel.2011.07.012. –––, p. elkie, a. rodgers, and m. gluck. 1997. timber-management and natural 10 populations and harvest of moose in bc – kuzyk alces vol. 52, 2016 disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61: 517–524. doi: 10.2307/ 3802610. resources information standards committee (risc). 1998. ground-based inventory methods for selected ungulates: moose, elk and deer. standards for components of british columbia’s biodiversity no. 33. version 2.0. ministry of environment, lands and parks, resources inventory branch, victoria, british columbia. –––. 2002. aerial-based inventory methods for selected ungulates: bison, mountain goat, mountain sheep, moose, elk, deer and caribou. standards for components of british columbia’s biodiversity no. 32. version 2.0. british columbia ministry of sustainable resource management, victoria, british columbia, canada. ritchie, c. 2008. management and challenges of the mountain pine beetle infestation in british columbia. alces 44: 127–135. serrouya, r. 2013. an adaptive approach to endangered species recovery based on a management experiment: reducing moose to reduce apparent competition with woodland caribou. phd thesis, university of alberta, edmonton, alberta, canada. –––, b. n. mclellan, s. boutin, d. r. seip, and s. e. nielsen. 2011. developing a population target for an overabundant ungulate for ecosystem restoration. journal of applied ecology 48: 935–942. doi: 10.1111/j.1365-2664.2011.01998.x. shackleton, d. 1999. hoofed mammals of british columbia. royal british columbia museum handbook. university of british columbia press, vancouver, british columbia, canada. spalding, d. j., and j. lesowski. 1971. winter food of the cougar in southcentral british columbia. journal of wildlife management 35: 378–381. doi: 10.2307/3799618. stent, p. 2009. management unit 4-34 moose inventory. ministry of forests, lands and natural resource operations, nelson, british columbia, canada. –––. 2012. management unit 4-03 moose inventory. ministry of forests, lands and natural resource operations, nelson, british columbia, canada. timmerman, h. r., and m. e. buss. 2007. population and harvest management. pages 559–615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. white, g. c., and b. c. lubow. 2002. fitting population models to multiple sources of observed data. journal of wildlife management 66: 300–309. doi: 10.2307/3803162. alces vol. 52, 2016 kuzyk – populations and harvest of moose in bc 11 provincial population and harvest estimates of moose in british columbia study area methods results discussion acknowledgements references alces20_95.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces18_186.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces18_258.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces15_80.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alcessupp1_112.pdf alces15_19.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces18_94.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces19_191.pdf alces vol. 13, 1983 alces vol. 13, 1983 alces vol. 13, 1983 alces vol. 13, 1983 alces vol. 13, 1983 alces vol. 13, 1983 alces vol. 13, 1983 alces17_111.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 f:\alces\supp2\pagema~1\rus2s.pdf alces suppl. 2, 2002 abaturov and smirnov moose density effects 1 effects of moose population density on development of forest stands in central european russia boris d. abaturov 1 and konstantin a. smirnov 2 1institute of evolutionary animal morphology and ecology, russian academy of science, 117071 moscow, russia; 2laboratory of forestry research, russian academy of science, 143030, uspenskoe, moscow region, russia abstract: when moose population density is high (3–5 individuals per 1,000 hectares), deciduous trees, in particular aspen, are depressed, and cutovers are rapidly overgrown with spruce. higher moose densities can result in the depression of spruce and degradation of the tree stands. in the near future, preservation of a high population of moose may cause the aspen to disappear and prevent regeneration of pine, oak, and mountain ash. when moose density is 2–3 individuals per 1,000 hectares, the development of stands follows its usual pattern. the composition and structure of modern forests is a function of the pattern of tree stand development on cutovers. according to modern theory, following the removal of coniferous trees on cutovers, deciduous stands are formed, and regeneration of the main coniferous forests is extended over a period of more than 100 years. as a result, coniferous forests are ubiquitously replaced by deciduous. alces supplement 2: 1-5 (2002) key words: aspen, birch, cutover, forest, moose, population density, spruce during recent years, due to increasing moose populations and associated activit i e s , t h e n a t u r a l p a t t e r n o f c u t o v e r overgrowth is substantially modified. a large number of studies of the effect of moose on the formation of stands indicate that selective removal and damage of young forests and species composition of stands have been changed (dinesman 1961, elsky et al. 1975, dunin 1979, tikhonov 1980, smirnov 1987). the present study is an attempt to estimate the consequence of different moose population densities on forests. the fate of different tree species and the pattern of stand development were inferred from comparing stands on cutovers of different ages, for which a set of different-age cutovers were selected. the research was performed in the kostroma, moscow, and yaroslavl regions, characterized by different moose population densities: 2–3 individuals per 1,000 ha (the kostroma region), 4–5 individuals per 1,000 ha (the moscow region), and 6–10 individuals per 1,000 ha (the yaroslavl region). the most important forage habitats of moose are at present associated with increasingly overgrown cutovers, where young stands of early successional trees provide the necessary food for the animals. the cutovers are primarily overgrown with birch (betula spp.) and aspen (populus tremula) that, in the absence of moose, after 5–10 years form a deciduous canopy, which for a long time depresses regrowth of primary coniferous species. this pattern of overgrowth of cutovers in the regions under consideration was observed in the 1950s and 1960s, when the moose population density was low. results and discussion currently the overgrowth of cutovers, under increasing activities of moose, proceeds differently in various regions. in the moose density effects abaturov and smirnov alces suppl. 2, 2002 2 moscow region, the cutovers are actively overgrown with deciduous trees, the aspen predominating. of all the cutovers under study, the aspen was superior or slightly inferior in the number of stems compared to other trees. during the first years, aspen root shoots substantially exceeded other species in growth. at a 2-year-old cutover, the average height of aspen reached 1.6 m, while in other species it did not exceed 70 cm, and birch, which ranked second in abundance, was equal to only 25 cm (table 1). some individual groups of aspens formed shoots, or under weak activity of moose at this cutover, an aspen canopy would develop in several years. however, as early as the first year, the aspen was under heavy browsing pressure by moose. almost all the aspen forests were browsed by moose at a height of about 1 m (table 1). characteristically, browsing did not affect the state of aspen, and new regrowth averaged 60 cm in length. overall, 93% of the aspens were browsed, while other species were virtually intact (except the mountain ash). a similar pattern was observed in other overgrowing cutovers. at a 4-year-old cutover, where there were fewer aspens, most of the trees were removed annually by the moose. hence, despite the age of 4 years, the height of aspens was only 87 cm. nevertheless, their condition was still good, and the annual regrowth of the top shoot reached 61 cm (table 1). during subsequent years, an active selective removal of the aspen continued. in all the cutovers, the aspen stems were removed by moose at a constant height, and the aspen formed regrowth of about 1 m in height, irrespective of age. at a 17-yearold cutover, as a result of annual removal, all the aspens were depressed and weakened, 90% of the stems were dry, and the entire aspen forest was a growth of semidry stems. despite the great age of this aspen (7–14 years), its average height was 93 cm, and the average size of annual increment was only 33 cm, the stem diameter being 12 mm (table 1). thus, despite its great resistance to removal, the aspen was weakened and dried in all of the sites, yet continued to be among the regenerating species of the cutover. in addition to the aspen, less abundant trees, such as oak (quercus spp.), mountain ash (sorbus aucuparia), and maple (acer spp.), were also controlled by moose. due to active browsing of these trees by moose, these trees remained in the zone of influence and were not involved in the formation of the stand. the overgrowth of the cutover by other species followed a different pattern. birch, substantially inferior to aspen during the first years, subsequently became dominant. at a 17-year-old cutover, the birch was substantially superior to aspen in height, and became a dominant species (table 1). quite a number of species were not included in the zone of moose influence. nevertheless, 80% of the birch proved injured, and the average height of these trees was only 144 cm (table 1). hence, the birch formed only rare stands. of special interest in these conditions was the regeneration of spruce (picea spp.), which occurs in the clearings mainly in the form of forest plantations and which is the primary species of these forests. at the clearings examined, spruce regrowth was virtually undamaged by moose. the spruce surpassed the height of the deciduous trees, getting out from under the canopy of the stand. at a 17-year-old cutover, it considerably outstripped other early successional species in growth and reached the developing stand. in the yaroslavl region, where the population density of moose is higher, moose still substantially affect the development of stands. during the 1950s, when the moose alces suppl. 2, 2002 abaturov and smirnov moose density effects 3 t a b le 1 . t h e s ta te o f re g ro w th a t c u to v e rs o f d if fe re n t a g e i n t h e m o sc o w r e g io n . s p e c ie s a g e # s te m s s te m s b ro w se d h e ig h t l e n g th s te m (y e a rs ) p e r h a b y m o o se (c m ) o f a n n u a l d ia m e te r s h o o t (c m ) (m m ) # % 2 y e a ro ld c u to v e r (1 9 8 7 ) a s p e n 2 7 1 ,3 2 0 ± 1 8 ,0 6 0 6 6 ,2 2 0 ± 1 ,6 2 4 93 1 5 9 ± 2 .4 62 1 0 .7 ± 0 .2 7 b ir c h 3 8 ,0 0 0 ± 4 ,2 1 8 0 0 2 5 ± 1 .8 2 .1 ± 0 .2 1 l in d e n 1 3 ,3 2 0 ± 2 ,2 3 0 0 0 7 1 ± 2 .8 7 .3 ± 0 .5 0 w il lo w 0 0 1 ± 5 .2 m o u n ta in a sh 56 6 9 ± 5 .4 s p ru c e 2 ,6 6 0 ± 3 3 0 0 0 3 0 ± 3 .0 2 .6 ± 0 .2 4 4 -y e a ro ld c u to v e r (1 9 8 5 ) a s p e n 4 1 0 ,0 0 0 ± 2 ,9 4 0 7 ,8 4 0 ± 2 ,8 9 4 78 8 7 ± 2 .6 6 1 ± 3 .1 8 .8 ± 0 .3 9 b ir c h 1 -4 1 6 ,9 3 4 ± 4 ,2 1 8 1 ,2 8 0 ± 6 1 6 8 8 0 ± 3 .5 6 .2 ± 0 .3 1 w il lo w 1 -4 7 ,8 6 6 ± 2 ,2 3 0 2 ,1 4 0 ± 1 ,1 1 6 27 7 9 ± 4 .2 6 .4 ± 0 .4 7 m o u n ta in a sh 1 -4 ra re 40 7 5 ± 3 .3 5 .8 ± 1 .1 6 s p ru c e 9 3 4 ± 3 3 0 0 0 6 7 ± 8 .8 1 2 .4 ± 2 .5 5 1 7 -y e a ro ld c u to v e rs ( 1 9 7 2 ) a s p e n 7 -1 4 1 6 ,4 4 0 ± 3 ,2 8 4 1 5 ,8 8 0 ± 3 ,1 8 0 97 9 3 ± 1 .9 3 3 ± 1 .5 1 1 .9 ± 0 .3 3 b ir c h 1 5 -1 7 7 ,3 4 0 ± 2 ,1 2 0 5 ,8 8 0 ± 1 ,4 0 4 80 1 4 4 ± 8 .9 1 5 .3 ± 1 .0 8 w il lo w 4 4 4 ± 3 4 6 3 3 3 75 8 3 ± 9 .7 7 .5 ± 1 .8 9 m o u n ta in a sh si n g le 1 0 0 6 0 ± 1 0 .0 9 .5 ± 1 .5 0 s p ru c e 3 ,3 3 4 ± 8 4 0 ra re 1 5 0 ± 1 1 .5 2 3 .7 ± 1 .9 5 moose density effects abaturov and smirnov alces suppl. 2, 2002 4 population was low, the cutovers were being actively overgrown with birch and aspen. the spruce remained in the second layer. but, as early as the 1960s, when the moose population density reached 5–6 individuals per 1,000 ha, the aspen was damaged by moose, and lagged behind in growth. at a 19-year-old cutover, the aspen did not exceed 2 m in height, and was behind the growth of other species. in the 1970s, the population density of moose was even higher (up to 10 individuals per 1,000 ha), and the aspen began to be fully removed from the developing stand. at a 7-year-old cutover, there were only dead aspen stands. the birch was also badly injured there. the percentage of injured stems was 80%; hence, birch does not develop a closed canopy, and at the cutover site there were a large number of clearings overgrown with grass. in these conditions, the regrowth of spruce was also impaired, but it was browsed only rarely and spruce development was not depressed. as a result of the absence of important competition on the part of deciduous trees, the spruce grew very well. at a 7-year-old cutover, its average height was 1 m, at a 19-year-old cutover, 5 m, and by 15– 20 years of age, it escaped from the zone of moose influence. under these conditions, the spruce, because of moose activities, rapidly occupied the dominant position in the stand, but it should be remembered that with growth of the spruce and with increase in the diameter of the stem, spruce bark may be extensively damaged by moose. injured in this way, the trees are infected by stem rot and are thus removed from the stand (smirnov 1987). in the final analysis, the activities of moose at such population density may cause degradation of the developing stand. in kostroma region, where the population of moose is low, they do not notably affect developing stands. during the first years, the cutovers are overgrown with deciduous trees, the birch, willow, and aspen predominating. aspen contributed to the primary stand only little, and because of this, only single aspen trees were observed at the cutovers. there was no effect of moose on the early successional species, which explains their rapid growth and abundance. at the 3-year-old cutover, their height already reached 130 cm, and at 7–8year-old cutovers, their height ranged between 1.5 to 2.0 m. good growth of mountain ash, which is depressed in other regions with high moose population density, is noteworthy. in this case, the number of individuals at cutovers of different age remains consistently high (2.5–3.5 thousand individuals). in the final analysis, by as early as 7–8 years, a dense deciduous canopy is formed. in such conditions, spruce is sharply lagging behind the deciduous trees. by the age of 7–8 years, it reaches the height of only 70–80 cm (as opposed to 200 cm in the birch), remaining under the canopy of deciduous trees. thus, the kostroma region, with a low density moose population, exhibits a common pattern of overgrowth of cutovers, leading to replacement of primary coniferous species by deciduous remains. conclusions the above gives grounds to conclude that in regions with increased density of moose populations, the moose has drastically changed the pattern of development of stands in clearings, which was manifested in the depression of growth of deciduous, in particular, the forage species, i.e., the aspen, and to a lesser extent the birch. as a result of that, the aspen is entirely eliminated from the developing stand. it can be expected that with preservation of the present-day population of moose, in another 30–40 years, the aspen may completely disappear from these forests as a forestforming species. the same applies to some alces suppl. 2, 2002 abaturov and smirnov moose density effects 5 other species, most vulnerable to the effects of moose. this holds for the pine, which has long ceased being a regenerative species. the oak, increasingly damaged by moose, no longer regenerates. the mountain ash, that has until recently been widely distributed and yielding fruit in the forests, at present in regions with high moose density, is represented by injured and depressed individuals, and it practically yields no fruit. being the least damaged species, the birch predominated at cutovers, but it cannot form a closed canopy there either. the spruce, the primary species of these forests, is the least affected. liberated by the moose from the depressive effect of the deciduous trees, it grows fast, outstripping other species in relation to growth and reaches the first layer. in this case, the moose causes the regeneration of primary spruce stands to accelerate. since some of the birches are preserved, being involved in the formation of the stand, there are grounds to believe that at the site of cutovers, mixed spruce-birch stands will be formed. characteristically, all these features were previously noted for other regions; in particular, for the forests of byelorussia, leningrad region, and siberia (elsky et al. 1975, dunin 1979, tikhonov 1980). this suggests the phenomenon observed is quite common. where moose density is low (under 2– 3 individuals per 1,000 ha) (kostroma region), its activity does not noticeably affect the formation of stands. however, where the moose population density is very high (>5 individuals per 1,000 ha), as has been the case in the yaroslavl region, moose activity depresses the growth of aspen and also that of spruce and birch, which results in degradation and decomposition of stands. under such conditions, the result of moose activities cannot in its complexity of effects be easily predicted. references dinesman, l. g. 1961. the effect of wild mammals on the development of forest stands. ussr academy of sciences, moscow, russia. (in russian). dunin, v. f. 1979. moose and forest regeneration. lesnoye khozyaistvo 7:65–67. (in russian). elsky, g. m., a. s. shishkin, and v. y. shvetsov. 1975. ecological assessment of the effect of herbivorous mammals on the development of young stands. pages 112–124 in conservation and management of the forests of the krasnoyarsk region. krasnoyarsk, russia. (in russian). smirnov, k. a. 1987. the role of moose in biocenoses of the southern taiga. nauka, moscow, russia. (in russian). tikhonov, a. a. 1980. on the formation of spruce stands at regular and gradual cutting areas under conditions of high moose density. lesnoi zhurnal 6:28-31. (in russian). alces16_1.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 f:\alces\supp2\pagema~1\rus 25s alces suppl. 2, 2002 minaev – telemetry of domesticated moose 89 use of telemetry to study behavior of domesticated moose alexander n. minaev institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: a telemetry system was designed to assist in the study of moose behavior at the kostroma moose farm in russia. two telemetry systems were used to locate instrumented animals and to capture physiological data from some. the activity rhythm of moose could be generalized from records of heart and respiration rates without the need for visual observation. microcomputer software was designed to process heart–rate data previously recorded on a strip chart. alces supplement 2: 89-92 (2002) key words: behavior, kostroma moose farm, moose, respiration rate, telemetry moose behavior was studied at the kostroma experimental moose farm in russia. hand–reared moose on the farm were studied with minimal interference because the animals were habituated to the presence of people. experimental animals were able to range over large areas that resembled natural moose habitat outside the farm setting. telemetry was used to study movements and home range, as well as certain physiological parameters. this paper generalizes our experience with using telemetry to study moose behavior at the kostroma moose farm over a 10–year period. methods and equipment two telemetry systems were designed; “los–2”, a simple radio–tracking transmitter that provided positional data only, and “los–3”, a system that transmitted physiological data in addition to position information (table 1). both systems operated between 166.7 and 167.5 mhz, with output less than 30 mw. transmitter range was 2 – 15 km, depending on terrain. the 330g version of los–2 had a life expectancy of 670 days (30 days for the 60g version). the los–3 transmitter could transmit for 40 days (transmitter weight = 330g). the 60g table 1. technical specifications of “los–2” and “los–3” moose telemetry systems. “los–2” “los–3” band (mhz) 166.7–167.5 166.7–167.5 maximum output power (mw) 30 30 signal range depending on the terrain (km) 2–15 2–15 ecg frequency band (hz) 0.2–300 range of ecg transmission (km) 1.5–10 expected life (days) 330g transmitter 670 40 60g transmitter 30 1 precision of radio direction finding (degrees) 1–2 1–2 telemetry of domesticated moose – minaev alces suppl. 2, 2002 92 rates of a newborn calf. portable telemetry equipment was always used when gathering physiographic data. two stationary receivers were used occasionally to determine the initial location of animals for the physiological work. moose movement and home range studies conducted at the kostroma moose farm are reported in other publications. more than 1,000 hours of telemetric records and several thousand animal locations (fixes) were made during 10 years of work at kostroma. f:\alces\supp2\pagema~1\rus9s.pdf alces suppl. 2, 2002 chalyshev and badlo nutrient composition of milk 41 nutrient composition of milk from domesticated taiga moose during the lactation period aleksandr v. chalyshev and larisa p. badlo institute of physiology, komi scientific center, ural division of the russian academy of sciences, 167610, syktyvkar gsp, komi republic, russia abstract: changes in the concentrations of mineral and organic substances in moose milk during lactation were examined in relation to lactation stage and nutritional (dry matter) intake by domesticated taiga moose. initial increases in nutrient and fat concentrations were documented in the milk during the lactation peak in june (lactation days 1–25), concurrent with the availability of high quality forage. subsequent measures of mineral element concentrations in moose milk gradually decreased (lactation days 26–100). alces supplement 2: 41-44 (2002) key words: calves, domestication, forage, milk, moose, nutrition it has been previously documented that moose (alces alces) milk has a high nutritional value and is important to the survival of neonates (franzmann et al. 1976, gelbert and kargina 1989, ivanova et al. 1991). however, little is known about the nutrient content of moose milk during the lactation period. the purpose of this paper is to document the concentrations of mineral and organic substances in the milk of domesticated taiga moose during lactation. methods we examined the nutrient composition (i.e., mineral and organic substances) of milk produced by lactating, domesticated female moose on a pechora-ilych reservation farm. data were collected over a 2year period from 8 animals of differing ages (2 – 12 years) and milk productivity (140 – 350 kg total production). during the lactation period moose were fed birch, willow, rowan-tree, and aspen branches, in addition to a maximum of 10 kg of potatoes and 3-3.5 kg of concentrates (mixed feed) per animal per day. samples were collected during every milking event 7-10 days after the calving and 3 – 4 times during each of the following periods: peak lactation (20 – 25 days post-calving), the decreasing production period (45 – 55 days post-calving), and the completion of lactation (80–100 days post-calving), in proportion to milking yield amounts. results and discussion generally, colostrum appears in the nipples a day or two before calving and the colostrum period finishes a day or two postcalving. during the next period (10–25 days post-calving) female moose achieve their highest milk productivity. these levels coincide with increases in the availability of green forage plants, which have high mineral and organic substance concentrations (potassium estimated maximum of 25 g/kg of dry matter; calcium estimated maximum of 11 g/kg of dry matter; magnesium estimated maximum of 3 g/kg of dry matter; phosphorus estimated maximum of 2.5 g/ kg of dry matter; and, protein estimated maximum of 25 g/kg of dry matter). it appears that maximum milk productivity and nutritional quality are related to the nutrient composition of milk chalyshev and badlo alces suppl. 2, 2002 42 timing of calving and the production of nutritious forage in the spring (table 1). for example, on the kostroma farm, the highest milk yield was observed during the first days of female moose lactation (gelbert and kargina 1989). the colostrum produced during the first two lactation days is rich in protein, after which time there is a decreasing percentage of fat in the milk. the milk composition of pechora taiga female moose is characterized by the amount of protein exceeding that of fat until the end of the lactation peak period. calcium and phosphorus levels are greatest during the peak of lactation, although these levels are only 20% of those found in the colostrum (table 2). sodium, potassium, and magnesium concentrations in domesticated taiga moose milk remains relatively unchanged over the entire lactation period. the maximum amino-nitrogen level (free amino acids) in the milk occurs during the peak of lactation, however concentrations of nitrogen and protein are low during this period. therefore, we suggest that because the milk composition of junecalving females has the greatest nutrient concentrations on days 8 through 11 of lactation (20th through the 25th days for may-calving females), lactation levels may be dependent upon forage availability and quality. calcium and amino-nitrogen levels in the moose milk decreased 45 – 55 days post-calving by approximately 27 and 40%, respectively. however, dry matter, protein, and nitrogen levels increased by 15 – 16%. in general, the nutritional elements of the milk decreased by an average of 1.5 from the peak of lactation. the completion of lactation (80 – 100 days post-calving) is characterized by an increase in levels of most of the mineral elements by approximately 10–20% over the previous period. in total, domesticated female taiga moose produced an average of 320-350 kg of milk t a b le 1 . n u tr ie n t c o n c e n tr a ti o n ( g /k g d ry m a tt e r) o f d o m e s ti c a te d t a ig a m o o s e m il k . l a c ta ti o n d ry t o ta l a m in o d a y m a tt e r n it ro g e n n it ro g e n p ro te in s o d iu m p o ta ss iu m c al ci u m m a g n e si u m c h lo ri d e p h o sp h a te 1 2 0 0 .0 1 3 .7 0 .2 8 7 .3 0 .3 1 .6 2 .9 0 .2 0 .8 2 .1 2 1 8 0 .0 1 2 .6 0 .2 8 0 .2 0 .3 1 .6 3 .0 0 .2 0 .9 2 .2 3 1 7 8 .0 1 3 .2 0 .2 8 4 .4 0 .3 1 .6 3 .3 0 .2 0 .9 2 .3 4 -5 1 6 0 .0 1 2 .7 0 .2 8 0 .1 0 .3 1 .5 3 .4 0 .2 0 .9 2 .3 6 -1 2 1 5 7 .0 1 2 .5 0 .4 7 9 .6 0 .3 1 .5 3 .6 0 .2 1 .0 2 .3 1 5 -2 5 1 5 2 .0 1 1 .3 0 .4 7 2 .3 0 .3 1 .5 3 .7 0 .2 0 .9 2 .4 4 5 -5 5 2 1 8 .0 1 1 .6 0 .2 7 4 .2 0 .4 1 .4 2 .7 0 .3 0 .8 2 .8 8 0 -1 0 0 2 5 4 .0 1 3 .5 0 .1 8 6 .0 0 .4 1 .6 3 .4 0 .2 0 .7 2 .7 alces suppl. 2, 2002 chalyshev and badlo nutrient composition of milk 43 t a b le 2 . n u tr ie n t c o n te n t (g /d a y ) o f d o m e s ti c a te d t a ig a m o o s e m il k . l a c ta ti o n t o ta l a m in o d a y n it ro g e n n it ro g e n p ro te in s o d iu m p o ta s s iu m c a lc iu m m a g n e s iu m c h lo ri d e p h o s p h a te 1 2 0 .9 4 0 .0 3 1 3 3 .6 0 0 .4 9 2 .3 2 4 .4 8 0 .3 2 1 .2 5 3 .2 7 2 3 5 .0 7 0 .0 6 2 2 3 .8 0 0 .8 7 4 .3 9 8 .7 7 0 .5 2 2 .5 8 6 .2 5 3 4 3 .5 5 0 .0 7 2 7 7 .7 0 1 .2 0 5 .7 6 1 2 .0 4 0 .7 0 3 .5 9 8 .5 3 4 -5 4 2 .0 0 0 .0 7 2 7 1 .5 4 1 .0 0 4 .9 8 1 1 .1 4 0 .6 6 3 .0 1 7 .6 8 6 -1 2 4 2 .3 0 0 .1 4 2 6 9 .8 0 1 .0 3 4 .9 4 1 2 .2 6 0 .7 0 3 .2 1 7 .9 9 1 5 -2 5 5 4 .6 6 0 .1 9 3 4 8 .5 0 1 .3 8 7 .0 0 1 7 .6 8 1 .0 9 4 .4 3 1 1 .7 6 4 5 -5 5 3 9 .9 3 0 .0 7 2 5 4 .5 0 1 .2 4 4 .6 9 9 .5 2 1 .0 0 2 .6 0 9 .3 7 8 0 -1 0 0 3 9 .8 3 0 .0 3 2 5 3 .7 0 0 .4 3 1 .9 8 3 .8 9 0 .2 1 0 .8 6 3 .1 8 (kudryavtzeva 1976, sivoha 1991) and up to 200 g of sodium, 700 g of potassium, 1,300 g of calcium, 110 g of magnesium, 300 g of chlorine, 1,000 g of phosphorus, and 2,500 g of nitrogen. therefore, it appears that the milk of domesticated taiga moose possesses its highest nutritional value during the peak of lactation, although protein levels are high through to the completion of lactation. in contrast with the nutritional content of milk from other domesticated ruminants, which do not exhibit sharp changes in the nutrient levels, the nutritional quality of milk from wild and domesticated female moose appears to be dependent upon the lactation period and the quantity and quality of available forage. references franzmann, a. w., a. flynn, and p. d. arneson. 1976. moose milk and hair elements levels and relationships. journal of wildlife diseases 12:202-207. gelbert, m. d., and m. s. kargina. 1989. moose as a productive animal: about ratio of fat and protein in the milk. agricultural biology 2:95-98. (in russian). ivanova, g. m., m. v. kojhyhov, and a. f. simakov. 1991. biochemical characteristics of moose milk. pages 100–107 in biological studies in pechora–ilych reserve, syktyvkar. works of the komi scientific center, ural division, russian academy of sciences, no. 116. (in russian). kudryavtzeva, e. n. 1976. changes of fat amount of female moose milk during the lactation. pages 134–142 in works of pechora–ilych reserve, part 13. komi publishing house, syktyvkar, russia. (in russian). sivoha, i. n. 1991. milk productivity of domesticated moose. pages 126–130 in biological studies in pechora–ilych nutrient composition of milk chalyshev and badlo alces suppl. 2, 2002 44 reserve, syktyvkar. works of the komi scientific center, ural division, russian academy of sciences, no. 116. (in russian). alces14_attendeessvii.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces20_27.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces16_398.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces17_282.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces14_68.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 moose modify bed sites in response to high temperatures bryce t. olson1, steve k. windels1, ron a. moen2, and nicholas p. mccann3 1national park service, voyageurs national park, 360 highway 11 e, international falls, minnesota 56649; 2natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811; 3great lakes indian fish and wildlife commission, p.o. box 9, odanah, wi 54861 abstract: moose (alces alces) employ physiological and behavioral mechanisms to enable them to dissipate excess heat when ambient temperature is above the upper critical temperature of their thermoneutral zone. in this note, we describe 2 cases where gps radio-collared female moose modified summer bed sites as a potential thermoregulatory response to high temperatures. the first case occurred on 18 21 july 2011 when ambient temperatures averaged 25 °c (8 °c above the upper critical temperature of moose) and reached 32 °c and 96% relative humidity. based on field observations of the bed site immediately after use, the moose cleared litter and duff to expose 3 m2 of mineral soil under a closed-canopy balsam fir (abies balsamea) stand. the moose spent 64% of the time bedded during a 4-day event, with ≤11 individual bedding events in the same bed site. a second case was observed on 5 july 2013 during similar weather conditions (29 °c and 70% relative humidity) when a different moose cleared a bed site and used it continuously for 10 hours. alces vol. 52: 153–160 (2016) key words: alces alces, bedding, bed site, behavior, moose, temperature, thermoregulation moose (alces alces) have a low upper critical temperature estimated at 14 °c (renecker and hudson 1986) and 17 °c (mccann et al. 2013) in summer during calm conditions, and under windy conditions mccann et al. (2013) estimated an upper critical temperature of 24 °c. to accommodate these physiological constraints, moose can be expected to rely on behavioral plasticity. for instance, to maintain thermal balance when summer temperatures are high, moose can be expected to reduce daytime activity and increase nocturnal activity (dussault et al. 2004). when daytime temperatures are high, moose can also be expected to select habitats that serve as thermal refuges, including areas with thick vegetation and dense canopy cover (demarchi and bunnel 1995, dussault et al. 2004, van beest et al. 2012, melin et al. 2014, street et al. 2016), and shift nocturnal habitat use to openings (mccann et al. 2016). moose spend approximately half of each summer day bedded when they ruminate and rest (renecker and hudson 1989b, moen et al. 1997). proximity to browse and predator avoidance are commonly considered to be the primary drivers that influence where moose locate beds (mysterud and østbye 1999, van moorter et al. 2009). however, beyond these primary factors that influence general space use, thermal conservation can drive finer scale behaviors and influence where moose locate their bed sites, as moose use shaded bed sites with wet substrates when daytime temperatures are hot (mccann et al. 2016). moose could increase thermal conduction to the ground on hot summer days by exposing soil under the litter layer. bedding reduces energy expenditure (relative corresponding author: bryce olson, 360 highway, 11 e international falls, mn 56649, office (218) 283-6694, fax (218) 285-7407, bryce_olson@nps. 153 mailto:bryce_olson@nps. to walking; renecker and hudson 1989a), and increases thermal conduction to the ground (moen 1973, gatenby 1977). the litter layer has lower thermal conductivity than bare soil (hanks et al. 1961, barkley et al. 1965), retains moisture, and stops most incoming radiation (kelliher et al., 1986, schaap and bouten 1997, ogée et al., 2001) which dampens the fluctuations of soil temperature (johnson-maynard et al. 2004). ungulates other than moose are known to modify bed sites to enable cooling. mule deer (odocoileus hemionus) dig into the soil to modify bed sites when temperatures are high (sargeant et al. 1994) and wapiti (cervus elaphus) select bedding areas with bare soil (e.g., merrill 1987, millspaugh et al. 1997). these behaviors, however, have not been documented for moose. in this case study, we describe modification of bed sites by moose as a potential thermoregulatory response to extreme thermal conditions in summer. study area we studied moose in voyageurs national park (vnp; utm 15u 508618 / 5371882) in northcentral minnesota along the ontariocanada border. the climate in this region is characterized by long winters (mean january temperature = �15 °c) and short summers (mean july temperature �19 °c; noaa 2010), with vegetative communities comprised of a mix of southern boreal and laurentian mixed conifer-hardwood forests (faber-langendoen et al. 2007). common species include quaking aspen (populus tremuloides), paper birch (betula papyrifera), white spruce (picea glauca), and balsam fir (abies balsamea) in uplands, and northern white cedar (thuja occidentalis), black spruce (picea mariana), and black ash (fraxinus nigra) in forested lowlands. moose density is low in vnp (<0.15 moose/km2; windels and olson 2016). methods we immobilized adult female moose in mid-winter via dart gun from helicopters (quicksilver air inc., fairbanks, alaska, usa) using 1.2 ml (4.0 mg/ml) carfentanil citrate and 1.2 ml (100 mg/ml) xylazine hcl for immobilization. we subsequently used a combination of 7.2 ml (50 mg/ml) naltrexone hcl and 3 ml (5 mg/ml) yohimbine hcl as an antagonist. a gps radiocollar that obtained a location every 20 min was fitted to each immobilized moose (sirtrack limited, hawkes bay, new zealand). animal capture and handling protocols met the guidelines recommended by the american society of mammalogists (sikes et al. 2011) and were approved by u.s. geological survey and university of minnesota animal care and use committees. the first case of a modified bed site (utm 15u 503608/5368384) was observed while checking on the status of a female moose that had moved minimally for several days (assessed from gps locations via argos upload). the second case occurred when a bed site was found (utm 15u 518511/ 5369230) opportunistically while conducting other field work. gps location, photos, and size measurements of each bed site area were taken and a 50 m radius area was searched for additional bed sites, browsing, and other moose sign. we verified that the locations of the modified bed sites corresponded with gps locations of each collared female moose. to do this, we created a 25 m radius buffer around each bed site’s gps location. this buffer equaled the 95% circular error probable for stationary tests of the collars we used (r. moen, unpubl. data). we then examined gps locations from all collared moose in vnp using arcgis 10 (esri 2011) to confirm that locations from other collared moose did not overlap the 25 m buffer. because thermal conditions were predicted to influence bed site modification, 154 moose bed site modification – olson et al. alces vol. 52, 2016 we obtained ambient air temperature, relative humidity, and solar radiation data for the period of use from the vnp meteorological station nearest the bed sites (mean 10.1 ± 1.3 km, sullivan bay castnet). previous research found that temperatures from nearby weather stations were correlated with temperatures recorded by other gps radiocollars on moose (ericsson et al. 2015). results case 1 an area of 2 m x 1.5 m was nearly cleared entirely of litter and duff leaving the underlying mineral soil exposed (fig. 1), and moose hoof prints and hair were found within the bed site. further investiga‐ tion revealed multiple hoof scratch marks suggesting deliberate removal of the litter and duff layers as indicated by it being piled at the edges of the bed site. no other bed site was found within a 50 m radius around the modified bed site, and evidence of recent intensive browsing was within 50 m of the bed site. the bed site and the surrounding area was within the deciduous shrubland cover type, with >90% canopy coverage of balsam fir overstory immediately above the bed site. soil type at the bed site was sandy loam and relatively dry at the time of observation. sixty-four percent of gps locations over a 96-h period (18 21 july 2011) were within the 25 m radius buffer around the modified bed and accurately preceded the period when the freshly disturbed site was observed fig.1. modified bed site from case 1 showing litter and duff cleared and piled to the edges by gps radio-collared female moose in northern minnesota, usa. the modified bed site in case 2 was nearly identical. alces vol. 52, 2016 olson et al. – moose bed site modification 155 on 3 august 2011. during this period, the moose made 8 short (< 75 m) and 3 longer movements (200-400 m) to and from the bed site. the shorter movements were probably foraging events or movements to nearby water; most were within the area of intensive browsing (fig. 2). mean temperature during the 96-h period was 25 °c, with a maximum daily high of 32 °c (fig. 2). relative humidity averaged 67%, with a maximum of 96%. skies were mainly clear with only a single, short precipitation event during this time period. case 2 the second case of a modified bed site was similar to case 1 but the cleared area was smaller (1.5 m x 1.1 m). moose hoof prints and hair were found within the bed site and no other bed sites were found within a 50 m radius surrounding the bed site; little browsed vegetation was observed. the bed site and surrounding area were in the boreal forest cover type, with >90% canopy coverage of balsam fir above the bed site. soil type at the bed site was sandy loam and it was relatively dry at the time of observation. all gps locations were in the 25 m buffer around the bed site location consecutively for 10 h on 5 july 2013. one movement of ~50 m followed by 1 location back at the bed site was recorded before the moose left the area without returning. mean temperature during the 10-h event was 25 °c, with a maximum daily high of 29 °c (fig. 3). relative humidity averaged 62%, with a maximum of 70% (fig. 3). skies were mainly clear and no precipitation was recorded. discussion the environment selected by animals influences their ability to maintain thermal homeostasis when temperatures are outside of their thermoneutral zone, and animals exhibit a wide range of behavioral responses to fig. 2. ambient air temperature (°c; solid black line) and relative humidity (%; dotted black line) during long bedding events for gps radio-collared female moose on 18 21 july 2011 (case 1). gray bars indicate when the moose was within 25 m of the bed site and assumed to be bedded. cross-hatched bars indicate when the moose moved from the bed site to nearby standing water. moose locations in other areas are indicated by the white. air temperatures were recorded at the sullivan bay castnet meteorological station near ash river, minnesota, usa. 156 moose bed site modification – olson et al. alces vol. 52, 2016 balance energy gains and losses (moen 1968). bed site modification and selection of bed site locations in response to high temperatures has been documented in other ungulates, including mule deer and wapiti (sargeant et al. 1994, merrill 1987, millspaugh et al. 1997). moose commonly exhibit thermoregulatory behavior (demarchi and bunnel 1995, dussault et al. 2004, van beest et al. 2012, melin et al. 2014, street et al. 2016) and it should not be surprising for them to exhibit these same behaviors. moose cleared away duff and litter to expose mineral soil under a closed canopy. exposing soil would have improved thermal conductivity at bed sites and reduced internal heat loads during periods of extreme thermal conditions, thereby helping moose maintain thermal balance. it is highly probable that increasing thermal conductivity was important in both cases because of the extreme thermal conditions. habitats around each bed site fit the description of thermal refuges that moose select for bed sites during hot temperatures (mccann et al. 2016). the bed site modification behavior we describe may be most common when moose bed in drier soils as it is known that moose use bed sites with higher moisture content during times of high temperatures (mccann et al. 2016), and bed in wet soils in hot weather even when standing water exists (renecker 1987). energy expenditure was reduced 9% and respiration rate 29% when moose bedded in wet versus dry soils (renecker 1987). in both cases we describe, the soils were sandy loam that was relatively dry at the time of use. removing the insulating litter should have improved conductivity by exposing the cooler soil underneath (hanks et al. 1961, barkley et al. 1965, kelliher et al. 1986, schaap and bouten 1997, ogée et al. 2001). however, it is possible the soil was moist when the bed site was modified and dried out prior to our inspections. in both cases, the hoof marks extending to the edge of the bed site strongly suggest this material was deliberately cleared while standing. during times when temperatures fig. 3. ambient air temperature (°c; solid black line) and relative humidity (%; dotted black line) during long bedding events for gps radio-collared female moose on 5 july 2013 (case 2). gray bars indicate when the moose was within 25 m of the bed site and assumed to be bedded. moose locations in other areas are indicated by white. air temperatures were recorded at the sullivan bay castnet meteorological station near ash river, minnesota, usa. alces vol. 52, 2016 olson et al. – moose bed site modification 157 were above the upper critical temperature, moose in alberta were observed bedded in shade with legs extended (renecker and hudson 1989a, 1990). this repeated leg extension and process of lying down and getting up could shift some vegetative material to the edges of the bed site, though the litter would likely remain in the middle of the bed site. in the cases we describe, the entire bedding area, similar in size to other moose bed sites investigated in the area (nps unpubl. data), was cleared of litter and duff. moose have been viewed scratching at the ground prior to bedding in the kenai peninsula of alaska, though meteorological conditions and habitat characteristics were not documented (j. crouse, pers comm). addison et al. (1993) recorded cow moose creating small “scratch holes” during the calving season in central ontario. these scratch holes were attributed to territorial markings due to their positioning near calving sites in a high density moose area. the bed sites we describe were created outside the calving season and moose densities were low in our study area (windels and olson 2016), suggesting they reflect thermoregulatory behavior not territorial marking. the behavior we describe appears to be uncommon in northern minnesota as mccann et al. (2016) visited 57 bed sites within 51 days of use by gps radio-collared moose (n = 8) and never observed similar modification of bed sites in the larger study area. additionally, they visited another 155 bed sites in 2012 that were made the previous summer and did not observe bed site modifications. though some evidence of modification could have been obscured during the ≥10 months (e.g., by fall litter deposition or vegetative growth the following spring) between use of a bed site and the subsequent visitation, we believe that the deliberate exposition of soil would have been evident at some sites. while sample size limits the inferences which can be drawn from this case study, we describe a thermoregulatory behavior undocumented for moose, though not uncommon for other ungulates. further research might better explain how this behavior varies across moose range and in response to different environmental conditions. these studies should focus on determining the potential thermal benefits of bed site modification in various habitat types, the frequency of occurrence under different climatic regimes and landscape compositions, and potential reasons for bed site fidelity. references addison, e. m., r. f. mclaughlin, d. j. h. fraser, and m. e. buss. 1993. observations of preand post-partum behavior of moose in central ontario. alces 29: 27–33. barkley, d. g., r. e. blaser, and r. e. schmidt. 1965. effect of mulches on microclimate and turf establishment. agronomy journal 57: 189–192. demarchi, m. w., and f. l. bunnell. 1995. forest cover selection and activity of cow moose in summer. acta theriologica 40: 23–36. dussault, c., j. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioural responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. ericsson, g., h. dettki, w. neumann, j. m. arnemo, and n. j. singh. 2015. offset between gps collar-recorded temperature in moose and ambient weather station data. european journal of wildlife research 61: 919–922. esri 2011. arcgis desktop: release 10. environmental systems research institute, redlands, california, usa. faber-langendoen, d., n. aseng, k. hop, m. lew-smith, and j. drake. 2007. vegetation classification, mapping, and monitoring at voyageurs national park, 158 moose bed site modification – olson et al. alces vol. 52, 2016 minnesota: an application of the u.s. national vegetation classification. applied vegetation science 10: 361–374. gatenby, r. m. 1977. conduction of heat from sheep to ground. agricultural meteorology 18: 387–400. hanks, r. j., s. b. bowers, and l. d. boyd. 1961. influence of soil surface conditions on net radiation, soil temperature, and evaporation. soil science 91: 233–239. johnson-maynard, j. l., p. j. shouse, r. c. graham, p. castiglione, and s. a. quideau. 2004. microclimate and pedogenic implications in a 50-year-old chaparral and pine biosequence. soil science society of america journal 68: 876–884. kelliher, f. m., t. a. black, and d. t. pierce. 1986. estimating the effects of understory removal from douglas fir forest using a two-layer canopy evaporation model. water resources research 22: 1891–1899. mccann,n.p.,r.a.moen,andt.r.harris. 2013. warm-season heat stress in moose (alces alces). canadian journal of zoology 91: 893–898. ________, ________ s. k. windels, and t. harris. 2016. bed sites as thermal refuges for a cold-adapted ungulate in summer. wildlife biology 22: 228–237. melin, m., j. matala, l. mehtatalo, r. tiilikainen, o. tikkanen, m. maltamo, j. pusenius, and p. packalen. 2014. moose (alces alces) reacts to summer temperatures by utilizing thermal shelters in boreal forest – an analysis based on airborne laser scanning of the canopy structure at moose locations. global change biology 20: 1115–1125. merrill, e. h. 1987. habitat ecology and population dynamics of mount st. helens elk. ph. d. dissertation, university of washington, seattle, washington, usa. millspaugh, j. j., k. j. raedeke, g. c. brundige, and c. c. willmott. 1997. summer bed sites of elk (cervus elaphus) in the black hills, south dakota: considerations for thermal cover management. the american midland naturalist 139: 133–140. moen, a. n. 1968. the critical thermal environment: a new look at an old concept. bioscience 18: 1041–1043. __________. 1973. wildlife ecology: an analytical approach. w. h. freeman and company, san francisco, california, usa. moen, r. a., j. pastor, and y. cohen. 1997. a spatially explicit model of moose foraging and energetics. ecology 78: 505–521. mysterud, a., and e. østbye. 1999. cover as a habitat element for temperate ungulates: effects of habitat selection and demography. wildlife society bulletin 27: 385–394. noaa (national oceanic and atmospheric administration). 2010. climatological data for international falls, minnesota. national climatic data center, ashville, north carolina, usa. ogée, j., e. lamaud, y. brunet, p. berbigier, and j. m. bonnefond. 2001. a long-term study of soil heat flux under a forest canopy. agricultural and forest meterology 106: 173–187. renecker, l. a. 1987. bioenergetics and behavior of moose (alces alces) in the aspen dominated boral forest. ph. d. thesis, university of alberta, edmonton, canada. ____________, and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. ____________, and ____________. 1989a. ecological metabolism of moose in aspen-dominated boreal forests, central alberta. canadian journal of zoology 67: 1923–1928. ____________, and ____________. 1989b. seasonal activity budgets of moose in aspen-dominated boreal forests. journal of wildlife management 53: 269–302. alces vol. 52, 2016 olson et al. – moose bed site modification 159 renecker, l. a, and r. j. hudson. 1990. behavorial and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspendominated forests. alces 26: 66–72. schaap, m. g., and w. bouten. 1997. forest floor evaporation in a dense douglas fir stand. journal of hydrology 193: 97–113. sargeant, g. a., l. e. eberhardt, and j. m. peek. 1994. thermoregulation by mule deer (odocoileus hemionus) in arid rangelands of southcentral washington. journal of mammology 75: 536–544. sikes, r. s., w. l. gannon, and the animal care and use committee of the american society of mammalogists. 2011. guidelines of the american society of mammalogists for the use of wild mammals in research. journal of mammalogy 92: 235–253. street, g. m., j. fieberg, a. r. rodgers, m. carstensen, r. moen, s. a. moore, s. k. windels, and j. d. forester. 2016. habitat functional response mitigates reduced foraging opportunity: implications for animal fitness and space use.landscape ecology: 1–15. sullivan bay castnet. meteorological station data. http://ard-request.air-resource. com/data.aspx (accessed february 2015). van beest, f. m., b. van moorter, and j. s. milner. 2012. temperature-mediated habitatuse and selection by a heat-sensitive northern ungulate. animal behavior 84: 723–735. van moorter b., j. m. gaillard, p. d. mcloughlin, d. delorme, f. klein, and m. s. boyce. 2009. maternal and individual effects in selection of bed sites and their consequences for fawn survival at different spatial scales. oecologia 159: 669–678. windels, s. k., and b. t. olson. 2016. voyageurs national park moose population survey report 2016. natural resource data series nps/voya/nrds-2016/ 1031. national park service, fort collins, colorado, usa. 160 moose bed site modification – olson et al. alces vol. 52, 2016 http://ard-request.air-resource.com/data.aspx http://ard-request.air-resource.com/data.aspx moose modify bed sites in response to high temperatures study area methods results case 1 case 2 discussion references f:\alces\supp2\pagema~1\rus 23s alces suppl. 2, 2002 polehzaev and korolev – moose in the komi republic 109 structure of the moose population and its utilization in the komi republic n. m. polehzaev and i. c. korolev institute of biology, komi scientific center, ural division of the russian academy of sciences, 167610, syktyvkar gsp, komi republic, russia abstract: cutover stands comprise a large proportion of the forested area within the komi republic. the natural succession of these stands leads to a decrease in productivity of moose populations. we describe characteristics of moose hunting in the komi republic and discuss methods for managing harvest by means of age– and sex–specific hunting licenses. poaching, decreased habitat productivity due to forestry, and predation all impact moose populations adversely and must be addressed by managers to ensure future harvests of moose. alces supplement 2: 109-112 (2002) key words: forest productivity, forest succession, forestry, harvest, hunting, predation intensive forestry drastically affects the landscapes of the european north. presently, in the komi republic, young stands occupy more than 4.8 million ha, or 17% of all the area covered by forests. further to the north, there is a more intensive process of coniferous species replacement by deciduous species (larin and pautov 1989). a significant mosaic is, as a rule, observed on the territories affected by concentrated timber harvest. these factors result in a decrease in the biological activity of moose (alces alces). the number of moose in the northeastern portion of the european part of the ussr suffered great changes. at the beginning of the 20th century, moose were a rare species, but by 1939–1940 they were common. in 1947, 12,000 individuals inhabited this territory. by 1951–1952, as a result of exhaustive hunting, the abundance of moose decreased to 4,000, but by 1969 it again increased up to 10,000 individuals (ostroumov 1953, 1972). preservation and improvement of the forage base in the cutovers stimulated the growth of the population. according to the komi hunting organization, in 1984 the total number of moose was about 19,000 individuals, in 1985 it increased to 20,000, and in 1986 it increased again to 25,000 animals. at present [1990], the total number is stable at about 26,000 moose, but in some regions of the republic there are significant variations. the analyses of hunting license data showed that among the animals hunted, males prevail in all the age groups. in southern regions, the percentage of females increases among adult animals and there is a decrease in the abundance of young animals in the harvest. on the rest of the territory the percentage of young individuals is about 20%. the body weight given in license reports is rather variable: adult males weighed 276–320 kg. the lowest body weight of young animals during the harvesting season is observed in the north of the region (inta region); there are no reliable data on the variation of body weight among adult animals. many hunters in their questionnaire emphasize “rejuvenation” of the population, which results from an intensification of hunting effort. they also note the decrease in the size of animals. after moose in the komi republic – polehzaev and korolev alces suppl. 2, 2002 110 accepting a differentiated price of a license (dependent on animal age), hunting of young individuals increased. in the northwestern part of the region, where moose hunting is based on migrants, the harvest of calves amounts to 12.5% of the total harvest; in the regions of most intensive timber harvest (ustkulom region) this figure is 25%. harvest of adult animals by these license holders amounts to 60–70% (in the northwest it ranges up to 74.7%). the carcass weight of the harvested animals indicates that most of the adult males harvested are in the 3.5–5.5 year–old age classes. for females there is more variation in age. among adult females, 65–75% (but ranging from 56–100%) have calves, about half of them with twins. in the southern part of the komi republic the percentage of twins is lower (24–40%). moose cows with 3 calves are rare (0.07%). in the upper pechora region, 53.8% are females; the percentage of the harvest comprised of adult animals is 58.9%, 26.7% 1–year–old animals, and 14.4% calves. the mean index of fertility in moose females with calves is 1.5+0.5 embryos. moose females with twins make up 37.2% of all females harvested and cows without calves comprise 19.8%. the accuracy of body weights of moose reported from license returns is low. it is necessary to have one unique, reliable criterion for a more objective evaluation. the mass of the heart, easily estimated, can become such a criterion. according to the weight of this organ, the mass of an animal can be estimated more precisely. with the constantly growing demand in licenses, it is necessary to ask hunters to pay serious attention to such duties as estimation of the heart mass, the number of embryos, and placental scars, etc. with all the data available, it is possible to obtain objective information about the state of the moose population in the region. most effective moose harvesting is at the end of the season. in 1983–1984, the hunting season lasted through the end of january. in october 1983, 5.8% of license holders hunted. in november that figure was 15.3%, and in january it was 50%. in the 1988–1989 season, moose harvesting lasted until mid–february. in october, 1.5% of license holders hunted. in november that estimate was 19.5%, in december and january it was 21%, and in february it was 27%. this shift of the main hunting season to a later period results from climatic factors. common methods of moose hunting in the komi republic are individual shooting and stalking. the results depend on the quality of snow cover and weather. successful harvesting of migrants depends on the depth of snow. winter moose migration starts, as a rule, when the depth of snow approaches 1.5 m (usually in december). five days are required to shoot one animal. it is not easy to transport moose carcasses within the komi republic; sometimes the distance from the place of harvest to the point of marketing is more than 1,000 km. snow–cars (“buran”) and rifled guns are used for the most part on state–industrial– farms. at the places where hunters use smooth–bore guns there are more animals with old wounds. productivity of moose harvesting is also improved by the possibility of conducting hunting on remote territories, but to ensure timely meat transportation all this becomes possible only after tracks are open. that is why the time of hunting must be set by taking into consideration regional environmental characteristics. in the komi republic, natural regeneration is the preferred method of reforestation. more than 90% of all cutovers are regenerated that way. as a result of forest succession, the region is characterized by the substitution of pinus sylvestris by picea obovata. in winter, p. obovata is an alces suppl. 2, 2002 polehzaev and korolev – moose in the komi republic 111 important food item for moose only in the subpolar urals; on the rest of the territory it is not reported as significant. evidently this is the reason that young stands of coniferous trees are not badly damaged by moose. the animals intensively browse young aspen (populus tremula) growth on cutovers, which results in their shading by birch. with the onset of deciduous forest stands dominating the cutovers, the forage capacity of moose habitats will decrease. the primary wintering areas for moose are partial cutovers dominated by spruce. as a result, the browsing intensity on such forage plants as young larch (larix spp.) and rowan tree (sorbus aucuparia) increases. it leads to their degradation – drying out and disappearance from the stand. this is followed by a decrease in winter carrying capacity. in the middle taiga subzone, the lack of winter habitats is felt already. naturally it affects the state of the moose population. with the size of the known moose population at present, licensed harvesting of the animals is about 12%. such a harvest does not seriously affect the stability of the population, but annual losses increase due to poaching. according to the opinions of the authorities on hunting farms and hunters themselves, 16–17 animals are harvested for every 10 individuals reported through a hunting permit. officially, not more than 10 violations are registered every year. in this way, anthropogenic removal is increased by up to 5,000 animals annually, which is more than 19% of the population. most intensive moose harvesting is observed along highways and railways, where changes in the structure of the moose population and decreases in density are most evident. the impact of predators such as brown bear (ursus arctos), wolf (canis lupus), and skunk bear (wolverine, gulo gulo) should also not be ignored. moose are preyed on by bears during the rut (yazan 1972) and also in spring, when bears leave their dens. wolves are responsible for the largest part of moose population losses; the percentage of wolves, by the way, is also rather high. after aerial hunting of predators, it was confirmed that in the northern subzone in winter, moose are the primary prey of wolves. moose hair was found in the digestive tracts of all the animals investigated, and all the wolves were in good physical condition. the examination of wolf carcasses showed that every adult female had 6 embryos. that is why the problem of wolf control is one of the most important in the komi republic; its success would save quite a number of ungulates. in some regions of the komi republic, the effects of anthropogenic factors account for changes in the density and structure of the moose population. for the rational utilization of the moose stock, hunting should be differentiated in such a way that commercial harvest of the population would correspond to its density, food quality, and habitat conditions. to stabilize the existing quantity of moose hunting licenses, it is necessary to limit non–licensed harvesting, to take certain measures to limit the wolf population, and regenerate partial cutovers in the middle taiga. when setting the timing of a hunting season, it is necessary to take into consideration biological, climatic, and economic factors. some steps should be taken to collect objective information about the state of the population. for this purpose hunters need to accurately report their activities on their licenses. references larin, v. b. 1979. natural and artificial reforestation on concentrated cuttings of the northeast of the european part of the ussr. pages 5–23 in ecology of growth and development of pine and spruce in the northeast of the european part of the ussr. komi science center of the russian academy of sciences, moose in the komi republic – polehzaev and korolev alces suppl. 2, 2002 112 syktyvkar, russia. (in russian). , and y. a. pautov. 1989. formation of young coniferous stands on cutovers of the northeast of the european part of the ussr. 144 pp. (in russian). ostroumov, n. a. 1953. wildlife of taiga. pages 38–40 in productive resources of the komi assr. (in russian). . 1972. wildlife of the komi assr: vertebrates. syktyvkar, russia. (in russian). yazan, y. p. 1972. moose. pages 178– 187 in commercial animal species of pechora taiga. kirov, russia. (in russian). alces16_238.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 f:\alces\supp2\pagema~1\rus6s.pdf alces suppl. 2, 2002 bogomolova and kurochkin parturition activity 27 parturition activity of moose ekaterina m. bogomolova and yuriy a. kurochkin anokhin research institute of normal physiology, russian academy of medical science, moscow, russia abstract: behavior of female moose (alces alces) during parturition was studied in 1977–1990 on the kostroma experimental moose farm. we found that moose parturition behavior is organized on the systems principle and aimed at the calf’s survival. the corresponding system is formed only at the time of parturition on the basis of inherent elements of behavior. we report the results of our investigation of cardiac and respiratory dynamics (as indicators of emotional states) during parturition. alces supplement 2: 27-31 (2002) key words: behavior, heart rate, moose, parturition, telemetry the concept of systemogenesis suggested by anokhin (1948) reveals aspects in an organism’s prenatal development that specifically prepare it for the postnatal encounter with species-specific factors of the environment. as a result of this series of events, the bases of functional systems of newborn behavior are formed that must help the newborn adequately fit into its new environment (bogomolova and kurochkin 1984, 1985, 1987). our data revealed that the adaptive abilities of the newborn are necessary, but obviously insufficient, for its survival and normal development after birth. even such precocial newborns as moose calves need their mother’s specific care during the first period of their lives. this species-specific maternal care, along with the newborn’s innate adaptive behavior, is designed strictly to prepare the calf for the critical timing of its main biological tasks. there have been few studies of the parturition behavior of moose (altmann 1963, stringham 1974, cederlund 1987). we designed the present study to investigate ethological and physiological aspects of moose activity, especially those aspects that enhance the normal species-specific development of newborn calves. we used the systems approach before, during, and just after labor. methods this study was conducted in 1977–1990 on the kostroma experimental moose farm in natural or near-natural environments. the behavior and heart rate (hr) characteristics of moose cows before, during, and after labor were studied, as well as those of their fetuses and newborn calves. emotional reactions of animals were characterized on the basis of hr and respiratory dynamics in biologically important situations. the study animals were 30 female moose, which had many births during the investigation period, and 225 newborn calves. behavior before, during, and several hours after parturition was investigated in large (0.5–2 ha) enclosures with natural vegetation or directly in the forest. we observed the entire parturition and postpartum interaction of mother and newborn in 62 situations. in 32 other cases, the observation of mother-newborn interaction began shortly after labor. radiotracking systems “los-2” and bioradiotelemetric systems “los-3” (departurition activity bogomolova and kurochkin alces suppl. 2, 2002 28 signed by f. m.minaev) were used, as well as night-vision apparatus and radio communication between observers by means of portable radiostations (“lastochka-m”). animal responses were captured using still photography, film, and videorecording, and moose vocalizations were recorded on magnetic audiotape for subsequent sonographic analysis. to estimate heart rate dynamics, hr diagrams were made either by averaging the 3–5 rr-intervals or by calculating without averaging. results and discussion our findings demonstrated the seasonal character of moose parturition. in the kostroma region, moose generally give birth in may; 70% of calves are born during the first half of may (bogomolova and kurochkin 1984). this is a cornerstone of the entire systemic organization of moose cow parental behavior. in early may the growth of green plants begins, which are the main food for moose calves and their mothers. thus, the offspring gain access to forage as early as possible and have the opportunity to procure the optimal mass to survive during winter. moose in the last period of pregnancy revealed the same specific alterations in their behavior and physiological state described in other ungulates: swelling of the udder, appearance of colostrum, and refusing preferred food. additionally, we found an increase in the locomotor activity of cow moose 2–3 days before labor. having the opportunity to observe the behavior of radio-tagged animals, we found 2 typical features of moose prepartum behavior. first, moose seem to remember the location of previous births and every year, before their next labor, they returned to the same area of the forest. we found 4 subsequent birthing locations of radio-tagged moose that were less than 200 m from each other (the whole area of the cow’s home range is about 60 km2). second, 1–3 days before labor, cows became aggressive to yearlings and drove them away. at the same time, however, cows not pregnant continued to stay with their yearlings for several more months. these behaviors, well-timed to the moment of labor, increase the probability of newborn calf survival, therefore, they are useful for the species. breathing rate (br) of moose in the late period of pregnancy fluctuated between 15 and 60 respirations per minute; often there were breathing delays for 10–25 seconds. generally the hr of cows in normal circumstances was relatively regular. there were short bursts (5–30 sec) of rapid hr, increasing up to 100–105 beats per minute (bpm) during strong muscular efforts and in response to significant external signals. the hr and br indices of pregnant cows in a quiet state were much more variable than moose that were not pregnant. fetal hr in the last month before parturition was about 2 times higher and more variable than the mother’s hr. short-term (2–5 sec), sudden increases of fetal hr up to 190 bpm were observed episodically. they do not correlate with the mother’s and other fetus’ hr alterations. fetal ecg revealed long-term (up to 30–40 min) periods of rather stable basic hr; for example, 100–120 or 120–130 bpm. relative independence of mother and fetus hr was found by comparisons in different situations. such autonomy of fetal heart activity may provide the optimal conditions for its development. one peculiarity of moose prepartum behavior, important for successful calf survival, is the tendency to find the most solitary place for labor. only 2 of 140 labors that we observed happened in the presence of other cow moose. such behavior of a solitary forest animal is obviously purposeful from the biological point of view, especially because the other cows may be agalces suppl. 2, 2002 bogomolova and kurochkin parturition activity 31 showed excitation, displayed threatening facial expressions, and sometimes kicked their calves with the front legs. each calf’s survival is guaranteed not only by its mother’s behavior, but also by the set of its own inherent reactions (oral automatism, following response, etc.). references altmann, m. 1963. naturalistic studies of maternal care in moose and elk. pages 233-255 in h.l. rheingold, editor. maternal behavior in mammals. john wiley & sons, new york, new york, usa. anokhin, p. r. 1948. system organization for the general conformity of evolutionary processes. bulletin of experimental biology and medicine 26:81–90. (in russian). bogomolova, e. m., and y. a. kurochkin. 1984. parturition in a moose-cow. behavior of a moose-cow and a newborn calf. zoological journal 63:1713– 1724. (in russian). , and . 1985. system organization of behavioral acts of animals in natural existence conditions. bulletin of the academy of medical science 2:79–85. (in russian). , and . 1987. system organization of behavioral acts. pages 353–369 in functional systems of organisms. medicina, moscow, russia. (in russian). cederlund, b. m. 1987. parturition and early development of moose (alces alces l.) calves. swedish wildlife research supplement1:399–422. stringham, s. f. 1974. mother-infant relations in moose. naturaliste canadien 101:325–369. alces14_157.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty richard b. harris1, michael atamian2, howard ferguson3, and ilai keren1 1washington department of fish and wildlife, 600 capital way north, olympia, washington 98504; 2washington department of fish and wildlife, 2315 n. discovery place, spokane valley, washington 99216; 3retired abstract: the state of washington was historically considered to be unoccupied by moose (alces alces) with initial colonization in the 1920s primarily in the northeastern quarter of the state. all evidence indicates a steadily increasing population since, with moose and moose hunting now firmly established. given the expectation that washington's moose population will face increasing challenges in the coming decades, our monitoring objective is to move from index-counts to valid estimates of abundance. we documented environmental covariates as an adjunct to simple counts from annual helicopter-based surveys in 2002–2012, and examined the performance of existing moose sightability models on these data. while acknowledging our inability to compare modeled estimates with actual abundance, we reasoned that if existing models converged on similar results, this would suggest that moose sightability is a sufficiently general phenomenon that the cost of developing a specific local model might not be justified. however, despite using similar covariates, the sightability models applied to our data produced widely disparate abundances and estimates with poor precision. specifically, where coniferous forest cover renders expected detection probability low, sightability models tend to behave erratically. we also used covariate data bearing on sampling variation to refine our estimate of population trend. multiple regression analyses revised the linear rate of increase associated with the raw counts of the instantaneous rate of growth, r = 0.084 (se = 0.019) to an adjusted estimate of r = 0.077 (se = 0.075). while incapable of transforming an index into a population estimate, accounting for variables likely to affect raw counts may be useful to refine estimates of trend. the use of an approach that avoids the autocorrelation inherent in a simple regression of counts on time better reflects true uncertainty. alces vol. 51: 57–69 (2015) key words: aerial survey, alces alces, moose, regression, sampling variation, sightability models moose (alces alces) are generally considered to have colonized the northeastern portion of the u.s. state of washington in the 1920s, but did not become well established until the 1970s (base et al. 2006). the population evidently increased in the latter part of the 20th century, with limitedentry hunting initiated by the then washington department of game in 1977, and increasing to approximately 130 permits drawn annually in 2012 (wdfw 2013). evidence available from hunters suggests that moose have increased since 2001, at least within areas open to hunting. the mean annual number of moose observed/day/hunter (as documented via a mandatory, webbased reporting system) increased from 2001 to 2012 (linear regression of raw counts on time: β = 0.086, se = 0.034, n = 11, t = 2.57, p = 0.030; marginal decline in days required/successful hunt: β = −0.295, se = 0.164, n = 11, t = −1.8, p = 0.102) at the same time that hunter success rate (average = 93%) increased (β = 0.009, se = 0.001, 57 n = 11, t = 4.27, p = 0.002). however, with the decline of many moose populations in adjacent jurisdictions due to forest maturation, increases in parasites, increases in predators, and the effects of climate change, the status of washington's moose population has elicited increased concern among the public. the problem of estimating abundance and trends of moose populations has vexed biologists and managers (e.g., gasaway et al. 1985), just as surely as the public at large – and hunters in particular – have expressed the expectation that such figures be available. as an often solitary, generally forestdwelling and invariable shy animal that eschews large aggregations, moose share with white-tailed deer (odocoileus virginianus) characteristics that make it among the more difficult of north american ungulates to survey. due to the logistical challenges of estimating abundance over large areas, surveys from fixedor rotary-wing aircraft have become the staple among north american wildlife management agencies (timmermann 1993, although see rönnegård et al. 2008, månsson et al. 2011, and boyce et al. 2012 for alternatives to aerial survey). but, it has long been recognized that even raw counts of animals from aerial surveys are often insufficient to estimate either abundance or trends. among approaches used to move from raw index counts of moose to population estimates are double-sampling (gasaway et al. 1986), conventional distance sampling (dalton 1990), sightability models (i.e., logistic regression based upon detectability of marked animals; anderson and lindzey 1996, drummer and aho 1998, quayle et al. 2001, guidice et al. 2012), application of infra-red thermal imagery to doubly-sampled units (bontaites et al. 2000), mark-recapture distance sampling (nielson et al. 2006), and independent double-observer surveys (cumberland 2012). since 2002, the washington department of fish and wildlife (wdfw) has conducted standardized moose surveys from helicopters in both the colville and spokane districts to produce indices of population abundance. although referenced when devising hunting seasons and harvest limits, these surveys have not been used to estimate abundance. rather, management has been based on informal evaluation of these surveys in combination with hunting statistics, implicitly assuming that these indices track population abundance. over the years, the number and distribution of survey units flown has varied, as have biological and environmental attributes (documented elsewhere) that influence detection probability. in short, understanding the various environmental factors that interfere with a simple equating of animals observed to animals present has become indispensable to our understanding of the count data. sightability models (samuel and pollock 1981) have been used for a variety of hunted ungulate species in western north america (e.g., elk [cervus elaphus; samuel et al. 1987, gilbert and moeller 2008] and mule deer [odocoileus hemionus; ackerman 1988]), and have been the focus of considerable efforts by wdfw (mccorquodale 2001, rice et al. 2009). biologists in the wdfw spokane district have gathered physical attribute data commonly assumed to affect detection of moose. although not collected for application to a specific sightability model, these covariate data allowed us to apply existing sightability models retrospectively. a number of situation-specific sightability models for moose have been developed, including those in wyoming (anderson and lindzey 1996), michigan (drummer and aho 1998), british columbia (quayle et al. 2001), alberta (peters 2010), interior alaska (christ 2011), minnesota (guidice et al. 2012), and coastal alaska (oehlers et al. 58 estimating moose abundance – harris et al. alces vol. 51, 2015 2012). although differing in details, these models are all notable in their common finding that vegetative cover (typically, coniferous forest) was the most important, and in some cases, the only covariate affecting detection probability of moose groups. other putative variables bearing on detection probability (e.g., snow cover, group size, weather conditions, individual observers) were generally unimportant. thus, we were motivated by the following notion: given the similarity of covariates shown to be predictive of detection in sightability models, might it be the case that “moose detectability” is a general-enough phenomenon that existing models can be applied in northeastern washington to estimate abundance, obviating the need to develop a local model? we did not attempt to validate or recalibrate any one model, but reasoned that a first approximation to answering the question of generalizability would result from comparing the performance of alternative models on an identical data set. if they generated similar results, this would suggest that the probability of moose detection is a generalizable phenomenon. if results diverged widely, it would suggest that moose detection is situationspecific, and that a novel sightability model would be required for site-specific and survey-specific data. additionally, if our analysis suggested that adopting an existing sightability model in eastern washington was unwarranted, we wondered if covariate data could refine our estimates of population trend. methods aerial winter surveys (decemberfebruary) of moose using a helicopter have been conducted annually by wdfw staff in the spokane district since 2002. these surveys were not designed to generate population estimates, but rather were considered as index counts that correlated positively with true abundance. we identified 51 survey blocks based on field landmarks; average block size was 13.5 km2 ranging from 9.0–17.8 km2. prior to each annual survey, each block was categorized into 1 of 3 population density strata (low, medium, high) based on the previous years’ survey, or if lacking, general field knowledge. the annual selection of blocks followed a stratified random design: all high density blocks were surveyed each year, whereas a random selection of medium and low density blocks was flown, depending on available funding. survey coverage (i.e., proportion of all mapped survey blocks included within that year's survey) averaged 33% ranging from 18–44%. flight lines within blocks were not mapped prior to the survey, nor were they strictly controlled. rather, flight paths were designed to maximize coverage within each block, reflecting the shape and topographic features within, and were generally ∼400 m apart. flight lines and locations of each moose group were recorded using hand-held gps units. all surveys were conducted with a robinson r-44 helicopter; typically 2 experienced observers were used (front left, rear right), and observations by the helicopter pilot were also counted. surveys were timed to coincide with good weather occurring shortly after a snowfall to the extent practical, avoiding patchy snow cover, and occurred as early as 8 december and as late as 3 february. surveys generally occurred over 3–6 days each winter in response to weather conditions. although never applied formally in a model setting, covariates hypothesized to influence detection of moose groups were collected in the same manner and with the same definitions each year (samuel et al. 1987). in addition to group size (and sex/age composition), these were activity (bedded, standing, moving), percent snow cover, percent obstructing vegetative cover (visually estimated to alces vol. 51, 2015 harris et al. – estimating moose abundance 59 nearest 5%), and an index of terrain type (flat or hilly). to assess the behavior of existing sightability models when applied to these data, we programmed the sightability model package (fieberg 2012) in r 3.1.1 (r development core team 2011) to replicate the models developed in the 3 closest geographic regions to eastern washington: wyoming (anderson and lindzey 1996), british columbia (quayle et al. 2001), and minnesota (guidice et al. 2012). in addition, we obtained the parameters for an additional model produced for a different geographic region of british columbia, but not published at the time (j. quayle, british columbia ministry of environment, victoria, british columbia, pers. comm.). these 4 models, hereafter referred to as “wyoming”, “bc”, “bc-2”, and “minnesota”, defined and categorized vegetative cover slightly differently; therefore, we binned our continuous data into the categories needed for each of the 4 models. we summarized data using jmp v. 11.1 (sas institute, cary, north carolina, usa). we further explored our covariate information to determine if it could improve our estimate of the rate of change of the moose population, even if it failed to find application in existing sightability models. a simple regression of the natural logarithm of raw index counts on time would ignore the effects of environmental variation on detection completely, as well as the existence of temporal correlation in population indices, thereby under-estimating true process variance and creating a sense of false precision. thus, we adopted the approach suggested by dennis et al. (1991:120; see also morris and doak 2002:68 and mills 2007:109) in which natural logarithms of the ratios of successive raw counts are regressed on the intervals between surveys, forced through the origin. we added to the basic regression model a suite of covariates hypothesized to affect detection probability or reflect survey effort, and thereby influence population trend estimates. we took as covariates the percent forest cover information used in the sightability models above (in this case, using the mean annual percent cover for all observed moose, weighted by moose group size). we further added other covariate data collected during 2002–2012 that did not enter into the top sightability models. these were the weighted mean annual percent snow cover recorded at each moose observation, the weighted mean annual index of moose activity of each observed moose group (1 = sitting, 2 = standing, 3 = moving), and the number of survey units entering the survey in each year (table 1). each of these varied annually and was a plausible candidate as a covariate that affected our interpretation of index counts. because the response variable in each case was the ratio of the natural logarithm of raw counts in successive years, we used as independent variables the ratios of the natural logarithms of the putative explanatory variables in those same 2 years. thus, our models took the form: y ¼ a þ b1x1 þ . . . þ b4x4 þ e ð1þ where: y = ln (count (t+1) / count (t)), the index counts of moose counted in each year (t), α = the intrinsic, annual growth rate (because the interval between successive counts were all 1 year in our case), β = coefficients to be estimated from data for each covariate hypothesized to affect sightability, x1 = ln(percent forest cover (t+1) / percent forest cover (t)), x2 = ln(percent snow cover (t+1) / percent snow cover (t)), x3 = ln(activity index (t+1) / activity index (t)), x4 = ln(units surveyed (t+1) / units surveyed (t)), and e = error, assumed normally distributed with constant variance. 60 estimating moose abundance – harris et al. alces vol. 51, 2015 we then assessed the strength of evidence for each of the 16 possible additive models (all possible combinations, plus a null model with no covariates) using aicc. in testing for significant pairwise correlations (p = 0.05), we did not detect any evidence of collinearity in the above set of predictors. our best estimate of the rate of growth during the time period was the model averaged estimate, â. results the number of moose observed annually ranged from 81 (2002) to 185 (2012). in total, 810 moose groups were observed in the 11 years, with a mean group size of 1.69 (sd = 0.98, range = 1–10). snow cover was generally high, and percent vegetative cover at observation sites ranged from 0–100% (annual range = 23–52%) (table 1). as expected, the 4 models generated point estimates of more moose than observed as raw counts, both because of imperfect detection and incomplete sample coverage. however, the 4 abundance estimates using identical data sets varied considerably within each year (fig. 1). point estimates produced by the wyoming model averaged 5.6x higher (range = 3.2–6.9) than those produced by the minnesota model. point estimates produced by the 2 bc models produced similar results (‘bc-2’ model not shown for clarity), and were generally closer to the wyoming than minnesota model. there was considerable annual fluctuation in abundance estimated by these models; in some cases, annual increases far exceeded the biological capability of even the most productive moose population (e.g., more than doubling between 2006 and 2007 in all 4 models). in addition, most abundance estimates had wide confidence intervals, especially the wyoming and bc models. the annual confidence intervals expressed as a % of the point estimates averaged 157% for the minnesota model, 356% for the wyoming model, and 368% for the bc model. the proportion of total variance due to the table 1. basic data used in application of ancillary data to refine trend estimates of moose abundance, spokane district in northeastern washington, winters 2002–2012. shown are number of moose seen during annual helicopter flights; number of survey units flown each year; the mean activity index of observed moose (weighted by group size) where 1 = bedded, 2 = standing, 3 = moving; weighted mean percent snow cover near observed moose, and weighted mean percent vegetation cover near observed moose. year raw count (moose observed) number of units surveyed mean moose activity index mean percent snow cover mean percent vegetation cover 2002 81 12 1.91 78.56 51.57 2003 59 17 1.69 84.07 35.85 2004 114 16 1.71 99.84 23.46 2005 74 9 1.54 76.42 50.88 2006 94 18 1.45 83.24 28.46 2007 112 13 1.42 100.00 49.46 2008 116 20 1.41 100.00 42.33 2009 124 20 1.45 95.56 48.02 2010 168 20 1.64 99.58 51.01 2011 117 20 1.55 70.09 45.81 2012 185 22 1.49 96.73 44.41 alces vol. 51, 2015 harris et al. – estimating moose abundance 61 model itself was much lower in the minnesota model than the other 3 models (table 2), whereas the proportion associated with incomplete sampling was much higher. examined over the 11-year time period, even the coarse population trends implied by application of the 4 sightability models were inconsistent. why did the wyoming and bc models project so many more moose than the minnesota model, given that they used a similarly defined covariate and an identical data set? graphical illustration of the core relationships underlying the 3 models (fig. 2) revealed the influence that a seemingly minor difference in the regression coefficient associated with detection probability relative to vegetative cover translated upon the estimates. when visual obstruction (forest canopy cover) is ∼30%, the models behave similarly; however, when visual obstruction approaches ≥50% the model estimates diverge (fig. 3). for example, figure 3 reconfigures the sightability curves in terms 0 1,000 2,000 3,000 4,000 5,000 6,000 7,000 8,000 9,000 10,000 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 po in t es ti m at e survey winter fig. 1. population trends of moose in northeastern washington (2002–2012) based on application of 3 moose sightability models. shown are point estimates produced by identical data sets each year by each model. purple triangles and dashed lines = minnesota model; green diamonds and dashed lines = bc model; blue squares and dashed line = wyoming model. table 2. point estimates, upper and lower 95% confidence bounds, variance components (sampling, sightability, model) from application of 4 moose sightability models to observation data from helicopterbased moose surveys, spokane district, northeastern washington, winters 2002–2012. model wyoming bc bc-2 minnesota point estimate 2,488 1,323 1,866 558 lower 95% 985 574 746 353 upper 95% 9,479 4,597 7,283 1,049 sampling variance 0.07 0.12 0.07 0.64 sightability variance 0.10 0.11 0.12 0.16 model variance 0.83 0.77 0.81 0.20 62 estimating moose abundance – harris et al. alces vol. 51, 2015 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 10 20 30 40 50 60 70 80 90 100 d et ec ti on p ro ba bi lit y visual obstruc�on fig. 2. smoothed representations of functions relating the probability of detecting a moose group given that it is present (vertical axis) to the percent vegetation capable of obstructing observation from a helicopter (horizontal axis). the lines were generated using sightability models developed in wyoming (anderson and lindzey 1996; dashed blue line), british columbia (qualye et al. 2001; solid red line), and minnesota (guidice et al. 2012; dot-dash green line). 0 2 4 6 8 10 12 14 16 18 20 0 10 20 30 40 50 60 70 m ul ti pl ie r visual cover a b 0 50 100 150 200 250 300 0 20 40 60 80 100 m ul ti pl ie r visual cover fig. 3. smoothed representations of mean expansion factors applied to each moose group observed based on the percent vegetation capable of obstructing observation from a helicopter (horizontal axis); shown are a) visual cover up to 70%, and b) visual cover up to 100%. the lines were generated using sightability models developed in wyoming (anderson and lindzey 1996; dashed blue line), british columbia (qualye et al. 2001; solid red line), and minnesota (guidice et al. 2012; dot-dash green line). alces vol. 51, 2015 harris et al. – estimating moose abundance 63 of the ‘expansion factor’ or ‘multiplier’ applied to each moose from each of the models. at 50% cover, the minnesota model projects ∼2 moose for each observed, whereas the wyoming model projects ∼4 (fig. 3a). with visual cover >60%, differences in multipliers applied to individual observations increasingly diverge (fig. 3b); for example, in the wyoming model at 80% cover, a single moose projects to about 50 and a cowcalf pair to 100 moose. figure 3b illustrates that under dense canopy, minor fluctuations in how field investigators code this covariate can produce substantial differences in detection probability, and ultimately the estimated abundance. regression of the natural logarithm of raw counts on time yielded a naïve estimate of the intrinsic growth rate (r) of 0.084 (se = 0.019), suggesting an average annual discrete growth rate (λ) of ∼1.09. the topranked regression model incorporating covariates and accounting for autocorrelation contained mean percent snow cover only (table 3). the model including the mean activity index and snow cover had similar support (δ aicc = 1.12), and together, these 2 models absorbed most (86%) of the akaike weight (table 3). the model averaged slope, accounting for all possible models and representing the estimate instantaneous growth rate r was 0.077 (approximate se = 0.075), table 3. competing models of the effects of annual covariates hypothesized to influence detection of moose observed during helicopter surveys, spokane district in northeastern washington, winters 2002–2012. all models are of the form shown in equation 1. shown are the point estimate of the intrinsic growth rate r, its standard error (se), the number of parameters in the model (k), δaicc, and the aicc weight. variables: snow = weighted mean percent snow cover near observed moose; activity = mean activity index of observed moose weighted by group size where 1 = bedded, 2 = standing, 3 = moving; units = number of survey units flown each year; and vegetation = weighted mean percent vegetation cover near observed moose. model predictors r se k δaicc aicc weight 1 intercept + snow 0.053 0.074 3 0.000 0.547 2 intercept + snow + activity 0.112 0.063 4 1.117 0.313 3 intercept only 0.084 0.114 2 4.489 0.058 4 intercept + activity 0.158 0.105 3 5.753 0.031 5 intercept + snow + vegetation 0.054 0.078 4 6.909 0.017 6 intercept + snow + units 0.055 0.080 4 7.291 0.014 7 intercept + vegetation 0.078 0.114 3 8.648 0.007 8 intercept + units 0.072 0.121 3 9.621 0.004 9 intercept + snow + vegetation + activity 0.113 0.064 5 10.370 0.003 10 intercept + activity + vegetation 0.154 0.101 4 11.019 0.002 11 intercept + snow + units + activity 0.112 0.068 5 11.654 0.002 12 intercept + units + activity 0.147 0.110 4 12.419 0.001 13 intercept + snow + units + vegetation 0.078 0.077 5 15.151 0.000 14 intercept + units + vegetation 0.010 0.122 4 15.365 0.000 15 intercept + units + vegetation + activity 0.169 0.109 5 20.918 0.000 16 intercept + snow + units + vegetation + activity 0.131 0.060 6 23.599 0.000 mean 0.077 0.075 64 estimating moose abundance – harris et al. alces vol. 51, 2015 slightly lower than the 0.084 with raw counts unadjusted for covariates and autocorrelation. that is, our best estimate of population trend (λ) that incorporated autocorrelation, our suite of visibility covariates, and model uncertainty was ∼1.08 during the 2002–2012 time period. however, none of the top-ranking regression models, nor the modeled averaged estimate (table 3) provided evidence that would reject the conventional null hypothesis that r = 0 at the customary type i error rate of α = 0.05. our data seemed to suggest a constant rate of growth throughout the time period. however, various alternative shapes may enjoy greater support than a linear trend. for example, the wyoming or bc models could arguably support a concave down function rather than a constant growth rate with high variance. to examine this alternative, we modeled a simple quadratic regression, using year and year2 as predictors in addition to the top covariates, allowing for a curving of the previously straight line (fig. 4). as suggested by harris et al. (2007), we assessed the strength of evidence for these 2 competing models using aicc, the aicc weight, and the significance of the quadratic term. models including the quadratic term were invariably less parsimonious than simple linear models, and quadratic terms were not significant (table 4). thus, fig. 4. trend of moose over time in northeastern washington (2002–2012) illustrating the effects of accounting for sightability covariates in a regression context, and of using a quadratic term. blue diamond symbols = natural logarithm of raw index counts of moose, 2002–2012; red square symbols = index counts predicted by top-ranked model accounting for mean annual moose activity and mean annual snow cover; bold solid red line = linear prediction from best model; bold dashed red line = quadratic prediction from best model; aic supports the linear over the quadratic. approximate 95% confidence limits surrounding quadratic prediction are shown in light dashed lines. alces vol. 51, 2015 harris et al. – estimating moose abundance 65 although the quadratic model will always be disadvantaged when compared with the linear model (by virtue of having an extra parameter), these results suggest that through 2012, counts provided no evidence of any moderation in the population growth rate. discussion most authors describing the development of sightability models caution about extrapolating coefficients beyond the conditions under which they were developed, and we concur. although all 4 sightability models used very similar covariates, subtle differences in their coefficients led to dramatically disparate estimates when identical data sets were applied to them. although sightability modeling is a well-explored and valid approach to estimating detection probability, it is vulnerable to extrapolation beyond site-specific conditions. we have concluded that should we wish to employ a valid sightability model for moose in northeastern washington, we have little choice but to develop one de novo using radiocollared animals. even then, the relatively dense conifer cover that characterizes most moose habitat in northeastern washington may, at best, yield a sightability model sensitive to errors in assigning covariate scores and have low precision. when we accounted for detectionrelated covariates and autocorrelation of counts, our best estimate of the rate of population growth was lowered slightly, but this also clarified that simple linear regression provided a misleading assessment of precision. although available evidence from both approaches suggested a positive trend, the addition of plausible sightability covariates (dennis et al. 1991) showed that data were not yet sufficiently precise or abundant to rule out an unchanging (or even negative) trend with time. the fact that the standard error of r (ln λ) exceeded its point estimate under the model lacking any sightability covariates (model 3, table 3) suggests that most of the difference in analyses came from accounting for autocorrelation rather than adding covariates. the estimates of r (i.e., ln λ) were identical (0.084), but the se in the dennis et al. (1991) regression approach (0.115) was much higher than the 0.019 returned by the simple linear regression model. however, by examining the suite of table 4. traditional models regressing ln(counts) on time, using the top ranking suite of covariates from table 3, comparing the fits of linear and quadratic relationships with time. support for the quadratic over the linear model would suggest that r, the intrinsic rate of growth, increased or decreased during the period. predictors time year se t p year2 se t p aicc aicc weight intercept + snow linear 0.075 0.015 5.05 0.001 0.722 0.902 quadratic 0.073 0.014 5.25 <0.001 0.007 0.013 1.45 0.190 5.164 0.098 intercept + snow + activity linear 0.095 0.014 6.72 0.002 1.148 0.995 quadratic 0.099 0.020 4.98 −0.003 0.007 −0.36 0.7289 11.909 0.005 intercept only linear 0.084 0.018 4.50 0.001 2.661 0.928 quadratic 0.084 0.020 4.26 0.003 0.002 0.007 0.30 0.775 7.780 0.072 intercept + activity linear 0.099 0.023 4.22 0.003 6.541 0.958 quadratic 0.115 0.031 3.74 0.007 −0.010 0.012 −0.85 0.423 12.789 0.042 66 estimating moose abundance – harris et al. alces vol. 51, 2015 models incorporating the sightability covariates, we gained insight into the potential that annual variation of one or another was the true driver underlying the apparent trend. rather than choosing only a single “best” model, our model averaging embraced and accounted for uncertainty while making use of the less-informative covariates. that said, its ambition was modest; it attempted to correct a trend index rather than to estimate true abundance. it did not provide a basis for scaling our estimated population increase in real numbers of moose, and provided only a relative, not an absolute measure of detectability. without the latter, we remain unable to estimate moose abundance. one surprising finding in our use of ancillary data to refine our estimate of population trend was that models incorporating vegetation cover, invariably identified as the most important covariate in sightability models (anderson and lindzey 1996, drummer and aho 1998, quayle et al. 2001, guidice et al. 2012, oehlers et al. 2012), were not ranked highly (table 3). one possible reason is that in this analysis, unlike with sightability models, we were not assessing the influence of vegetation cover ability to prevent detection, but rather its relationship with animals already detected. our regression approach was limited to data from animals that were observed. also, our regression analysis necessarily used the means of all covariates assessed across all animals observed in each year, in contrast to their use in sightability models where they are assessed from observations of each animal group. our regression approach could have been biased by covariates that were not quantified and/or included in our models. first, our annual selection of survey blocks may have been subconsciously biased to increasingly favor those with higher moose density as we gained experience in survey techniques and increased our qualitative understanding of moose distribution. we find such a bias unlikely because, except for the ∼1/3 of blocks categorized as “high density” (always surveyed), medium and low density blocks were selected on the basis of a random algorithm. secondly, we might imagine that observer expertise increased with time, such that moose detection increased independently of environmental covariates; if so, our estimate of population trend would be biased high. lacking marked animals and/or double-observer “markrecapture” data, an assessment of this source of bias was not possible. thirdly, we cannot rule out the possibility that unknown covariates affected detection probability. if so, and if these exhibit a trend with time, the resultant trend estimate could be biased. we quantified, but did not include in regression models, the group size of observed moose. although often an important predictor of elk sightability, group size in moose rarely exceeds 3 (usually 1 or 2) and has never been identified as an important predictor of detection. we are hardly plowing new ground by reiterating that models are only as useful as the reliability with which their assumptions align with intended use. we provide no basis for doubting the usefulness and accuracy of sightability models as a whole, but interpret our exercise as a sensitivity analysis applied to a similar (not identical) situation in which these models were developed. in this context, we find the divergence in estimates compelling evidence that extrapolation beyond their intended use or without proper and tested re-calibration is unwarranted. that said, the time and effort to collect ancillary data likely to be relevant to detection probability may be worthwhile. in our case, we used ancillary data in a regression environment to provide additional assurance that population trends suggested by raw index counts were unlikely to have been solely alces vol. 51, 2015 harris et al. – estimating moose abundance 67 artifacts of varying environmental or sampling conditions. in so doing, we clarified that even with 11 years of data and ancillary data related to sightability, aerial surveys if interpreted in isolation, were not capable of removing uncertainty about the actual population trend during the decade-long study period. acknowledgements we thank s. knapp for providing the necessary programming for this analysis. safe flying was provided by inland helicopters, spokane, washington; in particular we thank pilots d. valenti and j. snyder. funding for this work was provided by proceeds of auction and raffle hunts administered by wdfw and by the inland northwest wildlife council. administrative support was provided by the state of washington and pittman-robertson funds. j. fieberg, c. anderson, and j. quayle provided additional information and insight into their sightability models. we thank d. base and s. hansen for suggestions and improvements to the manuscript. literature cited ackerman, b. r. 1988. visibility bias of mule deer aerial census procedures in southeast idaho. ph.d. dissertation, university of idaho, moscow, usa. anderson, c. r., and f. g. lindzey. 1996. moose sightability model developed for helicopter surveys. wildlife society bulletin 24: 247–259. base, d. l., s. zender, and d. martorello. 2006. history, status and hunter harvest of moose in washington state. alces 42: 111–114. bontaites, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69–75. boyce, m. s., p. w. j. baxter, and h. p. possingham. 2012. managing moose harvest by the seat of your pants. theoretical population biology 82: 340–347. christ, b. 2011. sightability correction for moose population surveys. alaska department of fish and game. final wildlife research report. adf&g/ dwc/wrr-f-2011-1. juneau, alaska, usa. cumberland, r. e. 2012. potvin doublecount aerial surveys in new brunswick: are results reliable for moose? alces 48: 67–77. dalton, w. j. 1990. moose density estimation with line transect survey. alces 26: 129–141. dennis, b., p. l. munholland, and j. m. scott. 1991. estimation of growth and extinction parameters for endangered species. ecological monographs 61: 115–143. drummer, t. d., and r. w. aho. 1998. a sightability model for moose in upper michigan. alces 34: 15–19. fieberg, j. 2012. estimating population abundance using sightability models: r sightability model package. journal of statistical software 51: 1–20, 34: 15–19. gasaway, w. c., s. d. dubois, and s. j. harbo. 1985. biases in aerial transect surveys for moose during may and june. journal of wildlife management 49: 777–784. ———, ———, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska 22. gilbert, b. a., and b. j. moeller. 2008. modeling elk sightability bias of aerial surveys during winter in the central cascades. northwest science 82: 222–228. guidice, j. h, j. r. fieberg, and m. r. lenarz. 2012. spending degrees of freedom in a poor economy: a case study of building a sightability model for moose in northeastern minnesota. the journal of wildlife management 76: 75–87. harris, r. b., g. c. white, c. c. schwartz, and m. a. haroldson. 2007. population 68 estimating moose abundance – harris et al. alces vol. 51, 2015 growth of yellowstone grizzly bears: uncertainty and future monitoring. ursus 18: 167–177. månsson, j., c. e. hauser, h. andrén, and h. p. possingham. 2011. survey method choice for wildlife management: the case of moose alces alces in sweden. wildlife biology 17: 176–190. mills, l. s. 2007. conservation of wildlife populations: demography, genetics, and management. blackwell publishing, malden, massachusetts, usa. mccorquodale, s. m. 2001. sex-specific bias in helicopter surveys of elk: sightability and dispersion effects. journal of wildlife management 65: 216–225. morris, w.f., and d. f. doak. 2002. quantitative conservation biology: theory and practice of population viability analysis. sinauer associations, sunderland, massachusetts, usa. nielson, r. m., l. l. mcdonald, and s. d. kovach. 2006. aerial line transect survey protocols and data analysis methods to monitor moose (alces alces) abundance as applied on the innoko national wildlife refuge, alaska. technical report prepared for us fish and wildlife service, mcgrath, alaska, usa. oehlers, s. a., r. t. bowyer, f. huettmann, d. k. person, and w. b. kessler. 2012. visibility of moose in a temperate rainforest. alces 48: 89–104. peters, w. e. b. 2010. resource selection and abundance estimation of moose: implications for caribou recovery in a human altered landscape. m.s. thesis, university of montana, missoula, montana, usa. quayle, j. f., a. g. machutchon, and d. j. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–54. r development core team. 2011. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. version 2.13.1, isbn 3-900051-07-0. http:// www.r-project.org/. rice, c. g., k. j. jenkins, and w. chang. 2009. a sightability model for mountain goats. journal of wildlife management 73: 468–478. rönnegård, l., h. sand, h. andrén, j. månsson, and å. pehrson. 2008. evaluation of four methods used to estimate population density of moose alces alces. wildlife biology 14: 358–371. samuel, m. d., e. o. garton, m. w. schlegel, and r. g. carson. 1987. visibility bias during aerial surveys of elk in northcentral idaho. journal of wildlife management 51: 622–630. ——— , and k. h. pollock. 1981. correction of visibility bias in aerial surveys where animals occur in groups. journal of wildlife management 45: 993–997. timmerman, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29: 35–46. washington department of fish and wildlife (wdfw). 2013. 2012 game status and trend report. wildlife program, washington department of fish and wildlife, olympia, washington, usa. alces vol. 51, 2015 harris et al. – estimating moose abundance 69 http://www.r-project.org/ http://www.r-project.org/ estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty methods results discussion acknowledgements literature cited alces20_223.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces15_iiipreface.pdf alces vol. 15, 1979 alces14_247.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces17_veditorialcommittee.pdf alces vol. 17, 1981 alces17_180.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces17_56.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces16_489.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces18_xdistmoosebio.pdf alces vol. 18, 1982 alces19_178.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces19_271.pdf alces vol. 19, 1983 alces16_51.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces15_187.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alcessupp1_177.pdf alces18_iiiwelcomingaddress.pdf alces vol. 18, 1982 alcessupp1_115.pdf f:\alces\supp2\pagema~1\rus15s. alces suppl. 2, 2002 kuznetsov – moose and forest problems in russia 65 moose and forest problems in russia german v. kuznetsov institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: this article presents an analysis of the moose–forest relationship in russia characterized by utilization of land by humans and its consequences for moose and the forest. it provides a general overview of the research approaches regarding russia’s damaged forests by moose. in the early 1950s, the moose population increased sharply, primarily due to enlargement of the cutover area and the ensuing increased forage resource. devastation to pine and oak are emphasized amid a backdrop of damage to silviculture that cost millions of rubles. other northern countries were undergoing similar destruction by moose to their forests. three main research approaches are distinguished: determination of the damage by moose to stands, estimation of the effects of moose on the structure of forest phytocenoses, and the effects of moose on the productivity of particular plant species and forest phytocenoses. this well–documented article correlates various moose population densities with specific effects on different ecosystems and emphasizes the fact that trophic activity of moose is one of several factors affecting the structure and succession of forest phytocenoses of various natural zones. alces supplement 2: 65-70 (2002) key words: biocenosis, ecology, forest, moose, phytocenoses, population density, productivity, succession, silviculture, trophic activity in the ussr, a moose and forest problem originated in the early 1950s, when the moose population increased sharply, primarily due to enlargement of the cut over area and, subsequently, the increased forage resource. in this situation the moose began to detrimentally affect young forest growth and regrowth opportunity, especially those of pine and oak. the damage inflicted by moose on silviculture was estimated at millions of rubles. a negative effect of the moose on the forest was also recorded in other countries, such as sweden and norway (yurgenson 1979, filonov 1983). analysis of the studies conducted in the ussr on the moose and forest problem shows 3 main approaches: (1) determination (%) of the degree of damage by moose to forest stands; (2) evaluation of the effect of moose on the structure of forest phytocenoses and succession processes; and (3) investigation of the effect of moose on the productivity of particular plant species and forest phytocenoses. these 3 main approaches to a certain extent reflect the history of the development of research into estimation of the functional role of moose in forest biocenoses. in most studies of the first approach, the damage estimate of moose on the forest was based on the criterion of the level of damage to the shoots of trees (%). however, on the basis of such data, the effect of moose on the productivity of forest phytocenoses and related components cannot be estimated. nevertheless, an examination of studies on this subject is valuable to address the large body of evidence already collected from these data. an inverse relationship between the area of pine plantations and the rate of their damage by moose was established; i.e., the less area of pine plantations on cutovers moose and forest problems in russia – kuznetsov alces suppl. 2, 2002 66 being overgrown, the greater they are affected by moose (dinesman 1961). it follows that plantations should have a great area to resist the trophic pressure of moose. these results were supported by further studies. in fact, a number of authors (kozlovsky 1960, kaletskaya and kudinov 1987) revealed that a ratio of 20–30 ha of forage grounds (forest plantations) per moose reduces the detrimental role of moose to a minimum and, conversely, a ratio of <10 ha per moose sharply increases the damage to plantations by moose. to some degree this index indicates the expected effect of moose on forest plantations. for instance, in the nizhny novgorod region, the area of young pine per moose is 112 ha, while in the novgorod region, it is only 4.3 ha (chervonnyi 1975). naturally, in the novgorod regions the effect of moose on forest plantations is more substantial. judging from the fact that dense plantations of pine are damaged by moose to a much lesser extent, borodin (1959b) proposed a biological approach to pine protection from moose through an increase in the density of its plantations. numerous authors (borodin 1959a, dinesman 1961, kheruvimov 1969, padaiga 1980, smirnov 1987) have demonstrated that intensive trophic activity of moose delays the growth of pine, oak, ash, aspen, birch, fir, and other species, which is reflected in the quantity of timber. subsequently, a number of researchers revealed that the intensity of forest damage by moose is a function of their population density. in f a c t , i n s m a l l f o r e s t m a s s i f s o f t h e tsentralno–chernozemny reserve, the trophic pressure of moose is fairly heavy (gusev 1988). it is understandable that the degree of moose effect on forest plantations largely depends on the absence, in some cases, of a clear cut relationship between the population density of moose and plantation productivity over large areas (kuznetsov 1980). a large body of evidence was obtained on the development of stands under the effect of moose in reserves of the ussr; prioksko–terrasny, oksky, darvinsky, etc. ( k a l e t s k a y a 1 9 5 9 , k o z l o v s k y 1 9 6 0 , timofeeva 1974, chervonnyi 1975, dunin 1975, zablotskaya and zablotskaya 2002). the effect of moose is one of a series of factors affecting the health of the stand. in mixed plantations, it is important that such factors as growth conditions, the ratio of different trees, and the distribution of available forage should be taken into account. necessary cutovers can promote the formation of high–quality timber. the recent discussion on the moose–forest problem in t h e j o u r n a l o k h o t a i o k h o t n i c h y e khozyaistvo (hunting and game management) demonstrated that there are many problems yet to be solved on rational use of moose and the forest (pavlinov 1983, dunin 1984, perovsky 1984). a second direction of research is that it is known that moose promote replacement of the main species of regrowth and underbrush and the specificity of the effect of moose is associated with the features of succession in a particular region. for instance, on valdai, as a result of the effect of moose, there occurred gradual replacement of pine by deciduous trees and spruce. since spruce forests form under natural succession, there are grounds to believe that the moose promoted acceleration of this process. under conditions of the southern taiga, moose under different population density exert a dissimilar effect on the development of forest stands. when moose numbers are low, spruce–birch stands develop, but when moose numbers are high, primary spruce stands recover. an excessive population of moose degrades stands (smirnov 1987, abaturov and smirnov 2002). in another region in tulskie zaseki, according to our data, moose damaging oak trees alces suppl. 2, 2002 kuznetsov – moose and forest problems in russia 67 promote acceleration of the growth of ash, in particular the linden and filbert. there, the moose acts as a factor promoting the development of shady linden forests. thus, the trophic activity of moose is one of the factors affecting the structure and succession of forest phytocenoses of various natural zones. another approach to the moose–forest problem is associated with the investigation of the role of moose in the productivity of particular species of arboreous and herbaceous plants and phytocenosis. the main data were obtained through comparison of the state and productivity of plants under the effect of moose and in isolation from them. under conditions of forest–steppe, moose affect the growth and development of broad–leaved forests and decrease their productivity (zlotin and khodashova 1974, gusev 1989). in tulskie zaseki, oak and ash respond differently to the effect of moose. under conditions of a constant effect of moose, the oak reduces its productivity even under a small level of removal (13%). conversely, the ash is more resistant to removal of its phytomass by moose, which appears to be due to the biological properties of this species; on average, the annual increment of the leaf phytomass exceeds tenfold the respective increment of shoot phytomass. the shoots of ash are protected by the mass of foliage, and, hence, the moose utilizes them to a less extent. in valdai under conditions of the southern taiga, the annual increment in the scots pine in isolation was twice as high as in the pines accessible to moose; the admissible removal of annual shoots under which productivity is not reduced is close to 50%. it should be noted that removed phytomass of the pine is used by the moose fairly effectively and forage remains about 5%. it should also be emphasized that the state and productivity of pines is a function of not only the effect of moose but also of a set of other factors. for instance, the productivity of pines in an elevated plot is twice as high as in a lowland bogged area, despite the fact that the trophic pressure of moose in the former case is tenfold higher (kuznetsov 1980). the effect of moose on spruce is determined to a great extent by penetration of insect pests through injured parts of the stalk and infection by timber rots, which results in disintegration of the spruce layer of the stand rather than by removal of regrowth (smirnov 1987). willow and mountain ash can sustain 70–80% of removal of annual increment over many years, the productivity being maintained at a relatively constant level. the mountain ash is more resistant to phytomass removal by moose than birch and willow; favorite forage of moose is still moose resistant (kuznetsov 1980, chernyavsky and dubinin 1989). a cycle of observations over the natural response of arboreous plants to removal of their increment by moose is supplemented by data on experimental removal of phytomass in the pine and oak, and aspen (smirnov 1987). it was demonstrated that the productivity of the pine depends not only on the amount of removal of the increment phytomass but also on the method by which it is removed; i.e., removal of individual shoots does not bring about as sharp a reduction in productivity as does removal of the same amount of phytomass from each shoot. the moose normally browse only a part, about one third, of the shoots; it chooses the most judicious method of using the increment. also, as shown by defoliation and cutting the shoots, aspen sensitivity to removal of shoots can serve as a good indicator of the state of forage resources of moose (smirnov 1987). in general, we can conclude that deciduous trees, especially those that have root shoots, sustain greater trophic pressure by moose than coniferous moose and forest problems in russia – kuznetsov alces suppl. 2, 2002 68 and those without shoots. how moose actively affect the productivity of forest phytocenosis is not fully understood; however, the importance of this line of research is beyond doubt. we have demonstrated in valdai that such characteristics as annual increment, the long– term reserve of phytomass of herbaceous vegetation and shrubs, and the annual increment of heath under the effect of moose exceed the respective parameters under conditions of isolation. the increase in productivity of the pine in an open area (in a “range”) appears to be compensated for by an increase in the total increment of herbaceous plants and shrubs at the level of phytocenosis. however, to get the total balance of phytocenosis productivity, a number of other parameters are needed such as increment of the stalk loss due to trampling, etc. but still there are grounds to believe that moose, affecting the value of annual increment in some individual plant species, can to a considerable degree reduce the productivity of entire phytocenoses (kuznetsov 1980). thus, the line of research in question makes it possible to estimate the activity of moose as a component of forest ecosystems. the ecosystem approach regards the activity of moose as the same damage inflicted on the forest, since primary production will be transformed into secondary and the very notion of “damage” does not exist in the biological sense. conversely, the silvicultural approach presupposes obtaining high–quality timber and, hence, the moose and forest problem becomes realistic. in this connection, it is not by chance that the efforts of both zoologists and silviculturalists are aimed at investigating methods of forest protection from moose. the most harmless and accessible are the biotechnical methods, such as supplemental feeding, planting of fast–growing shrubs, and shoot cutting, etc. however, when moose numbers are high, these methods yield no positive effect. presumably, under these conditions it would be expedient to use such a powerful method of moose population control as removal. but in this country, only 10–12% of the moose population is removed, which is obviously insufficient. it will be remembered that in sweden they removed 40–50% of the population (dezhkin 1983). of interest are the approaches to the investigation of the intrapopulational structure and its role in the regulation of moose numbers (baskin 1984). mechanical methods, the exclosure of commercially important arboreous species, can be the most effective, but they are associated with great economic expense because exclosures must be efficient to achieve the desired outcomes. the optimal methods for protection of forest from moose depend on particular conditions in specific regions, and their application should be integrated. in recent years chemical methods of forest protection from moose have been developed (martynov 1980). the application of repellents can in some regions be very promising, but, unfortunately, there are no data available on the genetic control of their application. thus, through the efforts of scientists, the set of methods of forest protection from moose increases, but their implementation is lagging behind. it can be hoped that increasing interest in the problems of conservation of the natural environment will stimulate the solutions for these practical problems in the field of moose ecology. references abaturov, b. d., and k. a. smirnov. 2002. effects of moose population density on development of forest stands in central european russia. alces supplement 2:1-5. baskin, l. m. 1984. moose. pages 45–73 in v. e. sokolov, editor. the wildlife of alces suppl. 2, 2002 kuznetsov – moose and forest problems in russia 69 the southern taiga. nauka, moscow, russia. (in russian). borodin, l. p. 1959a. on the problem of moose role in silviculture. proceedings of the institute of silviculture, ussr academy of sciences 3:102–110. (in russian). . 1959b. increase in the density of pine plantations as a biological method of pine protection from the moose. proceedings of the institute of silviculture, ussr academy of sciences 13:124–126. (in russian). chernyavsky, p. b., and v. n. dubinin. 1989. moose in northeastern siberia. nauka, moscow, russia. (in russian). chervonnyi, v. v. 1975. winter forage resources and feeding of moose in the european rsfsr. proceedings of the oksky state reserve 11:321–389. (in russian). dezhkin, v. v. 1983. hunting and game management of the world. lesnaya promyshlennost, moscow, russia. (in russian). dinesman, l. g. 1961. the effect of wild mammals on the development of forest stands. ussr academy of sciences, moscow, russia. (in russian). dunin, v. f. 1975. the effect of moose on forest phytocenoses of byelorussian lakearea. pages 167–168 in ungulates of the ussr fauna. nauka, moscow, russia. (in russian). . 1984. both forest and moose. okhota i okhotnichye khozyaistvo 2:8– 9. (in russian). filonov, k. p. 1983. moose. lesnaya promyshlennost, moscow, russia. (in russian). gusev, a. a. 1988. admissible density of the population of wild ungulates and experience of its maintenance in the tsentralno–chernozemny reserve. pages 114–128 in population studies of animals in reserves. nauka, moscow, russia. (in russian). . 1989. animals in reserved territories. tsentralnoye chernozemnoye, moscow, russia. (in russian). kaletskaya, m. a., and a. a. kudinov. 1987. development of pine plantations of dense young stands damaged by moose. pages 189–215 in a. g. bannikov, editor. biology and harvest of moose. volume 3. moscow, russia. (in russian). kaletskaya, m. l. 1959. damage by moose to pine seedlings in the darwin animal preserve. proceedings of the institute of silviculture, russian academy of sciences 13:63–69. (in russian). kheruvimov, v. d. 1969. moose (comparative studies as exemplified by the tambov population). tsentralno– chernozyomnoye, moscow, russia. (in russian). kozlovsky, a. a. 1960. protection of f o r e s t s f r o m d a m a g e b y m o o s e . vnilm, moscow, russia. (in russian). kuznetsov, g. v. 1980. the role of ungulates in forest ecology (some insights and perspectives from research). pages 88–100 in herbivores in vegetative communities. nauka, moscow, russia. (in russian). martynov, e. n. 1980. the effect of chemical treatment of forest on the b i r d s a n d m a m m a l s . l e s n a y a promyshlennost, moscow, russia. (in russian). padaiga, v. i. 1980. estimate of the damage inflicted on silviculture by cervidae. m e t h o d o l o g i c a l r e c o m m e n d a t i o n . kaunas, lithuania. (in russian). pavlinov, n. 1983. moose through the eyes of a silviculturist. okhota i okhotnichye khozyaistvo 7:10–11. (in russian). perovsky, m. d. 1984. moose and forest: moose and forest problems in russia – kuznetsov alces suppl. 2, 2002 70 aspects of equlibrium. okhota i okhotnichye khozyaistvo 2:9–10. (in russian). smirnov, k. a. 1987. the role of moose in biocenoses of the southern taiga. nauka, moscow, russia. (in russian). timofeeva, e. k. 1974. moose: ecology, distribution, economic importance. leningrad state publishing house, leningrad, russia. (in russian). yurgenson, p. b. 1979. biological foundations of game management in forests. lesnaya promyshlennost, moscow, russia. (in russian). z a b l o t s k a y a , l . v . , a n d m . m . zablotskaya. 2002. anthropogenic effects on moose populations in the southern taiga. alces supplement 2:131135. zlotin, r. i., and k. s. khodashova. 1974. the role of animals in biological turnover of forest-steppe ecosystems. nauka, moscow, russia. (in russian). effects of variable fire severity on forage production and foraging behavior of moose in winter rachel lord1 and knut kielland1,2 1department of biology and wildlife, university of alaska, fairbanks, alaska 99775-7000, usa; 2institute of arctic biology, university of alaska, fairbanks, alaska 99775, usa abstract: the increasing frequency and extent of wildfires in alaska over the last half century has spurred increased interest in understanding the role of post-fire succession on vegetation establishment. our primary goal was to examine how wildfire affects production and distribution of winter forage for moose (alces alces) in interior alaska, and how these changes in forage availability control forage offtake. fire severity classification was based on post-fire depth of residual soil organic matter. we used a browse survey protocol to estimate the biomass of current year production (kg/ha) and overwinter offtake (kg/ha) by moose. under the assumption of homogenous effects of fire severity on regeneration, we estimated that moose consumed 36% of all forage (current annual growth) across the study area. however, we found that moose exhibited significantly higher browse consumption relative to browse production in high fire severity sites than in low severity sites (p < 0.05). when we adjusted our estimates of forage production and consumption by accounting for the significant differences in browse consumption between severity classes and their distribution across the burn, moose consumed approximately 49% of available forage. assessments of fire severity and its spatial distribution through remote sensing techniques and on-the-ground sampling provides improved projections of vegetation regeneration pathways following wildfires, and thus refined estimates of future browse production and habitat quality for moose. alces vol. 51: 23–34 (2015) key words: alaska, browsing, fire, foraging, functional response habitat, moose. fire is the primary disturbance in alaska’s boreal forest, burning on average more than one million hectares annually (dyrness et al. 1986). the post-fire landscape may be composed of a higher proportion of early successional stands, where successional pathways have led to deciduous species colonizing areas that were previously dominated by black (picea mariana) or white spruce (p. glauca). this mosaic of vegetation directly affects winter foraging and habitat use patterns of moose (alces alces). numerous studies have investigated the effects of fire on population dynamics, habitat, and foraging of herbivores (riggs and peek 1980, canon et al. 1987, kilpatrick and abendroth 2000) including moose (peek 1974, maccracken and viereck 1990, weixelman et al. 1998, maier et al. 2005). the purpose of this study was to assess the influence of fire severity, defined here as the amount of soil organic matter (som) remaining after the fire event, on the differential regeneration of plant species post-fire within the context of moose habitat. secondly, we examined these effects on moose forage consumption in winter within a burn. biotic and abiotic factors influence the spatial distribution of forest regeneration following wildfires (pastor et al. 1999, de groot et al. 2003, hellberg et al. 2003, wisdom et al. 2006), with fire severity playing an important role in post-fire secondary succession. fire events can increase diversity and density of plant species within the first 50 years after burning (kashian et al. 2005). 23 increased diversity in the vegetative community is due in part to differences in post-fire successional pathways within burn perimeters. severity is influenced by multiple interacting forces including the composition of the pre-fire vegetation community, weather patterns, fire behavior, and topographic variables (viereck et al. 1986, johnson 1992, schimmel and granstrom 1996, epting and verbyla 2005, johnstone and chapin 2006). post-fire vegetation establishment in the boreal forest generally follows 1 of 2 pathways: self-replacement or relay floristics (dyrness et al. 1986, landhäusser and wein 1993, johnstone and chapin 2006). in self-replacement succession, the same species within the pre-fire community re-establish after the disturbance, whereas relay floristics succession occurs in interior alaskan plant communities when the herbaceous (e.g., epilobium spp., oxytropis spp.) understory dominates immediately after fire, followed by shrub and deciduous tree establishment. deep soil organic horizons generally restrict germination of deciduous species in spruce-dominated boreal forests. selfreplacement by spruce is common during post-fire succession where fire intensity is low and a deep organic horizon remains (lebarron 1939, greene et al. 2004, johnstone and kasischke 2005, johnstone and chapin 2006). by contrast, relay floristics may take place where fire intensity is high and the organic layer is combusted to the extent that the mineral layer of the soil is exposed, allowing the germination of deciduous shrubs and trees (johnson 1992) such as willows (salix spp.), trembling aspen (populus tremuloides), and paper birch (betula neoalaskana). throughout winter moose are typically in a negative energy balance resulting in loss of body mass (schwartz et al. 1988). the main winter browse plants in interior alaska include twigs of several willow species, paper birch, and aspen. the abundance, availability, and quality of these browse species during winter represent, with predation, the primary limiting factors of moose populations in interior alaska (van ballenberghe and ballard 1998, boertje et al. 2007). several factors mediate moose use of burned areas including the generation of deciduous vegetation, pre-fire moose population densities and movement patterns, local predation rate, snow depths and movement corridors, and patches of unburned or lightly burned cover distributed among forage areas. peek (1974) found an increase in moose population density, specifically from increased immigration of yearlings, in the first 2 years following a large fire in northeastern minnesota. in contrast, gasaway et al. (1988) found no immigration into a 500 km2 burn in interior alaska 5 years post-burn, though moose in close proximity significantly increased their utilization of the burned areas during summer months and the pre-rut migration. immediately following the rosie creek fire near fairbanks, alaska in 1983, abundant regeneration of aspen, willow, and birch was present with active foraging in the area (maccracken and viereck 1990). the frequency of large fire years has increased since the 1950s in interior alaska’s boreal forest, and in the last 5 decades 33% of individual fires have burned >100,000 ha (kasischke et al. 2006). given the extent of land burned annually and increased forage production following fires, understanding the within-fire vegetation and herbivory dynamics coupled with a greater understanding of fire behavior and scope may gain managers important insight into future moose habitat in interior alaska. this study focused on forage production and use patterns by moose among different fire severities within a 1994 burn outside of delta junction, alaska. whereas studies of 24 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 captive moose have demonstrated a type-2 functional response to increased forage availability (e.g., renecker and hudson 1986), there is a knowledge gap pertaining to how spatial variation in forage production after disturbance affects herbivores in general (wisdom et al. 2006), and particular to moose, regarding foraging behavior and spatial organization. to examine the effect of variable fire severity on moose habitat, we hypothesized: 1) there would be more forage biomass produced in sites that were severely burned than in those which experienced lower severity burning, and 2) moose would preferentially use areas of high fire severity. study area the study area was in the flat tanana river valley which is within the tananakuskokwim lowlands ecoregion (kreig and reger 1982, jorgenson et al. 2001). we carried out field work within the hajdukovich creek burn, approximately 40 km se of delta junction, alaska (64.0° n, 145.4° w, hereafter denoted hc94, fig. 1). the fire burned from mid-june until september 1994 and consumed approximately 8900 ha (michalek et al. 2000). the pre-fire vegetation was dominated by stands of black spruce with a few aspen/mixed aspen-spruce stands throughout (michalek et al. 2000, johnstone and kasischke 2005). pre-fire soil organic layer depths in black spruce stands were estimated to be >25 cm (johnstone and kasischke 2005). the fire event was variable in its impacts on the black spruce forest; some areas experienced complete combustion of the organic layer while other areas had only small amounts of organic duff burned off (michalek et al. fig. 1. map of the 1994 hajdukovich creek burn in interior alaska. in the detail map (left), areas of high severity and low severity burning is indicated by dark and light shading, respectively. alces vol. 51, 2015 lord and kielland – fire severity and moose foraging 25 2000). fire severity classes were determined through post-fire satellite imagery and then field-checked for their correspondence with the depth of som remaining after the fire event (michalek et al. 2000). approximately 67% of the burn was classified as low and 33% as high severity (michalek et al. 2000). methods we sampled 17 sites among fire severity strata (11 high and 6 low) within the hc94 perimeter which had been established in a previous study of post-fire successional pathways (johnstone and kasischke 2005). sites were located using a handheld garmin etrex gps unit with an approximate accuracy of 5 m (garmin international, inc. olathe, ks, usa; coordinates in utm nad1983 zone 6). previous research had classified the sites as either high or low severity based on the amount of som remaining after the fire event (johnstone and kaschiske 2005, shenoy et al. 2011). the plots were distributed along the trail system within the burn scar and accessed by snow machine and snowshoes in late march 2007. three medium severity sites were also sampled, but were not statistically different from either high or low sites in any of the analyses, likely due to low replication; the results from these sites are not reported. plot biomass measurements a modified browse assessment protocol was used to estimate the biomass of forage production and removal (seaton 2002). at each site we established 30 m-diameter plots with the site gps coordinate as the center. random number tables were used to select distance and bearing to locate 3 plants within the accepted height range available to moose for winter browsing (0.5 – 3.0 m above ground) of each forage species (birch, aspen, and willows). willows were identified to species (i.e., salix scouleriana, s. bebbiana, s. glauca, and s. arbusculoides; simpson 1986, collet 2004) but were grouped as salix spp. in the analysis. for each plant we recorded 5 parameters: species, height, estimated number of current annual growth (cag) twigs, percent dead material by volume, and architecture class. plant architecture classes were defined by the percentage of the current growth (by volume) of the plant arising from any lateral branching that was due to moose browsing and were either unbrowsed (<5%), browsed (5–50%), or broomed (>50%). this classification provides a quick index for categorizing the browsing intensity on a plant throughout the course of its life (seaton 2002). the cag diameter was measured with dial calipers (nearest 0.1 mm) on 10 twigs (>1 cm long) per plant for a total of 30 twigs/forage species/plot. the diameter at point of browsing (dpb) was measured if the twig was browsed by moose. browsing by snowshoe hares was evident and we differentiated between their smooth-cut stems and the rough-edged browsing pattern of moose. if necessary, >3 plants were sampled if <30 twigs were available, until either 30 twigs or all available twigs in the plot were measured. stem densities were estimated within each plot using a 2 m × 30 m belt transect from a random starting point on the plot perimeter, running through the plot center. within this transect, stems of all forage species and non-forage tree species (picea spp.) above 0.5 m (typical snow depth by late winter) were counted. this sample density was then multiplied by the plot area (706.86 m2) to obtain an estimated stem count (density) within each plot. mass-diameter regressions twigs were collected to develop massdiameter regression equations for all forage species (table 1) except salix glauca of 26 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 which data from the delta junction area was provided by the alaska department of fish and game (t. paragi, adfg, unpublished data.). twigs were weighed immediately upon returning to the lab or kept frozen until subsequent weighing. they were clipped and weighed at each whole diameter interval, from 2–10 mm. samples of wet weight twigs from all diameter classes were then dried at 80 °c for 24–48 h. they were then reweighed to determine the percentage of dry mass by diameter class (lord 2008). the data were log transformed and a regression equation was fitted to relate dry mass to diameter (maccracken and van ballenberghe 1993, seaton 2002). we used software written in r language (r development core team 2008, version 2.1.1; code and instructions available under project 5.10 at ). dry mass calculations were then back transformed in r to obtain the original units of g of dry mass (paragi et al. 2008). biomass calculations biomass was calculated using the estimated dry mass from the mass-diameter regression equations. the formula used for estimating biomass production and removal was: bbk ¼ x j mjk mjk x i nijk nijk x h bzhijk ð1þ where bbk is the site estimate of removal or production biomass in grams. twigs are denoted by h, plants by i, species by j, and the sites by k. m and m are the total and sampled plants in each plot, n and n are the total and sampled twigs, and bz denotes individual twig biomass (seaton 2002). statistical analysis all statistical analysis was performed with sas software, version 8.0 (sas institute inc. 2002). linear regression was used to examine the influence of som depth on vegetation composition (proc reg). differences between cag and dpb diameters were tested using t-tests (proc ttest). one-way anovas were used to test for differences in vegetation composition and moose browse consumption between severity classes (proc glm). we used biomass as a habitat metric, and needed to estimate it across the study area. as is often done for management purposes, with our first method we took averages of production and removal across all of our study sites and extrapolated this across the entire study area. to assess the potential effects of variable fire severity within our study area on forage production and removal, we used a second method for extrapolating our data across hc94 in which we used production and removal biomass averages from the sampled fire severity strata. we used these averages and weighted them for the final study area estimate by the proportion of high (0.67) and low (0.33) severity areas within hc94. these 2 methods allowed us to compare the resulting biomass estimates for the study area, which differed only in whether or not table 1. regression coefficients to predict dry matter (g) from twig diameter (mm) of moose browse species in the hajdukovich creek burn near delta junction, alaska. species intercept slope mse n r2 betula neoalaskana 0.01 5.81 0.03 20 0.89 populus tremuloides 0.03 2.83 0.04 20 0.87 salix bebbiana 0.01 4.38 0.04 20 0.89 salix glauca 0.02 2.68 0.07 20 0.83 salix scouleriana 0.02 3.11 0.22 20 0.94 alces vol. 51, 2015 lord and kielland – fire severity and moose foraging 27 http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat http://www.adfg.alaska.gov/index.cfm?adfg=librarypublications.wildliferesearch#habitat they incorporated significant production and removal differences between fire severities. tukey’s adjustment for pairwise comparisons was used to test for differences among severity classes. values reported are means (± s.e.). all models were checked to ensure that they met basic assumptions of normality and homogeneity of variance. results the mean plant stem density of all principal forage species (aspen, willow, and birch) was greater (f2,17 = 7.44, p = 0.005) across all high severity (1.80 ± 0.57 stems/m2) than low severity sites (0.67 ± 0.34 stems/m2). there was a sharp decline in the number of these stems with increasing depth of som. in contrast, spruce stem density ranged from 0.02–0.15 stems/m2 with no difference between severity classes, and the mean number of spruce (non-forage) stems did not change with depth of som (fig. 2). past browsing resulted in 84% of forage plants exhibiting broomed architecture, with 13% classified as browsed and 3% unbrowsed. these proportions were similar across all severity classes as well as between forage species. an average of 190 ± 104 kg/ha of forage biomass was produced across all sites. however, high severity sites produced >3-fold more forage (225 ± 64 kg/ha) than sites of low fire severity (69 ± 48 kg/ha), and twig density was nearly 3-fold greater in high (35 twigs/m2) than low severity sites (13 twigs/m2). estimates of total biomass consumed/ha were larger (f2,17 = 8.92, p = 0.002) in high (104 ± 35 kg/ha) than low severity sites (17 ± 18 kg/ha). aspen and willow dominated the differences in consumption between fire severity classes. these species represented >95% of the forage consumed with greater (f2,17 = 7.34, p = 0.005) absolute biomass removal in high fire severity sites than low severity sites. offtake of forage relative to forage production was higher (f2,17 = 7.46, p = 0.005) in high (46 ± <1%) than low severity sites (19 ± <1%) (fig. 3). across all sites, an average of 36 ± <1% of cag was removed by moose over winter (table 2). when we used fire severity-specific estimates of production in conjunction with the remote sensing image of burn severity to account for the area covered by each severity class, the overall estimate of forage production was 128 kg/ha. by contrast, the production estimate generated in the absence of this correction (i.e., irrespective of differences in severity) was nearly 50% higher (190 kg/ha). discussion global climate change is predicted to increase the frequency and severity of large fig. 2. stem density of forage (aspen, birch, willow spp.) and non-forage (black spruce) plants between 0.5 – 3.0 m high, corresponding with residual som depth (cm) in interior alaska. the equation for the regression on forage data is: y = 2.13 * e−0.08 x. 28 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 wildfires across the boreal landscape (kasischke et al. 2010), creating the potential for increased moose habitat in the form of widely distributed deciduous stands. however, variable fire behavior creates a spectrum of fire severities, resulting in a range of depth in som that facilitates different successional pathways (shenoy et al. 2011). we found a sharp increase in aspen stem densities where the post-fire depth of som was <6 cm. a similar relationship between the som horizon and deciduous stem densities was found at the same sites by johnstone and kasischke (2005) who documented increased aspen density with increased fire severity. the hc94 fire occurred within game management unit (gmu) 20d which encompasses nearly 1.5 million ha of largely black spruce dominated boreal forest, as well as part of the alaska range alpine ecosystem. much of the boreal area in gmu 20d has burned since 1979, over two-thirds between 2001 and 2004 (blm 2005). the 2006 moose density estimate in southwestern 20d (2 moose/km2; dubois 2004) was among the highest in interior alaska, exceeding that on the tanana flats region in gmu 20a (1.1 moose/km2 in 2000; seaton 2002). in addition to predator control programs in the 1980s and low snow depths in gmu 20d (dubois 2004), it is possible that fires in this area have contributed to the increase in moose density by providing a substantial amount of high-quality winter browse. future management decisions will need to account for the distribution of forage resources throughout the gmu, which should include an assessment of how the fire regime and individual fire behavior have shaped and will continue to influence winter habitat. browsing by moose in the study area was high as demonstrated by the large proportion (80%) of plants with broomed architecture throughout, regardless of severity class. moreover, high forage offtake in interior alaska is associated with greater mortality of browse plants (butler and kielland 2008) and a significant shift in plant species composition towards less preferred species, such as alder and spruce (kielland et al. 1997, 2006). the increase in the amount of standing dead trees and evergreens suggest that browsing could accelerate fire return intervals due to increased flammability of fire severity 0.0 0.2 0.4 0.6 c on su m pt io n : p ro du ct io n ra tio low high fig. 3. consumption to production ratio of moose browse in the 1994 hajdukovich creek burn in sites of low (n = 6) and high (n = 11) fire severity in interior alaska; mean ± s.e. table 2. estimates of biomass production and removal given high and low fire severity across the entire hajdukovich creek burn in interior alaska. mean (s.e.) estimates represent all site data pooled. the severity-weighted estimates used the different estimates from high and low fire severity sites, coupled with burn-wide fire severity estimates derived from remote sensing, to weight the production and removal field estimates across the burn. severity production (kg/ha) consumption (kg/ha) c:p mean 190 (104) 69 (47) 36% severityweighted 128 64 49% alces vol. 51, 2015 lord and kielland – fire severity and moose foraging 29 the vegetation after browsing. intense browsing following wildfires may, in the short run, result in a significant numerical response by moose and high levels of consumption. however, these functional and numerical responses could cause plant mortality, reduced forage production, and ultimately have a negative feedback effect on the moose population in a relatively short period. examples of such unsustainable growth under high moose density are clearly reflected in demographic data including low twinning rates, delayed age of first reproduction, and low body mass of short-yearlings (boertje et al. 2007). in contrast, browsing at lower rates (when moose are at lower population density) may possibly extend the period of increased forage availability following high severity fires. we found that stand regeneration after high severity fires produced >3x forage biomass/ha than in lightly-burned areas 14 years post-fire, underscoring the importance of using spatial information to adjust estimates of the productivity of regenerating stands. moose responded to this heterogeneous environment by not only consuming more forage from high severity sites as predicted from the functional response, but by also consuming a much higher proportion of available browse. estimates with captive animals predict that moose have the capacity to double their winter consumption rate in the observed 5-fold range of browse availability (50 – 250 kg/ha) (renecker and hudson 1986). in light of the functional response based on these estimates, the approximately 3-fold increase in browse production from 69 kg/ha in low severity sites to 225 kg/ha in high severity sites had the potential to increase offtake rates ∼70% (from 17 to 29 kg/ha). however, we found that the actual offtake across this 3-fold increase in production resulted in a 6-fold increase in consumption or 104 kg/ha (fig. 4). this large discrepancy between predicted and observed offtake suggests that moose aggregated in areas of high browse availability and that changes in behavior at the population level (spatial organization) may supersede physiological constraints (consumption capacity) regarding browse offtake in the field. when evaluating possible indices useful to indicate density dependent nutritional limitation in moose populations, boertje et al. (2007) suggested that when offtake of browse is >30–35% of current production, a population may be nearing the carrying capacity of the habitat. indeed, landscape level offtake expressed as the ratio of consumption (kg/ha) to production (kg/ha) shows a strong inverse relationship with twinning rates even at browse removal rates of <20% (seaton et al. 2011). in our study, the average removal rate of browse across hc94 bracketed these offtake values, but we also provide evidence that there are hotspots of foraging over this large area that differ widely from the overall mean. fig. 4. observed (cobs) versus predicted (cpred) magnitudes of forage consumption by moose in relation to changes in forage production in sites exhibiting low and high fire severity in the 1994 hajdukovich creek burn in interior alaska. the stippled line connects estimated increase in forage offtake as a function of forage production predicted from the functional response (renecker and hudson 1986). 30 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 currently, black spruce represents the dominant tree species of the landscape in interior alaska (kasischke et al. 2010). wildfires in the interior boreal forest often cover 10,000s of ha and result in patchy regeneration of both spruce (through selfreplacement succession) and deciduous (through relay floristics) stands across the landscape (chapin et al. 2006). over 38 million ha are in post-fire secondary succession that resulted from fires ∼25–30 years ago (blm 2005). since the mid-1990s-2006, almost 75 million ha have burned in interior alaska, 63 million between 2001 and 2004 (blm 2005). the peak of post-fire succession, from the perspective of moose habitat, can last from 10-30 years following a fire event. these fires will directly impact moose habitat by facilitating an increasingly complex vegetation community across the landscape. coupling post-fire remote sensing data with field measurements of browse production and offtake should help managers better understand and predict the impacts of wildfire on moose habitat in boreal landscapes. acknowledgements the research was supported by the bonanza creek long-term ecological research program, funded jointly by nsf (deb-0423442) and usda forest service, pacific northwest research station (pnw01-jv11261952-231). we thank t. paragi, e. kasischke, and j. johnstone for their assistance. we appreciate help from k. spellman, c. williams, and d. vargas-kretsinger with field work, and c. brown for revision of the fire severity map. references boertje, r. d., k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antelerless harvests. journal of wildlife management 71: 1494–1506. doi: 10.2193/ 2006-159. blm (bureau of land managementalaska fire service). 2005. alaska fire history, 1950–2004. environmental resource institute of michigan. http:// agdc.usgs.gov/data/blm/fire/index.html . butler, l. g., and k. kielland. 2008. acceleration of vegetation turnover and element cycling by mammalian herbivory in riparian ecosystems. journal of ecology 96: 136–144. canon, s. k., p. j. urness, and n. v. debyle. 1987. habitat selection, foraging behavior, and dietary nutrition of elk in burned aspen forest. journal of range management 40: 433–438. doi: 10.2307/ 3899605. chapin iii, f. s., l. a. viereck, p. c. adams, k. van cleve, c. l. fastie, r. a. ott, d. mann, and j. f. johnstone. 2006. successional processes in the alaskan boreal forest. pages 100–120 in f. s. chapin iii, m. w. oswood, k. van cleve, l. a. viereck, and d. verbyla, editors. alaska’s changing boreal forest. oxford university press, new york, new york, usa. collet, d. m. 2004. willows of interior alaska. u.s. fish and wildlife service, yukon flats national wildlife refuge, fairbanks, alaska, usa. de groot, w. j., p. m. bothwell, d. h. carlsson, and k. a. logan. 2003. simulating the effects of future fire regimes on western canadian boreal forests. journal of vegetation science 14: 355–364. doi: 10.1111/j.1654-1103.2003.tb02161.x. dubois, s. d. 2004. unit 20d moose management report. alaska department of fish and game, juneau, alaska, usa. dyrness, c. t., l. a. viereck, and k. van cleve. 1986. fire in taiga communities of interior alaska. pages 74–86 in k. van cleve, f. s. chapin iii, p. a. flanagan, l. a. viereck, and alces vol. 51, 2015 lord and kielland – fire severity and moose foraging 31 http://agdc.usgs.gov/data/blm/fire/index.html http://agdc.usgs.gov/data/blm/fire/index.html c. t. dyrness, editors. forest ecosystems in the alaskan taiga: a synthesis of structure and function. springerverlag, new york, new york, usa. epting, j., and d. verbyla. 2005. landscapelevel interactions of prefire vegetation, burn severity, and postfire vegetation over a 16-year period in interior alaska. canadian journal of forest research 35: 1367–1377. doi: 10.1139/x05-060. gasaway, w. c., s. d. dubois, r. d. boertje, d. j. reed, and d. t. simpson. 1988. response of radio-collared moose to a large burn in central alaska. canadian journal of zoology 67: 325–329. doi: 10.1139/z89-047. greene, d. f., n. j. bergeron, m. roussau, and s. gauthier. 2004. the regeneration of picea mariana, pinus banksiana, and populus tremuloides along a fire severity gradient. canadian journal of forest research 34: 1845–1857. doi: 10.1139/ x04-059. hellberg, e., g. hornberg, l. ostlund, and o. zachrisson. 2003. vegetation dynamics and disturbance history in three deciduous forests in boreal sweden. journal of vegetation science 14: 267–276. doi: 10.1111/j.1654-1103. 2003.tb02152.x. johnson, e. a. 1992. fire and vegetation dynamics: studies from the north american boreal forest. cambridge university press, cambridge, uk. johnstone, j. f., and f. s. chapin iii. 2006. effects of soil burn severity on post-fire tree recruitment in boreal forest. ecosystems 9: 14–31. doi: 10.1007/s10021004-0042-x. ———, and e. s. kasischke. 2005. standlevel effects of soil burn severity on postfire regeneration in a recently burned black spruce forest. canadian journal of forest research 35: 2151–2163. doi: 10.1139/x05-087. jorgenson, m. t., c. h. racine, j. c. walters, and t. e. osterkamp. 2001. permafrost degradation and ecological changes associated with a warming climate in central alaska. climatic change 48: 551–579. doi: 10.1023/a:10056674 24292. kashian, d. m., m. g. turner, w. h. romme, and c. g. lorimer. 2005. variability and convergence in stand structural development on a fire-dominated subalpine landscape. ecology 86: 643–654. doi: 10.1890/03-0828. kasischke, e. s., t. s. rupp, and d. verbyla. 2006. fire trends in the alaskan boreal forest. pages 285–30 in f. s. chapin iii, m. w. oswood, k. van cleve, l. viereck, and d. verbyla, editors. alaska’s changing boreal forest. oxford university press, new york, new york, usa. ———, d. l. verbyla, t. s. rupp, a. d. mcguire, k. a. murphy, r. jandt, j. l. barnes, e. e. hoy, p. a. duffy, m. calef, and m. r. turetsky. 2010. alaska’s changing fire regime — implications for the vulnerability of its boreal forests. canadian journal of forest research 40: 1313–1324. doi: 10.1139/x10-098. kielland, k., j. p. bryant, and r. w. ruess. 1997. moose herbivory and carbon turnover of early successional stands in interior alaska. oikos 80: 25–30. doi: 10.2307/3546512. ———, ———, and ———. 2006. mammalian herbivory, ecosystem engineering, and ecological cascades in taiga forests. pages 211–226 in f. s. chapin iii, m. w. oswood, k. van cleve, l. viereck, and d. verbyla, editors. alaska’s changing boreal forest. oxford university press, new york, new york, usa. kilpatrick, s. a., and d. c. abendroth. 2000. aspen response to prescribed fire and wild ungulate herbivory. pages 387– 394 in w. d. shepperd, d. binkley, d. l. bartos, and t. j. stohlgren, editors. sustaining aspen in western landscapes: symposium proceedings rmrsp-18. department of agriculture, forest 32 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 service, rocky mountain research station, fort collins, colorado, usa. kreig, r. a., and r. d. reger. 1982. air photo analysis and summary of landform soil properties along the route of the trans-alaska pipeline system. geologic report 66. alaska department of natural resources, division of geological and geophysical surveys, juneau, alaska, usa. landhäusser, s. m., and r. w. wein. 1993. postfire vegetation recovery and tree establishment at the arctic treeline: climacticchange-vegetation-response hypothesis. journal of ecology 81: 665–672. lebarron, r. k. 1939. the role of forest fires in the reproduction of black spruce. proceedings of the minnesota academy of science 7: 11–14. lord, r. e. 2008. variable fire severity in alaska’s boreal forest: implications for forage production and moose utilization patterns. m. s. thesis, university of alaska-fairbanks, fairbanks, alaska, usa. maccracken, j. g., and v. van ballenberghe. 1993. mass-diameter regressions for moose browse on the copper river delta, alaska. journal of range management 46: 302–308. doi: 10.2307/ 4002462. ———, and l. a. viereck. 1990. browse regrowth and use by moose after fire in interior alaska. northwest science 64: 11–18. maier, j. a. k., j. m. ver hoef, a. d. mcguire, r. t. bowyer, l. saperstein, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics: effects of scale. canadian journal of forest research 35: 2233– 2243. doi: 10.1139/x05-123. michalek, j. l., n. h. f. french, e. s. kasischke, r. d. johnson, and j. e. colwell. 2000. using landsat tm data to estimate carbon release from burned biomass in an alaskan spruce forest complex. international journal of remote sensing 21: 323–338. doi: 10.1080/0143 11600210858. paragi, t. f., c. t. seaton, and k. a. kellie. 2008. identifying and evaluating techniques for wildlife habitat management in interior alaska: moose range assessment. federal aid wildlife restoration, project 5.10. final research technical report, grants w-33-4 through w-33-7. alaska department of fish and game, juneau, alaska, usa. pastor, j., y. cohen, and r. moen. 1999. generation of spatial patterns in boreal forest landscapes. ecosystems 2: 439– 450. doi: 10.1007/s100219900092. peek, j. m. 1974. initial response of moose to a forest fire in northeastern minnesota. american midland naturalist 91: 435– 438. doi: 10.2307/2424334. renecker, l. a., and r. j. hudson. 1986. seasonal foraging rates of free-ranging moose. journal of wildlife management 50: 143–147. doi: 10.2307/3801504. riggs, r. a., and j. m. peek. 1980. mountain sheep habitat-use patterns related to post-fire succession. journal of wildlife management 44: 933–938. doi: 10.2307/ 3808329. schimmel, j., and a. granstrom. 1996. fire severity and vegetation response in the boreal swedish forest. ecology 77: 1436–1450. doi: 10.2307/2265541. schwartz, c. c., m. e. hubbert, and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. journal of wildlife management 52: 26–33. doi: 10.2307/3801052. seaton, c. t. 2002. winter foraging ecology of moose in the tanana flats and alaska range foothills. m.s. thesis, university of alaska-fairbanks, fairbanks, alaska, usa. ———, t. f. paragi, r. d. boertje, k. kielland, s. dubois, and c. l. fleener. 2011. browse biomass removal and nutritional condition of moose, alces alces. wildlife biology 17: 55–66. doi: 10.2981/10-010. alces vol. 51, 2015 lord and kielland – fire severity and moose foraging 33 shenoy, a., j. johnstone, e. kasischke, and k. kielland. 2011. persistent effects of fire severity on early successional forests in interior alaska. forest ecology and management 261: 381–390. doi: 10.1016/ j.foreco.2010.10.021. simpson, d. t. 1986. key for the identification of salix in interior alaska. unpublished manual, 9 pp. van ballenberghe, v., and w. b. ballard. 1998. population dynamics. pages 223–245 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, new york, new york, usa. viereck, l. a., k. van cleve, and c. t. dyrness. 1986. forest ecosystem distribution in the taiga environment. pages 22–43 in k. van cleve, f. s. chapin iii, p. a. flanagan, l. a. viereck, and c. t. dyrness, editors. forest ecosystems in the alaskan taiga: a synthesis of structure and function. springerverlag, new york, new york, usa. weixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces 34: 213–238. wisdom, m. j., m. vavra, j. m. boyd, m. a. hemstrom, a. a. ager, and b. k. johnson. 2006. understanding ungulate herbivory-episodic disturbance effects on vegetation dynamics: knowledge gaps and management needs. wildlife society bulletin 34: 283–292. doi: 10.2193/00917648(2006)34[283:uuhdeo]2.0.co;2. 34 fire severity and moose foraging – lord and kielland alces vol. 51, 2015 effects of variable fire severity on forage production and foraging behavior of moose in winter study area methods plot biomass measurements mass-iameter regressions biomass calculations statistical analysis results discussion acknowledgements references alces20_107.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces18_197.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 181 a methodological comparison among dna source types for moose genotyping tessa l. unger1,2, ron a. moen1,2,3, and jared l. strasburg1,2 1department of biology, university of minnesota-duluth, 1035 kirby drive, duluth, minnesota, usa 55812; 2integrated biosciences program, university of minnesota-duluth, 1035 kirby drive, duluth, minnesota, usa 55812; 3natural resources research institute, 5013 miller trunk hwy, duluth, minnesota, usa 55811 abstract: population genetic analyses for moose have been based on dna extracted from blood and other body tissues. non-invasive sampling of fecal pellets is another potential source of dna. we compared dna extraction from blood, liver tissue, and fecal pellet samples from moose in minnesota and yellowstone national park, usa. extracted dna from all source types was sufficient for genotyping using 15 microsatellites. dna extracted from fecal pellets was of lower quality and quantity than dna extracted from blood and tissue. we provide comparisons of efficiency and effectiveness of dna extraction protocols for blood, tissue, and fecal pellets, and demonstrate the suitability of using dna extracted from non-invasively sampled material in moose. alces vol. 53: 181–197 (2017) key words: dna extraction, fecal pellets, genotyping, microsatellites, moose, polymerase chain reaction, population genetics an important advance in population and conservation genetics has been the ability to obtain dna from multiple biological source types (waits et al. 2005). blood and muscle are the dna source types used most often in population genetic research with ungulates. other tissue samples such as liver that are collected for other purposes can also be used for genetic analysis. samples obtained from hunter-harvested moose or moose killed in vehicular collisions are likely to be available in unlimited quantities on a relative basis. in contrast, samples may be limited if specimens are collected during moose capture operations, although samples collected during such operations have been used as dna sources in several population genetic studies of cervids (finnegan et al. 1999, coulon et al. 2004, kangas et al. 2013, wilson et al. 2015). more recently, fecal pellets collected non-invasively have been used as an alternative source of dna. these samples can be collected from multiple individuals over a broad geographic region and are relatively easy to collect, particularly if fieldwork is conducted in winter when fresh fecal pellets are visible on snow and cold temperatures limit degradation of dna. fecal pellets would be an ideal dna source in parks or other areas where non-invasive sampling would be preferred or required. non-invasive dna techniques are useful to study populations for which obtaining tissue samples is not logistically feasible. for example, fecal pellet dna has been used to estimate population size in roe deer (capreolus capreolus; ebert et al. 2012) and sitka black-tailed deer (odocoileus hemionus sitkensis; brinkman et al. 2011), two species that live in densely vegetated habitats that are difficult to survey. population size, survival rate, and rate of population unger et al. – dna source types in moose alces vol. 53, 2017 182 size change have been estimated for a protected subspecies of woodland caribou (rangifer tarandus caribou) in canada using dna from non-invasively collected fecal pellets (hettinga et al. 2012). dna has also been extracted from fecal pellets of mountain goats (oreamnos americanus; poole et al. 2011) and red deer (cervus elaphus; valière et al. 2007). the quality of dna extracted from fecal pellets can be problematic. typically, higher quantity and quality dna is obtained from body tissues than from fecal samples (waits and paetkau 2005, ball et al. 2007) because dna from feces is more degraded and more likely to be contaminated. the rate of dna degradation is also affected by the time since deposition and environmental conditions (kreader 1996, piggott 2004, brinkman et al. 2010b). high temperatures, rainfall, bacteria and fungi, and exposure to uv radiation increase the degradation rate of dna (piggott 2004, brinkman et al. 2010b, buś and allen 2014). dna can be extracted with higher success from moose pellets collected from snow in late spring versus pellets collected after snowmelt and temperatures warm (rea et al. 2016). dna extracted from fecal samples is further affected by the diet. fecal pellet samples from herbivores contain tannins and other substances of vegetative (diet) origin which increases the number of pcr inhibitors in extracted dna (kreader 1996), and they may also contain dna from plants. carnivore scat includes dna from prey species which may inflate dna concentration measurements, but will not affect results if genetic markers are species-specific (deagle et al. 2005). in addition to inherent factors involving the sample itself, proper collection and lab techniques, including sample collection, storage methods, and specified extraction protocols, are necessary to ensure quality results from all dna source types (waits and paetkau 2005, beja-pereira et al. 2009, buś and allen 2014). dna quantity and quality are important because they directly affect pcr amplification success, genotyping success, and genotyping error rates (taberlet et al. 1996, mckelvey and schwartz 2004, waits and paetkau 2005, ball et al. 2007, brinkman et al. 2010a). as dna degrades, nucleic acid residues undergo chemical changes and strands become fragmented (buś and allen 2014), and this fragmentation results in lower pcr amplification success and increased genotyping errors (taberlet et al. 1999). common pcr amplification problems using degraded dna include failure of dna to amplify due to absence of usable dna and genotyping errors (i.e., false alleles and allelic dropout). genetic techniques have been tested using dna from various source types for several species, but there are few direct comparisons among source types for ungulates (wehausen et al. 2004, valière et al. 2007), and none for moose. moose population genetic research to date has used either dna from blood or tissue, except for the recent tests with fecal pellets by rea et al. (2016). it would be useful to know the relative extraction and genotyping success from different dna sources because source type affects the time and resources required for population genetic studies. we compared dna extraction from existing samples of moose liver tissue, blood, and fecal pellets. we measured average dna yield, compared pcr amplification and genotyping success rates, and identified ways to improve extraction efficiency in the protocols. methods we compared dna from biological sources collected in northern minnesota alces vol. 53, 2017 unger et al. – dna source types in moose 183 (mn) and the northern range of yellowstone national park (ynp), usa. samples from mn had been archived, and samples from ynp were collected from free-ranging moose in winter 2013. study areas minnesota.— the study area of pellet origin includes northern minnesota which transitions from mixed conifer-deciduous forests, bog, and swamp in the east to an agricultural matrix in the west. the northeastern area is characterized by coniferdeciduous forests, conifer bogs and swamps, and numerous small lakes, peatlands, and wet forest throughout. the predominant tree species are white pine (pinus strobus), red pine (p. resinosa), quaking aspen (populus tremuloides), paper birch (betula papyrifera), white spruce (picea glauca), balsam fir (abies balsamea), and white cedar (thuja occidentalis). the northwestern area is relatively flat and dominated by aspen parkland and farmland. yellowstone national park.— the study area included the portion of the northern yellowstone elk winter range (houston 1982) located within ynp as well as some creek drainages located outside the park. vegetation consists primarily of sage steppe and grassland at low elevation (< 2,000 m), and coniferous forests at high elevation (>3,000 m). the most common conifers are lodgepole pine (pinus contorta), engelmann spruce (picea engelmannii), sub-alpine fir (abies lasiocarpa), douglas-fir (pseudotsuga menziesii), and whitebark pine (pinus albicaulis). willow (salix spp.) is present in drainages and other wet areas. samples minnesota.— blood samples (hereafter blood) were collected by the minnesota department of natural resources (mndnr) and stored on whatman fta® classic cards (whatman international ltd., maidstone, united kingdom). they were collected from hunter-harvested (n = 116), gps-collared (n = 132), and sick (n = 6) moose in 2011– 2013 (fig. 1). samples for hunter-harvested moose were not available in 2013 after the cancellation of the moose hunt (delgiudice 2014). liver tissue (n = 31) samples from sick moose were collected from 2009–2012 throughout northern mn and frozen after collection (fig. 1). moose health was determined by mndnr personnel based on a range of observations such as non-normal behavior associated with neurological impairment, emaciation, and inability to stand upright. sick moose were either found dead or were euthanized. sex of moose was determined by direct observation except for certain unidentified/unmarked samples: sample sizes were 148 males, 108 females, and 29 unknown. fta cards were stored at room temperature and frozen liver tissue was stored at -20 °c until analysis. yellowstone national park.— fecal pellet samples (n = 489) were collected primarily along drainages in ynp (fig. 2) during winter months with snow present on fig. 1. location of hunter-harvested (n = 117), gps-collared (n = 132), and sick (n = 36) moose used in genetic studies in minnesota, usa. manitoba minnesota wisconsin ontario unger et al. – dna source types in moose alces vol. 53, 2017 184 the ground. a minimum of 5 fecal pellets from each deposition pile of each sampled moose was collected and stored in whirlpak® or ziploc freezer bags. tissue samples (n = 2) collected opportunistically from dead animals were stored and frozen in whirlpak® bags. fecal pellets and tissue collected in ynp were kept frozen and sent to the university of minnesota-duluth for analysis. date and time of sample collection, location, and estimated age of the sample was provided for most samples. sex was determined directly, by collecting fecal pellet samples from an observed animal or inferred from physiological and behavioral clues, such as size of snow bed or presence of a calf. the age of fecal pellet samples at collection was estimated based on direct observation of the moose for certain samples. for fecal pellets found on snow without observing the moose, evidence from the fecal pellets, tracks, snowfall dates, or snow montana idaho wyoming fig. 2. location of non-invasively collected moose fecal pellet samples (n = 489) in the northern range of yellowstone national park, wyoming and montana, usa. alces vol. 53, 2017 unger et al. – dna source types in moose 185 cover was used to estimate the age of the sample. all fecal pellets were found on snow that had fallen in the current winter, and it is likely that most samples were much less than 4 months old. the maximum age of fecal pellets could not be estimated, but we know that all pellets were deposited in the year collected, and were on or in continuous snow cover until collection. tissue and fecal pellets were stored at -20 °c until analysis. extraction blood.—whole genomic dna was extracted from one drop of dried blood for each sample (n = 251) using the fermentas/ thermo scientific genejet whole blood genomic dna purification mini kit (thermo fisher scientific inc., pittsburgh, pennsylvania) according to the manufacturer’s protocol. each blood drop was taken from individual fta cards using a 4 mm hand punch. blood drops were categorized as small, medium, or large with diameters averaging 7.4 mm, 9.8 mm, and 11.8 mm, respectively. extractions followed kit protocol, with one modification to improve final dna yield, particularly for small blood drops. small blood drops were eluted twice with 100 µl of elution buffer and then pipetted back into the spin column after being centrifuged, producing a final volume of 100 µl. medium and large blood drops were also eluted twice, but with new elution buffer for the second elution step, resulting in a final volume of 200 µl. liver tissue.—dna was extracted from 0.02 g of frozen liver tissue (n = 33) using the thermo scientific genejet genomic dna purification kit (thermo fisher scientific inc., pittsburgh, pensylvania) and the manufacturer’s protocol with 4 modifications to improve final dna yield and purity: 1) overnight incubation with digestion solution and proteinase k instead of the manufacturer’s recommended 3–4 h, 2) a minute added to each of the highest speed centrifugation times, 3) a minute added to the elution buffer incubation time, and 4) a second elution step, resulting in a final elution volume of 400 µl. fecal pellets.— three dna extraction kits were tested during a pilot study using 30 fecal pellet samples: 1) qiaamp dna stool mini kit (qiagen inc., valencia, california), 2) thermo scientific genejet genomic dna purification kit (thermo fisher scientific inc., pittsburgh, pennsylvania), and 3) powerfecal dna isolation kit (mo bio laboratories inc., carlsbad, california). qiagen’s stool extraction kit produced sufficient dna yield and pcr success with our laboratory techniques (described below), and was chosen for large scale dna extractions with fecal pellet samples. dna was extracted from fecal pellets (n = 489) using two qiagen dna extraction kits because the manufacturer discontinued the first kit we used. the first method involved extracting dna from one whole fecal pellet (n = 301) using the qiaamp dna stool mini kit (qiagen inc., valencia, california) and a modified protocol designed to isolate dna from intestinal cells sloughed off onto the surface of fecal pellets (esteszumpf et al. 2014). inner fecal pellet material can contain pcr inhibitors that lead to increased variability in pcr amplification and genotyping success rates (flagstad et al. 1999, wehausen et al. 2004). to exclude this material from the extraction process, each fecal pellet was submerged in stool lysis buffer (buffer asl) from the qiaamp dna stool mini kit, and agitated to rinse cells off the outer surface instead of vortexing, which can break up the fecal pellet exposing the inner material. unger et al. – dna source types in moose alces vol. 53, 2017 186 the second method we used to extract dna from fecal pellets (n = 188) was with the qiaamp fast dna stool mini kit (qiagen inc., valencia, california) and the manufacturer’s protocol with several modifications to improve final dna yield and purity, including: 1) centrifuging after step 2 to reduce bubbles caused by vortexing, and 2) reducing centrifuge rates during step 14 to 6,000 x g instead of the manufacturer’s recommended 20,000 x g. this protocol required that a portion of a fecal pellet be used, rather than the whole pellet. we used a razor to slice thin layers from the outer fecal pellet material that contained the sloughed off intestinal mucosal cells. for both methods, fecal pellets were kept frozen until processing to prevent thaw and subsequent break up of the pellets. we did not use fecal pellet samples that were frozen together or samples with snow in the collection bag. sex determination sex was determined for each sample using the se47/se48 primer pair (brinkman and hundertmark 2009). this primer pair produces a single band for females and a double band for males by pcr amplifying xand y-specific alleles of the amelogenin gene. this method has been used previously with moose and other cervid species (brinkman and hundertmark 2009). pcr products for sex identification were visualized using gel electrophoresis on 1.5% agarose gels stained with 10 mg/ml of ethidium bromide. genotyping dna extracted from blood (n = 248) and fecal pellets (n = 269) was genotyped using one sex-linked and 15 autosomal microsatellites previously used for moose (table 1). all autosomal forward primers contain an m13 (-21) tail on the 5’ end (schuelke 2000), and pcr products were labeled by incorporating a universal fluorescently labeled m13 (-21) primer (fam, pet, or vic) during pcr. for dna extracted from blood sources, all microsatellites were amplified separately with a total volume of 13 µl containing sterile water, gotaq dna polymerase, and 1x gotaq buffer (promega corporation, madison, wisconsin), 2 mm mgcl 2 , 0.2 mm dntps, 0.08 µm forward primer, 0.8 µm reverse primer, 0.8 µm labeled primer, 1% bovine serum albumin (bsa), and 1 µl/reaction dna. bsa was added to all pcr to bind potential inhibitors and improve amplification specificity (kreader 1996). the addition of bsa to pcr was initially implemented to increase amplification success using dna extracted from fecal pellets. it was added to pcr using dna extracted from blood and tissue for consistency, and to potentially increase pcr amplification success for dna from each source type. dna extracted from fecal pellet sources was amplified similarly using the same set of microsatellite loci; however, an additional step was taken for pcr due to low amplification success rates in a pilot study. for dna extracted from fecal pellets, microsatellite loci were amplified using either single step or pre-amplification pcr methods (table 1). pre-amplification is a two-step pcr method designed to increase the amount of dna template for amplification and reduce genotyping error (piggott et al. 2004). because the success of the pre-amplification method has been questioned (hedmark and ellegren 2006, de barba and waits 2010), we conducted a pilot study to test this method using dna extracted from fecal pellets amplified with our microsatellites. we used pre-amplification methods described in piggott et al. (2004) with pcr mixtures modified for reduced total volume (tjepkes 2015). the pre-amplification method was used only on loci for which it increased pcr alces vol. 53, 2017 unger et al. – dna source types in moose 187 amplification success and subsequent genotyping success. analyses pcr products were analyzed at the university of minnesota biomedical genomics center using an abi 3730xl capillary genetic analyzer. genotypes were assigned using genemarker (v.2.6.0, softgenetics llc, state college, pennsylvania) to score alleles for each locus. failed or ambiguous allele scores were re-amplified and genotyped again to reduce scoring errors and missing data. dna quantification.— dna concentration for each sample was quantified using a nanodrop nd-1000 spectrophotometer. using dna concentration and total extraction volume, total dna yield from each extraction was calculated for comparison. the two protocols for dna extraction from fecal pellets were also compared to determine which method resulted in higher dna yield. dna yield was compared between blood, tissue, and fecal pellet samples using anova with bonferroni post hoc comparisons. sex determination.— sex determination success rates were compared between dna extracted from blood, tissue, and fecal pellet samples using fisher’s exact test. the accuracy of sex determination using genetic methods was determined by comparing genetically determined sex with recorded sex from direct field observations using dna from blood, tissue, and fecal pellet table 1. characteristics of one sex-linked and 15 autosomal microsatellites used in genetic analysis. autosomal microsatellites were pcr amplified with m13 fluorescently labeled primers, then combined into non-overlapping panels for genotyping. ta is the optimal annealing temperature. the designated pcr method is for dna extracted from moose fecal pellets; dna extracted from blood samples was all pcr amplified using the single step method. locus m13 t a (°c) size range (bp) pcr method reference se47/se48 53 224–260 single step [1] rt30 vic 54 212–232 single step [2] rt5 fam 54 168–180 single step [2] rt1 pet 47 247–255 pre-amp [2] rt9 pet 54 138–154 single step [2] bl42 fam 49 263–285 pre-amp [3] bm848 fam 54 358–382 single step [3] bm888 vic 50 191–209 single step [3] bm1225 fam 50 237–267 single step [4] bm2830 pet 50 127–139 single step [4] crfa vic 53 264–274 pre-amp [5] kcsn pet 50 206–218 pre-amp [5] igf-1 fam 54 123–127 single step [5] cervid14 fam 54 227–249 single step [6] nvhrt03 pet 54 124–138 pre-amp [7] nvhrt21 vic 50 174–190 pre-amp [7] references: [1] brinkman and hundertmark 2009; [2] wilson et al. 1997; [3] hundertmark 2009; [4] broders et al. 1999; [5] cronin et al. 2001; [6] wilson et al. 2003; [7] roed and midthjell 1998. unger et al. – dna source types in moose alces vol. 53, 2017 188 samples. additionally, genetically determined sex using dna from fecal pellet samples was compared to recorded sex from indirect field observations. pcr and genotyping success rates.— success of pcr amplification of autosomal microsatellites and genotyping success rates were calculated. the autosomal pcr success rate was the proportion of successful pcr amplification attempts used to estimate the amount of effort required in the laboratory for a sample. pcr attempts were classified as successful if they produced viable product that could be used to genotype individuals. autosomal pcr success rates using dna extracted from blood and fecal pellets were compared using a χ2 test for independence. the genotyping success rate allows for a comparison of overall success using dna from several dna source types; however, it does not reflect the amount of effort and resources required to obtain that success. for example, if all 15 loci were successfully pcr amplified on the first attempt, the overall genotyping success rate would be 100% (15 of 15 pcr amplifications were successful) and the autosomal success rate would also be 100%. if fewer than 15 loci were successfully pcr amplified in the first attempt, but then all 15 loci were successfully prc amplified in a second attempt, the overall genotyping success rate would also be 100%, but the second sample would have required twice as many pcr amplifications. the autosomal pcr success rate would be between 50% (15/30) and 97% (29/30) depending on how many loci were successfully amplified in the second sample. genotyping success rate is the proportion of microsatellites for which we were able to obtain genotypes for a given sample, given that a pcr amplification attempt was successful. pcr and allele scoring was attempted a second time for dna samples if the first attempt failed. individual genotyping success rates were then averaged for each dna source type. genotyping success rates using dna extracted from blood and fecal pellet samples were compared using two-sample t-tests. genotyping error.— a concern when working with lower quality and quantity dna in fecal samples is the increased risk of genotyping errors such as allelic dropout or false alleles. to estimate genotyping error, pcr and allele scoring was repeated for at least 24 randomly chosen individuals at each microsatellite locus using previously extracted dna from fecal pellets. however, because many of the original repetitions using dna extracted from fecal pellet samples did not produce a usable amplicon, 95 additional randomly chosen dna samples from fecal pellets were repeated at 6 microsatellite loci. genotyping error was calculated using these duplicated allele scores. we were interested only in estimating genotyping error and did not investigate observed genotyping errors further (i.e., amplifying conflicting loci a third time) as recommended by taberlet et al. (1996). predictors of genotyping success.— to identify ways to improve efficiency and success rates, dna yield and sex determination success were tested as predictors of genotyping success. two-sample t-tests were used to determine whether higher dna yield or pcr amplification success of se47/ se48 resulted in greater genotyping success. if dna yield or pcr amplification success leads to greater genotyping success, poor quality samples could be identified and censored, reducing the amount of time and effort spent on those samples. effect of fecal pellet age.— since numerous environmental factors degrade dna, the time between deposition and collection was investigated to determine an effect on dna quality and quantity. time alces vol. 53, 2017 unger et al. – dna source types in moose 189 since deposition was estimated for fecal pellet samples based on observations made in the field during collection, including visual fecal pellet characteristics, moose tracks, and snowfall. using these data, dna was separated into 3 fecal pellet age classes: < 24 h (n = 84), 24–48 h (n = 65), and > 48 h (n = 109) since deposition. dna yield for each fecal pellet age class was compared using anova to determine if time since deposition affected the amount of dna obtained from fecal pellets. sex-linked and autosomal pcr amplification success between age classes was compared using fisher’s exact and χ2 tests, respectively. finally, two-sample t-tests were used to determine if age since deposition affected overall genotyping success. because of potential uncertainty in differentiating fecal pellets deposited < 24 h and 24–48 h since deposition, pcr amplification and genotyping success rates between dna from fecal pellets collected < 48 and > 48 h after deposition were also compared. results dna yield we extracted dna from 251 blood samples, 33 liver tissue samples, and 489 fecal pellet samples. yield of extracted dna for blood, tissue, and fecal pellet samples was sufficient for genotyping. blood and liver tissue samples produced the highest average dna yield, and fecal pellets produced the lowest average dna yield (anova, f 3,769 = 113, p < 0.001). dna extractions using sliced fecal pellets had slightly higher average dna yield than dna extractions using whole fecal pellets (fig. 3). sex determination we determined sex with primer pair se47/se48 using dna extracted from all 3 sample sources. sex determination was successful for 100% of blood samples (n = 251), 91% of tissue samples (n = 33), and 83% of fecal pellet samples (n = 460). success was significantly greater using dna extracted from blood than liver tissue or fecal pellets (fisher’s exact test, p < 0.01). the success rates using extracted dna from tissue and fecal pellets were not significantly different (fisher’s exact test, p > 0.05). we calculated the accuracy of sex determination using these methods by comparing results determined genetically to field records from moose directly observed depositing fecal pellets. genetic and field sex determination was consistent for 219 of 225 blood samples (0.97), 24 of 25 tissue samples (0.96), and 64 of 67 fecal pellet samples (0.96) when analyzed separately. when sex was determined from indirect evidence in ynp, the geneticallydetermined and field-determined sex were consistent for 74 of 100 fecal pellet samples (0.74). the 26 fecal pellet samples that resulted in inconsistent sex determination results were from both males (n = 11) and females (n = 15). fig. 3. average dna yield extracted from multiple dna source types. dna sources were from moose in minnesota (blood, n = 251; tissue, n = 31) and yellowstone national park (tissue, n = 2; fecal pellet, n = 489). fecal pellets were extracted using either the whole fecal pellet (n = 188) or using outer slices of fecal pellet (n = 301). unger et al. – dna source types in moose alces vol. 53, 2017 190 pcr and genotyping success rates we genotyped 517 total samples using dna from blood (n = 248) and fecal pellets (n = 269) at 15 autosomal microsatellite markers. average autosomal pcr success rate was higher for blood than fecal pellets (0.81 vs. 0.63, respectively, χ1 2 = 57, p < 0.001). similarly, average genotyping success rate, the percent of microsatellites that we were able to use if pcr amplification was successful, was higher for dna extracted from blood than fecal pellets (0.82 vs. 0.76, respectively, t 506 = 6.04, p < 0.001). genotyping error the average genotyping error rate calculated by repeated pcr amplification and allele scoring using dna extracted from fecal pellet samples was 0.10 for re-analyzed microsatellite loci. pcr did not produce usable amplicons for 3 microsatellite loci (igf-1, rt5, and crfa) using dna extracted from fecal pellets. predictors of genotyping success dna yield and pcr amplification success of the sex-linked primer pair se47/ se48 were tested in order to determine whether they could be used as predictors of downstream success, particularly genotyping success. dna yield was not correlated with genotyping success using dna extracted from blood or fecal pellets (fig. 4). dna extracted from fecal pellets that successfully pcr amplified at the sex -linked loci had higher genotyping success at autosomal loci compared to those that failed to amplify at this locus (0.78 [n = 236] vs. 0.59 [n = 31], respectively, t 36 = -4.61, p < 0.001). this comparison could not be made for dna from blood or tissue samples because the pcr success of > 96% produced too few failed samples to test. effect of fecal pellet age time since deposition was estimated for 447 samples as either < 24 h (n = 178), 24–48 h (n = 114), or > 48 h (n = 155). average dna yield was not significantly different between different age classes (anova, f 2,255 = 1.73, p = 0.18; table 5). dna extracted from fecal pellets collected < 24 h and 24–48 h after deposition had the highest sex determination success rates and were not different from each other (0.85 and 0.92, respectively, fisher’s exact test, p = 0.10), whereas dna from fecal pellets collected > 48 h after deposition had lower success rate than fecal pellets collected < 48 h after deposition (0.74, fisher’s exact test, p = 0.0003; table 2). in addition, fecal pellet samples were compared to determine whether fecal a. blood b. fecal fig. 4. genotyping success rate for (a) blood and (b) fecal pellet samples of varying dna yield (µg) from moose in minnesota and yellow national park, usa. alces vol. 53, 2017 unger et al. – dna source types in moose 191 pellet age at collection influenced pcr amplification or genotyping success rates. similar to sex determination success rates, fecal pellets collected < 24 and 24–48 h after deposition had the highest pcr success rates and were not different from each other (0.67 and 0.69, respectively), whereas the success rate for fecal pellets collected > 48 h after deposition was significantly lower (0.56) (table 2). autosomal pcr success rate for fecal pellets in the < 24 and 24–48 h age classes was not different ( χ 2 2 = 1.31, p = 0.25), but was different from fecal pellets collected > 48 h after deposition ( χ 2 2 = 46, p < 0.001). genotyping success rate was the highest for dna from fecal pellets collected < 24 and 24–48 h after deposition (0.82 and 0.83, respectively) (fig. 5). fecal pellets in the < 24 and 24–48 h age classes were not different (t 136 = -0.10, p = 0.92), but were different from fecal pellets collected > 48 h after deposition (t 191 = -5.29, p < 0.001). fecal pellets collected > 48 h after deposition had the lowest average genotyping success rate (0.69), which was significantly different from the two other age classes. similar results for pcr amplification and genotyping success rates were obtained when dna extracted from fecal pellets collected < 48 h after deposition was compared to dna extracted from fecal pellets collected > 48 h after deposition. discussion dna was successfully extracted from blood, liver tissue, sliced fecal pellets, and whole fecal pellets. tissue samples produced the greatest amount of dna per extraction and fecal pellets produced the smallest average dna yield. because fecal pellet samples had less available sample and required a greater amount of pcr, these dna samples were more likely to be exhausted. there is also variability in dna quantity and quality in fecal pellet samples, even among fecal pellets from the same individual and the table 2. effect of moose fecal pellet age on dna yield, sex-linked microsatellite pcr amplification success (sex-linked pcr success), and autosomal microsatellite pcr amplification success (autosomal pcr success) using 15 microsatellite markers using fecal pellets collected >24 (n = 84), 24–48 (n = 65), and >48 h (n = 109) after deposition. fecal pellets in > 48 h age class ranged from 48 h to 4 months since deposition. time since deposition (h) average dna yield (µg) sex-linked pcr success autosomal pcr success <24 2.54 0.85 0.67 24–48 2.45 0.92 0.69 >48 2.08 0.74 0.56 fig. 5. effect of fecal pellet age at collection (<24 h, n = 84; 24–48 h, n = 65; >48 h, n = 109) on individual genotyping success rate from moose in yellow national park, usa. age of fecal pellets in the > 48 h age class was unknown, but pellets were collected from on top of snow in the current winter. the age could have been up to 4 months, but was likely less. unger et al. – dna source types in moose alces vol. 53, 2017 192 same fecal pellet group (taberlet et al. 1996). therefore, it would be best to maximize fecal pellet collection in the field. dna yield was not a reliable predictor of genotyping success as expected. dna yield estimates the amount of dna in a sample, but it does not provide information on dna quality, presence of pcr inhibitors, or presence of foreign dna. samples with high dna yield may have dna degradation, and indeed, we found evidence supporting this because pellets > 48 h old had similar dna yield as pellets < 48 h old, but lower genotyping success. dna extracted from blood had the highest average pcr amplification success for both sex-linked and autosomal microsatellites. this suggests higher quality dna was extracted from blood than liver tissue, as liver tissue samples had higher average dna yield. dna extracted from fecal pellets had the lowest sex determination success, suggesting lower dna quality and/or increased presence of pcr inhibitors. genotyping and pcr amplification success greater effort was required for genotyping success with fecal pellet samples than blood samples. this was because autosomal pcr amplification success was lower using dna from fecal pellets compared to blood, a consequence of the lower quality dna. for regions or populations where non-invasive genetic sampling is the only feasible option, fecal pellet samples are an alternative source of dna, but only if collected at the optimal time of year. if higher quality samples such as blood or liver tissue are available, these samples are easier to process in the laboratory; however, the effort required to obtain samples is an important consideration. obtaining samples of blood, muscle, or other tissues can require more effort than obtaining fecal pellet samples, unless collection is in conjunction with other projects. reported pcr amplification success using dna from feces has been variable (wehausen 2004, broquet et al. 2007), and is likely due to factors beyond the dna source. for example, methods of sample collection, storage, and extraction affect dna quality and downstream success (roon et al. 2003, wehausen 2004, waits and paetkau 2005). the effectiveness of these methods has been tested, but without clear consensus (luikart et al. 2006, beja-pereira et al. 2009). the lack of consensus is likely influenced by inherent variation among species and environmental variables (waits and paetkau 2005); therefore, it is essential to conduct a pilot study before beginning large-scale extractions (taberlet et al. 1999). improvements and recommendations we made many attempts to improve pcr success using dna from fecal pellets, including testing multiple dna extraction kits and protocols, adding bsa to remove pcr inhibitors, using the pre-amplification method for pcr, and testing a variety of pcr conditions (optimizing pcr master mix ingredients and pcr temperature profiles) for each microsatellite locus. the average amount of dna extracted from sliced fecal pellets was greater than that from whole fecal pellets using the surface washing method. a similar comparison using whole fecal pellets and sliced outer fecal pellet material from bighorn sheep found no difference in extracted dna yield (wehausen 2004). however, the differences could be due to the substantially larger sample size in our study, and differences in species, environmental conditions, and extraction protocols. we found slicing fecal pellets to be a more time consuming process with increased probability of contamination through exposure to multiple laboratory alces vol. 53, 2017 unger et al. – dna source types in moose 193 surfaces (e.g., cutting surface, razor, and forceps), even though protocols were in place to minimize such. additionally, accidental inclusion of inner fecal pellet may increase pcr amplification failure and variation among samples (wehausen 2004). therefore, although the sliced fecal pellet method resulted in higher average dna yield, we prefer the whole fecal pellet surface washing method. although not quantified, we felt that pcr success using dna extracted from fecal pellets improved with the inclusion of bsa, although it was not as beneficial when attempting to amplify dna extracted from blood or tissue. this is because bsa does not have a noticeable effect on pcr amplification success using dna with low levels of pcr inhibitors (kreader 1996). thus, the difference in the effect of including bsa provides evidence for increased levels of pcr inhibitors in dna extracted from fecal pellets compared to either blood or tissue. the pre-amplification method for pcr was beneficial for 6 microsatellites. however, this method required twice the number of pcrs, and thus more resources (time and reagents). in some cases the pre-amplification method caused non-specific amplification or amplification patterns that created difficulty for allele scoring, thus, we recommend caution with this method. time since deposition affected genotyping success of fecal pellets. dna is degraded by several environmental conditions (e.g., high temperature, precipitation, uv radiation, and microorganisms), thus the longer it is subjected to adverse conditions, the less likely a fecal pellet will contain usable dna. selecting fresh fecal pellets (< 48 h after deposition) will likely result in better quality and quantity of dna, and improved pcr amplification and genotyping success rates. success rate was lower when pellets of unknown age were collected, and fecal pellets collected after march had low genotyping success (rea et al. 2016). however, samples collected in dry or protected areas may have a substantially larger window for collection (brinkman et al. 2010b). if precise estimates of fecal pellet age were available, it could be possible to determine the age at which fecal pellets should not be collected/used for dna analysis. variability does exist in dna quality and quantity from fresh fecal pellets. poor quality samples can be removed from analysis based on pcr amplification success using primer pair se47/se48, and other microsatellites could be used for this purpose. however, using a sex-linked microsatellite enables identification of sex, which is often missing from non-invasively collected samples. an additional benefit of using a sex -linked microsatellite is to check for errors in data records, and to evaluate accuracy of identifying the sex of animals not directly observed. in fact, we identified such errors in data from ynp. if the sex-linked microsatellite fails, then it is likely that dna extraction will not be successful, and the sample could be censored from the data set without investing additional resources of time or materials. future work could include testing the effectiveness of sample collection and storage methods. blood and tissue samples were collected and stored before our study began, and reported methods from other non -invasive genetic studies for collection and storage were used (carr et al. 2010, ebert et al. 2012). however, collection and storage methods might have influenced our results, as for example, liver tissue samples from sick moose that were found dead or euthanized. death and subsequent post-mortem decomposition result in dna degradation, particularly in liver tissue (alaeddini et al. 2010). samples unger et al. – dna source types in moose alces vol. 53, 2017 194 were stored in -20 °c freezers, but this has not proven entirely effective at eliminating dna degradation (dawson et al. 1998), especially for extended periods. storage beyond 6 months reduces both dna yield and pcr amplification success for multiple dna source types (roon et al. 2003). extracted dna quantity and quality from tissues we used may have been improved using samples collected more recently, if dna had been extracted immediately following collection, or if samples had been stored at colder temperatures. however, it is also significant that despite the potential issues that could have confounded successful dna extraction, we were able to extract dna from most of these samples with reasonably high success. conclusion population genetic studies of moose have traditionally used dna extracted from tissue, blood, or a combination of these sources. the source, methods, and available resources (time and money) which influence the quantity and quality of data typically vary in genetic research with moose. similar to that found with other ungulate species (luikart et al. 2006, brinkman et al. 2010a, ebert et al. 2012), we have shown that moose fecal pellets are viable sources of dna. importantly, fecal pellets can be collected non-invasively which increases the sampling potential in genetic studies of moose. we provide guidelines concerning the optimal collection, storage, and laboratory procedures for using fecal pellets in population genetic studies with moose. acknowledgements we would like to thank the university of minnesota-duluth for funding, mndnr for providing minnesota moose samples, and our collaborators k. and l. koitzsch from k2 consulting, llc for funding and fieldwork performed in ynp. partial support for rm on this project was provided by the minnesota environment and natural resources trust fund as recommended by the legislative-citizen commission on minnesota resources (lccmr). two anonymous reviewers also provided helpful comments on the manuscript. literature cited alaeddini, r., s. j. walsh, and a. abbas. 2010. forensic implications of genetic analyses from degraded dna a review. forensic science international: genetics 4: 148–157. ball, m. c., r. pither, m. manseau, j. clark, s. d. petersen, s. kingston, n. morrill, and p. wilson. 2007. characterization of target nuclear dna from faeces reduces technical issues associated with the assumptions of low-quality and quantity template. conservation genetics 8: 577–586. beja-pereira, a., r. oliveira, p. c. alves, m. k. schwartz, and g. luikart. 2009. advancing ecological understandings through technological transformations in noninvasive genetics. molecular ecology resoures 9: 1279–1301. brinkman, t. j., and k. j. hundertmark. 2009. sex identification of northern ungulates using low quality and quantity dna. conservation genetics 10: 1189–1193. _____, d. k. person, f. s. chapin, w. smith, and k. j. hundertmark. 2011. estimating abundance of sitka black-tailed deer using dna from fecal pellets. journal of wildlife management 75: 232–242. _____, _____, m. k. schwartz, k. l. pilgrim, k. e. colson, and k. j. hundertmark. 2010a. individual identification of sitka black-tailed deer alces vol. 53, 2017 unger et al. – dna source types in moose 195 (odocoileus hemionus sitkensis) using dna from fecal pellets. conservation genetics resoures 2: 115–118. _____, m. k. schwartz, d. k. person, k. l. pilgrim, and k. j. hundertmark. 2010b. effects of time and rainfall on pcr success using dna extracted from deer fecal pellets. conservation genetics 11: 1547–1552. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8: 1309–1315. broquet, t., n. menard, and e. petit. 2007. noninvasive population genetics: a review of sample source, diet, fragment length and microsatellite motif effects on amplification success and genotyping error rates. conservation genetics 8: 249–260. buś, m. m., and m. allen. 2014. collecting and preserving biological samples from challenging environments for dna analysis. biopreservation and biobanking 12: 17–22. carr, n. l., a. r. rodgers, s. r. kingston, p. n. hettinga, l. m. thompson, j. l. renton, and p. j. wilson. 2010. comparative woodland caribou population surveys in slate islands provincial park, ontario. rangifer 20: 205–217. coulon, a., j.f. cosson, j. m. angibault, b. cargnelutti, m. galan, n. morellet, e. petit, s. aulagnier, and a. j. m. hewison. 2004. landscape connectivity influences gene flow in a roe deer population inhabiting a fragmented landscape: an individual–based approach. molecular ecology 13: 2841–2850. cronin, m. a., j. c. patton, r. courtois, and m. crete. 2001. genetic variation of microsatellite dna in moose in quebec. alces 37: 175–187. dawson, m. n., k. a. raskoff, and d. k. jacobs. 1998. field preservation of marine invertebrate tissue for dna analyses. molecular marine biology and biotechnology 7: 145–152. de barba, m., and l. p. waits. 2010. multiplex pre-amplification for noninvasive genetic sampling: is the extra effort worth it? molecular ecology resoures 10: 659–665. deagle, b. e., d. j. tollit, s. n. jarman, m. a. hindell, a. w. trites, and n. j. gales. 2005. molecular scatology as a tool to study diet: analysis of prey dna in scats from captive steller sea lions. molecular ecology 14: 1831–1842. delgiudice, g. d. 2014. 2013 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. ebert, c., j. sandrini, b. spielberger, b. thiele, and u. hohmann. 2012. noninvasive genetic approaches for estimation of ungulate population size: a study on roe deer (capreolus capreolus) based on faeces. animal biodiversity and conservation 35: 267–275. estes-zumpf, w. a., s. e. zumpf, j. l. rachlow, j. r. adams, and l. p. waits. 2014. genetic evidence confirms the presence of pygmy rabbits in colorado. journal of fish and wildlife management 5: 118–123. finnegan, l. a., p. j. wilson, g. n. price, s. j. lowe, b. r. patterson, o. flagstad, k. røed, j. e. stagy, and k. s. jakobsen. 1999. reliable noninvasive genotyping based on excremental pcr of nuclear dna purified with a magnetic bead protocol. molecular ecology 8: 879–883. flagstad, o., k. roed, j. e. stacy, and k. s. jacobsen. 1999. reliable noninvasive genotyping based on excremental pcr of nuclear dna purified with a magnetic bead protocol. molecular ecology 8: 879-883. hedmark, e., and h. ellegren. 2006. a test of the multiplex pre-amplification approach in microsatellite genotyping of wolverine faecal dna. conservation genetics 7: 289–293. unger et al. – dna source types in moose alces vol. 53, 2017 196 hettinga, p. n., a. n. arnason, m. manseau, d. cross, k. whaley, and p. j. wilson. 2012. estimating size and trend of the north interlake woodland caribou population using fecal dna and capture-recapture models. journal of wildlife management 76: 1153–1164. houston, d. b. 1982. the northern yellowstone elk: ecology and management. macmillan, new york, new york, usa. hundertmark, k. j. 2009. reduced genetic diversity in two introduced and isolated moose populations in alaska. alces 45: 137–142. kangas, v. -m., l. kvist, s. laaksonen, t. nygren, and j. aspi. 2013. present genetic structure revealed by microsatellites reflects recent history of the finnish moose (alces alces). european journal of wildlife research 59: 613–627. kreader, c. a. 1996. relief of amplification inhibition in pcr with bovine serum albumin or t4 gene 32 protein. applied and environmental microbiology 62: 1102–1106. luikart, g., s. zundel, d. rioux, c. miquel, k. a. keating, j. t. hogg, b. steele, k. foresman, and p. taberlet. 2006. low genotyping error rates and noninvasive sampling in bighorn sheep. journal of wildlife management 72: 299–304. mckelvey, k. s., and m. k. schwartz. 2004. genetic errors associated with population estimation using non-invasive molecular tagging: problems and new solutions. journal of wildlife management 68: 439–448. piggott, m. p. 2004. effect of sample age and season of collection on the reliability of microsatellite genotyping of faecal dna. wildlife research 31: 485–493. _____, e. bellemain, p. taberlet, and a. c. taylor. 2004. a multiplex pre amplification method that significantly improves microsatellite amplification and error rates for faecal dna in limiting conditions. conservation genetics 5: 417–420. poole, k. g., d. m. reynolds, g. mowat, and d. paetkau. 2011. estimating mountain goat abundance using dna from fecal pellets. journal of wildlife management 75: 1527–1534. rea, r. v., c. j. johnson, b. w. murray, d. p. hodder, and s. m. crowley. 2016. timing moose pellet collections to increase genoytyping success of fecal dna. journal of fish and wildlife management 7: 461–466. roed, k. h., and l. midthjell. 1998. microsatellites in reindeer, rangifer tarandus, and their use in other cervids. molecular ecology 7: 1771–1788. roon, d. a., l. p. waits, and k. c. kendall. 2003. a quantitative evaluation of two methods for preserving hair samples. molecular ecology notes 3: 163–166. schuelke, m. 2000. an economic method for the fluorescent labeling of pcr fragments. nature biotechnology 18: 233–234. taberlet, p., s. griffin, b. goossens, s. questiau, v. manceau, n. escaravage, l. waits, and j. bouvet. 1996. reliable genotyping of samples with very low dna quantities using pcr. nucleic acids research 24: 3189–3194. _____, l. p. waits, and g. luikart. 1999. noninvasive genetic sampling: look before you leap. trends in ecology and evolution 14: 323–327. tjepkes, t. l. 2015. genetic analysis of moose populations from minnesota and yellowstone national park. m. s. thesis, university of minnesota-duluth, duluth, minnesota, usa. http:// conservancy. umn.edu/handle/11299/177051. valière, n., c. bonenfant, c. toïgo, g. luikart, j. m. gaillard, and f. klein. 2007. importance of a pilot study for non-invasive genetic sampling: genotyping errors and population size estimation in red deer. conservation genetics 8: 69–78. waits, l., and d. paetkau. 2005. noninvasive genetic sampling tools for wildlife biologists: a review of applications and http://conservancy.umn.edu/handle/11299/177051 http://conservancy.umn.edu/handle/11299/177051 alces vol. 53, 2017 unger et al. – dna source types in moose 197 recommendations for accurate data collection. journal of wildlife management 69: 1419–1433. wehausen, j. d., r. r. ramey, and c. w. epps. 2004. experiments in dna extraction and pcr amplification from bighorn sheep feces: the importance of dna extraction method. journal of heredity 95: 503–509. wilson, g. a., c. strobeck, l. wu, and j. w. coffin. 1997. characterization of microsatellite loci in caribou rangifer tarandus, and their use in other artiodactyls. molecular ecology 6: 697–699. wilson, p. j., s. grewal, a. rodgers, r. rempel, j. saquet, h. hristienko, f. burrows, r. peterson, and b. n. white. 2003. genetic variation and population structure of moose (alces alces) at neutral and functional dna loci. canadian journal of zoology 81: 670–683. wilson, r. e., s. d. farley, t. j. mcdonough, s. l. talbot, and p. s. barboza. 2015. a genetic discontinuity in moose (alces alces) in alaska corresponds with fenced transportation infrastructure. conservation genetics 16: 791–800. scat-detection dogs survey low density moose in new york heidi kretser1, michale glennon1, alice whitelaw2, aimee hurt2, kristine pilgrim3, and michael schwartz3 1wildlife conservation society, north america program, 132 bloomingdale avenue, saranac lake, new york, usa 1298; 2working dogs for conservation, 52 eustis road, three forks, montana, usa 59752; 3united states forest service rocky mountain research station, national genomics center for wildlife and fish conservation, 800 east beckwith, missoula, montana 59801 abstract: the difficulty of collecting occurrence and population dynamics data in mammalian populations of low density poses challenges for making informed management decisions. we assessed the use of scat-detection dogs to search for fecal pellets in a low density moose (alces alces) population in the adirondack park in new york state, and the success rate of dna extraction from moose fecal pellets collected during the surveys. in may 2008, two scat-detection dog teams surveyed 20, 4-km transects and located 138 moose scats. in 2011 we successfully amplified dna from 39 scats (28%) and were able to uniquely identify 25 individuals. improved storage protocols and earlier lab analysis would increase the amplification success rate. scat-detection dogs proved to be a reasonable, non-invasive method to collect useful data from the low density moose population in the adirondack park. alces vol. 52: 55–66 (2016) key words: adirondack park, dna, fecal pellets, moose, new york, scat-detection dog. moose (alces alces) were nearly extirpated from the northeastern united states in the late 1800s, but have recently undergone natural recolonization in the region (alexander 1993, bontaites and gustafson 1993, wattles and destefano 2011) and the adirondack park in new york (hicks 1993, reeves and mccabe 1997, jenkins and keal 2004). moose have no natural predators in new york other than possibly black bears (ursus americanus) that prey upon neonatal calves, but concerns about over-browsing of regenerating forests, trampling of vacuum tubing in sugar maple (acer saccharum) stands, and the potential for moose to pose roadway hazards have prompted calls for a hunting season. the recent population decline in minnesota suggests that moose at the southern extent of their range may face thermoregulatory stress that could possibly translate to poor body condition, malnutrition, and energy loss making them more susceptible to parasites (lenarz et al. 2010). although state wildlife biologists recognize the need to understand their population dynamics and structure, moose in northern new york occur at low density in small, widely-scattered groups that challenge the collection of meaningful population data. moose biologists from the region met in 2003 to discuss potential research and management methods to study the low density population in the adirondack park (kretser et al. 2014). gps radio-collaring of 10 females, aerial surveys, deer hunter surveys, and other non-invasive approaches were presented as viable options for studying this low density population. at the time, cost of gps corresponding author: heidi kretser, wildlife conservation society, 132 bloomingdale avenue, saranac lake, ny 12983, hkretser@wcs.org 55 mailto:hkretser@wcs.org radio-collars and the logistical difficulty in capturing moose were considered prohibitively expensive, especially in this heavily forested region dominated by dense coniferous and mixed forest with minimal road access. deer hunter surveys began in 2005 with low participation rates, and flyovers occurred when helicopter availability and weather conditions aligned, albeit, not frequently enough to collect robust data. the wild center, a local natural history museum, offered pilot funding to test other non-invasive methods, specifically, using scat-detection dogs to collect population data. the initial objective of this study was to assess the use of scat-detection dogs to locate moose scat efficiently as a potential technique to estimate moose abundance in the dense forests of the adirondack park. measuring occurrence and abundance of a wide-ranging mammal at low density poses challenges for biologists desiring to make informed management decisions (mackay et al. 2008). at low abundance, the effort required to observe or capture individuals may exceed the resources available to obtain adequate and useful data. methods such as camera trapping and track stations can supply presence/absence information, and in some cases, information about population structure (e.g., identifying males, females, and juveniles in photographs); however, these methods do not produce dna samples. non-invasive techniques such as hair snares and scat sampling are often good alternatives for obtaining dna samples. hair snares work well in situations where bait and lure are used to attract an animal to the site (e.g., woods et al. 1999), or when surveying areas such as feeding sites or habitat features where species congregate or visit regularly (kendall and mckelvey 2008). scat collection does not require luring a species to a specific site, rather, an efficient means of locating scat in a natural setting. recent studies have used fecal dna to identify individuals, evaluate kinship, and describe distributions and sex ratios in wild populations (taberlet et al. 1997, lucchini et al. 2002, eggert et al. 2003, bellemain et al. 2005). because human detection of scats is challenging in a low density population, scat-detection dogs are often used to increase efficiency (smith et al. 2003, long et al. 2008). combining dna analysis with the use of scat detection dogs eliminates the need to capture, handle, or observe individual animals and minimizes the field time required to collect samples (kohn and wayne 1997, kohn et al.1999). obtaining dna from wild animals provides for a variety of uses and approaches to extract relevant population data. noninvasive genetic samples can be used in a population genetic framework to understand effective population size, gene flow, genetic diversity, and kinship across multiple populations (schwartz and monfort 2008). mitochondrial dna (mtdna) can be used to identify individual species (foran et al. 1997), nuclear dna (often using microsatellites) can identify individuals, and sex identification is possible by focusing on specific genes that determine gender (schwartz and monfort 2008). these data can be obtained from dna extracted from blood, tissue, hair, or scat. sampling of high quality template dna samples (i.e., tissue and blood) is often invasive, requiring physically handling animals which may entail high cost, physiological stress, and/or injury. to date, most research involving scatdetection dogs has focused on carnivores (long et al. 2008). we sought to assess the feasibility of using these dogs to search for moose fecal pellets in the adirondack park and to determine whether dna extraction from moose fecal pellets was feasible. several factors influence whether dna can be extracted from a scat sample including diet, environmental conditions at collection, storage methods, and the specific extraction 56 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 method. this method has been used successfully to empirically address a variety of questions about carnivores (smith et al. 2005, beckmann 2006), and other organisms ranging from right whales (eubalaena glacialis; rolland et al. 2006) to invasive plants (goodwin et al. 2010). ungulate scat has been successfully amplified in a different ecosystem where scat remained frozen throughout the study period (wasser et al. 2011). our two primary objectives were to 1) evaluate if scat-detection dogs could efficiently locate moose scat in a low density population in the adirondack park, and 2) to determine the efficacy of extracting dna from moose fecal pellets collected in this ecosystem. study area the adirondack park (park) in northern new york is a 24,000-km2 mountainous area with more than 3,000 lakes and ponds and 45,000 km of waterways (fig. 1). elevation ranges from 305-1671 m and the dominant forest types are northern hardwood, conifer, and boreal upland forests. northern hardwoods include american beech (fagus grandifolia), yellow birch (betula alleghaniensis), and sugar maple, with red spruce (picea rubens) balsam fir (abies balsamea) forests at higher elevations and rare alpine vegetation above 1500 m. more than 280 bird, mammal, amphibian, and reptile species inhabit the landscape, alongside 130,000 fulltime human residents in 103 rural communities. nearly half of the land within the park boundary is privately owned and managed; the public land is permanently protected from development by the new york state (nys) constitution. the local economy is based on year-round tourism, commercial forest industry (private land), and governmental services (jenkins and keal 2004). the adirondack park agency oversees and regulates activities on the privately-owned portions of the park, and management of the wildlife resources on both public and private land rests with the new york state department of environmental conservation (dec), including hunting and responding to humanwildlife conflicts. methods scat detection the wildlife conservation society (wcs) conducted a pilot test of scat-detection dogs in the northern adirondacks in partnership with working dogs for conservation, inc. (wdc, three forks, montana, usa). initially, wcs staff worked with local individuals to locate scat samples from multiple moose throughout the park. scats were collected on public and private lands including private parcels undergoing active or recent logging that provided contrast to the protected and unlogged state lands. we sent scat samples to the wdc for training dogs on scents associated with scats that reflected the diet of moose in the park; dogs were trained for 6 weeks followed by a 2-day in situ training. maps and aerial photographs were used to establish 20 line transects of ~4 km at each site prior to deployment of the dog team. in the challenging terrain, a 4-km transect was considered a reasonable distance for the team to traverse during a one-day session. each team included one dog handler, one orienteer, and one dog. each dog wore a gps unit attached to a work vest to track their movements; likewise, each orienteer carried a gps to track their movements. handlers kept the dogs under voice command within 100 m of the transect, and we measured both human and dog tracks by summing the distance between track points recorded every 15 sec. each transect was labeled with a unique identifier code and described by date, start time, end time, duration, dog and handler, orienteer, temperature, weather at the start of the transect, human track (km), and dog track (km). we recorded the number of alces vol. 52, 2016 kretser et al. – scat-detection dogs survey moose 57 moose scats, bear scats, and unknown scats; we included bear in the survey effort because both dogs were trained previously on black bear scat. orienteers used latex gloves and plastic zip-lock bags to collect and store scats in an attempt to maintain a sterile environment and reduce the potential for crosscontamination of samples. each scat was assigned a unique identification number and described by species, dog ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! !! ! ! _̂ _̂ _̂ _̂ tupper lake lake placid saranac lake lake george ingraham pond east mountain hardwood hill ragged mountain 16 6 3 scats located dna amplified individual moose identified 6 20 california road champion land 14 1 1 9 hatch brook 1 11 sporting hill 9 6 5 wolf pond mountain catamount mountain 15 12 6 base of whiteface vic moose pond grass pond 1 1 2 deer pond huntington cherry patch 12 1 1 south meadow 15 1 1 boreas i 15 9 7 boreas ii 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 fig. 1. map of transects surveyed by working dog crews and the number of scats detected, amplified, and identified as individual moose at each transect. dotted line denotes straight line path connecting three transects (wolf pond, champion, and boreas ii) where the same individual moose was identified. inset: location of adirondack park within new york state. 58 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 or human found, date, time, zone, easting, northing, elevation in feet, canopy (i.e., open (>50%) or closed), forest type (i.e., hardwood, softwood, or mixed), and water characteristic of the site (i.e., wetland, within 100 m of a wetland, or upland). we assigned each scat to 1 of 3 condition categories: 1) excellent – pellets well-formed, moist or wet, and dark color (described as “fresh” in smith et al. 2003, mondol et al. 2009), 2) good – pellets becoming unformed and starting to have discoloration, or 3) poorpellets not formed, dry, light in color, sometimes moldy. scats were stored in ziploc bags in the field, subsequently transferred to 5 ml plastic vials and covered with ethanol, and stored in a cool dark basement; duplicates were frozen in ziploc bags. extremely moist samples were kept in open, brown paper bags (9×16 cm) to air dry (franzen et al. 1998, piggott and taylor 2003) for 24–48 h prior to storage (maudet et al. 2004). scats in ethanol were submitted to the united states forest service rocky mountain research station wildlife genetics laboratory (rmrs) in march 2010. dna extraction using moose reference samples from the northeast, we optimized a panel of 9 variable microsatellites in an attempt to uniquely identify individuals from the park (wilson et al. 2003, schmidt et al. 2008, wilson pers. commun.). reference blood and tissue samples were obtained from collections and sampling of harvested and vehicular-killed moose from 4 northeastern states (new york, new hampshire, maine, and vermont) and 4 canadian provinces (new brunswick, ontario, nova scotia, and quebec). the nova scotia samples included moose from the mainland and cape breton island. we performed an initial dna extraction on 140 fecal pellet samples (stored in ethanol) using a standard protocol developed for ungulate fecal pellets (maudet et al. 2004, schwartz et al. 2007). we then performed a dna extraction on these samples with a pellet swab to test the efficacy of this approach; after two failed attempts, we repeated this process on frozen duplicate samples (n = 42). dna from reference samples was amplified at the following 8 microsatellite loci: nvhrt21, bm1225, bm4516, fcb193, map2c, rt5, rt9, and rt30 (wilson et al. 2003, schmidt et al. 2008, wilson pers. commun.). the reaction volume (10 μl) contained 1.0 μl dna, 1× reaction buffer (applied biosystems), 2.0 mm mgcl2, 200 μm of each dntp, 1 μm reverse primer, 1 μm dyelabeled forward primer, 1.5 mg/ml bsa, and 1u taq polymerase (applied biosystems). the pcr profile was (94 °c/5 min, 94 °c/1 min, 55 °c/1 min, 72 °c/30 s) × 45 cycles. the resultant products were visualized on a li-cor dna analyzer (li-cor biotechnology). all non-invasive samples were initially amplified twice using the multitube approach (eggert et al. 2003, schwartz et al. 2004), and allele scores were entered only when consistent for both amplifications. microsatellite data were checked for genotyping errors (false alleles, allelic dropout and scoring errors) using the program dropout (mckelvey and schwartz 2005, schwartz et al. 2006). microsatellite data was also errorchecked with the program micro-checker (van oosterhout et al. 2004) to identify loci with possible genotyping errors leading to homozygote excess. we calculated the probability of identity (pid) and probability of identify given siblings (psib) from these samples. we used chi-square with a cochran’s test of linear trend to assess the relationship of scat condition, forest type, and wetland proximity to successful dna extraction. we then evaluated all possible general linear model (glm) combinations of these factors in systat with akaike information alces vol. 52, 2016 kretser et al. – scat-detection dogs survey moose 59 criterion (aic) to assess their relative importance on our ability to successfully extract dna. results the 20 transects were sampled on 10 field days in may 2008. dogs located 191 scats and 4 additional scats were located by the orienteers; 134 (69%) were moose, 56 (29%) bear, and 5 (2%) were unknown. the proportions of moose scat relative to condition were 18% excellent, 45% good, and 36% poor; 63 and 37% of scats were collected on private and state lands, respectively. dogs traveled 134 km and orienteers 114 km during the surveys. we genotyped 270 tissues and hair samples from 10 locations at 8 variable microsatellite loci (table 1). we obtained quality multi-locus genotypes from 28 pellet samples using the initial dna extraction from fecal pellets. the swabbing method yielded quality dna from 9 additional pellet samples, and the duplicate samples yielded 2 additional samples. dna was successfully amplified from 28% (39 of 137) of collected scats using the 8 loci. errors were identified in 3 samples at loci rt9 and nvhrt21. dna from these samples was reanalyzed following the approach of schwartz et al. (2006) until no errors occurred. we identified 25 unique individuals. there was a 1 in 9,737 chance of identifying 2 individuals as identical (pid = 1.03×10−4), and a 1 in 64 chance of identifying 2 siblings as identical (psib = 1.57×10−2). in 8 cases we identified the same moose from multiple scat, and one moose was identified on 3 different transects located >40 miles distant (fig. 1). successful amplification was achieved in 52% of excellent, 31% good, and 12% of poor scats; of the total amplified sample (n = 38), 34% were excellent, 50% good, and 16% were poor scats (table 2). scat condition (χ2 = 7.928, p < 0.001) and forest type (χ2 = 7.928, p < 0.05) affected our ability to successfully extract dna; excellent scats and scats located in hardwoods had the highest success rates. location was not related to successful extraction (table 2). the glm table 1. origin and number of tissue samples received and analyzed to create markers for use in the dna extraction of moose fecal scats, adirondack park, new york. country location received used us new york 16 16 us new hampshire 30 30 us maine 41 41 us vermont 31 31 ca nova scotia – mainland 31 29 ca nova scotia – cape breton island 9 9 ca new brunswick 46 46 ca ontario 29 26 ca quebec – rfpl 13 13 ca quebec – plc 29 29 total 275 270 table 2. the rate (%) of successful extraction of dna (yes = 38, no = 99) from moose scat relative to scat condition, forest type, and microhabitat location in the adirondack park, new york. sample sizes in parentheses. variable yes no scat condition* excellent (25) 52% 48% good (62) 31% 69% poor (50) 12% 88% forest type† hardwood (74) 36% 50% softwood (45) 22% 78% mixed (18) 6% 94% location‡ immersed in water (2) 50% 50% near water (24) 21% 79% upland (111) 29% 71% *χ2= 13.782, p < 0.001. †χ2= 7.928, p = 0.019. ‡χ2= 1.131, p = 0.568. 60 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 underscored the relative importance of scat condition and forest types. the top four glms included scat condition, with the top model indicating that 69.7% of model weight was associated with scat condition and forest type (table 3). discussion we demonstrated that scat-detection dogs were effective at locating moose scat in a low density moose population in the dense forests of the adirondack park. scatdetection dogs are more frequently used in carnivore research because of their obvious advantage in sampling wide-ranging and low abundance populations (mackay et al. 2008); their use in ungulate research is less common. wasser et al. (2011) used these dogs to locate a variety of species including moose and woodland caribou (rangifer tarandus caribou) in areas proximal to the alberta tar sands, and successfully extracted dna from scats of both. similarly, dogs located scats of a variety of deer species (mazama spp.) in brazil and outperformed human searchers; humans located zero scats whereas dogs located 0.21 scats/km (de olivera et al. 2012). the success rate of our dogs was ~1.4 samples/km, confirming our supposition that dogs could efficiently ‘sample’ a low density moose population which cannot be easily observed/sampled. although we successfully amplified dna from moose fecal pellets, our success rate was relatively low (<30%) but was explained by pellet condition and location. age and environmental factors (e.g., precipitation, temperature) affect the quality of collected scats (brinkman et al. 2010), and these factors also affect detection rate (reed et al. 2011). these factors undoubtedly affected the quality of our samples as we collected scat in the spring, relatively soon after snowmelt. environmental conditions are different within the 3 forest types, with mixed and softwood stands moister at the forest floor which would presumably degrade scat faster; in fact, extraction rates were higher in scats collected in hardwood forest (table 2). collection of fresh scats across seasons and repeat sampling would improve our sampling protocol. for example, in an area of winter concentration of moose, repeat sampling would minimize exposure of fecal pellets to the elements (see brinkman et al. 2010). future work may also take advantage of in situ photographs of scats to compare condition across sites, and relate condition to amplification success more objectively. storage methods and storage time had strong influence on our success rate of dna amplification. we stored scats in ethanol based on the best information available at the time, and although a common storage method, it was not ideal in our study. researchers examining relative success rates associated with various storage media hesitate to provide overall table 3. model selection results using aic for general linear models to predict successful dna extraction from moose scat collected in the adirondack park, new york. model rank variables aic δaic aic weights 1 scat condition + forest type 157.53 0.00 0.697 2 scat condition + forest type + water 159.98 2.45 0.204 3 scat condition 162.07 4.54 0.072 4 scat condition + water 164.51 6.98 0.021 5 forest type 168.43 10.90 0.003 6 forest type +water 168.65 11.12 0.003 7 water 175.46 17.93 0.000 alces vol. 52, 2016 kretser et al. – scat-detection dogs survey moose 61 recommendations because of inconsistency among studies (schwartz and monfort 2008); however, ethanol was considered the worst storage medium among 3 alternatives (soto-calderón et al. 2009). the relatively long storage time of our samples (2 years) was probably the major reason for our low success rate with amplification; however, we still recovered dna from ~25% of the highly variable sample and >50% of excellent pellets. schwartz and monfort (2008) suggest processing samples immediately because dna is more stable in laboratory buffers than fecal material, but the logistical issues of fieldwork would often preclude this approach. however, dna swabbed from moose fecal pellets and processed within a few weeks of collection yielded high amplification rates (90%; k. pilgrim, rmrs, unpublished data). field swabbing of fresh samples also results in higher amplification rates compared to swabbing frozen samples (rutledge et al. 2009). scat-detection dogs offer many advantages but with certain considerations. cost is a concern for any field-based project and substantial costs are associated with dog and handler selection and training, as well as field time for executing transects (mackay et al. 2008). in this study we spent $25,000 to hire wdc to find 191 scats, at a per-unit cost of ~$130 per sample. numerous scats are required to address population studies and depending upon animal density, may require substantial field time and cost. researchers must understand and work within the physical limitations of the dog and recognize that detection rates often vary among dog/handler teams. the handler must ensure that the dog focuses on the desired scat and is not inadvertently trained to non-target scat; this is particularly salient for handlers when target and non-target species have morphologically similar fecal pellets. lastly, this method may result in real or perceived potential conflict with wildlife (mackay et al. 2008), and the presence of dogs in a given environment may result in unforeseen conflicts with local wildlife. despite certain limitations, there are numerous and obvious advantages in using scat-detection dogs. in comparison to human searchers, dogs are highly efficient and effective at locating scats (de olivera et al. 2012), and in this study, covered ~20% more ground than humans and collected all but 4 scats. dogs create minimal sampling bias (mackay et al. 2008), allowing for quick confirmation of occupancy of the study area by target species. collection of scats ultimately provides for discrimination between species and individuals, and has proven applicable to a wide variety of species and habitat types. collection of scat not only allows for subsequent assessment of population structure, it also provides opportunities to explore additional factors such as stress levels and diet. the charismatic and broad public appeal of using dogs should not be discounted as an opportunity for public outreach and engagement (mackay et al. 2008, woollett et al. 2014). one of the lasting impacts of using scat-detection dogs was the creation of two high definition videos describing the project and highlighting the dogs in the field. according to the wild center staff, these two films continue to capture audiences largely due to the appeal of the dogs performing in the field. our pilot research in the adirondack park of new york state is one of a limited number of studies in which scat-detection dogs have been used in ungulate research, and these dogs provided a viable method for sampling a low density moose population. we also found that forest type, the condition of fecal pellets, storage method, and storage time influenced the efficacy of dna amplification. the impact of these factors can be controlled through improved study design that addresses temporal sampling, field swabbing, shorter storage time, and performing 62 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 extractions soon after sample collection. in particular, we recommend swabbing the scats using synthetic swabs (e.g., dacron swabs), at least two swabs per sample, and storing them dry in envelopes or vials. fecal pellets can be stored in vials of 95% ethanol at room temperature, frozen in secure plastic bags, or air dried and kept at room temperature. ideally, the swabs and scats would be submitted for dna extraction and analyses within a few days of collection. this approach works for a variety of carnivores and would be an improved protocol for moose research (reed et al. 2004, mckelvey et al. 2006, rutledge et al. 2009, anwar et al. 2011). our data and that collected in subsequent and future surveys provide an important foundation to understand habitat use and population dynamics of moose in the park, and conduct genetic research to determine relationships among park and regional moose. given the increased interest and funding available for moose research in new york state, we encourage continued use of scatdetection dogs, in concert with other techniques, to monitor and study the low density moose population in the adirondack park. acknowledgements funding for this project was provided by the wild center, the northeastern states research cooperative, and the new york state department of environmental conservation. we thank e. reed, new york department of environmental conservation, for providing transect site maps, our orienteers g. lee and b. kitchen, and many individuals who provided access to private lands and assisted with the data collection and general field work details: upland forestry, p. bogdanovich, the nature conservancy, m. carr, j. fogarty, and s. moody and a. brown. finally, to moose biologists in the northeast who assisted with collection of tissue samples for dna markers: k. hynes, new york department of environmental conservation; c. alexander, vermont department of fish and wildlife; k. rines, new hampshire fish and game department; l. kantar, maine department of inland fish and wildlife; a. r. rodgers – ontario ministry of natural resources; p. wilson – trent university, ontario; c. dussault, d. jean, s. lefort, and b. beaudoin – quebec ministère des ressources naturelles et de la faune; k. craig and d. sabine new brunswick department of natural resources; s. mcburney and t. nette nova scotia dnr and the canadian cooperative wildlife health unit. references alexander, c. e. 1993. the status and management of moose in vermont. alces 29: 187–195. anwar, m. b., r. jackson, m. s. nadeem, j. e. janečka, s. hussain, m. a. beg, g. muhammad, and m. qayyum. 2011. food habits of the snow leopard panthera uncia (schreber, 1775) in baltistan, northern pakistan. european journal of wildlife research 57: 1077–1083. beckmann, j. p. 2006. carnivore conservation and search dogs: the value of a novel, non-invasive technique in the greater yellowstone ecosystem. pages 28–34 in a. wondrak biel, editor. greater yellowstone public lands: a century of discovery, hard lessons, and bright prospects. proceedings of the 8th biennial scientific conference on the greater yellowstone ecosystem. yellowstone center for resources, yellowstone national park, wyoming, usa. bellemain, e., j. e. swenson, d. tallmon, s. brunberg, and p. taberlet. 2005. estimating population size of elusive animals with dna from huntercollected feces: four methods for brown bears. conservation biology 19: 150–161. alces vol. 52, 2016 kretser et al. – scat-detection dogs survey moose 63 bontaites, k., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29: 163–167. brinkman, t. j., m. k. schwartz, d. k. person, k. l. pilgrim, and k. j. hundertmark. 2010. effects of time and rainfall on pcr success using dna extracted from deer fecal pellets. conservation genetics 11: 1547–1552. de oliveira, m. l., d. norris, j. f. m. ramírez, p. h. de perez, m. galetti, and j. m. b. duarte. 2012. dogs can detect scat samples more efficiently than humans: an experiment in a continuous atlantic forest remnant. zoologica 29: 183–186. eggert, l. s., j. a. eggert, and d. s. woodruff. 2003. estimating population sizes for elusive animals: the forest elephants of kakum national park, ghana. molecular ecology 12: 1389–1402. foran, d. r., k. r. crooks, and s. c. minta. 1997. species identification from scat: an unambiguous genetic method. wildlife society bulletin 25: 835–839. frantzen, m. a. j., j. b. silk, j. w. h. ferguson, r. k. wayne, and m. h. kohn. 1998. empirical evaluation of preservation methods for faecal dna. molecular ecology 7: 1423–1428. goodwin, k. m., r. e. engel, and d. k. weaver. 2010. trained dogs outperform human surveyors in the detection of rare spotted knapweed (centaurea stoebe). invasive plant science management 3: 113–121. hicks, a. c. 1993. using road-kills as an index to moose population change. alces 29: 243–247. jenkins, j. c., and a. keal. 2004. the adirondack atlas. syracuse university press, syracuse, new york, usa. kendall, k. c., and k. s. mckelvey. 2008. hair collection. pages 141–182 in r. a. long, p. mackay, w. j. zielinski, and j. c. ray, editors. non-invasive survey methods for carnivores. island press, washington d. c., usa. kohn, m. h., and r. k. wayne. 1997. facts from feces revisited. trends ecological evolution 12: 223–227. ———, e. c. york, d. a. kamradt, g. haught, r. m. sauvajot, and r. k. wayne. 1999. estimating population size by genotyping faeces. proceedings of the royal society b: biological sciences 266: 657–663. krester, h. e., m. g. glennon and z. p. smith. 2014. ngos enhance state wildlife agencies’ capacity to meet public trust doctrine obligations. human dimensions of wildlife 19: 437–447. lenarz, m. s., j. fieberg, m. w. schirage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. long, r. a., p. mackay, w. j. zielinski, and j. c. ray. 2008. non-invasive survey methods for carnivores. island press, washington d. c., usa. lucchini, v., e. fabbri, f. marucco, s. ricci, l. boitani, and e. randi. 2002. noninvasive molecular tracking of colonizing wolf (canis lupus) packs in the western italian alps. molecular ecology 11: 857–868. mackay, p., d. a. smith, r. a. long, and m. parker. 2008. scat detection dogs. pages 182–222 in r. a. long, p. mackay, w. j. zielinski, and j. c. ray, editors. non-invasive survey methods for carnivores. island press, washington d. c., usa. maudet, c., g. luiket, d. dubray, a. von hardenberg, and p. taberlet. 2004. low genotyping error rates in wild ungulate faeces sampled in winter. molecular ecology notes 4: 772–775. mckelvey, k. s., and m. k. schwartz. 2005. dropout: a program to identify problem loci and samples for noninvasive genetic samples in a capture64 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 mark-recapture framework. molecular ecology notes 5: 716–718. ———, j. von kienast, k. b. aubry, g. m. koehler, b. t. maletzke, j. r. squires, e. l. lindquist, s. loch, and m. k. schwartz. 2006. dna analysis of hair and scat collected along snow tracks to document the presence of canada lynx. wildlife society bulletin 34: 451–455. mondol, s., k. u. karanth, n. s. kumar, a. m. gopalaswamy, a. andheria, and u. ramakrishnan. 2009. evaluation of non-invasive genetic sampling methods for estimating tiger population size. biological conservation 142: 2350–2360. piggott, m. p., and a. c. taylor. 2003. extensive evaluation of faecal preservation and dna extraction methods in australian native and introduced species. australian journal of zoology 51: 341–355. reed, j. e., r. j. baker, w. b. ballard, and b. t. kelly. 2004. differentiating mexican gray wolf and coyote scats using dna analysis. wildlife society bulletin 32: 685–692. ———, a. l. bidlack, a. hurt, and w. m. getz. 2011. detection distance and environmental factors in conservation detection dog surveys. journal of wildlife management 75: 243–251. reeves, h. m., and r. e. mccabe. 1997. of moose and man. pages 1–75 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north america moose. smithsonian institution press, washington d. c., usa. rolland, r. m., p. k. hamilton, s. d. kraus, b. davenport, r. m. gillett, and s. k. wasser. 2006. faecal sampling using detection dogs to study reproduction and health in north atlantic right whales (eubalaena glacialis). journal of cetacean research management 8: 121–125. rutledge, l. y., j. j. holloway, b. r. patterson, and b. n. white. 2009. an improved field method to obtain dna for individual identification from wolf scat. journal of wildlife management 73: 1430–1435. schmidt, j. i., k. j. hundertmark, r. t. bowyer, and k. g. mccracken. 2008. population structure and genetic diversity of moose in alaska. journal of heredity 100: 170–180. schwartz, m. k., s. a. cushman, k. s. mckelvey, j. hayden and c. engkjer. 2006. detecting genotyping errors and describing american black bear movement in northern idaho. ursus 17: 138–148. ———, g. luikart, r. s. waples. 2007. genetic monitoring as a promising tool for conservation and management. trends in ecological evolution 22: 25–33. ———, and s. l. monfort. 2008. genetic and endocrine tools for carnivore surveys. pages 238–262 in r. a. long, p. mackay, w. j. zielinski, and j. c. ray, editors. non-invasive survey methods for carnivores. island press, washington d. c., usa. ———, k. l. pilgrim, k. s. mckelvey, e. l. lindquist, j. j. claar, s. loch, and l. f. ruggiero. 2004. hybridization between canada lynx and bobcats: genetic results and management implications. conservation genetics 5: 349–355. smith, d. a., k. ralls, b. l. cypher, and j. e. maldonado. 2005. assessment of scat-detection dog surveys to determine kit fox distribution. wildlife society bulletin 33: 897–904. ———, ———, a. hurt, b. adams, m. parker, b. davenport, and j. e. maldonado. 2003. detection and accuracy rates of dogs trained to find scats of san joaquin kit foxes (vulpes macrotis mutica). animal conservation 6: 339– 346. soto-calderón, i. d., s. ntie, p. mikala, f. maisels, e. j. wickings, and n. m. anthony. 2009. effects of storage type and time on dna amplification success alces vol. 52, 2016 kretser et al. – scat-detection dogs survey moose 65 in tropical ungulate faeces. molecular ecological resources 9: 471–479. taberlet, p., j. j. camarra, s. griffin, e. uhres, o. hanotte, l. p. waits, c. dubois-paganon, t. burke, and j. bouvet. 1997. noninvasive genetic tracking of the endangered pyrenean brown bear population. molecular ecology 6: 869–876. van oosterhout, c., w. f. hutchinson, d. p. wills, and p. shipley. 2004. micro-checker: software for identifying and correcting genotyping errors in microsatellite data. molecular ecology notes 4: 535–538. wasser, s. k., j. l. keim, m. l. taper, and s. r. lele. 2011. the influences of wolf predation, habitat loss, and human activity on caribou and moose in the alberta oil sands. frontiers of ecological environment 9: 546–551. wattles, d.w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. wilson, p. j., s. grewal, a. rodgers, r. rempel, j. saquet, h. hristienko, f. burrows, r. peterson, and b. n. white. 2003. genetic variation and population structure of moose (alces alces) at neutral and functional dna loci. canadian journal of zoology 81: 670–683. woods, j. g., d. paetkau, d. lewis, b. n. mclellan, m. proctor, and c. strobeck. 1999. genetic tagging freeranging black and brown bears. wildlife society bulletin 27: 616–627. woollett (smith), d. a., a. hurt, and n. l. richards. 2014. the current and future roles of free-ranging detection dogs in conservation efforts. pages 239–264 in m. e. gompper, editor. free-ranging dogs and wildlife conservation. oxford university press, cambridge, england. 66 scat-detection dogs survey moose – kretser et al. alces vol. 52, 2016 scat-etection dogs survey low density moose in new york study area methods scat detection dna extraction results discussion acknowledgements references alces15_32.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces16_429.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 recruitment of winter ticks (dermacentor albipictus) in contrasting forest habitats, ontario, canada e. m. addison1,2, r. f. mclaughlin3, p. a. addison4, and j. d. smith5 1wildlife research and monitoring section, ontario ministry of natural resources and forests, 2140 east bank drive, peterborough, ontario, canada k9j 7b8; 2present address: ecolink science, 107 kennedy st. w., aurora, ontario, canada l4g 2l8; 3r.r. #3, penetanguishene, ontario, canada l0k 1p0; 4northwest region, regional operations division, ontario ministry of natural resources and forests, 173 25th sideroad, rosslyn, ontario, canada p7k 0b9; 558 chemin rheaume, val des monts, quebec, canada j8n 6l5. abstract: recruitment of winter tick larvae (dermacentor albipictus) was studied in a forest opening and a closed canopy deciduous forest to evaluate their potential as sources of tick infestation to moose (alces alces). engorged female ticks were set out in early may at each site and monitored to measure the proportions of females producing larvae and the number of larvae recruited per g of surviving female. recruitment was higher in the forest during the hotter, drier summer of 1983, primarily due to fewer engorged females producing larvae in the opening, and was much higher (>2 x) in the opening during the cooler, damper summer of 1984. recruitment in the field was 20–40% of that under laboratory conditions. desiccation of eggs and/or larvae was the probable cause for the annual variation in recruitment in the opening. most larvae were recruited earlier in the opening than in the forest site. neither weight nor date of detachment of engorged female ticks influenced when larvae first ascended vegetation. weather, especially temperature, and site structure and composition affect abundance of the free-living stages of the winter tick and larvae available for transmission to moose. open sites should support more winter tick larvae than densely forested sites except in years of particularly hot and dry weather. alces vol. 52: 29–40 (2016) key words: dermacentor albipictus, winter tick, moose, recruitment, weather, habitat, alces the overall impact of winter ticks (dermacentor albipictus) on moose (alces alces) populations varies annually (samuel 2004). understanding the reasons for this variation is key to predicting when and in what habitats moose are most likely to acquire winter ticks. survival of eggs and larvae of ixodid ticks are affected most markedly by desiccation (sonenshine 1970), and since temperature and moisture vary annually, reproductive potential of ixodid ticks can also vary annually within the same habitat (patrick and hair 1979, fleetwood et al. 1984). production and survival of winter tick eggs and larvae from wapiti (cervus canadensis) have been studied in 2 habitats in oklahoma, usa (patrick and hair 1975), and from moose in a variety of habitats in central alberta, canada (drew and samuel 1986, aalangdong et al. 2001). objectives of this study were to establish if adult females, eggs, and larvae of the winter tick could survive in 2 contrasting forest habitats in ontario, canada, and to document factors related to variation in recruitment between years and habitats. study area two study sites 13.3 km apart in algonquin provincial park, ontario were selected for their contrasting habitats: 1) a forest opening and 2) a closed canopy deciduous forest. the opening site was an old lumber camp 29 (45° 42′ 13″ n, 78° 15′ 9.5″ w) with ground vegetation <0.3 m high and comprised of scattered grasses, sweet fern (comptonia peregrina), bracken (pteridium aquilinum), and blueberry (vaccinea sp.); trees were limited to red pine (pinus resinosa) and white spruce (picea glauca) <3 m high and present in 11% of plots. the deciduous forest site (45° 35′ 6″ n, 78° 18′ 53″ w) was on a northeasterly aspect sloping away from the summer sun, with a closed canopy dominated by sugar maple (acer saccharum) with scattered white birch (betula papyrifera). sub-canopy vegetation was limited and ground vegetation was predominantly sugar maples 1–1.5 m high. methods weather data were collected from “an historical climate analysis, version 2.2” which uses latitude and longitude, environment canada data, and topographical source data to calculate location specific weather (cross et al. 2012). specific weather data (e.g., temperature, precipitation, snow accumulation) were recorded occasionally at the study sites. we attempted to minimize and control for the influence of outside factors that could affect the outcomes of live specimens. specifically, we: 1 divided into 2 treatment groups winter ticks that detached on the same date from moose with similar exposure history and infestation level to account for immunocompetence that can suppress fecundity in ticks (mcgowan et al. 1980, 1981, chiera et al. 1985), 2 allocated similar-sized engorged ticks between treatment groups to account for the relationship between weight and fecundity (drummond et al. 1969, addison and smith 1981, drew and samuel 1987, addison et al. 1998), and further, calculated fecundity relative to weight, 3 used control plots and documented dissemination from points of deposition to account for ingress and egress of adult female and larval ticks, 4 assumed that losses to predation and disease were constant (addison et al. 1989, tuininga et al. 2009), and 5 used sites where the height of vegetation conformed with the known ascension height of winter ticks (0.5–1.9 m; drew and samuel 1985, mcpherson et al. 2000). definitions the following terms are used in the paper: 1 reproductive efficiency index (rei) is the number of eggs produced per g of engorged adult female tick (ef) (drummond and whetstone 1970). 2 larval efficiency index (lei) is the number of larvae recovered from a plot per g of ef placed on that plot, and was calculated only for plots from which larvae were recovered. 3 flagging is the action of dragging flannel sheets over vegetation to sample larvae available for transmission. 4 recruitment is a measure of tick larvae available for transmission by flagging and is comprised of both lei and proportion of efs producing larvae. 5 season of transmission is the time from first to last collection of larvae by flagging. 6 minimum free-living period is time (d) from detachment of an ef to first recovery of larvae from that ef by flagging. 7 vapor pressure deficit (vpd) is a measure of the difference (or deficit) between the pressure (mm hg) exerted by the moisture currently in the air and the pressure at saturation. engorged female ticks (efs) detached efs were collected on 1–20 april, 1983 and 15 march–27 april, 1984 from captive moose experimentally infested with ticks (see mclaughlin and addison 1986). prior to placement in the field, they were maintained at ambient temperature within cardboard boxes containing moist sand and soil overlaid with leaves. a total of 100 efs were marked on the abdomen with nail polish (revlon nail enamel, ottawa, canada k1g 3n1), and one set of 50 was placed at flagged locations in the 30 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 forest and opening sites in early may. their dispersal distances were measured in early july. an egg group (3.775 g) equivalent to ~54,000–63,000 eggs (see addison et al. 1998) was deposited at a single point in 2 grass plots in the opening site. from point of deposition, concentric 1-m wide rings were flagged 4 m outward to measure dispersal of larvae in late september through midoctober. fecundity and hatching forty efs were selected at random and each placed in a 2 cm2 cell within wood frame trays approximately 3 cm deep, enclosed on top and bottom by insect screening. one tray was placed on the soil surface and covered with ground vegetation in each study site in early may. eggs were collected from individual efs and counted weekly from 8 june until deposition ended. fifty efs were placed in each of 6, 60 cm2 wooden trays with fine screening on top and bottom. one tray was placed in each of 3 locations within each study site during mid-may and examined weekly to document initial presence of larvae. larval recruitment in 1983 (5–7 may), 453 and 451 efs were placed on the ground singly, 5–10 m apart, in the opening and forest sites, respectively. numbered plastic tape identified each plot with date of detachment, moose of origin, and weight of the ef; placement location was the center of a 2.5 m radius plot that was the source of data for larval recruitment. in 1984 (3–4 may), 213 and 206 efs were placed similarly in the opening and forest sites, respectively. control locations (without ticks) were also established 10 in the opening and 8 in the forest site. plots were flagged (sampled) with white, flannel sheets (2.3�2.6 m) every 9–13 (1983) and 11–15 (1984) days from august until after snowfall (3–22 november). one end of the flannel sheet was wrapped around a stick (2.4�0.05�0.02 m) until about 30 cm of the sheet was rolled in. the plot was flagged by dragging the sheet over the vegetation such that no person walked within 2.5 m of the plot center; the center area (~0.5 m out from the center point) was flagged once separately. the sheet was examined and if few larvae were present, they were counted, killed, and removed. if numerous larvae were collected, the sheet was folded by 2 people, bagged, and labeled. larvae were subsequently counted as they were evacuated (�40 kpa relative pressure) from the sheet. sheets were washed and dried in an electric dryer prior to reuse. in 1983, all plots were flagged and larvae counted through to the end of the season. plots were considered positive (i.e., efs had survived) if >10 larvae were recovered. in 1984, plots were identified as positive or negative in mid-september and a random subsample of positive plots was used for further measurements of total larval count. survival was measured by flagging these and the remaining plots and categorizing them positive or negative. data analysis placement of efs on plots and subsequent flagging of plots to establish recruitment occurred in both 1983 and 1984. timing of incubation of eggs, presence of larvae on control plots, and dispersal of efs and larvae were studied only in 1984. the r statistical package (r core team 2013) was used to analyze data on eggs, rei, and duration of the non-parasitic period. egg production and rei data were not normally distributed (shapiro-wilk test) and variances were not homogenous. since transformations did not normalize data, the mann-whitney-wilcoxon test was used to test for differences in number of eggs alces vol. 52, 2016 addison et al. – recruitment of winter tick larvae 31 produced and rei between habitats. duration of the non-parasitic period was tested for normal distribution with the shapirowilk test. transformations harmonized the variances (bartlett’s test) and while normal distributions were not realized, the consequences of this were acceptable to use factorial anova (glass et al. 1972, harwell et al. 1992, lix et al. 1996). three-way contingency tables and chi-square tests were used to evaluate frequency of positive plots between years and habitats (zar 1982). square root transformation was applied to the number of larvae produced per g for efs that produced larvae (lei); this transformation stabilized the variance (bartlett’s test) and normalized the distribution (kolmogorov-smirnov test). data for the first appearance of larvae on vegetation, size of efs, and date of detachment of efs could not be normalized through transformation because of the discontinuous nature of the sampling protocol both years: 9–13 and 11–15 days in 1983 and 1984, respectively (fig. 4, 5). consequently, scatter diagrams were used to evaluate possible relationships. results average monthly temperature and precipitation varied annually (table 1). may was wetter, june and july (the period of egg-laying and incubation) were warmer and drier, and september (period of larvae ascending) was warmer and wetter in 1983 than 1984. temperature was intermittently low and snow fell late in the experiment. for example, in 1983 the morning air temperature at the forest site was �6 °c and the ground was solid with heavy frost on 15 october, 10–15 cm of snow accumulated in the forest on 8 november and was gone by 10 november, and finally 10–15 cm of snow permanently covered the ground on 22 november. dispersal and fecundity mobility was limited for both efs and larvae. in the opening, 44 of 45 efs were recovered within 67 cm, and in the forest 41 of 42 were recovered within 34 cm of the site of deposition. in the opening, 98.5% of 26,485 larvae moved <2 m from the site of egg deposition. a similar number of efs held in trays laid eggs in the opening (64%) and forest (70%) in 1984. however, efs in the opening laid more (p < 0.05) eggs (median = 6226) than in the forest (median = 3632), and the rei in the opening (median = 8150) was greater (p < 0.05) than that in the forest (median = 4620). further, efs in the opening completed ovipositing on 27 july, a month earlier than in the forest (30 august). larval recruitment larvae hatched about 2 weeks earlier in the opening (4 august) than the forest (17 august) in 1984, were recruited earlier in the opening than the forest both years, but not until 2–4 weeks post-hatch in 1984 (table 2, fig. 1). most opening plots became positive for larvae later in 1983 than 1984 (fig. 1); in contrast, forest plots were positive at the same time both years (fig. 1). approximately 80% of all larvae were recruited in the opening by late september, about 2 weeks prior to the forest (fig. 2). minimal numbers table 1. average monthly temperatures and precipitation at opening and forest study sites in algonquin park, ontario, 1983–1984. temperature (°c) precipitation (mm) month 1983 1984 1983 1984 may 8.8 9.1 167.5 116.3 june 16.8 16.6 51.9 83.7 july 19.6 18.0 62.4 102.8 august 18.8 18.7 97.8 96.6 september 14.7 11.7 104.2 87.7 october 6.3 8.5 112.9 64.1 32 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 of larvae were collected during flagging in early to mid-november both years (figs. 2, 3). in the opening, <13 larvae were recovered in 9 of 10 control plots; 1 plot yielded 5,283 larvae. in the forest, 2 control plots had <10 larvae and 6 plots had none. a higher proportion of efs produced larvae in the forest than the opening in 1983 (p < 0.05); no difference was found in 1984 (p > 0.05) (table 2). at both sites fewer plots yielded larvae in 1983 (p < 0.05) (table 2). recruitment of larvae ranged from 10– 7,347 larvae per surviving ef. the mean lei was similar (p > 0.05) both years in the forest; however, the lei in the opening was higher (>2x) in the cool, wet summer of 1984 than the hot, drier summer of 1983, as well as both years in the forest (p < 0.05) (table 2, fig. 3). in both habitats in 1983, lei from efs from which larvae were available in september (early samples) was higher (p < 0.05) than from efs from which larvae were first recovered in october (late samples) (table 2, fig. 3). minimum free-living period the minimum free-living period ranged from 122–215 days (n = 699, μ = 162). effects of habitat, year, and the interaction between habitat and year on the minimum free-living period were significant (p < 0.05), but neither weight nor date of detachment of efs influenced duration of the freeliving period (fig. 4, 5). discussion winter tick larvae were available in both contrasting habitats, hence, are likely available for transmission in most terrestrial habitats frequented by moose in the great lakes – st. lawrence forest ecosystem. annual differences in fecundity and recruitment of larvae and between habitats reflected the influence of weather, habitat structure and composition, and their interactions. many tick studies have documented higher temperature, lower rh, and higher vpd in open (e.g., fields) compared to forest habitats during summer; e.g., in alberta (aalangdong et al. 2001), oklahoma (patrick and hair 1975), virginia (sonenshine 1970), eastcentral texas (fleetwood et al. 1984), and connecticut (bertrand and wilson 1996). wind also has a drying effect, open habitats being usually drier than closed habitats (schütte and king 1965), and sites with table 2. recovery of dermacentor albipictus larvae from contrasting habitats of the great lakes – st. lawrence forest ecosystem, algonquin park, ontario, 1983–1984. open field deciduous forest 1983 1984 1983 1984 1st larvae flagged sept 6 aug 28 sept 13 sept 18 % engorged females (efs) with larvae 38 (n = 453) 88 (n = 193) 67 (n = 450) 85 (n = 196) lei1 for efs with early2 larvae 1502 ± 1186 (n = 71) 1520 ± 1112 (n = 50) lei for efs with late3 larvae 346 ± 506 (n = 19) 868 ± 711 (n = 19) total lei 1258 ± 1176 (n = 90) 3214 ± 1734 (n = 49) 1340 ± 1054 (n = 69) 1463 ± 1315 (n = 50) 1larval efficiency index = number of larvae/gram of engorged female from which larvae recovered). 2larvae first flagged in september. 3larvae first flagged in october. alces vol. 52, 2016 addison et al. – recruitment of winter tick larvae 33 aspect towards the summer sun being hotter and drier (londt and whitehead 1972). we had no weather data at the microsite level (i.e., in the duff layer where eggs and larvae reside). however, because microsites in open habitats have higher temperature, lower relative humidity (rh), and higher wind speed (londt and whitehead 1972, daniel et al. 1977, daniel 1978), we assumed that our flat opening had lower rh, and higher temperatures, wind speeds, and vapor pressure deficits (vpds) at the microsite level than the forest habitat with dense canopy cover and northeast sloping aspect. weather and desiccation both eggs and larvae lose moisture under conditions of vpd, with temperature the main influence on vapor pressure. desiccation of eggs is detrimental because eggs cannot reabsorb water even at high rh (rechav and maltzahn 1977). desiccation of eggs (dermacentor variabilis) occurred when rh was reduced 6–8 h daily, fewer eggs hatched at 85% than 95% rh (sonenshine and tigner 1969), and weight and hatchability of boophilus decoloratus eggs were reduced at lower rh (rechav and maltzahn 1977). high vpd is correlated positively with weight loss in boophilus microplus and b. annulatus eggs, with a strong negative relationship between weight loss and percent hatched (teel 1984). the ~50% lower recruitment of larvae in the opening in 1983 versus 1984 presumably reflected higher desiccation and mortality of eggs during oviposition and incubation in the hotter and drier june and july of 1983. desiccation also threatens tick larvae (knülle 1966). for example, survival of amblyomma hebraeum larvae dropped precipitously at 70% rh compared to higher rhs (londt and whitehead 1972), ixodes ricinus larvae died sooner in open, grassy than in forest habitats (daniel et al. 1977), and a. americanum larval survival was shorter fig. 2. proportion of total seasonal recruitment of larvae of dermacentor albipictus acquired during individual sampling periods in opening and forest habitats, algonquin park, ontario, autumn 1983 and 1984. fig. 3. larvae per gram of engorged female (lei) dermacentor albipictus recruited from opening and forest habitats, algonquin park, ontario during early and late autumn 1983 and autumn 1984. fig. 1. accumulation of tick plots with larvae of dermacentor albipictus in opening and forest habitats, algonquin park, ontario, autumn 1983 and 1984. 34 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 (10–19 d) in meadow than forest habitat (33–106 d) in oklahoma (patrick and hair 1979). incubation of winter tick eggs occurs more rapidly at higher than lower temperatures (addison et al. 1998). in 1984, larvae first appeared in early august and presumably earlier in 1983 given the hotter june and july. thus, although desiccation in july 1983 may have impacted larval survival, the higher average monthly september temperature (3 °c) than in 1984 may also have contributed to reduced larval survival and recruitment in 1983. overall desiccation of larvae may have been moderated by higher precipitation in september 1983 and the relatively lower vpd associated with cooler weather in late august and september. we conclude that reduced recruitment of larvae due to desiccation in the opening in 1983 likely had a minimal role in overall reduced lei, particularly relative to desiccation of eggs. fecundity while the reduced proportion of efs with larvae in the opening in 1983 could be attributed to higher desiccation and mortality of eggs and larvae, it is possible that high mortality occurred in efs prior to egg-laying in the hotter and drier june and july of 1983. in 1984, the rei values in cells within the forest site were ~50% lower than in the opening; in contrast, drew and samuel (1986) reported similar rei values for efs in open and closed habitats. in 1984, the rei values in the opening were similar (median = 7538) to those produced under laboratory conditions (μ = 7097–9443) (addison et al. 1998), indicating that conditions in the opening were highly favorable for egg production and more so than at the forest site (may mean monthly temperature of 9.1 °c) where the microsite temperature was presumably lower. egg production was also lower at 15 °c than at 20 and 25 °c in rhipicephalus sanguineus (sweatman 1967). in general, it is anticipated that overall fecundity of efs would be lower in cooler than warmer habitats in northerly parts of winter tick range, that more open habitats are the likely nidus for transmission in cooler summers, and that fecundity might be correlated with latitude within winter tick range. fig. 4. scatter gram of weight of detached engorged female dermacentor albipictus versus first date of appearance of larvae on vegetation in opening and forest habitats, algonquin park, ontario, autumn 1983 and 1984. fig. 5. scatter gram of date of detachment of engorged female dermacentor albipictus from moose versus first date of appearance of larvae on vegetation in opening and forest habitats, algonquin park, ontario, autumn 1983 and 1984. alces vol. 52, 2016 addison et al. – recruitment of winter tick larvae 35 rate of development oviposition occurred earlier in the opening than in the forest. this is consistent with earlier oviposition by winter ticks at 24 °c than at 20 °c under laboratory conditions (addison et al. 1998) and oviposition of other species of ticks occurring more quickly in open field than in forested sites (daniel et al. 1977, dusbabek et al. 1979, patrick and hair 1979, koch 1984). incubation of winter tick eggs is also accelerated at higher temperatures (addison and smith 1981, addison et al. 1998), hence, the accelerated incubation in the opening as compared to the forest was likely due to higher microsite temperatures. as with earlier oviposition, this could have contributed to the earlier appearance of larvae in the opening than in the forest in both years. similarly, the incubation period for d. albipictus in alberta was shorter in an open grassland than in an aspen forest (drew and samuel 1986), and shorter incubation periods in open fields as compared to forests have been reported for a. americanum in oklahoma (patrick and hair 1979) and d. reticularis in slovakia (dusbabek et al. 1979). recruitment of larvae lei, one measure of recruitment, was based only on those efs that produced larvae. total recruitment also is affected by the proportion of efs producing larvae. the leis in 1983 and in the forest in 1984 were <20% of the reis of egg-laying efs in laboratory conditions (see addison et al. 1998) indicating about 80% attrition of the maximum potential recruitment from efs producing larvae; the 1984 lei in the opening was ~40% of the maximum potential. total recruitment in 1984 was higher at both sites, particularly in the opening where twice as many plots (efs) were productive compared to 1983. conversely, total recruitment in 1983 was highest in the forest habitat where 29% more efs produced larvae. these results are consistent with the hypothesis that weather and ground conditions are primary influences on which habitats have highest potential to contribute viable larvae in any given year. no compensatory increased production of larvae was observed from plots not producing larvae until late in the season. thus, earlier ascent onto vegetation should be a major advantage if increased availability (time) for transmission to moose is a primary selective advantage; however, desiccation of larvae is also more likely in warmer weather characteristic of earlier transmission in early september. winter tick larvae do not descend vegetation to rehydrate like certain tick species (see knülle and rudolph 1982). after ascending, they remain on vegetation (patrick and hair 1975, drew and samuel 1985) and must employ alternative strategies to avoid desiccation and death when vpd is high. in 1984, the first larvae to ascend vegetation in the open plots were recovered 3–4 weeks after the initial appearance of larvae in closed cells at both sites. this delayed ascent was synchronous with the transition to cooler night air and abundant dew on vegetation (lower vpd) that would help rehydrate larvae permanently exposed after ascending vegetation, and might also account for why larvae appeared on vegetation earlier during the cooler september of 1984 as compared to 1983. timing of ascent by winter tick larvae varies annually with weather (samuel and welch 1991) and habitat in the same area (patrick and hair 1975, current study), and between ecosystems at the continental scale. for example, winter tick larvae experienced 4–8 month post-hatch delays before appearing on vegetation or hosts during october– november in warm climates of texas, oklahoma, and parts of coastal central california (table 3). ascent of vegetation at a south-facing site near kamloops, british columbia occurred as early as in central 36 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 alberta and central ontario, but not until ~2 months post-hatch (wilkinson 1967). in central alberta and central ontario, larvae were first reported on vegetation after a minimum post-hatch delay of 2–4 weeks (table 3). although the timing of transmission of winter tick larvae may seem to be an adaptation coincident with increased movement of moose during the mating season (drew and samuel 1985), the wide variation in the presence of larvae before ascending vegetation likely reflects a weather-induced evolutionary adaptation to reduce desiccation and mortality. one disadvantage of delayed ascent could be a truncated transmission period that would occur more frequently in colder and more northern moose range; the opposite may explain more frequent epizootics in southern moose range. we collected limited larvae following cold weather and substantial snowfall in 1983 and, like drew and samuel (1985), observed that larvae were much slower in response to movement and less effective in attaching to a flagging sheet at colder temperatures. summary this study compared recruitment of winter tick larvae in 2 different habitats, a forest opening and a closed canopy, deciduous forest, by measuring survival of known adult female ticks and their productivity relative to weather. recruitment varied annually both within and between habitats indicating that weather and microsite conditions influence recruitment of winter tick larvae. this influence was most important during the egglaying and incubation periods in summer and table 3. timing of ascent of dermacentor albipictus larvae. location 1st larvae on vegetation 1st larvae on host post-hatch delay in ascent edwards plateau, texas (1) early nov. san antonio/kerrville texas (2) oct. eastern oklahoma (3) open meadow nov. 15 151 d forest nov. 15 133 d central california (4) sept. 15 -sept. 18 -aug. 20–27 ~ 4–8 mon kamloops, british columbia (5) sept. 19 ~ 2 mon oct. 3–6 central alberta (6) early sept. ~ 2 wk 1985 (7) oct. 3 1986 (7) sept. 30 1987 (7) sept. 6 1988 (7) sept. 11 central ontario (8) 1983 opening sept. 6 1983 forest sept. 13 1984 opening aug. 28 17–24 d min 1984 forest sept. 18 25–32 d min numbers in parentheses refer to references: (1) parish and rude 1946; (2) drummond 1967; (3) patrick and hair 1975; (4) howell 1939; (5) wilkinson 1967; (6) drew and samuel 1986; (7) samuel and welch 1991; (8) current study. alces vol. 52, 2016 addison et al. – recruitment of winter tick larvae 37 when larvae ascended vegetation in autumn. open and warmer habitats are presumably the nidus for transmission of larvae to moose except in years of hot, dry weather in summer and autumn that increases egg and larval desiccation. the end of the diapause that occurs between hatching and ascent of vegetation appears synchronous with cooler, more humid conditions that would reduce desiccation of larvae both on the ground and questing on vegetation in late summer and autumn. acknowledgements we greatly appreciate the efforts of d. fraser, s. fraser, s. gadawaski, a. jones, s. mcdowell, l. berejikian, k. long, k. paterson, l. smith, d. bouchard, v. ewing, j. jefferson, m. van schie, a. macmillan, a. rynard, n. wilson, c. pirie, m. mclaughlin, d. carlson, c. mccall, and p. methner for their strong commitment to some or all of capturing, raising, and maintaining moose calves and collection of ticks. field work was conducted at the wildlife research station in algonquin park where the omnr staff was of great help. we thank d. joachim for technical assistance, g. smith and c. macinnes for strong administrative support. r. addison, w. samuel, m. lankester and p. pekins provided much appreciated constructive review of the manuscript. references aalangdong, o. i., w. m. samuel, and a. w. shostak. 2001. off-host survival and reproductive success of adult female winter ticks, dermacentor albipictus in seven habitat types of central alberta. journal of the ghana science association 3: 109–116. addison, e. m., d. g. joachim, r. f. mclaughlin, and d. j. h. fraser. 1998. ovipositional development and fecundity of dermacentor albipictus (acari: ixodidae) from moose. alces 34: 165–172. ——, and l. m. smith. 1981. productivity of winter ticks (dermacentor albipictus) collected from road-killed moose. alces 17: 136–146. ——, r. d. strickland, and d. j. h. fraser. 1989. gray jays, perisoreus canadensis, and common ravens, corvus corax, as predators of winter ticks, dermacentor albipictus. canadian fieldnaturalist 103: 406–408. bertrand, m. r., and m. l. wilson. 1996. microclimate-dependent survival of unfed adult ixodes scapularis (acari: ixodidae) in nature: life cycle and study design implications. journal of medical entomology 33: 619–627. chiera, j. w., r. m. newson, and m. p. cunningham. 1985. cumulative effects of host resistance on rhipicephalus appendiculatus neumann (acarina: ixodidae) in the laboratory. parasitology 90: 401–408. cross, j., d. kaukinen, r. sitch, s. heringer, a. smiegielski, d. hatfield, g. macisaac, and t. marshall. 2012. historic climate analysis tool [digital application] version 2.5. ontario ministry of natural resources, northwest science and information, thunder bay, ontario, canada. daniel, m. 1978. microclimate as a determining element in the distribution of ticks and their developmental cycles. folia parasitologica (prague) 25: 91–94. ——, v. cerny, f. dusbabek, e. honzakova, and j. olejnicek. 1977. influence of microclimate on the life cycle of the common tick ixodes ricinus (l.) in an open area in comparison with forest habitats. folia parasitologica (prague) 24: 149–160. drew, m. l., and w. m. samuel. 1985. factors affecting transmission of larval winter ticks, dermacentor albipictus (packard), to moose, alces alces l. in alberta, canada. journal of wildlife diseases 21: 274–282. 38 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 ——, and ——. 1986. reproduction of the winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64: 714–721. ——, and ——. 1987. reproduction of the winter tick, dermacentor albipictus, under laboratory conditions. canadian journal of zoology 65: 2583–2588. drummond, r. o. 1967. seasonal activity of ticks (acarina: metastigmata) on cattle in southwestern texas. annals of the entomological society of america 60: 439–447. ——, and t. m. whetstone. 1970. oviposition of the gulf coast tick. journal of economic entomology 63: 1547–1551. ——, ——, s. e. ernst, and w. j. gladney. 1969. biology and colonization of the winter tick in the laboratory. journal of economic entomology 62: 235–238. dusbabek, f., v. cerny, e. honzakova, m. daniel, and j. olejnicek. 1979. differences in the developmental cycle of dermacentor reticulatus in two closely situated biotypes. recent advances in acarology ii: 155–157. fleetwood, s. c., p. d. teel, and g. thompson. 1984. seasonal activity and spatial distribution of lone star tick populations in east central texas. the southwestern entomologist 9: 109–113. glass, g. v., p. d. peckham, and j. r. saunders. 1972. consequences of failure to meet assumptions underlying fixed effects analyses of variance and covariance. review of educational research 42: 237–288. harwell, m. r., e. n. rubinstein, w. s. hayes, and c. c. olds. 1992. summarizing monte carlo results in methodological research: the one-and two-factor fixed effects anova cases. journal of educational and behavioral statistics 17: 315–339. howell, d. e. 1939. the ecology of dermacentor albipictus (packard). proceedings of the sixth pacific science congress 4: 439–458. knülle, w. 1966. equilibrium humidities and survival of some tick larvae. journal of medical entomology 2: 335–338. ——, and d. rudolph. 1982. humidity relationships and water balance of ticks. pages 43–70 in f. d. obenchain and r. galun, editors. current themes in tropical science. volume 1. physiology of ticks. pergamon press, oxford, united kingdom. koch, h. g. 1984. survival of the lone star tick, amblyomma americanum (acari: ixodidae), in contrasting habitats and different years in southeastern oklahoma, u.s.a. journal of medical entomology 21: 69–79. lix, l. m., j. c. keselman, and h. j. keselman. 1996. consequences of assumption violations revisited: a quantitative review of alternatives to the one-way analysis of variance f test. review of educational research 66: 579–619. londt, j. g. h., and g. b. whitehead. 1972. ecological studies of larval ticks in south africa (acarina: ixodidae). parasitology 65: 469–490. mcgowan,m.j.,r.w.barker,j.t.homer, r. w. mcnew, and k. h. holscher. 1981. success of tick feeding on calves immunized with amblyomma americanum (acari: ixodidae) extract. journal of medical entomology 18: 328–332. ——, j. t. homer, g. v. o’dell, r. w. mcnew, and r. w. barker. 1980. performance of ticks fed on rabbits inoculated with extracts derived from homogenized ticks amblyomma maculate koch (acarina: ixodidae). journal of parasitology 66: 42–48. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. mcpherson, m., a. w. shostak, and w. m. samuel. 2000. climbing simulated alces vol. 52, 2016 addison et al. – recruitment of winter tick larvae 39 vegetation to heights of ungulate hosts by larvae of dermacentor albipictus (packard) (acari: ixodidae). journal of medical entomology 37: 114–120. parish, h. e., and c. s. rude. 1946. ddt to control the winter horse tick. journal of economic entomology 39: 92–93. patrick, c. d., and j. a. hair. 1975. ecological observations on dermacentor albipictus (packard) in eastern oklahoma (acarina: ixodidae). journal of medical entomology 12: 393–394. ——, and ——. 1979. oviposition behaviour and larval longevity of the lone star tick, amblyomma americanum (acarina: ixodidae), in different habitats. annals of the entomological society of america 72: 308–312. r core team. 2013. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. http://www.r-project. org/ (accessed november 2015). rechav, y., and h. c. von maltzahn. 1977. hatching and weight changes in eggs of two species of ticks in relation to saturation deficit. annals of the entomological society of america 70: 768–770. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m., and d. a. welch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27: 169–182. schütte, k. h., and w. h. king. 1965. a rapid and simple method for measuring evaporating power of the air. journal of south african botany 31: 127–131. sonenshine, d. e. 1970. current studies on tick biology in relation to disease in the americas. miscellaneous publications of the entomological society of america 6: 352–358. ——, and j. a. tigner. 1969. oviposition and hatching in two species of ticks in relation to moisture deficit. annals of the entomological society of america 62: 628–640. sweatman, g. k. 1967. physical and biological factors affecting the longevity and oviposition of engorged rhipicephalus sanguineus female ticks. journal of parasitology 53: 432–445. teel, p. d. 1984. effect of saturation deficit on eggs of boophilus annulatus and bmicroplus (acari: ixodidae). annals of the entomological society of america 77: 65–68. tuininga, a. r., j. l. miller, s. u. morath, t. j. daniels, r. c. falco, m. marchese, s. sahabi, d. rosa, and k. c. stafford iii. 2009. isolation of entomopathogenic fungi from soils and ixodes scapularis (acarai: ixododae) ticks: prevalence and methods. journal of medical entomology 46: 557–565. wilkinson, p. r. 1967. the distribution of dermacentor ticks in canada in relation to bioclimatic zones. canadian journal of zoology 45: 517–537. zar, j. h. 1982. biostatistical analysis, 2nd edition. prentice hall, englewood cliffs, new jersey, usa. 40 recruitment of winter tick larvae – addison et al. alces vol. 52, 2016 http://www.r-project.org/ http://www.r-project.org/ recruitment of winter ticks (dermacentor albipictus) in contrasting forest habitats, ontario, canada study area methods definitions engorged female ticks (efs) fecundity and hatching larval recruitment data analysis results dispersal and fecundity larval recruitment minimum free-iving period discussion weather and desiccation fecundity rate of development recruitment of larvae summary acknowledgements references alces18_276.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces15_119.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces14_1.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces19_204.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces20_47.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces15_349.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces16_11.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces20_299workshop.pdf alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 alces vol. 20, 1984 f:\alces\supp2\pagema~1\rus 3s. alces suppl. 2, 2002 antipov et al. comparative anatomy of hearts 7 the morphological characteristics of the sinoauricular and auriculo-ventricular nodes in moose, cattle, pig, and human hearts1 nicholas v. antipov2, v. a. golovko3, and a. f. sinev4 2medical institute donetsk, ukraine; 3institute of physiology, komi scientific center, ural division of the russian academy of sciences, 167610, syktyvkar gsp, komi republic, russia; 4sri of cardio-vascular surgery, moscow, russia abstract: comparative-anatomic studies on morphometric characteristics of the heart, cardiac circulatory system (ccs), and qualitative estimation of specific volumes of ccs structural components were conducted. it was found that the histologic structure of ccs in the investigated species (moose, calf, pig, and man) differs in size, blood supply, innervation, and in the correlation of specific volumes of conducting and contracting cardiocytes of connective tissue. the original data collected on the structure of moose will help to explain functional characteristics in the moose electrocardiogram. alces supplement 2: 7-10 (2002) key words: calf, comparative anatomy, circulatory system, heart, man, moose, pig there are data in the literature concerning the anatomy of the cardiac circulatory system (ccs) in cows (abe 1987), pigs (lopes 1976), and man (sinev and krymsky 1987). however, the structure of the moose circulatory system has not been studied, although ecg recordings were made long ago (roshchevsky et al. 1976). because of this lack of study, a problem with the comparative anatomy of the cardiac circulatory system in ungulates living in different ecological conditions still appears in biology, physiology, and zoo-veterinary science. the purpose of this study was to investigate the morphometric characteristics of the moose heart, structural and comparative anatomy of its circulatory system, and to quantitatively estimate specific volumes of structural components of the cardiac circulatory system (conducting cardiocytes, hemocirculatory path, conductive tissue, and stromatic elements) in moose, calf, pig, and man. materials and methods the objects of the investigation were moose heart (n = 10), calf heart (n = 12), and pig (n = 20). the measurements were made on the heart mass, wall thickness in different parts of the myocardium, length of inflow and outflow parts of the right heart, and the area of the great vessels. paraffin serial histotopograms (approximately 7 microns) were stained according to the method of van gizon, vergoff, and mallory in masson’s modification. to estimate peculiarities of the circulatory system, blood supply, and structure of the heart, bone coronaroangiography and rentgenography were performed on 12 preparations. specific volumes of structural components of the cardiac circulatory system were studied on histopreparations with the application of 1editor’s note: this manuscript was submitted without references and due to difficulties contacting the author the manuscript was published without them. alces16_preface.pdf alces vol. 16, 1980 alces14_89.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces19_272.pdf alces vol. 19, 1983 alcessupp1_1.pdf alces18_iipreface.pdf alces vol. 18, 1982 alces 45 (2009) contents in memoriam john lykke...................................................................................... i foreword.............................................................................................................. ii status of regional moose populations in european and asiatic russia........................................................................................leonid m. baskin 1 history and status of the population dynamics of moose in the steppe zone of ukraine....................................................anatolii m. volokh 5 the status and management of moose in the murmansk region, russia............................................olga a. makarova and anatoly m. khokhlov 13 regional populations and migration of moose in northern yakutia, russia...................................................................valeriy m. safronov 17 history and status of moose in the rostov region, russia............ ................viktor a. minoranskiy, viktor v. sidelnikov, and elena i. simonovich 21 fragmentation of eurasian moose populations during periods of population depression.........taras p. sipko and marina v. kholodova 25 status of reintroductions of three large herbivores in russia ...........................................................................................................taras p. sipko 35 improving population management and harvest quotas of moose in russia..............................................................vladimir m. glushkov 43 the influence of moose on tree species composition in liesjärvi national park in southern finland.......................................... ......................................................................risto heikkilä and marita tuominen 49 moose conservation in the national wildlife refuge system, usa ...........................................................................................................robin l. west 59 wood quality of birch (betula spp.) trees damaged by moose...... ........................................sauli härkönen, arto pulkkinen, and henrik heräjärvi 67 deer ked (lipoptena cervi) dermatitis in humans – an increasing nuisance in finland........................................................................................... ..................sauli härkönen, maria laine, martine vornanen, and timo reunala 73 status and review of the vector-borne nematode setaria tundra in finnish cervids..................sauli laaksonen and antti oksanen 81 deer ked, an ectoparasite of moose in finland: a brief review of its biology and invasion...arja kaitala, raine kortet, sauli härkönen, sauli laaksonen, laura härkönen, sirpa kaunisto, and hannu ylönen 85 a program to monitor moose populations in the dehcho region, northwest territories, canada...................................nicholas c. larter 89 (continued on inside back cover) compensatory shoot growth in trembling aspen (populus tremuloides michx.) in response to simulated browsing............... ..........................................allan w. carson, roy v. rea, and arthur l. fredeen 101 the effects of human activity on summer habitat use by moose .........................................odd n. lykkja, erling johan solberg, ivar herfindal, jonathan wright, christer moe rolandsen, and martin g. hanssen 109 how moose select forested habitat in gros morne national park, newfoundland.......................................brian. e. mclaren, s. taylor, and s. h. luke 125 reduced genetic diversity in two introduced and isolated moose populations in alaska..................................................kris j. hundertmark 137 assessment of a line transect field method to determine winter tick abundance on moose..............meghan sine, karen morris, and david knupp 143 6th international moose symposium......................................................... 147 previous meeting sites................................................................................. 149 distinguished moose biologist past recipients............................... 150 distinguished moose biologist award criteria ............................... 151 editorial review committee........................................................................ 152 additional copies available from: lakehead university bookstore, thunder bay, ontario, canada p7b 5e1 alces 39-45 price $40.00 canadian or u.s. each (including supplementary issues) alces 24-38 price $38.00 canadian or u.s. each (including supplementary issues) alces 17-23 price $20.00 canadian or u.s. each proceedings of the north american moose conference and workshop 8-16 price $20.00 canadian or u.s. each make cheques, money orders or purchase orders payable to lakehead university bookstore. all prices include 5% g.s.t., mailing and handling costs. prices are subject to change. acknowledgements brooke pilley worked long hours formatting and typesetting manuscripts. alces home page further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our website: http://bolt.lakeheadu.ca/~alceswww/alces.html alces18_xiworkshopsessions.pdf alces vol. 18, 1982 alces16_255.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 f:\alces\supp2\pagema~1\rus 26s alces suppl. 2, 2002 roshchevsky – moose relations with biotic communities 113 moose relations with biotic communities and man in samarskaya luka national park jury k. roshchevsky samarskaya luka national park, 446350, zhiguliovsk, samara region, russia abstract: during 1985–1990 the density of the moose (alces alces) population in the peninsular territory (samarskaya luka national park and zhiguliovski reserve in the middle volga region) was studied. it was found that the natural cyclicity of the population differed between regions. the lowest intensity of oscillation in moose density was in the population nucleus, where the wavelength equaled 5 years. in the interflow areas the wavelength was 2–4 years. oscillatory dependence of the moose density in the nucleus on the density of other trophic level mammals such as the mountain hare (lepus timidus), red fox (vulpes vulpes), wolf (canis lupus), and pine marten (martes martes) was also evident. alces supplement 2: 113-117 (2002) key words: moose, natural cycle, population dynamics in the forests of the samarskaya luka national park prior to 1985, the moose p o p u l a t i o n w a s s t u d i e d o n l y i n t h e zhiguliovski reserve (1/7 of the regional area). information about moose density, number, and mortality in the reserve was scarce (belanin 1977, 1980, 1981; korotaev 1983). study area the study was conducted in samarskaya luka national park. this is a peninsular territory in the mid volga region (fig. 1). it is surrounded by the volga river and reserv o i r s . t h e z h i g u l i o v s k i r e s e r v e (zapovednik) (fig. 1b, 1e) and samarskaya luka national park lie on the peninsula. the reserve was established in 1966 and the park in 1985. the total area is 160,000 ha. the area of the reserve is 23,000 ha; the area of park is 128,000 ha. the geography of the peninsula is quite varied. there are low mountains (zhiguli) which do not rise above 375m in elevation. these mountains cover an area of 22,152 ha of which 41.3% are in the reserve (central zhiguli, fig. 1b) and 58.7% are in the national park (western, fig. 1a and eastern zhiguli, fig. 1c). zhiguli is the mountain forest area. t h e a r e a o f t h e f o r e s t p l a t e a u (charokayski forest) is 41,069 ha (fig. 1e, 1d). only some of this region (13,010 ha) is included in the reserve (fig. 1e). other natural areas of the samarskaya luka lie only in the park. the area of the bolshoi riazanski forest is equal to 2,703 ha (fig. 1i). the area of the bolshoi chuvashski forest is 3,702 ha (fig. 1j). the well– balanced forest–steppe (46% forested) covers 12,757 ha, including the bahilovski forest–steppe (fig. 1f) and eastern forest– steppe (fig. 1f, 1g). the askulskaya forest–steppe, which occupies 51,814 ha (fig. 1h), is more open (15.5% forested). the mean forest stand size here is 53 ha. the flood plain areas (fig. 1m) occupy 7,427 ha (65% forested). the podgorki area (fig. 1k) is volga lowland and is not moose habitat. the total area of the national park forest is 62,499 ha, comprised of lime (29,181 ha), oak (15,443 ha), aspen (10,802 ha), and, alces suppl. 2, 2002 roshchevsky – moose relations with biotic communities 117 symposium on ungulates of ussr (moscow, 24–26 december 1979). nauka, moscow, russia. (in russian). . 1981. mammals of zhiguliovski state preserve. pages 89–103 in ecological and faunistic research in preserves. moscow, russia. (in russian). korotaev, g. p. 1983. the status of the population of ungulates and large predators in zhiguliovski state preserve. pages 108–110 in problems with the management and protection of natural areas in samarskaya luka. kyibyshev, russia. (in russian). lubvina, i. v. 1982. the impact of oil production on the functional status of biocenoses. pages 27–43 in ecology and animal protection. kyibyshev, russia. (in russian). mirkin, b. m. 1989. further considerat i o n s a b o u t t h e o r g a n i s m i n phytocenology. botanical journal 74:3– 13. (in russian). different trophic levels and the formation and oscillations of free populations. acknowledgements v. zhukov, i. belikov, g. belikova, n. savin, m. kochetkov, v. vehnik, t. s a z o n o v a , v . s h e b a r s h e n k o , e . shebarshenko, and other research workers of samarskaya luka national park and zhiguliovski reserve conducted ground and aerial surveys. the ground surveys were conducted by kyibyshev (samara) university students (the detachment for nature protection). references belanin, v. n. 1977. ungulates of zhiguliovski state preserve. pages 21– 24 in hunting and management of preserves. moscow, russia. (in russian). . 1980. data on ungulate mortality in zhiguliovski state preserve. pages 122–123 in ungulates of ussr fauna. proceedings of the second all–union alces14_178.pdf alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 alces vol. 14, 1978 f:\alces\supp2\pagema~1\rus7s.pdf alces suppl. 2, 2002 bogomolova et al. home ranges and migrations of farm moose 33 home ranges and migrations of the kostroma farm moose ekaterina m. bogomolova1, yuriy a. kurochkin1, and alexander n. minaev2 1anokhin research institute of normal physiology, russian academy of medical science, moscow, russia; 2institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: the published literature about moose home ranges is rather contradictory. we studied summer (april–september) home ranges, degree of their variability, and features of moose migrations on the kostroma moose farm during 1977–1989. the home range of a free-living, handreared female moose encompassed an aggregated area of 57 km2 over 8 years, with the yearly home range varying from 15 to 44 km2. home ranges of her 4 female offspring proved to be about the same as their natal ranges. we also detected short-term (1–3 day) migrations of cows during the breeding season. alces supplement 2: 33-36 (2002) key words: home range, migrations, moose the published literature about moose home ranges is rather inconsistent. this fact may be explained by the natural variability of moose home ranges dependent upon specific, local environments, and by different estimation methods used by investigators. the results of our work provide evidence that at least some hand-reared and wild female moose in the region of the kostroma moose farm are highly sedentary. their home ranges are rather stable. home ranges of young females in the first years of independent life are almost the same as their natal areas. besides wellknown spring migrations of yearlings, we detected short-term autumn migrations of cows beyond the usual home range with a rather quick return rate of 1–3 days. we assume that the most reliable results may be obtained by long-term studies of radiotagged animals (cederlund and okarma 1988, bogomolova et al. 1989). study area we used the kostroma moose farm for our study area. using this facility made locating radiocollared moose convenient, and we were able to study moose in their forest habitats. methods we studied summer (april–september) home ranges, degree of their variability, and features of moose migrations on the kostroma moose farm during 1977–1989. in this work we used the radio-tracking system “los-2,” (designed by a. n. minaev), and radio-communication among biologists. we repeatedly located free-ranging, radiotagged animals with a portable receiver and observed these moose in their forest habitats. the coordinates of their locations were recorded with accuracy up to 50 m. each year we studied a freeranging, hand-reared moose cow, her 4 female offspring (which had their own home ranges) at different ages up to 4 years of age, 5–15 milk moose, and 10–15 yearlings, and other young moose. we used 2 methods to estimate the home range area: (1) we considered the home range to be a convex polygon containhome ranges and migrations of farm moose bogomolova et al. alces suppl. 2, 2002 34 ing all the points where an animal was located (corresponding area is identified as s1); and (2) we divided the entire investigated territory into squares 250 x 250 m and summed the areas of those squares (s2) in which an animal was found 1 or more times. the important methodical point is the reliability of results of such investigations. we believe it is better to give not only the estimation of the home range area itself but also the number of animal locations and the number of observation days (if an animal could be observed more than once daily). results and discussion the home range of the free-ranging cow, lastochka, and her calves encompassed approximately 57 km2 over 8 years, with the yearly home range area varying from 15 to 44 km2 (table 1). we also studied home ranges of lastochka‘s 4 radiotagged wild female offspring (table 2). their home range areas were approximately as large as those of lastochka in the first years of her free life. unfortunately, the study of these cows was relatively short; at least 2 of them were killed by table 1. summer home range areas of the moose-cow lastochka and her offspring. year s1 s2 number of number of (km2) (km2) observations days located 1982 calves 16.4 5.38 248 86 1983 calves 17.6 6.13 214 70 1984 yearlings 17.0 2.94 90 49 1985 calves 43.9 6.13 279 106 1986 yearlings 34.2 8.25 241 122 1987 calves 25.1 4.94 172 135 1988 calves 15.5 1.81 40 36 1989 yearlings 16.4 1.19 20 17 table 2. summer home range areas of the wild moose cows born to a hand-reared moose cow lastochka. name year period number of home range area of of days located s1 (km2) s2 (km2) birth observation devochka 1983 1985 64 49 17.6 malenkaja 1985 1987 90 85 12.3 malenkaja 1985 1988 34 28 13.8 malenkaja 1985 1989 17 17 9.3 malenkaja 1985 1987–89 141 130 21.2 bolshaja 1985 1987 96 94 14.7 lusa 1987 1988 31 28 10.8 lusa 1987 1989 14 12 8.3 lusa 1987 1988-89 45 40 13.8 alces suppl. 2, 2002 bogomolova et al. home ranges and migrations of farm moose 35 poachers, 1 was lost because of antenna breakage, and 1 more disappeared under unknown circumstances. we found that during the entire period of observation lastochka‘s female offspring hardly went out of the natal area boundary (table 3). in spite of her attachment to a particular territory, lastochka, having been transported some kilometers away from her home range in winter to a wood-cutting area with much food, did not return home in spring and instead stayed in the new location. there are 2 possible explanations for this fact. the first was that she did not want to return and formed a new home range. the other explanation was that lastochka might lose her way to her habitual home range. however, the next year after having been transported in winter to the same wood-cutting area, lastochka did find her way “home.” on her first attempt at finding her home range, lastochka walked the wrong direction and returned to a woodcut, but the next day we located her already within the bounds of her usual home range. according to many experienced moose farmers, animals of various ages often leave their home ranges and do not return. the most noticeable are the spring migrations of yearlings and young milk-moose cows. they have been found many kilometers away from the farm. some moose, especially young males, leave the farm in autumn. biologists still do not fully understand the reasons for these migrations. several cases of migrations may be explained by individual characteristics of some moose that have weak attachments to their home ranges. it is more probable that, having gone occasionally past the limits of its’ home range and venturing further, an animal must spend more and more energy to return. the “cost of return” may prove to be too high, and the animal will stay in its newly chosen location. there may be various reasons for “border crossing,” such as predators, interactions among males in autumn, or an abundance of gnats in spring. during the breeding season, some young males left the farm. twice we observed yearling bulls leaving their mother, mostly during breeding time, while their sisters stayed with their mother until the next spring. autumn migrations of moose cows may be related to quick displacements of moose cows beyond their usual home ranges during the breeding season. we observed such displacements many times. without being chased by a bull, very excited estral cows made sudden 10–15 km trips (in autumn they generally walk no more than 2 km per day). as a rule, during these trips the cows returned in 1–3 days. the results of our work provide evidence that at least some hand-reared and wild female moose in the region of the kostroma moose farm are highly sedentary. their home ranges are rather stable. table 3. comparison of home range and natal areas of the wild lastochka’s female offspring. name home natal area home range area outside % of range overlap with bounds of the natal area natal area area lastochka (km2) (km2) s1 (km2) devochka 17.6 21.1 3.5 1.7 malenkaja 21.2 44.5 1.9 4.3 bolshaj 14.7 44.5 0.4 0.9 lusa 13.8 21.2 1.9 9.0 home ranges and migrations of farm moose bogomolova et al. alces suppl. 2, 2002 36 home ranges of young females in the first years of independent life are almost the same as their natal areas. besides wellknown spring migrations of yearlings, we detected short-term autumn migrations of cows beyond the usual home range with the rather quick return rate of 1–3 days. references bogomolova, e. m., y. a. kurochkin, and a. n. minaev. 1989. summer home range and daily movements of a moosecow with calves of different age. pages 119-120 in ecology, morphology, utilization and protection of wild ungulates. abstracts. moscow, russia. (in russian). cederlund, g. n., and h. okarma. 1988. home range and habitat use of adult female moose. journal of wildlife management 52:336–343. f:\alces\supp2\pagemaker\rus10s alces suppl. 2, 2002 dvornikov moose population control in the urals 45 moose (alces alces) population control in the urals mikhael g. dvornikov kirov agricultural institute, kirov, russia abstract: i discuss the spatial dynamics of moose numbers, contemporary hunting practices, and migrations based on the structure of the landscape and subsequent human modification. i conclude that a single population of moose, comprised of subpopulations of non-migratory and migratory animals, inhabits the urals. an optimal ratio of those groups allows full usage of habitats and high population densities. existing methods of using the resources of the urals are considered. forms, terms, norms, and structural characteristics of removal of animals belonging to different demographic classes are recommended for population productivity to be increased. alces supplement 2: 45-47 (2002) key words: moose, population control, ural mountains rational usage of game resources demands a whole set of measures for their exploitation, reproduction, and protection. these measures are based on optimal population control methods, where the main goal for exploited populations is to ensure maximum, and if possible, stable harvest while preserving optimal population structure and numbers. research is required to achieve that goal via the following activities: managing game species on a population basis; estimating carrying capacity of habitat; revealing and eliminating limiting factors; controlling harvest; tailoring harvest quotas to population surpluses; and, orientation toward demographic classes with regard to the quantity and quality of exploitation. a number of specialized and generalized works (yazin 1972, filonov 1977, gordiyuk 1981, nikulin 1981, dvornikov 1984) on ecology of moose (alces alces) have been conducted in the ural mountains. until the present, however, the population approach to game management has not been well grounded, and no management strategies have been based on it. therefore, we are trying to define the population as a management unit and outline a strategy for conservation of moose for the time being (dvornikov 1989). i chose a population on the basis of data taken from the above-mentioned works. it was possible to trace the population history in a particular region, where animals functioned and responded to environmental changes as a whole, in accordance with their ecological characteristics. bone remnants of moose can be found in pleistocene deposits. dynamics of habitation of the urals by hoofed animals has been traced via paleontologic material. moose, as a rule, inhabited forest biotopes. at the same time, findings of material in unusual habitats for moose gave grounds for supposition that during periods of landscape change the animals migrated towards places with a mosaic distribution of vegetation and considerable supplies of food. contemporary natural populations were formed in the holocene. meridional mountain ridges are spotted with communities of mountain tundra and meadows, fir (abies spp.), spruce (picea spp.), mountain pine (pinus spp.), and pre-forest-steppe areas. at the foothills of the northern urals, fir, spruce, and mountain pine forest grow. in moose population control in the urals dvornikov alces suppl. 2, 2002 46 the eastern part of the southern and middle urals there are still some remnants of pine forest. clearings in the mountain forest started to appear in the 18th century, which contributed to the mosaic distribution of vegetation. therefore, the perpetuation of moose in the urals through periods of generally unfavorable conditions (filinov 1983) has been facilitated thanks to the mosaic of ecological conditions present. in the 20th century, transformation of the landscape by humans influenced the dynamics, numbers, and distribution of moose inhabiting the middle and southern urals. during the 1920s through the 1940s, moose were few in numbers in the urals. their migrations were but feebly noticeable. an intensive change in the forest structure during the 1940s through the 1960s contributed to increasing moose numbers. at present moose inhabit 64% of the basic forested areas. one can observe that moose favor biotopes situated in low places and mediumhigh mountains covered with saplings on clearings and with young, mature, and overmature coniferous forest. thus, the dynamics of forest formation influenced the number and distribution of moose. during the 1970s and 1980s the moose population became relatively stable and averaged 55-60,000 individuals. traditional migrations took place throughout the urals. in the absence of vegetation differences, animals may be observed across a range of 560,000 km2 (350 x 1,600km). in places of high concentration, including remnant pine forests, the density reached 50-80 individuals per 1,000 ha. at the same time the density in plains areas averaged 1-5 moose per 1,000 ha. the available food supply in mountain and pre-mountain areas averaged 400-2,000 kg/ha, in forest-steppe areas the average was 60-800 kg/ha (dry mass). the amount of food used in the latter places was higher. in addition, there is some specificity in feeding among groups of animals inhabiting different areas. no doubt it confirms the theory that moose migrate to mosaic vegetation areas with large amounts of diverse forage. the characteristics of traditional migrations are believed to be caused by the development of special mechanisms in response to the environment: the amount and the quality of forage and the depth of snow cover late in winter. once a stimulus is received, animals inhabiting large areas begin to migrate. the traditional migration of moose in the urals has been one of the main ecological events for a long time. those characteristics in ungulates are fixed genetically and through parents’ experience. the migrating groups are believed to appear as a common adaptive phenomenon, which later directionally changed the genetic structure of animals in accordance with the population cycle. from the existing evidence, i conclude that there is one population of moose in the urals. it includes subpopulations of nonmigratory and migrating individuals in the same area. their optimal ratio of abundance ensures their ability to use the forest fully and maintain high populations. at the same time one can see that the noticeable migration of moose is caused by the fact that in freezing weather moose prefer young coniferous stands as well as mature juniper stands over deciduous stands. large virgin tracts of forests remained throughout the north urals and in the middle and southern urals at an elevation of 500-1,200 m. thus, large groups of moose exist in the mountains where clearings border on large tracts of forest. we know that population production may be increased through intensive management of hunting. hunters take 6–7,000 moose in the region annually, 35–45% of them in november. the mortality rate including hunting is 22–31%. so taking into account the population of moose in the alces suppl. 2, 2002 dvornikov moose population control in the urals 47 north pre-urals and trans-urals, i recommend harvest rates of 10% in heavy coniferous forests in the mountain-, northand middle-taiga habitats, and 15% in the light coniferous forests in the middle-taiga, with a ratio of yearlings:immature:mature animals of 15:10:70, and a sex ratio of 50:50. in the middle and southern urals, as well as in the pre-urals and trans-urals, harvest rates can reach 15% in dark, light, and broad-leaved coniferous forest if the aforementioned ratios are met. in mountain pine, sub-taiga, and broad-leaved, dark coniferous forests, it is possible to harvest 20%, with age ratios of 20:15:65 and a sex ratio of 55:45 (males:females). in broadleaved and aspen (populus)-birch (betula) forest-steppes, the bag rate may amount to 25% with ratios of 20:15:65 and 50:50. in island pine-steppe forests, including foreststeppe reserves, the same harvest rate can be achieved with age and sex ratios of 25:18:57 and 60:40, respectively. it is reasonable to carry out hunting from 1 october to 30 november, sport hunting from 15 october to 15 december, hunting utilizing calling from 20 august to 20 september, and selective harvest from 15 december to 15 january. in some papers there are data on the validity and efficiency of biotechnical measures directed at the weakening of limiting factors. i believe that biotechnical measures are necessary in mountain pine preforest, steppe-pine, and birch forests, and broad-leaved and aspen-birch forest steppes of the middle and south urals. it is necessary to concentrate those measures in biotopes situated in middle parts of mountain ridges and in low places with mixed 80– 140-year-old forested stands with a diversity of young trees and high capacity of biological rotation. while carrying out these measures it is necessary to follow the norms given for the region. moose preservation and further observance of their sex and age structure by visual methods and hunting samples are the main conditions of their rational use. it is also necessary to envisage hunters returning one molar from their kill with a license in a special envelope. from the information received it will be possible to judge the status of the population and to correct the number and quality of animals harvested. references dvornikov, m. g. 1984. ecology and the biogeocentric role of ungulates in the ilmeny state preserve named after lenin. candidate scientific thesis, sverdlovsk, russia. (in russian). . 1989. ecological aspects of harvest, reproduction and protection of wild ungulates of the ural mountains. pages 75-80 in management of populations of wild ungulates. sverdlovsk, russia. (in russian). filonov, k. p. 1977. management of hunting. the dynamics of the number of ungulates and management of preserves. forest management, moscow, russia. (in russian). . 1 9 8 3 . m o o s e . l e s n a y a promyshlennost, moscow, russia. (in russian). gordiyuk, n. m. 1981. aspects of the ecology of ungulates in the bashkir preserve. candidate scientific thesis, sverdlovsk, russia. (in russian). nikulin, v. f. 1981. the moose of the upper kama river and its role in forest management and hunting. candidate scientific thesis, sverdlovsk, russia. (in russian). yazan, y. p. 1972. the animals of pechora taiga. volga-vyatskoye, kirov, russia. (in russian). alcessupp1_193.pdf susceptibility of winter tick larvae and eggs to entomopathogenic fungi beauveria bassiana, beauveria caledonica, metarhizium anisopliae, and scopulariopsis brevicaulis jay a. yoder1, peter j. pekins2, blake w. nelson1, christian r. randazzo1, and brett p. siemon1 1department of biology, wittenberg university, springfield, ohio 45501, usa; 2department of natural resources and the environment, university of new hampshire, durham, new hampshire 03824, usa abstract: an isolate of the soil fungus scopulariopsis brevicaulis was identified from the surface of female winter ticks (dermacentor albipictus) collected from recently dead moose (alces alces) calves in new hampshire in the northeastern united states. it was the sole isolate, and it matched with 98% nucits similarity (molecular systematics blast match) to s. brevicaulis species from soil and other tick species. inoculation of tick larvae and eggs with 108 spores/ml + 0.05% tween (aqueous inoculum) resulted in mortality, reduced survival time, and recovery of s. brevicaulis from within tick tissues. rapid water loss and death from dehydration were the pathogenic consequences of the fungal infection. three entomopathogenic fungal isolates from laboratory culture (beauveria bassiana, b. caledonica, and metarhizium anisopliae) inoculated concurrently at the same dose, were slightly less pathogenic to eggs than larvae of winter ticks. we conclude that s. brevicaulis imposes a limitation on the freeliving stages of the winter tick population in specific environmental conditions, but commercial fungal treatments as used in local situations to control ticks, are impractical as a means of controlling winter tick density across moose habitats. alces vol. 53: 41–51 (2017) key words: alces alces, dermacentor albipictus, moose, pathogenic fungi, scopulariopsis brevicaulis, survival, water balance, winter ticks introduction winter ticks (dermacentor albipictus) periodically cause high calf mortality in moose (alces alces) populations in the northeastern united states and southern canada due to numerous interactive stressors, including extreme blood loss, problems with thermoregulation, and incidence of pathogenic bacteria (mclaughlin and addison 1986, campbell et al. 1994, samuel 2004, musante et al. 2007, addison and mclaughlin 2014, jones 2016). as a one-host tick, it feeds, molts, and mates on the same individual moose from approximately late september to mid-april. once a female has mated and fully engorged, she drops to and crawls on the ground to lay eggs in a suitable, moisture-rich reprieve in soil and leaf litter. eggs hatch in about a month and the larvae enter a resting period during summer, regaining activity and questing for a host during autumn (drew and samuel 1985, addison and mclaughlin 1988). the moisture-rich microhabitats in the northeastern forest preferred for oviposition, hatching, and quiescence (yoder et al. 2016) expose adult females, eggs, and unfed larvae (the sole transmission stage) to numerous filamentous soil fungi (tuininga et al. 2009) that corresponding author: jay yoder, department of biology, wittenberg university, ward street at north wittenberg avenue, po. box 720, springfield, ohio 45501-0720, usa, jyoder@wittenberg.edu 41 mailto:jyoder@wittenberg.edu are principally saprobes and agents of decay, but could be infective to ticks. fungal infection is a primary source of mortality in ticks and is often the basis for their biological control as with the entomopathogenic fungi beauveria bassiana, metarhizium anisopliae, and scopulariopsis brevicaulis (kirkland et al. 2004, suleiman et al. 2013). these fungi are regular soil saprobes that produce copious amounts of spores (barnett and hunter 2003), live free in soil, and can invade a perfectly healthy tick. presumably, ticks come into contact and are infected by way of spores that adhere to the cuticular surface, germinate, and then colonize via hyphae that gain internal access via the mouth, anus, genital pore, cuticular gland openings, and between leg segments. infection often results from a single fungal species that exploits the tick (yoder et al. 2008), such that an infected tick is essentially a pure culture of the infectious agent. once the infection proliferates, the fungal hyphae typically protrude from the cuticular glands around the body. after death, the tick dries out, the body flattens, and fungal hyphae typically protrude from the mouth spreading over the front pair of legs, eventually enveloping the carcass within the fungal mycelium. an increased rate of water loss serves as an indicator of establishment and progression of the infection (cradock and needham 2011). not all entomopathogenic fungi are universally pathogenic to all ticks; for example, b. bassiana and m. anisopliae are particularly ineffective against the american dog tick (dermacentor variabilis; kirkland et al. 2004), a close relative of the winter tick. we suspected that a fungal mortality agent existed for winter ticks during routine handling of specimens used in related studies with engorged females collected from dead moose in new hampshire (yoder et al. 2016). under relatively moist and warm conditions (93% rh, 25 °c), some of the engorged females died and became moldy during storage. further, we observed healthy ticks die and became covered with a whitish mold when housed in the same storage containers with moldy ticks. the objectives of this study were to 1) isolate and determine the fungus that infected the engorged female ticks, 2) determine if the fungus was pathogenic to winter tick eggs and larvae (ground-dwelling stages), and 3) compare the relative pathogenicity of the isolated fungus with the entomopathogenic fungi b. bassiana and m. anisopliae in the context of biological control agents. methods study area winter ticks were collected from dead moose in eastern coos county, new hampshire, an area considered the best habitat and of highest moose density in new hampshire (see jones 2016). the area is dominated by mountainous terrain (elevation 300 to 1200 m) bordered by lowland valleys containing a myriad of lakes, ponds, and river systems. the dominant cover type is northern hardwood forest with a mix of american beech (fagus grandifolia), yellow birch (betula alleghaniensis), and sugar maple (acer saccharum); balsam fir (albies balsamea), red spruce (picea rubens), and white pine (pinus strobus) occur on more poorly drained sites. monthly precipitation, mean ambient temperature, precipitation, snow depth, and other weather variables were available from the national climatic data center (44827 'n, 71811 'w) weather station in berlin, new hampshire (#270690/99999) located centrally in the study area at 283 m elevation. annual ambient temperature ranged from 30 to �30 °c, annual precipitation from 91 to 123 cm, and maximum snow depth from 50 to 70 cm. tick collection aseptic technique was followed in the laboratory using materials that were autoclave-sterilized (121 °c, 19 psi, 15 min), 42 entomopathogenic fungi to moose ticks – yoder et al. alces vol. 53, 2017 flamed off using a bunsen burner, treated with 95% ethanol, or purchased sterile from the manufacturer. sterile, powder-free gloves were worn when necessary (microlex co., reno, nevada, usa). ticks were collected from 5, 10.5 monthold moose that had been captured and radiocollared in january 2015; death occurred 24-36 h prior to collection. all moose were emaciated and mortality was attributed to anemia and hypoproteinemia from excessive blood loss associated with >30,000 winter ticks/animal (jones 2016). ticks were collected from the neck, shoulder, abdomen, and rump of moose, and included nymphs and adults in various stages of feeding and unfed specimens. all ticks were identified as dermacentor albipictus from keys (brinton et al. 1965). fed females were placed individually into 50 ml polypropylene centrifuge tubes (fisher scientific, pittsburgh, pennsylvania, usa) within whirl-pak bags (nasco, salida, california, usa) and transported to the laboratory in 5 l coolers kept at ~15 °c with cold packs (koolit; fdc packaging, medfield, massachusetts, usa). in the laboratory, each fed female was transferred to a fresh tube and stored at 93% rh (sd ± 0.5% rh; winston and bates 1960) in 3000 ml glass desiccators at 25 ± 0.5 °c, 10l:14d (programmable incubator; fisher). subsequent hatched larvae were identified as d. albipictus using keys (20 slide-mounted larvae); no other species of larvae was identified. at the time of collection, all fed females were in healthy condition in that their body was plump and blood-filled, and they could crawl 5 body lengths. isolation and identification of fungi from tick cadavers the methods used to isolate fungi from moldy ticks, the preparation and spore concentration of inoculum, use of tween as an emulsifier (dispersing agent), treatment of larvae and eggs, and reisolation of fungi were modified from previous studies with ticks and entomopathogenic fungi (fernandes et al. 2004, kirkland et al. 2004, tuininga et al. 2009, suleiman et al. 2013). moldy engorged females were used to culture fungi from pieces of hyphae scraped from their carcasses. individual hyphae were plated individually onto solidified potato dextrose agar (pda; fisher) in disposable 100 � 15 mm petri plates (fisher) that were incubated in darkness at 25 °c. the fungus was purified with 3 rounds of subculturing hyphal tips, each utilizing the advancing edge of a 3-4 week old mycelium. pure cultures were identified at the university of alberta microfungus collection and herbarium (uamh) centre for global microfungal biodiversity at the gage research institute (toronto, ontario, canada). the gene sequenced for identification was nucits (internal transcribed spacer region) and primers its5/its4 were used for amplification. clustalx software in mega5 was used for aligning nucits sequences (gen bank) to compare with other species at the department of biological sciences, university of cincinnati (cincinnati, ohio, usa). preparation of fungal inoculum an aqueous inoculum was prepared from 1 month-old sporing pda cultures in phosphate buffered saline (pbs, ph 7.5) + 0.05% tween 20 (fisher). spores were scraped from the plates into pbs, purified, and the concentration was adjusted to 1.4 � 108 spores/ml with a 0.1% dye exclusion (ao spencer bright-line hemocytometer, st. louis, missouri, usa). identical preparation of aqueous inocula (each at 1.4 � 108 spores/ml in pbs + tween) was performed for beauveria bassiana, b. caledonica, and metarhizium anisopliae from the agricultural research service collection of entomopathogenic fungal cultures (arsef) associated with the united states depart‐ ment of agriculture-agricultural research service (usda-ars) (ithaca, new york, usa) (see table 1 for isolate number information). alces vol. 53, 2017 yoder et al. – entomopathogenic fungi to moose ticks 43 we used these entomopathogenic fungi as positive controls; pbs + 0.05% tween served as the negative control. treatment with inoculum ten larvae were placed into a 1.5 ml microcentrifuge tube (fisher) containing 1 ml of inoculum that was gently agitated for 2 min, and then poured onto filter paper (no. 3, whatman, hillsboro, oregon, usa). actively crawling larvae were collected and placed into a clean 1.5 ml microcentrifuge tube; a hole was punched through the tube lid and covered with mesh. it was placed at 80% rh (winston and bates 1960), 25 °c, and 10h:14h l:d cycle in a sealed glass desiccator. dead larvae were identified with a 40� light microscope; i.e., larvae with curled legs, deflated opisthosoma, and no movement. the identical inoculum treatment was used with eggs and death was assumed when the eggshell chorion showed sign of collapse; eggs in such condition fail to hatch based on water balance studies (yoder et al. 2016). the experiment was complete after all larvae died, including those in the pbs + tween control, and after egg hatching occurred. reisolation of fungi in accordance with koch's postulates as confirmation of pathogenicity, dead larvae and eggs were prepared for reisolation by fungus culturing. briefly, each was surface sterilized twice for 1 min in a mild bleach solution (18:1:1 ratio of deionoized water: absolute ethanol:5.25% naocl by volume), with a final rinse in water. it was then halved by scalpel and the portions were embedded in pda, each in its own plate, and incubated in darkness at 25 °c. tips of hyphae that could be traced as originating from the inter‐ nal body contents (40/45� microscopy) were removed as a 1 cm3 block for sub-culturing and identification. the fungus was identified with standard keys (barnett and hunter 2003) and pure culture comparison to the original fungus isolates used to prepare the inoculum. defining characteristics were identified by using colony obverse and reverse, and spore size and shape under oil (1000�). larvae were 4-6 weeks old and eggs were 2–3 weeks post-oviposition; all were healthy at the time of treatment. eggs were full and rounded, with a visible accumulation of white guanine through the eggshell that is a developmental landmark of regular embryonic development. larvae crawled about actively, could self-right, and crawl 5 body lengths. data are the mean ± se of 10 replicates from 10 specimens each (n = 100). an analysis of covariance was used to test data table 1. entomopathogenic fungi used in this study. the scopulariopsis brevicaulis isolate 11903 is the new isolate collected from dead winter ticks originating from new hampshire, usa. fungus and isolate# host origin deposited beauveria bassiana leptinotarsa decemlineata france, europe arsef1 149 (colorado potato beetle) beauveria caledonica hadenoecus cumberlandicus kentucky, usa uamh2 11821 (cave cricket) metarhizium anisopliae conoderus sp. north carolina, usa arsef 23 (click beetle) scopulariopsis brevicaulis dermacentor albipictus new hampshire, usa uamh 11903 (winter tick) 1arsef, usda-ars collection of entomopathogenic fungal cultures, ithaca, new york, usa. 2 uamh, uamh centre for global microfungal biodiversity, toronto, ontario, canada. 44 entomopathogenic fungi to moose ticks – yoder et al. alces vol. 53, 2017 (ancova; p = 0.05; spss 14.0, microsoft excel and minitab, chicago, illinois, usa). survival times were compared with the t statistic utilizing a kaplan-meier survival curve with a log rank test. an abbott correction for mortality data and logit-transformation for percentage data were used prior to analysis. water balance experiments eggs and larvae were analyzed similarly. after treatment with inoculum (4 d posttreatment), each specimen was weighed individually with a microbalance (sd ± 0.2 μg precision, ± 6 μg accuracy at 1 mg; cahn ventron co., cerritos, california, usa). this measurement was made in <1 min without enclosures or anesthesia. standard kinetic model equations were used to determine water balance characteristics based on the mass measurements (yoder et al. 2016). all specimens were predesiccated by 4-6% so that the change in mass reflected the change in body water content. the percent water content was determined by weighing the specimen (initial, fresh mass, f ), drying it to constant mass (dry mass, d) at 90 °c in a drying oven (< ± 2 ° c; blue m electric co., chicago, illinois, usa), and calculating the difference between these measurements: 100% (f � d)/f, where f � d is the water mass, m. the dehydration tolerance limit was de‐ termined by weighing the specimen, placing the specimen at 33% rh (winston and bates 1960) in a glass desiccator at 25 °c, and then reweighing the specimen at time intervals. the critical mass measurement for larvae was the point at which a larva was unable to self-right and crawl 5 body lengths, and for eggs when the eggshell chorion began to collapse. specimens at their critical mass were transferred to a 90 °c drying oven to determine dry mass (d). the difference between critical mass and dry mass was defined as the critical water mass (mc). the amount of water loss that was sustained between the initial water mass (m0) and mc was defined as the dehydration tolerance limit expressed as a percentage: 100% (mc � m0)/m0. the water loss rate (respiratory + integumental water loss) was measured at 0% rh because this is the only condition where the rate is exponential, allowing it to be derived from the slope of a regression line. the 0% rh condition was maintained with anhydrous calcium sulfate at 25 °c in a glass desiccator (drierite; 1.5 x 10�2% rh; w. a. hammond drierite company, xenia, ohio, usa). the specimen was weighed, placed at 0% rh, and reweighed 5 times at various intervals. the specimen was then transferred to a 90 °c drying oven to achieve dry mass (d) and water mass (m) was calculated by subtraction. the water loss rate (�kt) was determined from mt = m0exp-kt, where mt is the water mass at any time t, and m0 is the initial water mass. each water balance characteristic was based on a sample size of 100; 10 replicates of 10 specimens each. data (the mean ± se) were tested using ancova (p = 0.05). percentage data were logit-transformed prior to analysis, and regression lines were compared with a test to compare characteristics and slopes from multiple regression lines (spss 14.0, microsoft excel and minitab). results identification of fungus (s. brevicaulis) the fungus that was scraped and isolated from a moldy, fed female was identified as scopulariopsis brevicaulis (sacc.) bainier (uamh isolate 11903); this isolate originated from a single tick. the identification was based on sequence analysis and 99% similarity of the its region with other s. brevicaulis strains and the basic morphological characters of the group. the s. brevicaulis isolate 11903 is deposited at the uamc centre in toronto, ontario, canada (table 1). scopulariopsis brevicaulis was the most common isolate and was recovered in pure culture from 21 moldy females (identification based alces vol. 53, 2017 yoder et al. – entomopathogenic fungi to moose ticks 45 on pure culture comparison to authentic s. brevicaulis strain 11903). not all of the collected fed female ticks from moose had a fungal in‐ fection (n = 21 of >110), not all fed female ticks were collected from the same moose, and not all fed female ticks were collected from moose during the same week. the inoculums for testing were made from a single isolate culture (i.e., not batched from all 21 positive ticks). other fungi recovered less frequently (<15%) from fed female ticks included aspergillus spp., penicillium spp., and paecilomyces spp. effect of s. brevicaulis 11903 on survival on larval ticks and eggs survival was reduced in s. brevicaulistreated larvae (p < 0.05; fig. 1) that lasted 14.0 d (7.5 d for 50% of larvae) versus 18.0 d for control larvae (11.2 d for 50% of larvae). the recovery of s. brevicaulis from dead larvae at the end of the experiment was 90.4 ± 2.2% in the treated group and 6.3 ± 2.9% in the control group (p < 0.05). similarly, treated eggs had lower survival and hatching rate, and higher recovery of s. brevicaulis from unhatched eggs than the control group (table 2). effect of s. brevicaulis 11903 on water loss on larval ticks and eggs scopulariopsis brevicaulis-treated larvae lost water >2 � faster (p < 0.05) than con‐ trol larvae (4.39 ± 0.07%/h vs. 1.84 ± 0.04%/h; fig. 2). the recovery of s. brevicaulis was ~5� higher (p < 0.05) in the treated (86.7 ± 4.7%) than control group (17.8 ± 3.9%). scopulariopsis brevicaulis was re‐ covered internally from dead larvae in the treatment group, whereas only dead larvae tested positive for s. brevicaulis in the control group (p < 0.05). similarly, eggs had higher (p < 0.05) water loss rate in the treatment (1.12 ± 0.019%/h) than control group (0.71 ± 0.033%/h) (fig. 2). dead eggs also had higher (p < 0.05) internal recovery of s. brevicaulis in the treated (71.1 ± 2.9%) than control group (4.4 ± 3.9%). there was greater recovery of s. brevicaulis in larvae than eggs (p < 0.05). after 4 days post-treatment, the fresh mass, water mass, and % water content were similar for eggs and larvae in the s. brevicaulistreated and control groups (table 3), which reflected the similarity of the water:dry mass control m. anisopliae s. brevicaulis 0 20 40 60 80 100 0 2 4 6 8 10 12 14 16 18 0 20 40 60 80 100 0 2 4 6 8 10 12 14 16 18 b. caledonica b. bassiana time (days) su rv iv al o f � ck la rv ae ( % ) fig. 1. survivorship curves for unfed larvae of dermacentor albipictus after treatment with metarhizium anisopliae, scopulariopsis brevicaulis, beauveria caledonica, or b. bassiana ( in order from left to right on the graph). the control on both plots is the solid black line for comparison with the treatments. each point is the mean of 100 larvae (± se ≤ 2.1). 46 entomopathogenic fungi to moose ticks – yoder et al. alces vol. 53, 2017 ratios (m/d): 1.44 and 1.28 for larvae and 1.81 and 1.96 for eggs in the s. brevicaulis-treated and control groups, respectively. in all cases, water mass was a positive correlate of dry mass in larvae (r ≥ 0.90 for control and ≥ 0.94 for s. brevicaulis-treated) and eggs (r ≥ 0.89 for control and ≥ 0.91 for s. brevicaulis-treated) (p < 0.001). comparative observations with other entomopathogenic fungi on larval ticks and eggs larval survival was reduced (p < 0.05) by treatment with b. bassiana, b. caledonica, and m. anisopliae: 8.1, 7.0, and 8.3 d, respectively, compared to 11.2 days for 50% of control larvae (fig. 1). at the end of the table 2. mortality characteristics associated with beauveria bassiana, b. caledonica, metarhizium anisopliae, or scopulariopsis brevicaulis to eggs of dermacentor albipictus. values (the mean ± se; n = 100 eggs) followed by the same superscript letter within a column are not significantly different (p < 0.05). eggs in the control had an incubation time of 44.6 ± 2.1 days. days elapsed before treatment chorion collapsed (50% of eggs) %/100 eggs that hatched % dead eggs positive for test fungus control (pbs + tween) not observed (0.0)a 82.6 ± 5.5a 11.8 ± 2.0a test (108 spores/ml) b. bassiana 16.1 ± 3.1b 43.7 ± 5.2b 73.2 ± 2.0b b. caledonica 13.2 ± 1.7b 46.2 ± 3.0b 62.9 ± 1.4c m. anisopliae 11.0 ± 2.2c 44.1 ± 4.7b 80.4 ± 3.2d s. brevicaulis 8.4 ± 2.1d 36.6 ± 4.3c 76.2 ± 2.8b –0.25 –0.2 –0.15 –0.1 –0.05 0 0 2 4 6 8 10 ln (m t/ m 0) control +s. brevicaulis egg larva time (hours) –0.25 –0.2 –0.15 –0.1 –0.05 0 0 1 2 3 4 5 fig. 2. water loss rate of unfed larvae and eggs of dermacentor albipictus after treatment with scopulariopsis brevicaulis. the water loss rate is derived from the slope of the regression line; mt = water mass at time t; m0 = initial water mass. each point is the mean of 100 specimens. alces vol. 53, 2017 yoder et al. – entomopathogenic fungi to moose ticks 47 experiment, 86.4 ± 2.6% larvae tested positive for b. bassiana, 88.6 ± 3.1% tested positive for b. caledonica, and 92.9 ± 1.9% tested positive for m. anisopliae in their respective treatment groups; none was detected in the control groups. the 3 fungal treatment groups reduced egg survival and hatching compared to the control group (table 2). the s. brevicaulis treatment had more detrimental effect on hatching than the other fungal treatments. no dead eggs in the control tested positive for b. bassiana, b. caledonica, or m. anisopliae; however, 11.8% were positive for s. brevicaulis. the recovery of b. bassiana, b. caledonica, and m. anisopliae was consistently lower (p < 0.05) in treated eggs than treated larvae. discussion this study produced 2 novel findings: 1) that larvae and eggs of the winter tick are susceptible to fungal isolates of b. bassiana, b. caledonica, and m. anisopliae, and 2) that larvae and eggs of the winter tick are susceptible to s. brevicaulis, a common soil fungus. both b. bassiana and m. anisopliae are approved for tick control in the united states under various commercial formulation trademarks (stafford and allan 2010). beauveria caledonica is a pathogen of forest beetles and is used in formulated appli‐ cations for biological control of bark beetles (hylastes ater and hylurgus ligniperda) in new zealand (brownbridge et al. 2010). this is the first instance that an isolate of b. caledonica has been tested and shown to be pathogenic against ticks, suggesting promise for biological tick control. it is apparent that s. brevicaulis is a pathogen in the study area given its origin from host moose. although b. bassiana, m. anisopliae, b. caledonica, and s. brevicaulis were pathogenic against both winter tick larvae and eggs, larvae were more vulnerable to infection. it follows that application of an entomopathogenic agent would probably be most effective in the autumn when larvae are active. similarly, other investigators noted that eggs of other tick species are more resistant to entomopathogenic fungal infection than later life stages, and attribute this to the resistant properties of the eggshell chorion (fernandes et al. 2004). the entomopathogenic fungi tested were consistent in shared features with other ticks challenged with ento‐ mopathogenic fungi in laboratory studies (see fernandes et al. 2012). specifically, under warm temperature (25 °c) and moisture levels >80% rh favorable to ticks, there was: 1) confirmation of pathogenicity by koch's postulates, 2) suitable infection from a topical application, 3) high mortality with a 108 spores/ml concentration, and 4) post-treatment infection with 108 spores/ml causing death in approximately 10-12 days. scopulariopsis brevicaulis is distributed worldwide as a common mold in soil and table 3. water content and dehydration tolerance of unfed larvae and eggs of dermacentor albipictus that were treated with scopulariopsis brevicaulis. values (the mean ± se) followed by the same superscript letter within a column are not significantly different (p < 0.05); n = 100 specimens each. stage fresh mass (mg) water mass (mg) water content (%) dehydration tolerance (%) larva control 0.041 ± 0.008a 0.023 ± 0.006a 56.10 ± 1.43a 21.73 ± 0.62a + s. brevicaulis 0.044 ± 0.010a 0.026 ± 0.008a 59.09 ± 1.28a 24.06 ± 0.49a egg control 0.073 ± 0.012b 0.047 ± 0.005b 64.38 ± 1.23b 36.19 ± 0.82b + s. brevicaulis 0.068 ± 0.009b 0.045 ± 0.008b 66.18 ± 1.37b 38.44 ± 0.57b 48 entomopathogenic fungi to moose ticks – yoder et al. alces vol. 53, 2017 forest leaf litter, and is a common saprobe found on fur, hooves, and horns of small and large mammals (shubina et al. 2013). thus, seed ticks come into contact with s. brevicaulis spores when crawling or as eggs on the ground, or as larvae on host fur. scopulariopsis brevicaulis has also been isolated from moose dung in black spruce (picea mariana) forests in alberta, canada (listed as teleomorph m. brevicaulis uamh number 9458; isolator s. p. abbott), indicating a linkage among s. brevicaulis, moose, and moose habitat. results here for s. brevicaulis show at this inoculum dose (108 spores/ml), it kills winter tick. an extended period of wetness can trigger s. brevicaulis to proliferate and possibly induce tick mortality. for example, certain engorged specimen females died from s. brevicaulis infection after prolonged storage under high moisture (93% rh); effectively, s. brevicaulis proliferated and killed the ticks. exposure to a large inoculum of 108 spores/ml of the study set of fungi (b. bassiana, b. caledonica, m. anisopliae, and s. brevicaulis) and moisture >80% rh is a lethal combination to eggs, larvae, and adult female winter ticks. the biological significance of s. brevicaulis as a limitation on winter tick populations would presumably occur only during optimal environmental conditions (e.g., extended period of high moisture and temperature to enhance spore production), and the likelihood or frequency of such is unknown. conclusions due to the potential in mammalian nail and skin infections by s. brevicaulis (lee et al. 2012), precautionary measures should be taken if it is considered as a biological control agent. biological control of winter ticks is theoretically possible with b. bassiana and m. anisopliae at 108 spores/ml or higher formulations that are consistent with commercially available products. however, commercial applications are typically local (e.g., backyards), and arguably, impractical across moose habitat. fernandes et al. (2012) discuss concerns and issues concerning application of entomopathogenic fungi in natural settings for tick control. acknowledgements this project was made possible by grants from m. a. senich (midland, texas) to bwn and s. mcwilliam to bps, a gift from e. e. powelson (springfield, ohio) to the department of biology at wittenberg, and the new hampshire fish and game department (concord, new hampshire) to pjp. we thank y. guardiola and j. scott (uamh, toronto, ontario) for fungus identification, j. benoit (university of cincinnati, cincinnati, ohio) for sequence analysis, and h. jones and d. ellingwood of the university of new hampshire (durham, new hampshire) for collecting specimen ticks. references addison, e. m., and r. f. mclaughlin. 1988. growth and development of winter tick, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670–678. ———, and ———. 2014. shivering by captive moose infested with winter ticks. alces 50: 87–92. barnett, h. l., and b. b. hunter. 2003. illustrated genera of imperfect fungi. american phytopathological society press, st. paul, minnesota, usa. brinton, e. p., d. e. beck, and d. m. allred. 1965. identification of the adults, nymphs and larvae of ticks of the genus dermacentor koch (ixodidae) in the western united states. brigham young university scientific bulletin biological service 5: 83–96. brownbridge, m., s. reay, and n. cummings. 2010. association of entomopathogenic fungi with exotic bark beetles in new zealand pine plantations. alces vol. 53, 2017 yoder et al. – entomopathogenic fungi to moose ticks 49 mycopathologia 169: 75-80. doi: 10.1007/ s11046-009-9229-1. campbell g. d., e. m. addison, i. k. barker, and s. rosendal. 1994. erysipelothrix rhusiopathiae, serotype 17, septicemia in moose (alces alces) from algonquin park, ontario. journal of wildlife diseases 30: 436–438. doi: 10.7589/0090-3558-30.3.436. cradock, k., and g. r. needham. 2011. physiological effects upon amblyomma americanum (acari: ixodidae) infected with beauveria bassiana (ascomycota: hypocreales). experimental and applied acarology 53: 361–369. doi: 10.1007/ s10493-010-9405-5. drew, m. l., and w. m. samuel. 1985. factors affecting transmission of larval winter ticks, dermacentor albipictus (packard), to moose, alces alces l., in alberta, canada. journal of wildlife diseases 21: 274–282. fernandes, e. k. k. k., v. r. e. p. bittencourt, and d. w. roberts. 2012. perspectives on the potential of entomo‐ pathogenic fungi in biological control of ticks. experimental parasitology 130: 300–305. doi: 10.1016/j.exppara.2011.11. 004. ———, g. l. costa, a. m. l. moraes, and v. r. e. p. bittencourt. 2004. entomopathogenic potential of metarhizium anisopliae isolated from engorged females and tested in eggs and larvae of boophilus microplus (acari: ixodidae). journal of basic microbiology 44: 270–274. doi: 10.1002/jobm.200410392. jones, h. 2016. assessment of health, mortality, and population dynamics of moose in northern new hampshire during successive years of winter tick epizootics. m. s. thesis, university of new hampshire, durham, new hampshire, usa. kirkland, b. h., g. s. westwood, and n. o. keyhani. 2004. pathogenicity of entomopathogenic fungi beauveria bassiana and metarhizium anisopliae to ixodidae tick species dermacentor variabilis, rhipicephalus sanguineus, and ixodes scapularis. journal of medical entomology 41: 705–711. doi: 10.1603/0022-2585-41. lee, m. h., s. m. hwang, m. k. suh, g. y. ha, h. kim, and j. y. park. 2012. onychomycosis caused by scopulariopsis brevicaulis: report of two cases. annals of dermatology 24: 209–213. doi: 10. 5021/ad.2012.24.2.209. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. doi: 10.7589/00903558-22.4.502. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–111. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. shubina, v. s., d. y. alexandrov, and a. v. alexandrov. 2013. species composition of microsporic fungi in forest litter, burrows and fur of small mammals. mikologi i fitopatologi 6: 397–404. stafford iii, k. c., and s. a. allan. 2010. field application of entomopathogenic fungi beauveria bassiana and metarhizium anisopliae f52 (hypocreales: clavicipitaceae) for the control of ixodes scapularis (acari: ixodidae). journal of medical entomology 47: 1107–1115. doi: 10.1603/me10019. suleiman, e. a., m. t. shigidi, and s. m. hussan. 2013. activity of scopulariopsis brevicaulis on hyalomma anatolicum and amblyomma lepidum (acari: ixodidae). journal of medical science 13: 667–675. doi: 10.3923/jms.2013.667.675. tuininga, a. r., j. l. miller, s. u. morath, t. j. daniels, r. c. falco, m. marchese, s. sahabi, d. rosa, and k. c. stafford iii. 2009. isolation of entomopathogenic fungi from soils and ixodes scapularis (acari: ixodidae) ticks: 50 entomopathogenic fungi to moose ticks – yoder et al. alces vol. 53, 2017 prevalence and methods. journal of medical entomology 46: 557–565. doi: 10.1603/033.046.0321. winston, p. w., and d. s. bates. 1960. saturated solutions for the control of humidity in biological research. ecology 41: 232–237. doi: 10.2307/1931961. yoder, j. a., j. b. benoit, d. l. denlinger, j. l. tank, and l. w. zettler. 2008. an endosymbiotic conidial fungus, scopulariopsis brevicaulis, protects the american dog tick, dermacentor variabilis, from desiccation imposed by an entomopathogenic fungus. journal of invertebrate pathology 97: 119–127. doi: 10.1016/j. jip.2007.07.011. ———, p. j. pekins, h. f. jones, b. w. nelson, a. l. lorenz, and a. j. jajack. 2016. water balance attributes for offhost survival in larvae of the winter tick (dermacentor albipictus) from wild moose. international journal of acarology 42: 26-33. doi: 10.1080/ 01647954.2015. 1113310. alces vol. 53, 2017 yoder et al. – entomopathogenic fungi to moose ticks 51 susceptibility of winter tick larvae and eggs to entomopathogenic fungi -beauveria bassiana, beauveria caledonica, metarhizium anisopliae, and scopulariopsis brevicaulis introduction methods study area tick collection isolation and identification of fungi from tick cadavers preparation of fungal inoculum treatment with inoculum reisolation of fungi water balance experiments results identification of fungus (s. brevicaulis) effect of s. brevicaulis 11903 on survival on larval ticks and eggs effect of s. brevicaulis 11903 on water loss on larval ticks and eggs comparative observations with other entomopathogenic fungi on larval ticks and eggs discussion conclusions acknowledgements references alces16_69.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces 46, 2010 a journal devoted to the biology and management of moose edward m. addison ecolink science vince f. j. crichton manitoba conservation murray w. lankester lakehead university (retired) brian e. mclaren lakehead university printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 kristine m. rines new hampshire fish and game edmund s. telfer canadian wildlife service richard m. p. ward yukon department of renewable resources associate editors chief editor peter j. pekins university of new hampshire submissions editor gerald w. redmond maritime college of forest technology business editor arthur r. rodgers ontario ministry of natural resources alces vol. 34 (1), (1998) iv “albert w. franzmann and distinguished colleagues memorial award” june 2009 inspired by the passing of our beloved colleague, mentor, and friend al franzmann in february, 2009, and to honour all of those who have passed on and have contributed to our knowledge and understanding of moose biology and management, alces has established the “albert w. franzmann and distinguished colleagues memorial award.” the one-time award, valued at cdn $1,500, will be given annually to a graduate student entering or continuing1 in a masters or doctoral program at a recognized university in canada or the united states. the applicant’s research should be directed toward studies of the biology and management of moose within their circumpolar distribution or other ungulates or mammalian carnivores overlapping their range. applicants are required to submit: 1) an official (signed and sealed) academic transcript of their complete academic record; 2) an up-todate curriculum vitae; 3) a detailed description of the research to be undertaken (min. 4 pages, max. 10 pages) as would be prepared for a thesis advisory committee; and, 4) a short supportive letter from the student’s graduate supervisor. applications must be submitted2 by march 15 of the year in which the award is to be used. a committee of 3 past recipients of the distinguished moose biologist award will review applications and rank each submission. the recipient will be determined by consensus of the committee and their decision will be binding3. the recipient of the award will be announced at the annual north american moose conference and workshop4. recipients must acknowledge receipt of the award in any subsequent publications of their work and are strongly encouraged to publish at least some portion of their research in the journal alces prior to or following completion of their graduate program. donations to support the “albert w. franzmann and distinguished colleagues memorial award” should be made payable to lakehead university alces account # 50-1606-2051 and be sent to dr. art rodgers2. alces is not a registered charitable organization or incorporated as a not-for-profit corporation and cannot issue receipts for income tax purposes. 1 first-time applicants in the early phases of their research will be given preference but in the absence of suitable new applicants, consideration will be given to previous recipients pending submission of a progress report and recommendation by their graduate supervisor. previous recipients interested in reapplying should contact dr. art rodgers2 after march 15 to find out if the award is still available. the deadline for previous recipients to submit a progress report and letter of recommendation will be april 15 of the year in which the award is to be used. 2 applications must be submitted to; dr. arthur r. rodgers ontario ministry of natural resources centre for northern forest ecosystem research 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca 3 failed applicants are welcome to reapply and must do so to be considered in a future competition. 4 the recipient does not have to be in attendance but all applicants are encouraged to take advantage of the alces “newcomer’s travel award” to attend the annual meeting. alces15_213.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces19_14.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces16_124.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 distribution of winter browsing by moose: evidence of long-term stability in northern sweden r. thomas palo1, sara m. öhmark2, and glenn r. iason3 1department of wildlife, fish and environmental studies, swedish university of agricultural sciences, 90183 umeå, sweden; 2department of natural sciences, mid sweden university, 85170 sundsvall, sweden; 3the james hutton institute, craigibuckler, aberdeen ab15 8qh, scotland, united kingdom abstract: predicting spatial distribution of large herbivore foraging is important for successful management, but accurate predictions remain elusive against a background of multiple causes modified by environmental stochasticity. moose (alces alces) might prefer to browse areas with high plant density, but if snow depth co-varies with plant density, this could restrict access to these sites and force use of sites with lower plant density and snow depth. moose browsing was measured in 72 plots distributed within the subarctic birch (betula spp.) forest landscape at abisko in northern sweden in 1996. in 2010, the same plots were revisited and the measurements repeated. a generalized linear model predicted moose browsing on birch in 2010 from the browsing pattern on birch measured in 1996. the model suggested that neither total density of willow and birch stems nor snow depth were influential of foraging distribution of birch at multiple spatial scales. the spatial scale at which clustering of browsing on birch occurred, coincided with the scale of clustering of birch and willow (salix spp.) stems at distances of 1000–2500 m; at lesser distance browsing was distributed randomly. we concluded that moose demonstrate stability in spatial browsing patterns after 14 years which corresponds to 3–4 generations of moose, and that plant density represents a cue for moose only at certain scales. predictability of feeding sites is valuable for long-term moose and forest management, and conservation planning. alces vol. 51: 35–43 (2015) key words: alces alces, foraging distribution, moose, mountain birch, predictability, spatial scale, willow. introduction a key objective in the field of foraging ecology is to detect, quantify, and explain the patterns of spatial heterogeneity by feeding herbivores (bailey et al. 1996). moose (alces alces) browsing has high potential impact on the structure, dynamics, and composition of both natural and managed forests (pastor and naiman 1992, heikkilä and tuominen 2009). at the stand level in boreal forests, moose herbivory causes increased heterogeneity of vegetation due to patchy distribution of their browsing over the landscape (shipley and spalinger 1995, edenius et al. 2002). there are several possible underlying reasons for this spatial variation in foraging patterns. moose might seek the best available habitats (i.e., areas with comparatively high density and high quality of food) or they may avoid certain areas because of high intraor interspecific competition, high predation risk, or because landscape barriers prevent access (creel et al. 2005, van beest et al. 2011) resulting in a mosaic of browsing distribution. browsing density is an important predictor of home range size and browsing distribution by moose (van beest et al. 2011). although their large body size (400–600 kg) requires a large quantity of food (shipley 2010), moose tolerate a low quality diet, on a relative scale, because of the nutritional influence of allometric scaling (i.e., the bell-jarman-principle) (geist 1974, müller 35 et al. 2013). according to the snow-shrub interaction hypothesis, snow depth is positively related to leaf area index, stem diameter, and canopy height (sturm et al. 2001). deep snow may hinder moose from utilizing the best sites and force animals to feed in areas with less snow depth and poorer sites such as ridges and wind exposed areas (kelsall 1969). scale-dependency is evident in the geographical distribution of foraging sites, but also in terms of the hierarchical arrangements of plant tissues, individuals, populations, and communities upon which herbivores feed (palo et al. 1992, hodar and palo 1997). this scale-dependent heterogeneity is one fundamental factor that can restrict diet quality in a particular environment (bailey et al. 1996). both spatial and temporal variation in environmental characteristics, including food availability, influence patterns of herbivory (wiens 1989, horne and schneider 1995, kie et al. 2002, van beest et al. 2011). thus, a multi-scale perspective is useful to understand patterns of foraging distribution that underlie herbivoreplant interactions (owen-smith 2002, owen-smith et al. 2010). one approach is to determine the distribution of food resources, specifically their extent (geographical boundary), resolution (sites, plants, tissues), and complexity (diversity, interactions) across a landscape. distribution of “good patches” (i.e., areas with relatively high density of plants available to herbivores) often show spatial autocorrelation, meaning that nearby locations are more likely to have similar features than by chance alone (wagner and fortin 2005). it is expected that browsing will be clustered at the landscape scale since foraging is concentrated on certain valuable sites (shipley and spalinger 1995). the spatial scale at which foraging is clustered shows at what scale animals make foraging decisions, and it should coincide with the scale of clustering of forage that is tracked by the animal (saracco et al. 2004). since spatial distribution of high-value patches are predetermined by landscape structures such as soil nutrients, moisture, slope, and elevation, temporal consistency in foraging distribution would be expected (bjørneraas et al. 2012). further, previous observations of repeated browsing by moose on the same individual plant suggest that browsing patterns are consistent even at a landscape perspective since the same trees are visited repeatedly (danell et al. 1985). we tested this prediction by mapping the spatiotemporal distribution of winter browsing by moose in a subarctic landscape within unmanaged mountain forest in northern sweden during 2 winters, 14 years apart. we hypothesized that the same locations are used by moose repeatedly over time, and tested 2 possible explanations for spatial structuring of moose foraging: that browsing intensity at a particular site is 1) positively related to high tree species density and 2) negatively related to low snow depth. we also investigated how tree density covaries with snow depth which may hinder use of sites with relatively high tree density. study area this study was conducted in the abisko valley (68° 21′ n, 18° 49′ e) in northern sweden (fig. 1) where the low diversity of plant species (i.e., mostly birch and willow were prevalent in this scandinavian mountain ecosystem) provides herbivores with limited food choice, and consequently, study of feeding behaviour in a fairly simple system. moose occur year-round in the abisko valley but in higher numbers in winter, possibly because of its relatively low snow cover that may attract seasonally migratory moose (lundmark and ball 2008). the abisko valley receives less precipitation than the surrounding mountain region, with annual precipitation of 300 mm (abisko 36 distribution of winter browsing – palo et al. alces vol. 51, 2015 research facility). snow depth is about 50 cm at its peak in march (kohler et al. 2006), but varies spatially across the landscape. the study area is a mountain birch forest extending from lake torne träsk at 340 m to tree line at 670 m. mountain birch (betula pubescens sp. czerepanovii) is the dominant tree in the forest in mountainous areas of northern fennoscandia. it is limited in the south by boreal coniferous forest and in the north by tundra. shrubby willows (salix spp.) are also common in the region and some aspen (populus tremula) and scattered scots pine (pinus sylvestris) occur. moose have a notable impact on the vegetation as birch, young pines and willow in the abisko valley reflect continuous winter browsing (stöcklin and körner 1999). environmental conditions in the study area, as in similar subarctic environments, change abruptly within short distances, resulting in sharp contrasts in density of plants, levels of nutrients, and concentrations of plant secondary metabolites (karlsson 1991, hodar and palo 1997). mountain birch forests are typically not harvested, but infrastructural development has impacted nearby forest vegetation in the past century and natural disturbances such as insect outbreaks and prolonged frost drive vegetation change in the area (callaghan et al. 2013). tree and shrub cover have increased in the abisko valley from 1970 to 2010, mostly for mountain and dwarf birch (betula nana); willow response has been inconsistent (rundqvist et al. 2011). methods study design seven blocks ranging from 3–740 ha were placed within the 3000 ha study area in 1996. within each block, we established 10 randomly distributed 25 � 25 m sample fig. 1. location of the study sites at abisko with approximate position of blocks with sample plots in 1996 and 2010. dark grey areas are birch forest, light grey are mires, and white areas are bare ground or lakes; black is the village of abisko östra, sweden. the side of the largest block is 3 km. alces vol. 51, 2015 distribution of winter browsing – palo et al. 37 plots, with the exception of a single block with 12 plots. in this design the distance between plots increases systematically as block size increases, while the grain (i.e., the smallest pixel size is the plot: 25 � 25 m) and number of observations per block remain constant. all mountain birch and willow stems were counted in all sample plots in 1996; these measurements were repeated in 61 of 72 plots in 2010. we counted birches with accumulated bites by moose from last leaf fall to the time of investigation in february-march, hereafter denoted as “current season”. snow depth (cm) and all stem diameters at the snow surface (cm) were measured in each plot. annual snow depth in january-march was provided by the abisko research station from permanent plots in the area. we used only birch stems with diameter >1 cm for comparison between 1996 and 2010 to prevent bias in stem density due to variable annual snow depth; many of the thinnest stems are not visible in deep snow. relative moose density was estimated along 9, 1-km transects located systematically within the study area. moose tracks that intersected these transect were counted (# of tracks/km, table 1) within a 1-week period in february 1996 and 2010. statistical analyses in both sample periods (1996 and 2010) the distance matrix was used to calculate autocorrelation variograms (correlograms) at different spatial intervals (distance lags) using local moran’s i. this measurement is used to detect significant autocorrelation among factors varying over time or space (anselin 1995). values of local moran’s i were produced using the software sam (rangel et al. 2010). correlograms were drawn using r version 3.1.2 (r development core team 2014). distance lags were calculated using pairs of plots within distances of 0–5000 m with breakpoints every 500 m. distance lags beyond 5000 m contained a declining number of plot pairs and were excluded from analyses. the number of plot pairs included in each distance lag in the moran’s i analyses ranged from 290–294 for birch and 50–70 for willow. a generalized linear model (glm) was used to test for factors affecting moose browsing in the year 2010. we assumed that moose browsing in 2010 would reflect previous browsing history and current distribution of willows, because willow is preferred forage. we also assumed that variation in snow depth within the study area in 1996 reflected the conditions in 2010. we accounted for spatial dependence between sample plots with a set of 23 mean distances constructed from a distance matrix between pairs of plots within and between blocks. each data point for factors in the model is the mean at each of the 23 distances. mean distances between plots ranged within lags from 87–5133 m and the number of plot pairs within lags was 36–90. we used a backward stepwise procedure with variables removed from the model at p > 0.05. the best model was judged from akaike’s information criteria with correction for finite sample size (aicc). analyses were performed in systat 2013. table 1. densities of birch and willow, snow depth, and moose tracks/km measured in the study area near abisko, sweden, january–march 1996 and 2010. year 1996 2010 variable mean (sd) mean (±sd) p-value birch density (stems/m2) 0.13 (0.12) 0.43 (0.25) <0.001 willow density (stems/m2) 0.06 (0.06) 0.19 (0.3) <0.009 snow depth (cm) 66.9 (9.3)* 28.6 (9.6)* <0.0001 moose tracks (#/km) 16.4 (16.9) 2.2 (2.3) <0.05 *data provided by the abisko scientific research station. 38 distribution of winter browsing – palo et al. alces vol. 51, 2015 results basic forest characteristics differed between the 2 study periods as the density of birch and willow increased from 1996 to 2010 (table 1). the proportion of plots containing willow increased from 44 to 62%. there was considerable variation in stem density for both species over the landscape. birch density varied from 3–325 stems per plot in 1996 to 3–685 in 2010; similarly, willow ranged from 2–867 stems per plot. this reflected a heterogeneous landscape and a varied mosaic of forage distribution. mean snow depth and the estimated moose density (track counts) declined between 1996 and 2010 (table 1). no correlation was found between snow depth and birch and/or willow density. however, birch density in 1996 was positively correlated with birch density in 2010 (spearman’s rho = 0.39, p = 0.002); a similar correlation was not found for willow. temporal browsing on birch trees showed considerable variation ranging from 0.6–77% of birches in the sample plots in 1996 to 0.2–19% in 2010. browsing on birch was observed in 53% (sd = 17.4) of plots in 1996 and 35% (sd = 2.0) in 2010; corresponding browsing on willows was 84% and 70%, respectively. only 4% of available birch stems were browsed in 1996, declining to 0.5% in 2010. the most parsimonious model retained only browsing on birch in 1996 as a significant factor to predict browsing on birch in 2010 (table 2). the best model rendered the following equation: y ¼ 1:18 � x � 6:18; f-ratio ¼ 109:8; p < 0:0001; r2 ¼ 0:84 ð1þ where y = browsing on birch in 2010, and x = browsing on birch in 1996. the proportion of browsed birch stems was not correlated with birch density in 1996 or 2010. similarly, the proportion of browsed willows was not correlated (p > 0.05) with the density of willows, but almost all willows were browsed independent of their density. in general, for either year, browsing on willow stems did not depend on density of adult birch stems. according to moran’s i, the spatial variation of birch and willow densities, as well as of browsed birches, showed significant autocorrelation at distances between 1000– 2500 m (fig. 2). at greater distance, the number of browsed birch varied between plots more than expected by chance. spatial consistency between years was shown by moran’s i for birch densities which was consistent with the distribution of moose browsing, which was clearer in 1996 than in 2010. moran’s i for browsing on willow was inconsistent and randomly distributed at most distances (fig. 2). discussion there was a profound change in density of birch and willows in the abisko valley from 1996 to 2010 despite the absence of forest management. rundqvist et al. (2011) also found an increase in shrub and tree density in the area in recent decades. an increase in snow depth in the abisko region has occurred recently with higher precipitation table 2. stepwise selection of the best model with aicc for each proposed browsing model, northern sweden. variables are: a) moose browsing on birch in 2010, b) birches browsed in 1996, c) willows browsed in 2010, d) willow density in 2010, e) birch density in 1996, and f) snow depth in 1996. model aic (corr.) a= 1.2�b-c + d + 0.007�e-0.04�f 164.6 a= 1.2�b + 0.015�d + 0.196�e-0.04�f 160.4 a= 1.2�b + 0.196�e-0.04�f 156.7 a= 1.2�b + 0.196�e 153.3 a= 1.2�b 150.5 alces vol. 51, 2015 distribution of winter browsing – palo et al. 39 in the mountainous region (callaghan et al. 2010). contrary to this trend, snow depth was much lower in 2010 than 1996, but this did not attract more moose to the area. snow depth was not a significant parameter in the statistical model predicting browsing in 1996 or 2010. further, we did not find support for the hypothesis that birch and willow density were related to snow depth. although the average snow depth was less than the critical threshold that inhibits moose movement in both years (kelsall 1969, lundmark and ball 2008), the large variation across the landscape may result in certain sites having deep snow that hinders movement in any given year. our and the rundqvist et al. (2011) study suggest that more moose forage existed in the abisko valley in 2010 than in 1996. this difference may reflect several underlying causes including a decline in moose density that allowed increase in willow abundance and survival of young pine trees. other possible long-term factors include recovery of birch trees from insect outbreaks, change in human activities, or climate warming (emanuelsson 1987, tenow 1996, callaghan et al. 2010). an outbreak of the autumnal moth (epirrita autumnata) in the mid-1950s defoliated and killed a large proportion of the birch forest, and it is possible that birch biomass is still recovering to levels prior to the outbreak (tenow 1996). we were not able to distinguish the specific effect of moose on vegetation dynamics in the long-term, and despite large scale changes and varied snow depth in the landscape and between years, browsing distribution was unaffected and relatively unchanged in the study area. although profound changes 1000 3000 5000 –0 .5 0. 0 0. 5 1. 0 birch densities 1996 2010 a) 1000 3000 5000 –0 .3 –0 .1 0. 1 willow densitiesb) 1000 3000 5000 –0 .4 0. 0 0. 2 0. 4 browsed birches distance lags (meters) c) 1000 3000 5000 –0 .4 0. 0 0. 4 browsed willows distance lags (meters) d) lo ca l m or an ’s i lo ca l m or an ’s i fig. 2. local moran’s i for density of birches and willows (a, b) and number of browsed birches and willows (c, d) within 625 m2 sample plots in abisko, 1996 and 2010. black symbols indicate significant positive and negative autocorrelations at p < 0.05. dotted lines denote the expected value of moran’s i if observations were distributed randomly. 40 distribution of winter browsing – palo et al. alces vol. 51, 2015 occurred in vegetation cover, birch density was correlated between 1996 and 2010, without concurrent correlation for willow density. moose had high fidelity to feeding sites across the 14 years of the study, representing 3–4 generations of northern swedish moose (g. ericsson, swedish university of agricultural sciences, pers. comm.). our results correspond with that of månsson (2009) who also found that distribution of moose browsing was independent of the density of birch. although willow was browsed wherever it occurred in abisko, the presence of this preferred browse species did not result in higher browsing of birch at sites with high willow density. as expected from the theory of spatial autocorrelation, birch and willow density and moose browsing intensity had a trend of declining autocorrelation with increasing distance between plots (legendre and legendre 1998). at the smallest scale of observation, up to 500 m as indicated by moran’s i, browsing distribution is randomly dispersed. we observed a clustering of both density of birch and browsing of birch at lag distances of 1000–2500 m, indicating the spatial scale at which moose foraging choice occurs. at this lag distance, birch and willow density are cues that interact to facilitate spatial clustering, which is most visible in the 1996 data. wallgren et al. (2013) also found spatial autocorrelation for moose browsing on pine similar to birch and willow in this study. accurate prediction of animal distribution increases the probability of realizing management goals (månsson 2009). a prerequisite for making such predictions is to identify patterns of behaviour among animals that are stable over time, and to identify environmental factors that govern animal behaviour. although forage density did not explain moose browsing across all spatial scales considered in the model, the correlograms indicate that this is an important cue for moose at certain scales. the spatial distribution of preferred forage, and the animal’s perception and utilisation of it, were consistent and remained stable over a time period that spanned several generations of moose. our results indicate that spatial patterns of moose browsing across multiple scales can persist over a long time period despite changes in vegetation density and snow depth, specifically in unmanaged habitat like our study area. the ability to predict population distribution and utilization of resources relative to management strategies may be better is such situations because resource distribution remains stable for long periods of time. acknowledgments we express our gratitude to g. olsson who assisted with fieldwork in 1996 and to s. lundell and j. rauchfuss for assistance in 2010. b. g. jonsson and 2 anonymous reviewers provided valuable comments to the manuscript. the abisko scientific research station provided accommodation and logistics for the work, as well as financial support in 2010. the study was supported by a grant from sjfr (formas) in 1996 and by a faculty grant to r. t. palo and s. m. öhmark from mid sweden university. g. r. iason was supported by the scottish government’s rural and environment science and analytical services division. references anselin, l. 1995. local indicators of spatial association-lisa. geographical analysis 27: 93–115. bailey, d. w., j. e. gross, e. a. laca, l. r. rittenhouse, m. b. coughenour, d. m. swift, and p. l. sims. 1996. mechanisms that result in large herbivore grazing distribution patterns. journal of range management 49: 386–400. bjørneraas, k., i. herfindal, e.j. solberg, b.e. sæther, b. van moorter, and c. alces vol. 51, 2015 distribution of winter browsing – palo et al. 41 m. rolandsen. 2012. habitat quality in fluences population distribution, individual space use and functional responses in habitat selection by a large herbivore. oecologia 168: 231–243. callaghan, t. v, f. bergholm, t.r. christensen, c. jonasson, u. kokfelt, and m. johansson. 2010. a new climate era in the sub-arctic: accelerating climate changes and multiple impacts. geophysical research letters 37: 1–6. ———, c. jonasson, t. thierfelder, z. yang, h. hedenås, m. johansson, u. molau, r. van bogaert, a. michelsen, j. olofsson, d. gwynn-jones, s. bokhorst, g. phoenix, j. w. bjerke, h. tømmervik, t. r. christensen, e. hanna, e. k. koller and v. l. sloan. 2013. ecosystem change and stability over multiple decades in the swedish subarctic: complex processes and multiple drivers. philosophical transactions of the royal society b: biological sciences 368.1624: 1–18. creel, s., j. winnie, b. maxwell, k. hamlin, and m. creel. 2005. elk alter habitat selection as an antipredator response to wolves. ecology 86: 3387–3397. danell, k., k. huss-danell, and r. bergstrom. 1985. interactions between browsing moose and two species of birch in sweden. ecology 66: 1867–1878. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forests. silva fennica. 36: 57–67. emanuelsson, u. 1987. human influence on vegetation in the torneträsk area during the last three centuries. ecological bulletins 38: 95–111. geist, v. 1974. on the relationship of social evolution and ecology in ungulates. american zoologist 14: 205–220. heikkilä, r., and m. tuominen. 2009. the influence of moose on tree species composition in liesjärvi national park in southern finland. alces 45: 49–58. hodar, j. a., and r. t. palo. 1997. feeding by vertebrate herbivores in a chemically heterogeneous environment. ecoscience 4: 304–310. horne, j. k., and d. c. schneider. 1995. spatial variance in ecology. oikos 74: 18–26. karlsson, p. s. 1991. intraspecific variation in photosynthetic light response and photosynthetic nitrogen utilization in the mountain birch, betula pubescens ssp. tortuosa. oikos 60: 49–54. kelsall, j. 1969. structural adaptations of moose and deer for snow. journal of mammalogy 50: 302–310. kie, j. g., r. t. bowyer, m. c. nicholson, b. b. boroski, and e. r. loft. 2002. landscape heterogeneity at differing scales: effects on spatial distribution of mule deer. ecology 83: 530–544. kohler, j., o. brandt, m. johansson, and t. callaghan. 2006. a long-term arctic snow depth record from abisko, northern sweden, 1913–2004. polar research 25: 91–113. legendre, p., and l. legendre. 1998. numerical ecology. volume 24, second english edition. elsevier, amsterdam, netherlands. lundmark, c., and j. p. ball. 2008. living in snowy environments: quantifying the influence of snow on moose behavior. arctic antarctic and alpine research 40: 111–118. månsson, j. 2009. environmental variation and moose alces alces density as determinants of spatio-temporal heterogeneity in browsing. ecography 32: 601–612. müller, d. w. h., d. codron, c. meloro, a. munn, a. schwarm, j. hummel, and m. clauss. 2013. assessing the jarman-bell principle: scaling of intake, digestibility, retention time and gut fill with body mass in mammalian herbivores. comparative biochemistry and physiology-a molecular and integrative physiology 164: 129–140. 42 distribution of winter browsing – palo et al. alces vol. 51, 2015 owen-smith, r. n. 2002. adaptive herbivore ecology: from resources to populations in variable environments. cambridge university press, cambridge, united kingdom. ———, j. m. fryxell, and e. h. merrill. 2010. foraging theory upscaled: the behavioural ecology of herbivore movement. philosophical transactions of the royal society of london-series b: biological sciences 365: 2267–2278. palo, r. t., r. bergström, and k. danell. 1992. digestibility, distribution of phenols, and fiber at different twig diameters of birch in winter. implication for browsers. oikos 65: 450–454. pastor, j., and r. j. naiman. 1992. selective foraging and ecosystem processes in boreal forests. the american naturalist 39: 690–705. rangel, t. f., j. a. f. diniz, and l. m. bini. 2010. sam: a comprehensive application for spatial analysis in macroecology. ecography 33: 46–50. r development core team. 2014. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. rundqvist, s., h. hedenås, a. sandström, u. emanuelsson, h. eriksson, c. jonasson, and t. v. callaghan. 2011. tree and shrub expansion over the past 34 years at the tree-line near abisko, sweden. ambio 40: 683–692. saracco, j. f., j. a. collazo, and m. j. groom. 2004. how do frugivores track resources? insights from spatial analyses of bird foraging in a tropical forest. oecologia 139: 235–245. shipley, l. a. 2010. fifty years of food and foraging in moose: lessons in ecology from a model herbivore. alces 46: 1–13. ———, and d. e. spalinger. 1995. influence of size and density of browse patches on intake rates and foraging decisions of young moose and white-tailed deer. oecologia 104: 112–121. stöcklin, j., and c. körner. 1999. recruitment and mortality of pinus sylvestris near the nordic treeline: the role of climatic change and herbivory. ecological bulletins 47: 168–177. sturm, m., j. p. mcfadden, g. e. liston, f. stuart chapin, c. h. racine, and j. holmgren. 2001. snow-shrub interactions in arctic tundra: a hypothesis with climatic implications. journal of climate 14: 336–344. tenow, o. 1996. hazards to a mountain birch forestabisko in perspective. ecological bulletins 45: 104–114. wagner, h. h., and m.-j. fortin. 2005. spatial analysis of landscapes: concepts and statistics. ecology 86: 1975–1987. wallgren, m., r. bergström, g. bergqvist, and m. olsson. 2013. spatial distribution of browsing and tree damage by moose in young pine forests, with implications for the forest industry. forest ecology and management 305: 229–238. van beest, i. m. rivrud, l. e. loe, j. m. milner, and a. mysterud. 2011. what determines variation in home range size across spatiotemporal scales in a large browsing herbivore? journal of animal ecology 80: 771–785. wiens, j. 1989. spatial scaling in ecology. functional ecology 3: 385–397. alces vol. 51, 2015 distribution of winter browsing – palo et al. 43 distribution of winter browsing by moose: evidence of long-erm stability in northern sweden introduction study area methods study design statistical analyses results discussion acknowledgments references f:\alces\supp2\pagema~1\rus 27s alces suppl. 2, 2002 ulitin – moose hunting in russia 123 moose hunting in russia alexander a. ulitin central board, russian hunters and fisherman’s union, moscow, russia abstract: moose (alces alces) have become one of the popular big game species in russia, whereas only decades ago, low moose numbers precluded hunting. the rapid increase in moose numbers is primarily the result of forest harvest practices and intensive moose management policies. at present, according to the russia statistical committee, the moose population is stable at around 700,000 animals. use of intensive biotechnical moose management measures such as ashtree cutting, feeding of wood waste, and rock salt, combined with large scale protective measures have also favored this population increase. however, data collected by the all–union research institute show that moose density in some regions has exceeded the carrying capacity of game preserves for many years. this is the result of poor moose population estimates and low harvest rates. as a result of low harvest intensity, and in the absence of management actions aimed at increasing the carrying capacity on moose preserves, forest resources and habitat quality have been damaged in some economic regions and severely degraded in areas of the assr. the author suggests a winter feeding strategy for moose on hunting preserves that would use wood waste that is left after logging. this strategy would allow a more effective means of supplementing winter forage, but may be difficult to implement. alces supplement 2: 123-126 (2002) key words: biotechnical moose management, carrying capacity, hunting preserves, hunting societies moose (alces alces) have become one of the popular big game species in russia, whereas only decades ago, low moose numbers precluded hunting. in 1929, only 200 moose inhabited the moscow region. by 1959, moose had increased to 15,000. currently, the moscow society of hunters and anglers harvest over 1,000 moose annually from a total population of about 6,000. discussion the rapid increase in moose numbers is primarily the result of forest harvest practices and intensive moose management policies. extensive clear–cuts throughout forested regions have resulted in substantial increases in moose habitat quality. use of intensive biotechnical moose management measures such as ashtree cutting, feeding of wood waste, and rock salt, etc., combined with large scale protective measures, have also favored this population increase. at present, moose are found throughout all regions of the russian federation, with the exception of the caucuses and parts of the far east. moose numbers in the russian federation have increased considerably during the past 40 years. in 1950 there were 266,000 moose. by 1960 the moose population had almost doubled to 480,000. moose numbers continued to increase during the 1960s, although at a slower rate. according to glavokhota (personal communication) of the rsfsr, by 1966 the moose population had reached 730,000– 790,000. some researchers believe this is the highest moose abundance to occur in the past 200 years. at present, according to the russia statistical committee, the moose population moose hunting in russia – ulitin alces suppl. 2, 2002 124 is stable at around 700,000 animals. a large portion of the total moose population (253,900) inhabits the game preserves of hunting societies, mainly those of the russian hunter's and fisherman’s union (rhfu), which owns 14.5% of the hunting area in russia (table 1). in the last decade, moose in the european part of russia were concentrated in its northwestern, central, and northeastern regions. the highest densities are in central, eastern, and forest–steppe regions (table 2). on the preserves of hunter’s and angler’s societies, the increase in numbers progresses from northwest russia (leningrad region 2.0–3.2 moose/1,000 ha; novogorod region 3.8–4.6 moose/1,000 ha; kalinin region 3.6–6.4 moose/1,000 ha), through central russia (moscow, ryazan, and lipetsk regions 3.4–4.2 moose/1,000 ha), to the east (vladimir region 4.0–5.1 moose/1,000 ha; ivanovo and yaroslal regions 4.1–7.1 moose/1,000 ha); and to the south–east (mari and udmurt, assian soviet socialist republic (assr) 4.2–4.5 moose/1,000 ha; tatar assr 4.6–8.5 moose/1,000 ha; and penza region 5.0–5.6 moose/1,000 ha). moose numbers in forest–steppe regions have also increased during the last few years. in the kujbishev and orenburg regions, densities increased from 2.9 to 5.5 moose/1,000 ha and from 2.8 to 6.1 moose/ 1,000 ha, respectively. in the saratov region moose abundance increased from 4.3 to 5.5 moose/1,000 ha and then declined to 3.9 moose/1,000 ha. it should be noted that, according to the all–union research institute, moose density in some regions has exceeded the carrying capacity of game preserves for many years. this is the result of poor moose population estimates and low harvest rates. in many regions of the european part of russia, the average annual harvest of moose table 1. number of moose and moose hunts (x 1,000) in all of russia and on the preserves of the russian hunter’s and fisherman’s union (rhfu), 1976 – 1989. year number of number of moose moose hunts russia rhfu russia rhfu 1976 740 224.4 50.8 25.7 1977 750 217.4 56.3 25.5 1978 750 218.3 58.2 27.9 1979 770 213.5 57.2 27.9 1980 730 216.1 59.5 29.0 1981 790 206.0 56.8 26.6 1982 740 198.2 57.6 25.3 1983 780 190.7 58.4 26.0 1984 740 194.0 58.4 23.8 1985 780 192.7 61.1 26.5 1986 754 196.0 60.9 27.3 1987 – 201.6 – 30.8 1988 697.5 227.6 63.8 34.4 1989 253.9 36.2 table 2. average moose density in the game preserves of the russian hunters and fisherman’s union, 1985–1988. hunting regions/ density economic provinces (moose/1,000 ha)1 european north 1.7 – 1.4 southern taiga 3.6 – 4.1 central area 3.0 – 3.2 volga area 3.5 – 4.2 urals 4.4 – 4.4 caucuses 0.1 – 0.1 western siberia 2.2 – 2.4 eastern siberia 0.8 – 0.8 the far east 0.6 – 0.6 1according to the all–union research institute. alces suppl. 2, 2002 ulitin – moose hunting in russia 125 (post–commercial) generally ranged between 6.2 – 11.4% of the total population and did not exceed 25%. between 1980 and 1985 the number of moose harvested in the entire russian territory was only 3.5% greater than during the previous 5 years. as a result of low harvest intensity, and in the absence of management actions aimed at increasing the carrying capacity on moose preserves, forest resources and habitat quality have been damaged in some economic regions and severely degraded in some areas of the assr. the destructive effects of moose on forage resources are already extensive in the northern urals, central blacksoil, and particularly in the central and volga regions. as a consequence, only 19% of the area of moose preserves in the european part of russia is capable of sustaining a moose harvest without using intensive biotechnical management measures. habitat quality on 76% of the area of preserves fell below the average level and on 4.7% of the area in preserves, habitat quality declined below the level that moose hunting is economically viable. the present moose harvest level in the european part of russia and the condition of moose during winter leave no doubt that in the near future the quality of moose preserves will deteriorate. to prevent disastrous consequences, hunting preserves in some regions of the central and southern zones must intensively harvest moose to bring their numbers into balance with current carrying capacity. in some areas it may also be possible to increase carrying capacity and improve key habitats. the members of the central board of the rhfu are aware of their role in the sustainable management of moose. however, moose management is hampered by current management systems. hunters, in our case those of rhfu, do not have complete management authority for their preserves. the moose harvest is managed by republican bodies in charge of hunting. these management bodies often do not follow the recommendations of the scientific community. great problems have been caused by excessive regulation of the moose harvest along border zones between different administrative regions, in areas of seasonal concentration of moose, and along their migration routes. state control of the distribution of licenses and hunts is not effective. in my opinion, hunting societies could, however, determine reasonable moose harvest quotas and the allocation of the harvest among the various hunting groups. decisions on these issues could be made collectively at interregional meetings of the societies. analyses of current supplemental winter feeding strategies for moose on hunting preserves indicate that they provide a maximum of 24.8 – 27.6 feeding units per 1,000 ha during the winter. this is no more than 4% of the total winter forage requirement of 1 moose. a more effective means of supplementing winter forage would be to use wood waste that is left after logging. this could be done in conjunction with moose population control measures. the amount of waste that is annually destroyed at logging and storage sites is much greater than the amount of forage stored on hunting preserves. however, implementing such a strategy would be difficult, in spite of agreement between forestry and hunting preserves. as noted earlier, the preserves of the rhfu contain a large portion of the moose population in the european part of russia, and these areas account for over 50% of the moose harvest by republican hunters. on these preserves the annual moose harvest exceeds 15% of the population (table 1) compared to an average of less than 10% for all of russia. one third of moose shot are harvested moose hunting in russia – ulitin alces suppl. 2, 2002 126 under sport licenses. the remaining two thirds are harvested under commercial licenses. between 1986 and 1989, 128,700 moose were harvested on the preserves of the rhfu, producing 18,600,000 kg of valuable and delicious meat. i believe the use of the term “commercial hunt” is incorrect. hunting under both “sport” and “commercial” licenses is done by amateur hunters and according to accepted hunting ethics. the statements of some opponents that amateur hunters are engaged in commercial hunting are therefore quite erroneous. in this case, hunters of the society fulfill the social task of the state and supply local people with wild meat. f:\alces\supp2\pagema~1\rus 20s alces suppl. 2, 2002 moyseenko – components of red blood in young moose 93 components of red blood in young moose nelly a. moyseenko state university, 167 syktyvkar, komi republic, russia abstract: investigating moose domestication at the pechora–ilych reserve provided an opportunity to study the development of respiration blood activity. respiration (gas transportation) blood activity plays an important role in moose adaptation. characteristics of the gas– transporting function of blood of moose after birth are not synchronic, and this process is not completed by the age of 1 year. but the process of gas transporting in moose organs develops faster than that of domesticated hoofed animals and is dependent on their environment. alces supplement 2: 93-97 (2002) key words: agricultural animals, blood components, blood protein, hemoglobin, lactation, ontogenesis, physiological anemia, red blood cell, respiration activity materials and methods i experimented with 23 moose on the pechora–ilych reserve. blood was taken from the jugular vein and stabilized with geparin. morphofunctional parameters of red blood were determined by methods generally used in laboratory, clinical, and veterinary practice (kudrjavtsev 1952). electrophoresis of hemoglobin (hb) was done in 0.8–1.0% agar gel (agar “difko– bakto” without additional cleaning) according to the strekalov (1967a, b) method in kachmarchik (1973); modification if ph is 7.0 of k–phosphate buffer and if the ion force of k–phosphate buffer and the ion force of solution is 0.005 in gel and 0.05 in electrode vessel and in polyacrylamide gel if ph is 8.3 tris–glycine electrode buffer used for division of gel systems and buffer solutions n1 according to maurer (1971), modified for chemical polymerization. alkaline resistance of hb was determined by the modified zinger method (irzhak et al. 1985). results and discussion data from the first year of the study, including lactating moose and moose with calf–blood components, are given in table 1. erythrocyte (er) concentration in the blood of newborn moose and those 1.0–1.5 years of age was equal, but concentration was a little higher than that of lactating moose. it is also typical for intact adult animals; the characteristics of their red blood cells are given in the works of knorre and knorre (1959) and marma (1967). it is typical for mature moose. but hb concentration in blood of 1– day–old moose of postembryogenesis is 26– 27% lower than that of 1–year–olds and intact ones. this is the component that differs between moose and reindeer and domesticated, hoofed animals. we believe that this factor depends on the sedentary life, especially in postembryogenesis. fetal hemoglobin (hbf), which is considered to be the main component of total hb in blood of newborn moose, was first found by moyseenko and mochalov (1987) (fig. 1). concentration of hb is the factor that makes the affinity of newborn moose hb with oxygen higher than of adults; hb newborn moose p 50 hb is 25% higher (irzhak and gladilov 1981). during the first 2 weeks of life, microcell gipochrome anemia develops in the organs components of red blood in young moose – moyseenko alces suppl. 2, 2002 94 t a b le 1 . c h a ra c te ri s ti c s o f re d b lo o d i n t h e o n to g e n e s is o f m o o s e . d a ta a re m e a n ± s e ( ra n g e ). a g e a d u lt m o o se c h a ra c te ri st ic 1 d a y 1 w e e k 2 w e e k s 3 w e e k s 4 w e e k s 5 w e e k s 1 y e a r 1 .5 y e a rs l a c ta ti n g w it h c a lf , fe m a le s = 8 4 y e a rs e r c o n c e n tr a 6 .5 3 + 0 .2 0 4 .9 1 + 0 .2 4 5 .3 7 + 0 .1 4 5 .7 1 + 0 .2 5 6 .1 9 + 0 .1 1 6 .1 8 + 0 .0 9 6 .2 4 + 0 .1 6 6 .5 0 + 0 .2 8 4 .7 1 + 0 .1 1 5 .1 4 ti o n , m in /m m (5 .8 7 – 7 .8 5 ) (4 .3 0 – 5 .4 1 ) (4 .5 2 – 5 .9 1 ) (5 .4 6 – 6 .1 2 ) (5 .7 4 – 6 .6 9 ) (5 .9 0 – 6 .6 2 ) (5 .7 2 – 6 .7 6 ) (6 .1 1 -6 .7 5 ) (4 .0 0 – 5 .4 5 ) r e ti c u lo c y te s 2 6 .0 7 + 2 .2 4 4 5 .5 6 + 5 .7 8 7 4 .3 5 + 7 .4 3 7 4 .1 4 + 7 .3 1 6 7 .7 3 + 1 1 .3 3 1 .0 6 + 0 .3 0 5 .5 6 + 1 .7 0 2 .6 1 + 0 .4 2 8 .9 7 + 1 .9 4 1 2 .4 0 % (1 1 .9 0 – 3 3 .8 4 ) (1 4 .6 0 – 6 2 .5 0 ) (4 6 .7 1 – 1 0 9 .4 7 ) (4 5 .0 0 – 1 1 5 .0 0 ) (3 3 .8 5 – 1 2 0 .3 9 ) (0 .0 0 – 2 .2 6 ) (1 .9 0 – 1 3 .7 5 ) (1 .0 8 -4 .9 2 ) (2 .8 0 – 3 0 .5 4 ) h b c o n c e n tr a 9 .3 7 + 0 .3 6 6 .7 0 + 0 .1 3 6 .8 8 + 0 .2 7 7 .9 4 + 0 .2 2 9 .7 5 + 0 .1 0 1 1 .3 6 + 0 .2 0 1 2 .8 5 + 0 .2 6 1 2 .7 3 + 0 .0 8 1 0 .9 9 + 0 .1 9 1 1 .8 0 ti o n , g % (8 .0 0 – 1 1 .4 0 ) (6 .4 0 – 7 .2 0 ) (5 .2 0 – 7 .6 0 ) (7 .2 0 – 8 .8 0 ) (9 .3 0 – 1 0 .0 4 ) (1 1 .0 0 – 1 2 .6 0 ) (1 2 .0 0 – 1 3 .8 0 ) (1 2 .6 0 – 1 3 .0 0 ) (9 .6 0 – 1 1 .8 0 ) h e m a to c ri t, 3 4 .0 9 + 1 .1 2 2 3 .6 5 + 0 .2 0 2 4 .0 4 + 0 .9 1 2 8 .1 4 + 0 .6 4 3 2 .4 5 + 0 .5 4 3 4 .2 5 + 0 .7 7 3 6 .1 9 + 1 .1 8 3 6 .7 3 + 0 .2 3 3 3 .6 0 + 0 .7 8 3 6 .1 1 % (2 7 .5 0 – 4 0 .7 0 ) (2 2 .8 0 – 2 4 .3 1 ) (1 8 .8 0 – 2 7 .4 0 ) (2 4 .6 0 – 3 0 .6 0 ) (2 9 .8 5 – 3 5 .1 2 ) (3 2 .7 3 – 3 9 .0 5 ) (3 2 .6 7 – 4 1 .6 7 ) (3 6 .1 1 – 3 7 .4 6 ) (2 8 .4 0 – 3 8 .5 4 ) e r. v o lu m e, 5 2 .3 9 + 1 .5 9 4 8 .3 6 + 1 .8 3 4 4 .7 0 + 0 .9 9 4 9 .2 7 + 0 .9 9 5 2 .5 5 + 0 .9 8 5 5 .4 6 + 1 .0 9 5 7 .9 9 + 1 .2 1 5 6 .6 8 + 0 .9 6 7 1 .4 9 + 1 .2 8 7 0 .2 5 m k m (4 5 .0 1 – 5 9 .7 1 ) (4 3 .4 4 – 5 6 .5 3 ) (4 0 .6 1 – 4 8 .3 0 ) (4 4 .8 1 – 5 2 .0 4 ) (4 7 .4 3 – 5 6 .9 5 ) (5 1 .3 0 – 6 1 .6 9 ) (5 2 .7 9 – 6 7 .7 3 ) (5 4 .5 5 – 5 9 .9 5 ) (6 1 .8 3 – 7 7 .2 3 ) e r. d ia m et er , 6 .9 7 + 0 .0 9 6 .7 8 + 0 .0 5 6 .7 2 + 0 .0 3 6 .7 0 + 0 .0 2 6 .9 1 + 0 .0 7 6 .6 8 + 0 .0 6 6 .9 7 + 0 .0 6 6 .9 4 + 0 .0 8 7 .1 1 + 0 .0 4 7 .4 3 m km (6 .7 8 – 7 .2 5 ) (6 .6 4 – 7 .0 3 ) (6 .5 4 – 6 .8 6 ) (6 .6 3 – 6 .7 6 ) (6 .5 8 – 7 .0 6 ) (6 .2 2 – 6 .9 2 ) (6 .7 3 – 7 .0 8 ) (6 .8 6 – 7 .0 5 ) (6 .8 3 – 7 .4 3 ) e r. w id th , 1 .3 6 + 0 .0 4 1 .3 4 + 0 .0 5 1 .2 6 + 0 .0 2 1 .4 0 + 0 .0 2 1 .3 8 + 0 .0 4 1 .5 7 + 0 .0 3 1 .5 2 + 0 .0 2 1 .5 5 + 0 .0 1 1 .8 0 + 0 .0 4 1 .6 2 m km (1 .2 1 – 1 .4 5 ) (1 .2 6 – 1 .4 6 ) (1 .1 6 – 1 .3 8 ) (1 .3 0 – 1 .5 0 ) (1 .2 1 – 1 .4 9 ) (1 .4 9 – 1 .6 9 ) (1 .4 4 – 1 .6 2 ) (1 .4 8 – 1 .7 0 ) (1 .5 5 – 2 .0 5 ) e r. s p h e ri c it y 0 .1 9 5 + 0 .0 0 4 0 .1 9 9 + 0 .0 0 3 0 .2 0 0 + 0 .0 0 3 0 .2 0 0 + 0 .0 0 3 0 .2 0 0 + 0 .0 0 8 0 .2 8 3 + 0 .0 0 6 0 .2 1 8 + 0 .0 0 5 0 .2 7 7 + 0 .0 0 0 0 .2 5 4 + 0 .0 0 6 0 .2 1 8 in d ex (0 .1 7 6 – 0 .2 0 8 ) (0 .1 8 8 – 0 .2 0 8 ) (0 .1 7 3 – 0 .2 0 7 ) (0 .1 7 3 – 0 .2 0 7 ) (0 .1 9 6 – 0 .2 2 6 ) (0 .2 2 7 – 0 .2 7 1 ) (0 .2 0 2 – 0 .2 3 5 ) (0 .2 1 6 – 0 .2 1 7 ) (0 .2 1 8 – 0 .3 0 0 ) e r. s u rf a c e 9 1 .6 9 + 2 .1 7 8 6 .7 7 + 0 .9 4 8 5 .1 6 + 0 .9 4 8 4 .6 3 + 0 .4 4 8 9 .9 6 + 1 .7 4 8 5 .2 5 + 2 .3 3 9 1 .6 3 + 1 .6 3 8 8 .7 8 + 0 .0 7 9 5 .3 1 + 1 .0 7 1 0 3 .9 9 sq u a re , m k m (8 6 .5 9 – 9 9 .0 2 ) (8 4 .5 5 – 9 3 .1 4 ) (8 0 .5 7 – 8 8 .6 3 ) (8 2 .8 2 – 8 6 .0 6 ) (8 1 .5 4 – 9 3 .8 9 ) (7 2 .8 6 – 9 6 .0 0 ) (8 5 .3 0 – 9 6 .6 0 ) (8 8 .6 3 – 8 8 .9 2 ) (8 7 .8 6 – 1 0 3 .9 9 ) h b c o n c e n tr a 2 7 .5 5 + 0 .7 9 2 8 .3 4 + 0 .5 4 2 8 .6 5 + 0 .5 7 2 8 .2 6 + 0 .6 4 3 0 .8 7 + 0 .3 0 3 3 .3 3 + 0 .9 1 3 5 .6 7 + 0 .7 7 3 4 .6 7 + 0 .3 2 3 2 .7 0 + 0 .4 5 3 2 .6 8 ti o n , % (2 3 .3 2 – 3 0 .6 3 ) (2 6 .3 3 – 3 0 .6 4 ) (2 6 .2 8 – 3 1 .6 7 ) (2 5 .4 9 – 3 0 .5 1 ) (2 8 .4 7 – 3 1 .1 6 ) (2 8 .1 7 – 3 7 .1 0 ) (3 3 .1 2 – 3 7 .9 5 ) (3 3 .6 4 – 3 5 .4 9 ) (3 0 .1 0 – 3 5 .3 3 ) h b c o n te n t, 1 4 .3 1 + 0 .2 3 1 3 .6 2 + 0 .3 0 1 2 .7 7 + 0 .2 3 1 3 .9 0 + 0 .3 1 1 5 .8 0 + 0 .3 5 1 8 .3 9 + 1 .1 3 2 0 .6 2 + 0 .2 3 1 9 .6 4 + 0 .4 8 2 3 .4 3 + 0 .3 4 2 2 .9 6 p g (1 3 .0 9 – 1 5 .6 0 ) (1 2 .5 2 – 1 4 .8 8 ) (1 1 .5 0 – 1 3 .6 4 ) (1 3 .1 1 – 1 5 .7 1 ) (1 4 .0 5 – 1 7 .4 9 ) (1 7 .3 8 – 1 9 .0 3 ) (1 9 .9 1 – 2 1 .6 8 ) (1 8 .6 1 -2 1 .2 8 ) (2 1 .6 5 – 2 6 .5 0 ) components of red blood in young moose – moyseenko alces suppl. 2, 2002 96 osmotic resistance. the process of hbf synthesis is still going on. morphophysiological parameters of er are not finished during the first year of life and continue at various rates. the concentration of hb in blood increases during the first year. accordingly, hb saturation in er changes. we discovered higher concentration of hb in the blood of moose of different age groups than that described in the literature because we dealt with the result of 40 years of work on domestication of moose. nearly 40 years of selection in moose productivity and behavior should cause the changing of functionally connected organs and systems, although the latter could not have selectional symptoms. the process of hbf synthesis stops when the moose reaches the age of 1.0–1.5 months. for domestic cattle the same process lasts until the age of 1.5–2.0 and even to 5.0 months, depending on the strains of cattle and environment (mickle and merkurjeva 1963, sleptsov et al. 1977). we think that this process continues due to the rapid growth of moose after their birth, depending on environment. a high level of alkaline resistance is typical for other adult hoofed animals. but alkaline resistance of hb of moose is lower then that of reindeer. osmotic resistance of er of moose is lower than that of reindeer although the latter are larger. characteristics of the gas–transporting function of blood of moose after birth are not synchronic, and this process is not completed by the age of 1 year. but the process of gas transporting in moose organs develops faster than that of domesticated, hoofed animals and is dependent on their environment. respiration (gas transportation) activity of blood plays an important role in moose adaptation. references irzhak, l. i., and v. v gladilov. 1981. age characteristics of hb affinity with oxygen (alces alces). journal of biochemical and physiological evolution 17:66–69. (in russian). , , and n. a. moyseenko. 1985. blood respiration function during hyperoxygenation. medicina, moscow, russia. (in russian). kachmarchik, e. v. 1973. modification of zone electrophoresis of hb. agar gel laboratory work 5:308. (in russian). knorre, e. p., and e. k. knorre. 1959. investigations of some physiological characteristics of moose. proceedings o f t h e p e c h o r a – i l y c h r e s e r v e – syktyvkar 7: 133–167. (in russian). kudrjavtsev, a. a. 1952. the use of blood in veterinary diagnostics. state publishing house of agricultural literature, moscow, russia. (in russian). kushner, h. f. 1940. anemia of agricultural animals. report of the agricultural society of the ussr 27:167–170. (in russian). marma, b. b. 1967. veterinary and physiological observation of moose in the zoo. investigations of the pechora– ilych reserve–syktyvkar 12:74–86. (in russian). maurer, g. 1971. disk–electrophoresis. mir, moscow, russia. (in russian). mickle, s., and e. k. merkurjeva. 1963. fetal hb of meat cattle. scientific report of higher school, biological science 4:178–181. (in russian). moyseenko, n. a., and n. n. mochalov. 1987. ecology–physiological characteristics of red blood and energy expenditure at early postembryogenesis. investigations of the komi scientific center of the ural division of the ussr 89:135–144. (in russian). sleptsov, m. k., i. s. vasiljev, n. g. solomonov, r. t. sleptsova, and s. d. andreeva. 1977. polymorphism of yakutia agricultural animals blood proalces suppl. 2, 2002 moyseenko – components of red blood in young moose 97 tein. nauka, novosibirsk, russia. (in russian). strekalov, a. a. 1967a. the methods of hb electrophoresis. agar gel laboratory work 3:140–143. (in russian). . 1967b. improved cuvette for electrophoresis in agar gel. report of experimental biology and medicine 12:110–111. (in russian). alces15_148.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 44_front_cover v2.pdf 141 distinguished moose biologist award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public's attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (> alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual's interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3-8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special "distinguished moose biologist" presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http: //bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca alces19_222.pdf alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces vol. 19, 1983 alces 47 (2011) contents the founders: a tribute to pat karns and al elsey .............................. i giant liver fluke in north dakota moose .......... james j. maskey, jr. 1 moose experimentally infected with giant liver fluke (fascioloides magna) ............. murray w. lankester and william j. foreyt 9 the impact of human recreational activities: moose as a case study ............... wiebke neumann, göran ericsson, and holger dettki 17 aims-thermal a thermal and high resolution color camera system integrated with gis for aerial moose and deer census in northeastern vermont .................................................................................. ..... thomas l. millette, dana slaymaker, eugenio marcano, cedric alexander, and leif richardson 27 moose browsing and forest regeneration: a case study in northern new hampshire ................ daniel h. bergeron, peter j. pekins, henry f. jones, and william b. leak 39 status and management of moose in the northeastern united states ............................................. david w. wattles and stephen destefano 53 spatial and temporal characteristics of moose highway crossings during winter in the buffalo fork valley, wyoming ......................................................................... scott a. becker, ryan m. nielson, douglas g. brimeyer, and matthew j. kauffman 69 broccoli and moose, not always best served together: implementing a controlled moose hunt in maine ......... lee e. kantar 83 prudent and imprudent use of antlerless moose harvests in interior alaska ........................ donald d. young jr. and rodney d. boertje 91 using cover type composition of home ranges and vhf telemetry locations of moose to interpret aerial survey results in minnesota .. ron moen, michael e. nelson, and andy edwards 101 characteristics of post-parturition areas of moose in northeast minnesota ............................... amanda m. mcgraw, ron moen, and mike schrage 113 phenotypic variation in moose: the island rule and the moose of isle royale .................. rolf o. peterson, john a. vucetich, dean beyer, mike schrage, and jannikke räikkönen 125 (continued on inside back cover) compositional analysis of moose habitat in northeastern minnesota ............. mark s. lenarz, robert g. wright, michael w. schrage, and andrew j. edwards 135 assessment of crucial moose winter habitat in western wyoming ....................... megan a. smith, steve kilpatrick, brenda younkin, leigh work, and doug wachob 151 first nations moose hunt in ontario: a community’s perspect ives and reflections ................... joseph w. leblanc, brian e. mclaren, christopher pereira, mark bell, and sheldon atlookan 163 45th north american moose conference and workshop ................ 175 previous meeting sites................................................................................. 177 distinguished moose biologist michael w. schrage ..................... 178 distinguished moose biologist past recipients............................... 179 distinguished moose biologist award criteria ............................... 180 editorial review committee........................................................................ 181 additional copies available from: lakehead university bookstore, thunder bay, ontario, canada p7b 5e1 alces 39-46 price $40.00 canadian or u.s. each (including supplementary issues) alces 24-38 price $38.00 canadian or u.s. each (including supplementary issues) alces 17-23 price $20.00 canadian or u.s. each proceedings of the north american moose conference and workshop 8-16 price $20.00 canadian or u.s. each make cheques, money orders or purchase orders payable to lakehead university bookstore. all prices include 5% g.s.t., mailing and handling costs. prices are subject to change. acknowledgements brooke pilley worked long hours formatting and typesetting manuscripts. alces home page further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our website: http://bolt.lakeheadu.ca/~alceswww/alces.html alces15_362.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces17_xxiworkshopsessions.pdf alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 alces vol. 17, 1981 135 kjell danell distinguished moose biologist 2011 recipient the distinguished moose biologist award was presented to professor kjell danell at the 46th north american moose conference and workshop held in jackson hole, wyoming, may 23-26, 2011 in recognition of his many contributions to our understanding of moose biology and management. professor danell is senior researcher and lecturer in the department of wildlife, fish and environmental studies at the swedish university for agricultural sciences, uppsala, sweden. kjell graduated from umeå university in northern sweden with a master of science comprised of botany, chemistry, zoology, and environment protection in 1971. although kjell received formal qualifications as a teacher after his m.sc. degree, he instead moved into research and became a graduate student. he graduated with a phd in animal ecology with research focused on the ecology of the american muskrat that was an introduced species in sweden. afterward, kjell’s research quickly diversified and moose research and management became his primary interest. with the rapidly increasing number of moose in sweden that peaked at ~500,000 in the 1980s, moose became of major societal interest and importance, both pro and con. this unique ungulate population required intensified management with issues such as hunting practices, forest damage, and traffic accidents of critical importance in sweden, an extensive field of research that kjell embraced. his contributions to international moose research and management have been extensive, with numerous publications, students, and outreach activity marking his distinguished career. kjell was employed as professor in animal ecology in 1988, and is currently renowned as one of the most distinguished wildlife researchers in scandinavia. kjell has been a very productive researcher and writer, publishing well over 100 original research articles in more than 30 refereed international scientific journals including alces. importantly, he also contributes substantially to the popular literature and agencies, producing >15 articles annually. he has edited a number of books, and written book chapters and ~100 popular scientific papers in swedish and/or english. his magnum opus is the book vilt, människa samhälle (english translation: “wildlife, human, society”) on adaptive management in sweden that he recently edited with his long-time colleague and friend, professor roger bergström. this book, in many ways, comprises what kjell represents and believes in with regard to practical management and outreach; applied research that aims for sustainable resource management with broad societal acceptance. kjell has a discrete and sophisticated personality. his mind always moves in alternate ways, and his strategic approach is very result-oriented and determined. yet, so many have enjoyed and savoured their productive and enlightening sitdowns with him to discuss matters from research and practical management, to strategic passages and purely private perspectives and reflections in his exhilarating academic world. the north american moose conference and workshop is proud to recognize a humble and productive researcher that has devoted most of his life to science in general, and moose research and management in particular, professor kjell danell, the recipient of the distinguished moose biologist award in 2011. margareta stéen carl-gustaf thulin instructions for contributors to alces be used in the text for scientific names and statistical symbols. use the name-and-year system to cite published literature. cite references chronologically in the text. references – use large and small capitals for author’s last names and initials. do not use any abbreviations in the references. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. titles and all parts of tables must be typed doublespaced. tables must be constructed to fit the width of the page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use numerical superscripts to identify footnotes or asterisks for probabilities. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table columns must be generated with tab settings or a table editor. do not use spaces (i.e., the space bar). illustrations type figure captions on a separate page. identify each illustration by printing the author’s name and the figure number on the back in soft pencil. if necessary, also indicate the orientation of the illustration on the back. each illustration (either a photograph or linedrawn figure), must be of professional graphics quality, and reduced to fit into the area of either 1 (67 mm) or 2 (138 mm) columns of text by the author(s). letters and numbers on reduced figures must remain legible and be no less than 1.5 mm high after reduction. the same size and font of lettering should be used for all figures in the manuscript. photographs must be of high contrast and printed with a matte finish. typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original electronic graphics files or bitmap images (preferably as tagged image file format files) in an ibm-compatible format on 9-cm (3.5-inch) diskette or cd-rom. please submit manuscripts online at: http://alcesjournal.org if a problem is encountered, please contact: roy rea, submissions editor natural resources and environmental studies institute university of northern british columbia 3333 university way prince george, british columbia canada v2n 4z9 e-mail: reav@unbc.ca telephone: (250) 960 5833 editorial policy alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented at the annual north american moose conference and workshop, but works may be submitted directly to the editors at any time. reviewers judge submitted manuscripts on data originality, ideas, analyses, interpretation, accuracy, conciseness, clarity, appropriate subject matter, and on their contribution to existing knowledge. page charges current policies and charges are explained in a covering letter and invoice sent to authors with galley proofs. manuscript preparation authors should follow “manuscript guidelines for contributors to alces”, by rodgers et al. appearing in alces, vol. 34 (1): 1998 (available from the co-editors and associate editors). updates are posted on the alces web page; http://alcesjournal.org/publicdocs/manuscriptguidelines.pdf. copy – please provide an electronic copy of the manuscript in ms word to the submissions editor. this copy should maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. double-space and leftjustify all text. except for the first page, number all pages consecutively, including tables and figure captions. revisions should be handled similarly. corresponding author do not use a title page. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlespaced in the upper left corner of the first page. if available, the author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (<10 words) begins left justified on the next line. type the title in upper-case bold letters. do not use abbreviations or scientific names in the title. abstract & key words following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. do not use abbreviations or literature citations. type alces vol. 00: 000 000 (0000), right justified on the line following the abstract. after leaving a single blank line, provide 6-12 key words in alphabetical order. footnotes use only in tables and at the bottom of the first page to provide the present address of an author when it differs from the address at the time of the study. style accompany the first mention of a common name with its scientific name. do not use scientific names for the names of domesticated animals or cultivated plants. use système international d’unités (si) units and symbols. use digits for numbers unless the number is the first word of a sentence, in which case it is spelled out. italics should only alces vol. 45, 2009 härkönen et al. wood quality of birch and moose 67 wood quality of birch (betula spp.) trees damaged by moose sauli härkönen1,3, arto pulkkinen2, and henrik heräjärvi1 1finnish forest research institute, joensuu research unit, p.o. box 68, fi-80101 joensuu, finland; 2university of joensuu, faculty of forest sciences, p.o. box 111, fi-80101 joensuu, finland abstract: european white birch (betula pubescens) and silver birch (b. pendula) are important tree species for finnish pulp and wood-products industries. moose (alces alces) damage, however, reduces the quality of butt logs intended for high-quality plywood and saw logs. in addition to flaws in stem form, pith discoloration and color change outside the pith reduce quality and value of logs irrespective of their end use. our objectives were to 1) analyze the external and internal quality of birch trees damaged by moose, 2) measure whether the severity, type, and occurrence of damage differed between silver birch and european white birch trees, and 3) evaluate visual criteria that would enable a forest-owner to assess damage and future value of moose-damaged birch trees prior to the first commercial thinning. we sampled 4 stands with a known history of moose damage; 18 trees per stand were classified by visual evaluation into 3 damage categories. the severity and type of damage lowering the internal quality of logs from sample trees were classified into 5 grades. the proportion of all visible color defects and/or decay was 74% in silver birch trees and 67% in white birch trees. moose damage caused no visible color defect and/or decay in 35% of silver birch and 33% of white birch trees. the commercial quality and value of birch trees damaged by moose was reduced by the internal color defects and/or decay, even in certain trees without obvious external moose damage. nevertheless, forest-owners can evaluate the internal quality of most birch trees in order to remove those of low-quality in the first commercial thinning by using external quality indicators of moose-damaged stems (e.g., stem form and clear curve at the point of stem breakage). alces vol. 45: 67-72 (2009) key words: alces alces, betula pendula, betula pubescens, birch, browsing, damage, decay, discoloration, moose, wood quality. european white birch (betula pubescens) and silver birch (b. pendula) are the third and fourth most common tree species in finland. the pulp industries and wood-products industries used about 14.5 and 1.5 mill. m3 of the production of 16 mill. m3 of birch roundwood in 2006. birch species are considered as medium-preferred browse of moose (alces alces) and a high proportion of their annual browse consumption consists of birch owing to its widespread availability (bergström and hjeljord 1987). this browsing may cause substantial damage and financial loss in young birch stands. moose population density has increased in finland in recent decades (torvelainen 2007). the post-harvest moose population peaked in 2001 when it was estimated at 139,000 (mean moose density of 3.3 moose/10 km2). concurrently, increasing moose damage (i.e., twig-browsing, stem breakage, and bark stripping) was raising increased concern amongst forest-owners and associated forest industries. concern is based on the fact that, as a longterm consequence, moose damage reduces the quality of butt logs (i.e., merchantable timber that is intended as high-quality plywood or sawn timber), especially when main stems are broken (heikkilä et al. 1993, ingemarson et al. 2007, lilja and heikkilä 2007). in addition to flaws in stem form, pith discoloration and color change outside the pith reduce quality, hence 3present address: hunters central organization, fantsintie 13-14, fi-00890 helsinki, finland wood quality of birch and moose härkönen et al. alces vol. 45, 2009 68 the value of logs irrespective of their end use. as a consequence of damage risk from a high moose population, plantations of birch trees have declined markedly, especially in southern finland in the last decade (viiri 2007). the objectives of this study were to: 1) analyze the external and internal quality of birch trees damaged by moose, 2) determine whether any difference in severity, type, and occurrence of damage exists between silver birch and european white birch, and 3) determine selection rules based on visual evaluation that would enable a forest-owner to decide whether to remove or retain moosedamaged birch trees in the course of the first commercial thinning. material and methods data were collected in 2007 from 1 european white birch stand and 3 silver birch stands with a known history of moose browsing. all 4 stands were in central finland (kannonkoski, saarijärvi, and viitasaari municipalities, ca. 63˚n, 25–26˚e) and had reached the growth stage for the first commercial thinning (table 1). all stands had been planted and suffered from severe moose browsing damage at the sapling stage. the randomly selected sample trees (18 per site) were classified by visual evaluation into 3 damage categories: 1) 6 trees with no visible moose damage (trees were known to have had previous moose damage, i.e., stem breakage), 2) 6 trees with slight moose damage (i.e., slightly visible curve at the point of stem breakage), and 3) 6 trees with moderate moose damage (i.e., moderately visible curve at the point of stem breakage). consequently, this classification excluded birch with severe visible moose damage. we justified this approach because it is known that the inner quality of birch trees is reduced or lost when moose damage is highly visible (heikkilä et al. 1993, lilja and heikkilä 2007). sample trees were felled and two, 2 m-long logs were cut from each sample tree, yielding a total of 144 logs. the stem form of each log was measured as a maximum deviation from the center line of the log. the logs were sawn into one, 20 cm-long bolt and six, 30 cm-long bolts. next, each bolt was pith-centrally sawn into cants using a band saw. the severity and type of damage lowering the internal quality of log (i.e., color defects and/or decay) were classified into 5 grades: 1 = no damage (no visible color defects or decay); 2 = color defect in pith with a diameter <20 mm (slight color change); 3 = hard rot in pith with a diameter <20 mm (clear color change, usually as a result of chemical reaction or preliminary stage of decay); 4 = hard rot (the wood material dark but still hard, a condition caused by an infection related to a decaying fungus); 5 = soft rot (wood material is dark and soft, and wears away when scratched) (see schatz et al. 2008). in addition, moose damage was separated from other damaging agents (i.e., insects, voles, and others) by visual evaluation. the spreading distance of the moose-caused color defects and/or decay in the stem wood was measured both vertically and horizontally. the maximum spreading distance was used in the calculations. all statistical analyses were performed stand tree species planting year density (stems/ha) mean height (m) mean dbh1(cm) a silver birch 1991 1,725 15.8 12.0 b silver birch 1987 900 15.9 12.5 c silver birch 1987 1,450 14.0 11.1 d white birch 1987 1,950 10.7 10.6 table 1. stand characteristics of three silver birch stands and one white birch stand, central finland, 2007. 1 mean dbh = mean diameter at breast height (1.3 m). alces vol. 45, 2009 härkönen et al. wood quality of birch and moose 69 with spss package. the parametric tests (anova) were employed because the variables had normal distributions. results when combining all damage categories, the proportion of all visible color defects and/ or decay was 74% in silver birch trees and 67% in white birch trees (table 2). moose damage alone did not cause any visible color defects and/or decay in 35% of silver birch and 33% of white birch trees (table 3). in both species the most common damage type (>50%) was hard rot in pith with a diameter <20 mm; this exceeded or equaled the proportion of no damage (tables 2 and 3). in practice, this means a clear color change is visible in the stem wood from a chemical reaction or preliminary stage of decay. the moose-caused color defects and/or decay were spread both vertically and horizontally in damaged trees, as well as trees with no visible damage (table 4). there was no difference in the vertical spreading distance of moose-caused color defects and/or decay in the stem wood among different damage categories of moose-damaged silver birch trees (f = 0.17, p >0.05). in white birch trees, the vertical spreading distance increased considerably (>50%) with increasing damage level, but no difference was found (f = 1.20, p >0.05). all horizontal spreading distances were <40 mm and were not different (p >0.05). vertical spreading distance averaged 160 cm in the moderately damaged white birch trees. in silver birch trees, the maximum deviation from the center line of the log increased with increasing damage level (f = 5.38, p <0.01). discussion our data indicate that wood discoloration caused by different damaging agents remains at a lower level in european white birch than silver birch (table 2), although this difference was small. there was no moose-caused visible color defect and/or decay in 35% of silver birch and 33% of white birch trees based on the assessment of internal wood quality (table 3). thus, moose damage apparently does not always result in reduced wood quality because all trees were damaged by moose at some point. on the other hand, these data indicate that from a forest-owner perspective, the commercial quality and value of birch trees damaged by moose was substantially reduced due to internal color defects and/or decay. this was also the case for birch trees that were evaluated visually to have no external moose damage (table 4), hence, there was no indication of the compromised internal quality. importantly, these trees would normally be retained in the course of the first commercial thinning, therefore, visual assessments alone prior to the first commercial thinning will probably result in some low-quality birch trees being retained until maturity. such timber at harvest will necessarily be pulpwood with only 30–50% value in comparison to saw or plywood logs. nevertheless, for the worst tree species damage type no damage (%) color defect in pith (%) hard rot in pith (%) hard rot (%) soft rot (%) silver birch 25.9 11.1 35.2 18.5 9.3 (n = 54) white birch 33.3 5.6 44.4 5.6 11.1 (n = 18) table 2. the proportion (%) of damage type measured in logs cut from silver and white birch trees identified as damaged by moose browsing. logs were graded to the damage type that most lowered the internal quality (e.g., log with soft rot may also contain hard rot, hard rot in pith, or color defect in pith). these data reflect pooling of moose and other damage agents. wood quality of birch and moose härkönen et al. alces vol. 45, 2009 70 cases forest-owners can visually evaluate the internal quality of moose-damaged birch trees with external quality indicators (e.g., stem form, clear curve at the point of stem breakage), and be able to remove the lowest-quality trees in the first commercial thinning. if the first commercial thinning is carried out properly, potential damage and economic loss caused by moose browsing will be reduced. the horizontal spreading distance of moose-caused color defects and/or decay was relatively limited in size (<40 mm in different damage categories), as reported previously (heikkilä et al. 1993, lilja and heikkilä 2007). the vertical spreading distance of moosecaused color defects and/or decay was comparable to that measured by lilja and heikkilä (2007). the maximum deviations from the center line of the logs were relatively low in both birch species, and probably reflected the exclusion of sample trees with severe visible damage. this also indicates that the diameter of broken main stems at the sapling stage might have been relatively small in both species (cf., heikkilä et al. 1993). the internal quality of birch trees was reduced by factors other than moose (table 2). it is known that insects (annila 1979, lilja and heikkilä 2007) and voles (henttonen et al. 1994) are potential damaging agents, and other related data and our observations suggest such also (s. härkönen et al., finnish forest tree species damage type no damage (%) color defect in pith (%) hard rot in pith (%) hard rot (%) soft rot (%) silver birch 35.2 11.1 33.3 14.8 5.6 (n = 54) white birch 33.3 5.6 44.4 5.6 11.1 (n = 18) table 3. the proportion (%) of damage type measured in logs cut from silver and white birch trees identified as damaged by moose browsing. logs were graded to the damage type that most lowered the internal quality (e.g., log with soft rot may also contain hard rot, hard rot in pith, or color defect in pith). classification is based on the effect of moose damage; other damage agents were excluded but may have been present. tree species damage category vertical (cm) horizontal (mm) deviation (mm) silver birch none (7) 143±16 17±4 29±5a slight (13) 157±28 33±7 41±6ab moderate (15) 142±14 37±5 60±7b f 0.17 2.07 5.38 p 0.84 0.14 0.01 white birch none (3) 39±17 16±4 28±5 slight (5) 95±39 13±2 40±11 moderate (4) 160±72 26±6 32±8 f 1.21 2.54 0.45 p 0.34 0.13 0.65 table 4. damage categories, mean vertical and horizontal spreading distances of moose-caused color defects and/or decay in stem wood, and mean maximum deviation from the center line of the log in moose-damaged silver birch and white birch trees. the 3 damage categories were: none = trees with no visible moose damage, slight = trees with slight moose damage, and moderate = trees with moderate moose damage. sample sizes are in parentheses; means (± se) with the same letter are not different (anova, p >0.05). alces vol. 45, 2009 härkönen et al. wood quality of birch and moose 71 research institute, unpubl. data). in addition, pruning, sapping, and other wounds caused by careless thinning activities may also cause color defects in stem wood (nevalainen 2006, schatz et al. 2008). hallaksela and niemistö (1998) showed that planted silver birch trees may easily have stem discoloration from dead and broken branches, and that discoloration was connected with microbial invasion in 83% of their sample birch trees. we did not determine the microbes associated with damage, but different basidiomycotina and stain fungi species have been isolated in moose-damaged birch trees (heikkilä et al. 1993). it is evident that moose damage will lower the external and internal quality of birch trees. thus, preventive measures may be required depending on the level of moose browsing and the desired timber product, and whether a forest-owner wants to ensure high quality birch trees at harvest. various chemical repellents, visual and acoustic devices, and tree sheltering methods and devices have all been used to prevent moose damage in seedling and sapling birch stands. these methods are rather expensive, their effects are variable, and in many cases they have shown little promise for reducing moose damage on a large-scale or long-term basis. thus, development of cost-effective mechanical and/or chemical preventive methods is still needed to reduce the risk of moose damage in young birch stands. lower moose density may be the most cost-effective approach to reduce damage caused by moose, however, various moose-interest groups often have conflicting values and goals with respect to an ideal moose population density (aarnio et al. 2008). we suggest that moderate moose population densities would provide for both sustainable and profitable forestry producing high-quality timber, as well as socio-economically acceptable management of moose. acknowledgements we thank mr. pertti hokkanen and mr. hannu koivunen (finnish forest research institute, joensuu research unit) for their skillful technical help during fieldwork and laboratory tasks, respectively. upm forest and two private forest-owners kindly provided the sample trees for this study. we are also grateful to kristine m. rines, pete pekins, and 2 anonymous referees for their valuable comments on the manuscript. the finnish ministry of agriculture and forestry is gratefully acknowledged for their financial support. references aarnio, j., s. härkönen, l. petäjistö, and a. selby. 2008. hirvikannan nykyisen säätelyjärjestelmän ajanmukaisuus ja toimivuus riistanhoitopiirien hallitusten näkökulmasta. (the finnish moose man-(the finnish moose management system from the perspective of board members of game management districs.) metlan työraportteja /working papers of the finnish forest research institute 92. (in finnish). annila, e. 1979. damage by phyllobius weevils (coleoptera: curculionidea) in a birch plantation. communicationes instituti forestalis fenniae 97(3): 1-20. (in finnish with english summary). bergström, r., and o. hjeljord. 1987. moose and vegetation interaction in northwestern europe and poland. swedish wildlife research supplement 1: 213-228. hallaksela, a.-m., and p. niemistö. 1998. stem discoloration of planted silver birch. scandinavian journal of forest research 13: 169-176. heikkilä, r., a. lilja, and s. härkönen. 1993. recovery of young betula pendula trees after stem breakage. folia forestalia 809. (in finnish with english summary). henttonen, h., a. lilja, and j. niemimaa. 1994. myyrien ja hyönteisten aiheuttamat sieni-infektiot koivun taimien uhkana. (microbial infections in vole and insectdamaged birch trees.) metsäntutkimuslaiwood quality of birch and moose härkönen et al. alces vol. 45, 2009 72 toksen tiedonantoja 496: 125-129. (in finnish). ingemarson, f., s. claesson, and t. thuresson. 2007. costs and benefits of moose and roe deer populations. skogsstyrelsen, jönköping, sweden. rapport 3. (in swedish with english summary). lilja, a., and r. heikkilä. 2007. väriviat hirvien taittamissa koivuissa. (stem discolorations in moose-damaged birch trees.) metsätieteen aikakauskirja 2/2007: 127-129. (in finnish). nevalainen, s. 2006. discolouration of birch after sapping. in h. solheim and a. m. hietala, editors. forest pathology research in the nordic and baltic countries 2005. proceedings from the sns meeting in forest pathology at skogbrukets kursinstitutt, biri, norway, 28-31 august 2005. aktuelt fra skogforskningen 1/06: 32-36. schatz, u., h. heräjärvi, k. kannisto, and m. rantatalo. 2008. influence of saw and secateur pruning on stem discolouration, wound cicatrisation and diameter growth of betula pendula. silva fennica 42: 295-305. torvelainen, j. 2007. multiple-use forestry. pages 201-218 in a. peltola, editor. finnish statistical yearbook of forestry 2007. agriculture, forestry and fishery 2007. finnish forest research institute. viiri, h. 2007. syökö hirvi metsänuudista-2007. syökö hirvi metsänuudistamisen monimuotoisuuden? (does moose browsing threaten the biodiversity of forest regeneration?) metsätieteen aikakauskirja 2/2007: 133-136. (in finnish). alces18_17.pdf alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 alces vol. 18, 1982 f:\alces\supp2\pagema~1\rus 16s alces suppl. 2, 2002 kuznetsov faeces as indicators of moose activity 71 faeces as indicators of moose activity and role in ecosystems german v. kuznetsov institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: faeces can serve as a significant index of spatial distribution of moose and provide estimates of moose population composition. using faeces of moose to gather data for determining forage pressure and level of removal of plants, territorial distribution, population density, traces of activity, and other population indices can be regarded as a simple and reliable method for investigation of moose ecology. alces supplement 2: 71-76 (2002) key words: ecosystems, faeces, forage digestibility, moose, plant mass, valdai by consuming plants, excreting faeces and urine into the environment, maintaining metabolic processes of the body, and through diurnal and seasonal displacements, moose are involved in transfer of energy and metabolism of matter in forest ecosystems (kusnetsov 1976). faeces can serve as a significant index of spatial distribution (yurgenson 1961, 1970), the amount of consumed production (semenov-tyanshansky 1948, kuznetsov 1975), transfer of energy (kuznetsov 1976), and other forms of moose involvement in ecosystems. estimation of ungulate faeces as a method for the evaluation of the numbers of moose and spatial distribution first originated in the usa as early as the 1950s, and in the ussr this method was first applied in 1957 by p.b.yurgenson, who proposed an original variant based on the common principles of the macquane method adopting an almost constant number of defaecations per day in ungulates (yurgenson 1970). subsequently, faeces as an indicator were successfully used to determine the foraging pressure of moose and their winter distribution in various hunting grounds and habitats of european russia (pivovarova 1965, ivanova 1967). our studies in valdai based on an annual estimate of moose faeces along a 3 km transect with constant sampling sites (300 m2 every 50 metres), demonstrated varying pressure of moose on particular habitats. in fact, the amount of faeces, on average, on the sample site over the year reached 147 ± 46 pellets in the pine forest with cowberry, 85 ± 27 pellets in pine forest with green moss, 72 ± 11 in the pine forest with sphagnum, 60 ± 10 in spruce forest with herbs, 59 ± 9 in a mixed sphagnum forest, 49 ± 11 in a spruce forest with green moss, and 5 ± 1.5 in a mixed sphagnum forest (fig. 1). over the 20 years of observations (1970–1990), the greatest number of faeces in the pine forest with sphagnum was recorded > 7 times, in the pine forest with cowberry 5 times, in the pine forest with green moss 2 times, and in the spruce forest and other types of forest 1 time, which is indicative of different trophic and topical pressure of moose in these types of forest. interestingly, in valdai we recorded dissimilar rates of visitation by moose of particular habitats (fig. 2); this is true of other regions as well (the tula region). thus, depending on the amount of forage faeces as indicators of moose activity kuznetsov alces suppl. 2, 2002 74 pivovarova 1965). such data make it possible to estimate the involvement of moose in transfer of energy in different biocenoses. it is important to take into account the existence on the same ranges of the migrating and sedentary part of moose populations (knorre 1959, baskin 1984), which influences the complex biological nature of the role of moose in the ecosystems. the method for determining the amount of plant mass consumed by moose according to their faeces was first applied by semenov-tyan-shansky (1948). the use of this method in nature necessitates data on the amount of faeces (in dry weight) excreted by moose in a particular area, the period of time over which the moose defaecated, and the coefficient of forage digestibility. on the whole, it can be assumed that the mean annual coefficient of food digestibility in moose is close to 70%. it does not exceed 50–55% in winter and reaches up to 84% in summer (ivanova and simakov 1975; kuznetsov 1975, 1976). thus, on obtaining data on the amount of faeces excreted by moose (dry weight) over the year and in a definite area (the constant sample sites are preliminarily cleared of old faeces) and knowing the coefficient of forage digestibility, one can easily estimate the amount of forage consumed by moose over the year in a particular territory, using the formula for the digestibility coefficient (kuznetsov 1975, 1976). subsequently, abaturov (1980, 1984) proposed estimating the amount of vegetation consumed according to dry weight of the accumulated faeces and coefficient of forage digestibility by the formula: c = (fx100)/(100-b); where, c = consumed phytomass (dry weight), f = dry weight of the faeces, and b = coefficient of digestibility of the phytomass of the animal species (%); i.e., a mathematically proposed form is deduced for the digestibility coefficient. in this respect, the method of milner (1967) is also of interest. milner estimated the amount of forage consumed by animals from the ratio of the weight of faeces to the forage consumed as established in the laboratory. subsequently, by estimating the amount of faeces left by animals over a certain period, one can determine the amount of the forage consumed: c = f/(f/c) (dinesman and khodashova 1974). in the field, the amount of forage consumed by moose was determined from their faeces in the population of the byelorussian lake region (dunin 1989) and valdai (kuznetsov 1975, 1976). this method makes it possible to objectively evaluate the removal of forage by moose in a definite territory without estimation of their population. along with that, a number of authors (zlotin and khodashova 1974, abaturov 1984) indicate that the accuracy of the results can be affected by the activity of microorganisms and coprophagic animals, and also the effect of the physical factors of the environment in particular geographical zones. for instance, in the subarctic, conditions for the rate of decomposition of moose faeces is low. the mass of faeces over 4 years declines by 53% (malafeev and kryazhimsky 1990). according to our data, under valdai conditions, the mass of moose faeces decreases by 64% over 4 years, and by 89% over 6 years. one can estimate the population density of moose according to the amount of faeces deposited by a moose population over a certain time in a particular territory (yurgenson 1961, 1970; kuznetsov 1976). based on the fact that every adult moose over a year excretes 2 kg of faeces (dry weight) every day (this is in conformity with the data of knorre and knorre 1959), we conclude that the population density of moose in valdai, the amount of faeces excreted by moose in this area over 1 year being 2.9 kg/ ha, represents 1 individual per 252 ha; i.e., 4 individuals per 1,000 ha (kuznetsov 1976). faeces as indicators of moose activity kuznetsov alces suppl. 2, 2002 76 pages 281–287 in a. g. bannikov, editor. biology and harvest of moose. rosselkhozizdat, moscow, russia. (in russian). , and a. f. simakov. 1975. protein metabolism. pages 63–127 in comparative physiological study of protein and mineral metabolism in ruminants under northern conditions. viniti, moscow, russia. (in russian). knorre, e. p. 1959. ecology of moose. proceedings of the pechoro-ilych state reserve 7:5–22. (in russian). , and e. k. knorre. 1959. investigations of some physiological characteristics of moose. proceedings of the pechora–ilych reserve–syktyvkar 7: 133–167. (in russian). kuznetsov, g. v. 1975. estimation of the consumption of plant production by moose from their faeces. pages 176– 177 in ungulates of the ussr fauna. nauka, moscow, russia. (in russian). . 1976. the role of moose in energy transfer in forest biogeocenoses. pages 140–147 in soils and productivity of plant communities. volume 3. mgu, moscow, russia. (in russian). malafeev, y. m., and f. v. kryazhimsky. 1990. the rate of decomposition of moose faeces in the subarctic. page 68 in proceedings of the third international moose symposium. abstracts. syktyvkar, russia. (in russian). milner, r. s. 1967. the estimation of energy flow through populations of large herbivorous mammals. page 171 in secondary productivity of terrestrial ecosystems. warzawa-krakow, poland. padaiga, v. i. 1970. methods for the regulation of deer population in an intensive forest economy. kaunas, lithuania. (in russian). pivovarova, e. p. 1965. on winter distribution of moose in game management units of the vladimir and kaluga regions. pages 106–112 in a. g. bannikov, editor. biology and harvest of moose. volume 2. rosselkhozizdat, moscow, russia. (in russian). rukovsky, n. n. 1984. a path-finder hunter. fizkultura i sport, moscow, russia. (in russian). semenov-tyan-shansky, o. n. 1948. moose on the kola peninsula. proceedings of the lapland state reserve 2:91–162. (in russian). ustinov, s. k. 1964. some problems of the biology of the moose of the barguzin range. pages 142–153 in a. g. bannikov, editor. biology and harvest of the moose. rosselkhozizdat, moscow, russia. (in russian). yurgenson, p. b. 1961. census of moose and estimation of their winter activity in forests of the temperature zone by spring census of the number of defecations. proceedings of the priosko-terrasny reserve 3:48–71. (in russian). . 1970. winter census of game mammals and birds by their faeces. pages 287-288 in proceedings of the 9th international congress of game biologists. moscow, russia. (in russian). zlotin, r. i., and k. s. khodashova. 1974. the role of animals in biological turnover of forest-steppe ecosystems. nauka, moscow, russia. (in russian). alces 50, 2014 a journal devoted to the biology and management of moose chief editor peter j. pekins university of new hampshire submissions editor roy v. rea university of northern british columbia business editor arthur r. rodgers ontario ministry of natural resources associate editors edward m. addison ecolink science vince f. j. crichton manitoba conservation murray w. lankester lakehead university (retired) brian e. mclaren lakehead university ron moen university of minnesota gerald w. redmond maritime college of forest technology kristine m. rines new hampshire fish and game printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 the status and management of moose in north america – circa 2015 h. r. timmermann1 and arthur r. rodgers2 1rr #2 nolalu, ontario, canada pot 2k0; 2ontario ministry of natural resources and forestry, centre for northern ecosystem research, 103–421 james street south, thunder bay, onatario, canada p7e 2v6 abstract: both declining and increasing moose (alces alces) populations have been reported across north america over the last decade. we surveyed all jurisdictions with extant moose populations to determine the extent of these population trends. in 2014–2015, the north american moose population was estimated at ~1,000,000 animals distributed in 30 jurisdictions, which is unchanged since the turn of the century. populations occurred in 12 canadian provinces or territories, and in at least 18 states. in the past 5 years, moose density is believed to be increasing in 9, relatively stable in 8, and declining in 11 jurisdictions; estimates of change were unavailable in 2 jurisdictions. in 2014–2015, an estimated 425,537 licensed moose hunters harvested 82,096 moose in 23 jurisdictions. hunter numbers increased by 39,118, whereas total harvest remained virtually unchanged from a decade earlier. harvests by indigenous and subsistence users, although largely unquantified, are believed substantial and important to quantify in certain jurisdictions. a variety of active and passive harvest strategies used to manage moose are discussed. alces vol. 53: 1–22 (2017) key words: alces alces, distribution, harvest, hunter numbers, indigenous hunters, licensed qualifications, moose population status, national parks, seasons, subsistence over the last decade there have been several reports of declining moose (alces alces) populations across north america (lenarz et al. 2010, smith et al. 2011, decesare et al. 2014), but there have also been accounts of increasing numbers in other areas (wattles and destefano 2011, harris et al. 2015, laforge et al. 2016, tape et al. 2016). in this paper, we update the status and management of north american moose circa 2014–2015 from that reported in 2000–2001 (timmermann 2003) to determine the extent of these population trends across the continent. a comprehensive 9-page questionnaire (located at http://alcesjournal.org/index.php/ alces) similar to that employed previously (timmermann 1987, timmermann and buss 1995, timmermann 2003), and a literature review were used to update the status, population estimates, and harvest and non-harvest management strategies used in 23 jurisdictions with an annual licensed moose harvest. an additional 7 jurisdictions where hunting is currently prohibited were contacted to determine population status. tabulated data were returned for final perusal, edits, or corrections solicited. this paper reports on current (year 2014–2015) population status and strategies used to manage hunting harvest and nonharvest management of moose across north america. affiliations of those providing information through personal communication (pers. comm.) are provided in acknowledgements. 1 http://alcesjournal.org/index.php/alces http://alcesjournal.org/index.php/alces historical distribution and current status the distribution of moose in north america during the 20th century has been described by several authors including peterson (1955), telfer (1984), kelsall (1987), karns (1998), franzmann (2000), and rodgers (2001); 4 subspecies are recognized, namely a. a. gigas, andersoni, americana, and shirasi (peterson 1955). in the past 40+ years many have detailed expanding distributions of moose in both western and eastern states, provinces, and territories (kelsall and telfer 1974, compton and oldenberg 1994, karns 1998, peek and morris 1998, brimeyer and thomas 2004, toweill and vecellio 2004, base et al. 2006, thomas 2008, wolfe et al. 2010, matthews 2012, labonte et al. 2013, wattles and destefano 2011, 2013, decesare et al. 2014). periodic winter aerial surveys based on the gasaway method are used by most agencies to estimate moose populations and trends (gasaway et al. 1986, peterson and page 1993, timmermann 1993, smits et al. 1994, lynch and shumaker 1995, bisset 1996, lenarz 1998, timmermann and buss 1998, bisset and mclaren 1999, bontaities et al. 2000, ward et al. 2000, gosse et al. 2002, heard et al. 2008, larter 2009, moen et al. 2011a, cumberland 2012, fieberg and lenarz 2012, delgiudice 2013, kantar and cumberland 2013, millette et al. 2014, seaton 2014, harris et al. 2015). moose are considered among the more difficult ungulates to survey (harris et al. 2015) and estimating either abundance or population trends from raw counts obtained by aerial survey can be challenging. most agencies estimate total jurisdictional populations based on the cumulative total of specific management areas sampled every 3 or more years. such jurisdictional estimates are often considered relatively crude and are primarily used to assess population trends, recruitment, and distribution over time. real changes in population estimates are indicated by changes of ~20% or more between surveys (gasaway and dubois 1987). new hampshire, maine, and vermont rely heavily on surveys of moose observations by deer (odocoileus virginianus) hunters and vehicle collision rates to estimate population trends. new hampshire and vermont use these annual deer hunter surveys in a related regression formula developed from concurrent infrared aerial surveys in a 3-year new hampshire study (bontaites et al. 2000, millette et al. 2014). jurisdictions not employing formal methods of population assessment base their estimates on professional opinion. consequently, population estimates are not necessarily comparable across jurisdictions or years because of the high variation in methodology and quality of data. as with all survey data, absolute counts are not achievable and the data herein should be treated as providing an indication of trends rather than absolute population estimates; the direction of population change (decreasing, increasing, or stable) is more important than the magnitude of change since the last jurisdictional survey (timmermann 2003). eastern north america currently, moose (a. a. americana) appear to be still expanding and/or occupying former range in the states of maine, massachusetts, new york, and connecticut (kilpatrick et al. 2003, hickey 2008, labonte et al. 2013, wattles and destefano 2011, 2013, s. heerkens, l. kantar, a. labonte, and d. scarpitti, pers. comm. 2015; fig. 1, table 1). moose in vermont and new hampshire have reoccupied all suitable habitat and are currently considered to be in slow decline (musante et al. 2010, c. alexander and k. rines, pers. comm. 2015; fig. 1, table 1). factors believed responsible for lower densities in vermont include purposeful harvest to reduce specificregionalpopulations(andreozzi et al. 2014). in new hampshire, high abundance of winter ticks (dermacentor albipictus) due to shorter winters and possible increased 2 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers f ig . 1 . e st im at es o f 2 0 1 4 -2 0 1 5 p o st -h u n t m o o se p o p u la ti o n s in 3 0 n o rt h a m er ic an ju ri sd ic ti o n s. alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 3 t ab le 1 . n u m b er s o f sp o rt h u n te rs , h ar v es t, an d p o st -h u n t p o p u la ti o n es ti m at es fo r 2 4 n o rt h a m er ic an ju ri sd ic ti o n s su rv ey ed in 2 0 0 0 –0 1 (2 0 0 1 ) an d 2 0 1 3 –1 4 (2 0 1 4 ). s y m b o ls ar e as fo ll o w s: p o p u la ti o n tr en d : + , � , s in d ic at e in cr ea si n g , d ec re as in g , o r st ab le , re sp ec ti v el y ; — in d ic at es n o n o n -r es id en t se as o n an d n /a in d ic at es n o t av ai la b le . t o ta l h u n te rs n o n -r es id en t h u n te rs t o ta l es ti m at ed h ar v es t e st im at ed m o o se p o p u la ti o n a g en cy 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 y u k o n t er ri to ry 1 2 ,4 4 0 2 ,7 9 3 3 8 9 5 0 1 7 1 6 6 1 7 7 0 ,0 0 0 + 7 0 ,0 0 0 s n o rt h w es t t er ri to ri es 2 1 ,3 0 0 1 ,1 2 5 6 5 1 2 3 1 ,4 0 0 2 4 4 2 0 ,0 0 0 + n /a n u n av u t3 n /a n /a n /a n /a b ri ti sh c o lu m b ia 3 1 ,5 0 0 3 1 ,1 8 8 2 ,2 5 0 1 ,2 3 2 9 ,2 0 0 6 ,8 9 0 1 3 0 –2 0 0 k 1 6 2 ,5 0 0 s a lb er ta 2 0 ,2 4 9 2 1 ,5 6 0 1 ,1 3 9 1 ,0 4 0 7 ,9 7 1 7 ,7 4 8 9 2 ,0 0 0 s 11 5 ,0 0 0 � s as k at ch ew an 1 0 ,0 0 0 1 2 ,9 3 1 2 5 4 11 4 4 ,1 5 1 5 ,5 4 3 4 6 ,0 0 0 s 4 8 ,0 4 5 � m an it o b a4 5 ,4 0 9 3 ,2 5 2 1 0 0 1 4 3 1 ,0 0 0 6 0 0 2 3 5 ,0 0 0 + 2 7 ,0 0 0 � o n ta ri o 1 0 0 ,0 0 0 1 0 6 ,7 5 2 3 ,0 0 0 1 ,7 4 3 11 ,0 0 0 5 ,0 7 1 1 0 0 –1 1 0 k s 9 2 ,3 0 0 � q u eb ec 1 3 0 ,0 0 0 1 7 6 ,7 1 0 2 ,0 0 0 2 ,7 0 6 1 4 ,0 0 0 2 1 ,1 0 5 9 5 –1 0 5 k + 1 2 5 ,0 0 0 + n ew b ru n sw ic k 4 ,1 7 4 4 ,6 2 6 9 7 9 7 2 ,5 3 7 3 ,6 8 3 2 5 ,0 0 0 + 3 2 ,0 0 0 + n o v a s co ti a 2 0 0 3 4 6 — — 1 8 6 2 4 0 6 ,0 0 0 + 5 ,0 0 0 s n ew fo u n d la n d 4 0 ,4 4 9 2 6 ,7 7 0 3 ,0 4 4 3 ,9 3 0 1 9 ,3 2 2 1 8 ,2 2 6 11 5 –1 4 0 k s 11 4 ,0 0 0 � a la sk a 3 0 ,0 0 0 3 1 ,6 0 7 3 ,2 0 0 1 ,9 6 7 5 ,5 0 9 7 ,9 4 2 1 2 0 ,0 0 0 � 1 7 5 –2 0 0 k + w as h in g to n 6 9 1 3 4 — 2 6 4 11 8 1 ,0 0 0 + 3 ,2 0 0 + id ah o 1 ,0 11 8 6 3 1 0 1 8 6 7 7 4 6 6 2 1 5 ,0 0 0 + 1 0 ,0 0 0 � u ta h 1 8 2 1 3 7 7 1 0 1 7 5 1 2 8 3 ,4 0 0 + 2 ,6 2 5 s w y o m in g 1 ,3 7 9 4 6 1 1 9 9 9 0 1 ,2 1 5 4 1 5 1 3 ,8 6 5 + 4 ,6 5 0 � m o n ta n a 6 0 9 3 5 7 1 6 11 5 9 6 2 7 8 4 ,0 0 0 s 3 -5 ,0 0 0 � n o rt h d ak o ta 1 3 2 1 0 6 — — 11 7 9 3 7 0 0 + 8 5 0 � c o lo ra d o 7 4 2 5 5 — 2 5 6 4 2 0 9 1 ,0 7 0 + 2 ,4 0 0 + m in n es o ta 5 4 4 2 — — — 1 2 5 — 5 ,1 0 0 � 3 ,4 5 0 � m ai n e 6 ,0 0 0 3 ,0 9 5 3 0 0 3 1 0 2 ,5 5 0 2 ,0 2 2 2 9 ,0 0 0 + 6 0 ,0 0 0 + t ab le 1 co n ti n u ed . . . . 4 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers incidence of brainworm (parelaphostrongylus tenuis) due to higher deer densities are of concern. current moose populations in maine, however, appear to have more than doubled since 2001 (timmermann 2003, wattles and destefano 2011) and are second to only alaska in the united states (lichtenwalner et al. 2014). populations in the canadian provinces of new brunswick and quebec have increased, while those on cape breton, nova scotia are believed relatively stable or in slight decline since 2001 (beazley et al. 2008, smith et al. 2010, s. lefort, p. macdonald, and d. sabine, pers. comm. 2016; fig. 1, table 1). on mainland nova scotia, the current population estimated at 500 is likely still in decline, and is designated “endangered” under the nova scotia endangered species act (p. macdonald, pers. comm. 2016). moose numbers on the island of newfoundland have been decreasing since 2001, whereas moose in the labrador portion of the province appear to be increasing in recent years (j. neville, pers. comm. 2015). overabundant moose populations on the island of newfoundland, where densities remained higher than elsewhere in north america at the turn of the century, have led to habitat deterioration and localized population decline (mclaren et al. 2004). consequently, harvest quotas were adjusted to modify population size in an effort to reduce and sustain specific populations (mclaren and mercer 2005), and more recently to help address moose-human conflicts in select management units (j. neville, pers. comm. 2016). western north america moose populations (alces a. shirasi) are believed to have doubled in washington state since 2000–2001 (r. harris, pers. comm. 2015) and have dispersed into oregon (p. matthews, pers. comm. 2015; fig. 1, table 1). density has declined in idaho and wyoming, but is stable in utah (d. brimeyer, k. hersey,ta b le 1 co n ti n u ed t o ta l h u n te rs n o n -r es id en t h u n te rs t o ta l es ti m at ed h ar v es t e st im at ed m o o se p o p u la ti o n a g en cy 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 2 0 0 1 2 0 1 4 v er m o n t 2 1 5 3 4 2 2 2 3 4 1 5 5 1 7 1 3 ,5 0 0 + 2 ,2 0 0 � n ew h am p sh ir e 5 8 5 1 2 7 7 6 1 9 4 1 9 9 1 5 ,0 0 0 + 3 ,8 0 0 � t o t a l 3 8 6 ,4 1 9 4 2 5 ,5 3 7 1 6 ,1 5 8 1 4 ,1 8 3 8 3 ,2 4 6 8 2 ,0 9 6 9 3 5 ,6 3 5 –1 ,0 5 0 ,6 3 5 1 ,0 8 2 ,0 2 0 –1 ,0 8 9 ,0 2 0 1 y k in d ig en o u s h ar v es t is n o t in cl u d ed . 2 n w t in d ig en o u s h ar v es t is n o t in cl u d ed . 3 n u es ta b li sh ed 1 9 9 9 , fo rm er ly p ar t o f n w t , p o p u la ti o n 3 2 ,0 0 0 (2 0 11 ), 9 ,9 8 4 ,6 7 0 k m 2 . 4 m b 2 0 1 4 h ar v es t es ti m at e p en d in g su rv ey co m p le ti o n . 5 m n cl o se d m o o se se as o n b eg in n in g 2 0 1 3 . alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 5 and s. nadeau, pers. comm. 2015). wyoming populations declined from an estimated 13,865 in 2001 to 7,700 in 2008, and to 4,650 currently (timmermann 2003, brimeyer and thomas 2004, smith et al. 2011, d. brimeyer, pers. comm. 2015). populations have grown in colorado and remain relatively stable compared to a declining trend in montana (tyers 2006, decesare et al. 2014, n. decesare and a. holland, pers. comm. 2015). periodic dispersal into the central united states, primarily from north dakota and minnesota, is reported as far south as kansas and missouri (hoffman et al. 2006). moose (alces a. andersoni) populations in central british columbia have declined, but overall, the provincial population has remained relatively stable since 2000–2001 (kuzyk and heard 2014, kuzyk 2016). moose (alces a. gigas) in alaska have increased and those in the yukon territories have remained stable (b. dale, r. florkiewicz, and k. titus, pers. comm. 2015; fig. 1, table 1). moose on the arctic coastal plain in alaska have expanded and contracted their numbers and range twice in the past 25 years (b. dale and k. titus, pers. comm. 2015). recent research has linked range expansion in arctic alaska to warming and the associated increase in shrub habitat (tape et al. 2016). moose have also expanded their range in northern southeast alaska where first observed in haines in 1924, and now inhabit the gustavus forelands (1966; glacier bay national park). populations of moose now occur on all the major islands of the central southeast panhandle of alaska (b. dale and k. titus, pers. comm. 2015). central north america moose have expanded northward in nunavut and labrador, and are found as far north as 67˚ 31ʹ′ near kugluktuk in nunavut and richards island in the northwest territories mackenzie delta (v. crichton, pers. comm. 2015). population estimates for vast portions of the northwest territories and nunavut (formerly part of the nwt) are not currently available (m. dumond and a. smith, pers. comm. 2015; fig. 1, table 1). a further 6 of 8 jurisdictions in the midcontinent report recent, declining trends in moose populations (alces a. andersoni/ americana) including the adjacent provinces of alberta, saskatchewan, manitoba, and ontario, as well as the neighboring states of north dakota and minnesota (lenarz et al. 2010, r. corrigan, g. delgiudice, h. hristienko, g. lucking, l. mcinenly, j. smith, r. tether, pers. comm. 2015; fig. 1, table 1). laforge et al. (2016) report increasing moose populations in the farmlands of southern saskatchewan. similarly, manitoba populations have increased in the last 20 years in the southwest farmlands of the province where access is controlled and few predators exist (h. hristienko and k. rebizant, pers. comm.). these expansions appear linked to the reduction of small, privately owned farms being replaced by larger corporate farms, and a corresponding decline in undocumented harvest. however, the adjacent jurisdictions of ontario, manitoba, and minnesota give lower overall estimates than in 2001 (table 1). minnesota closed their harvest in the northwestern region in 1997 due to dramatic population decline from unknown causes (m. schrage, pers. comm. 2001, wünschmann et al. 2015). the estimated decline was from 4,264 in 1983 to 1,486 in 1995, to ~900 animals in 2001; essential collapse of this population occurred by the early 2000s. murray et al. (2006) concluded that the giant liver fluke (fascioloides magna) was largely responsible for this decline. a concurrent decline in adjacent northeastern north dakota was investigated by maskey (2011) who suggested other factors such as brainworm play a larger role in moose mortalities. minnesota closed its moose hunting season in 6 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers northeastern minnesota in 2013 after numbers dropped from ~8,500 in 2006 to 3,500 in 2014 (g. delgiudice and l. mcinenly, pers. comm. 2015). manitoba’s moose population is believed to have dropped from a historical high of 45,000 several decades ago to 27,000 in 2015 (h. hristienko, pers. comm. 2015). disease, over-harvest, and human development of landscapes are the primary factors thought responsible for the decline (crichton et al. 2004). recent surveys in northwestern ontario indicate a corresponding decline in certain moose populations (omnrf 2015, table 1). current populations are increasing in michigan, largely due to higher density estimates on isle royale (vucetich and peterson 2015). abundance in 2015 was estimated as 323 in the reintroduced population in the western upper peninsula, but low productivity and calf:cow ratios suggest population decline (dodge et al. 2004, d. beyer, pers. comm. 2015). populations in neighboring wisconsin, where moose regularly move in and out of northern michigan and minnesota, are currently estimated at <50 (k. wallenfang, pers. comm. 2015; fig. 1). the united states fish and wildlife service (usfws) is considering a listing under the endangered species act of the northwestern subspecies of moose (alces a. andersoni) that is purported inhabiting upper michigan, isle royale, minnesota, north dakota, and wisconsin (https://www.fws.gov/midwest/es/soc/pdf/ frbatch90dayfndngs03june2016piversion. pdf). in response, michigan and wisconsin submitted letters to the usfws indicating that moose in their jurisdictions originated from eastern moose populations (alces a. americana) (k. wallenfang, pers. comm. 2016). to summarize in 30 jurisdictions, current moose density is believed stable in 8, increasing in 9, decreasing in 11, with data unavailable in 2 (fig.1, table 1). in 22 jurisdictions (circa 2014–2015) for which population estimates are available, and in which an annual licensed harvest occurred in 2014, the total population estimate is 1,082,020 to 1,089,020 animals which is collectively similar to that reported in 2001 (table 1). remarkably little overall change has occurred despite the majority of jurisdictions reporting either increasing or decreasing populations. population estimates in 12 canadian jurisdictions totaled 790,845 in 2014 compared to a range of 734,000 to 849,000 in 11 jurisdictions in 2001 (table 1, fig.1, timmermann 2003). the total population increased in 17 states from 204,150–205,130 in 2001, to 274,768–302,268 in 18 states in 2014 (table 1, fig.1, timmermann 2003). of the 7 states where hunting is prohibited, 3 report expanding populations (oregon, michigan, new york), stable populations exist in wisconsin, massachusetts, and connecticut, whereas minnesota closed their season in 2013 due to significant population decline (moen et al. 2011b; fig. 1). factors affecting decreasing densities a host of factors are believed responsible for moose population declines including climate change, illegal harvest, habitat loss or degradation, parasites and disease, disturbance, moose-vehicular collisions, predators, and unregulated recreational and indigenous and subsistence harvests (west 2009). in the 11 of 30 (37%) jurisdictions that indicated a declining population trend, the “most important factors” were: parasites and disease (8 jurisdictions), predators (7), natural habitat loss (5), unregulated harvest (3), warmer summers/winters (2), increased access and vehicle technology (1), higher deer densities (1), over harvests by licensed hunters (1), increased hunting pressure (1), and variable factors (1). minnesota initiated a $1.2 m moose mortality study in 2013 to help determine factors responsible for the recent dramatic population decline. preliminary results provide evidence of the alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 7 https://www.fws.gov/midwest/es/soc/pdf/frbatch90dayfndngs03june2016piversion.pdf https://www.fws.gov/midwest/es/soc/pdf/frbatch90dayfndngs03june2016piversion.pdf https://www.fws.gov/midwest/es/soc/pdf/frbatch90dayfndngs03june2016piversion.pdf importance of parasites and disease and predators as mortality factors (wünschmann et al. 2015). in maine and new hampshire, similar research initiated in 2014 indicates that winter ticks remain a primary influence on calf mortality and adult cow productivity (l. kantar and k. rines, pers. comm. 2016). harvest management economic impact moose, a symbol of wilderness, are much valued by indigenous hunters, metis people, recreational hunters, and a host of nonconsumptive users (timmermann and rodgers 2005). licensed recreational hunting promotes substantial benefits to local economies valued in the $100s of millions annually. in the early 1990s, for example, legg (1995) estimated can $134.7 m in ontario for all hunter-related activities in 1993. more recently, maine estimated the economic impact of 3,095 resident and 310 non-resident hunters to represent us $11.9 m and $3.9 m in 2014 (l. kantar, pers. comm. 2015), and alaska valued its non-resident hunt at $11m in 2014 (b. dale, pers. comm. 2015). similarly, quebec estimated 176,710 residents and 2,707 non-residents generated can $204 m and $8.0 m in 2014 (s. lefort, pers. comm. 2015). harvest control objectives three territories and 9 provinces in canada, and 11 states in the united states administered a moose hunt in 2014 (table 1). collectively, 425,537 licensed hunters harvested an estimated 82,096 moose in 2014–2015; a decade earlier, the harvest was 83,246 moose by 386,419 licensed hunters (table 1). hunting regulations continue to become more restrictive and complex as the demand on moose populations and corresponding harvest success rates increase, due in part, to increased road access and use of mechanized equipment (timmermann and buss 1998). specific and strategic management of hunting is required to affect the desired allocation of moose harvest among licensed hunters, secure the sustainability of moose populations, and achieve other specified management objectives for a particular area. specific moose management plans, guidelines, or statements existed in 13 jurisdictions in 2000–2001 (maine, vermont, new hampshire, utah, colorado, wyoming, idaho, yukon territory, british columbia, alberta, saskatchewan, ontario and quebec; timmerman 2003). specific harvest policy is currently guided by an approved or draft management policy including goals and objectives in 13 jurisdictions; 3 employ unwritten or a generalized wildlife policy. for example, alaska’s constitution, statutes, and regulations direct management activities and objectives through a public process (b. dale, pers. comm. 2015), minnesota’s policy is guided by a research and management plan (mcgraw et al. 2010, minnesota dnrc 2011, moen et al. 2011b), and colorado uses a specific management plan for each of 10 herds (a. holland, pers. comm. 2015). british columbia has recently developed a provincial guidance and direction framework for sustainable moose management (british columbia fish &wildlife branch 2015). beginning in 2007, ontario conducted a 2-year broad review and wide consultation of their moose management program that produced a new set of policies and guidelines with objectives and strategies to address the declining population and harvest (omnrf 2008). two options to control calf harvests included a shorter calf season within the regular season and a draw for calf tags (bottan et al. 2002, timmermann et al. 2002, omnrf 2009a, b). a moose management plan has been developed in newfoundland and labrador that will help address human-wildlife conflicts (j. neville, pers. comm. 2015), and quebec currently employs a fourth iteration of a management plan spanning the period 2012–2019 8 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers (s. lefort, pers. comm. 2015). saskatchewan and manitoba are developing specific management plans (h. hristienko and r. tether, pers. comm. 2015). allocation of hunting opportunities moose are publicly owned and held in trust by provincial, territorial, and state wildlife agencies. the first priority of most agencies is to ensure the long-term conservation of moose populations and their habitats. harvest allocation is given prime consideration to subsistence use by indigenous people under treaty or other legal agreements in at least 20 of 23 jurisdictions that currently manage a harvest. resident hunters in 20 of 23 jurisdictions are typically favored over non-residents and non-resident foreigners (10 of 23) in allocating harvest opportunities. in 2014–2015, non-residents were eligible to hunt in 20 of 23 jurisdictions (table 1). additional controls such as increased license fees, resident-only seasons, guide requirements, and limited permits are commonly placed on non-resident hunters giving residents priority in allocation of hunting opportunities. a guide was required by 10 of 23 agencies, and at least 6 agencies required non-residents to register with a licensed tourist outfitter, and 8 required foreigners to do so to enhance safety and success, as well as provide local economic benefit. some agencies restrict or limit moose hunting opportunities including all states except alaska. washington and north dakota offer a single moose hunt per lifetime, and colorado, utah, and idaho limit hunters to one antlered animal per lifetime. others require a waiting period between hunts: 2 years in idaho, 3 years in new hampshire, maine, saskatchewan, and manitoba, 5 years in wyoming, vermont, and nova scotia, and 7 years in montana (if successful). hunters in alaska and 8 canadian jurisdictions may hunt annually within quotas regardless of previous harvest success. minnesota closed the entire state to moose hunting in 2013 after the northeastern population declined by half since 2006. manitoba legislated 3 conservation closure game hunting areas and 1 partial closure in 2011, coupled with a wolf reduction initiative to promote moose recovery (h. hristienko, pers. comm. 2015, v. crichton, pers. comm. 2016). ontario continues to offer moose hunting opportunities for physicallychallenged hunters in one wildlife management unit (armstrong and simons 1999). control concepts agencies employ a variety of strategies to regulate harvests and distribute hunting pressure (timmermann 1987, 2003). passive strategies include season length and timing, access restrictions, weapon requirements, and license qualification prerequisites; active measures include limiting license sales or specifying the sex, age, or number of animals taken by specific area. objectives often include the harvest of pre-determined numbers to sustain, increase, or reduce populations. in alberta, xu and boyce (2010) developed an age-sex matrix model for harvest quota management of moose populations that allows easy application by managers responsible for setting harvest quotas. antler-based hunting regulations in british columbia may have resulted in disrupted reproductive patterns and a consequent over-harvest of large bulls (child et al. 2010). in interior alaska, harvest restrictions on bull moose based on antler architecture allowed the recovery of bull:cow ratios from 26:100 to 32:100 after only 2 years of use (young and boertje 2008). conversely, liberal antlerless hunts in alaska are considered vital to control moose populations from reaching unsustainable densities in specific management units a general season harvest ticket is available to all residents/non-residents (young and boertje 2004, young et al. 2006, boertje et al. 2007, 2009, young and boertje 2011). alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 9 a specific controlled hunt to reduce a local moose population and impacts on cole crops was implemented in maine in 2009 (kantar 2011). new hampshire and vermont have increased regional/local harvest rates with higher antlerless quotas to alleviate browsing impacts on regenerating forests and vehicular collisions (c. alexander, pers. comm. 2002, andreozzi et al. 2014). in 2014–2015, 9 agencies offered unlimited selective or non-selective harvest opportunities, and all 23 jurisdictions restricted or limited harvests on a selective or non-selective basis in certain management areas (fig. 2). in addition, closed seasons were employed to prevent licensed harvest in specific areas, including certain provinces, territories, states, national parks, and wildlife refuges. license qualifications and fees in 2014, proof of hunting proficiency, either a previous license or completing a hunter safety education course, was required to obtain a moose hunting license in all jurisdictions. in 2015, canadian resident license fees averaged can $52.57 (range $5.00 in the yukon territory to $81.30 in new brunswick), and non-resident licenses averaged $374.45 (range $48.00 in quebec to $619.24 in new brunswick). resident fees in the states averaged us $162.00, (range = $25.00 in alaska to $ 413.00 in utah), and non-resident fees averaged $1,168.03 (range = $350.00 in vermont to $2,271.25 in idaho). some agencies, including alaska and maine, charged higher fees to non-resident foreigners. export permits or trophy fees are required, in addition to the license fee, to transport an animal from the yukon territory, northwest territories, british columbia, alberta, and ontario. currently, only new hampshire requires moose hunters to demonstrate shooting proficiency using conventional fire-arms prior to purchasing a hunting license, as described by buss et al. (1989). previously, new brunswick and newfoundland required hunters to pass a shooting and written test before qualifying for a big game hunting license (timmermann and buss 1995). alaska requires all archery and black powder hunters to pass a proficiency test (w. regelin, pers. comm. 2002). seasons season length and timing are used to manage the availability of hunting opportunity, hunter success relative to vulnerability based on moose behavior, and seasonal access. seasons are generally specific to weapon type (e.g., conventional firearms, black powder, or archery), and seasons tend to be longer in more remote areas and shorter closer to population centers. the most liberal season length (365 days, 1 july-30 june) occurs in 3 game management areas in nunavut (table 2). season lengths for all hunts in parts of alaska, idaho, wyoming, montana, north dakota, yukon, northwest territories, british columbia, alberta, saskatchewan, manitoba, ontario, quebec, and newfoundland equal or exceed 3 months; new brunswick, nova scotia, vermont, and new hampshire restrict season length to 5–9 days. early archery seasons are typically in addition to firearm seasons, and are offered in 17 jurisdictions (table 2). most firearm seasons begin during the latter portion of the rut period (wilton 1995) and many extend into november or december. split seasons (early vs. late fall) occur in at least 11 jurisdictions. minnesota closed their northeast moose hunt in 2013 after a continuous 41-year period of offering a limited non-selective hunt requiring all eligible hunters to apply in groups of up to 4 individuals (judd 1972). management areas and harvest strategies all agencies have subdivided their moose range into various sized management areas (wildlife, game, or moose management units) to facilitate specific harvest control 10 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers f ig . 2 . m o o se h ar v es t st ra te g ie s em p lo y ed b y 3 0 n o rt h a m er ic an ju ri sd ic ti o n s (c ir ca 2 0 1 4 – 2 0 1 5 ). n u m b er s in d ic at e m an ag em en t ar ea s o r su b d iv is io n s u n d er ea ch h ar v es t st ra te g y in ea ch ju ri sd ic ti o n . alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 11 strategies. the number of management areas ranges from 3 in nunavut to 443 in the yukon, and vary in size from 47 km2 in north dakota to 278,183 km2 in the northwest territories (table 2). all jurisdictions except nunavut continue to employ either a selective or non-selective limited hunter participation strategy, or a combination of both (fig. 2). most favor some form of limited selective or limited non-selective strategy to control sexand age-related harvests. alaska alone continues to employ registration hunts which require mandatory kill registration and season termination once a prescribed harvest is achieved. several agencies have developed harvest strategies that maximize hunter participation. for example, sharing a moose between 2 or more hunters optimizes hunting opportunities and accommodates hunters who wish to hunt with friends. perhaps the most liberal approach is in ontario which previously allowed all eligible hunters to hunt calves in any of 64 wildlife management units (timmermann table 2. characteristics of moose hunting seasons in north america, 2014–2015. number of management areas season length/timing size (km2) agency with moose min. max. with open season max days earliest latest yukon territory 443 64 2,919 360 92 aug. 01 oct. 31 northwest territories 6 56,270 278,183 61 123 sept. 01 jan. 31 nunavut 3 n/a n/a 1 365 july 01 june 30 british columbia 193 465 18,982 1771 107 aug. 15 nov. 30 alberta 177 99 26,079 1631 91 aug. 25 nov. 30 saskatchewan 82 232 82,443 801 92 sept. 01 nov. 30 manitoba 60 222 139,204 291 117 aug. 31 jan. 24 ontario 76 832 122,397 691 93 sept. 14 dec. 15 quebec 30 1,471 204,142 281 92 aug. 27 dec. 01 new brunswick 25 826 6,402 25 10 sept. 20 sept. 30 nova scotia 5 204 2,700 51 6 sept. 29 dec. 11 newfoundland 54 217 4,533 541 121 aug. 29 jan. 24 alaska 22 9,117 217,559 221 243 july 01 april 15 washington 11 747 2,857 111 60 oct. 01 nov. 30 idaho 70 220 7,843 461 86 aug. 30 dec. 01 utah 15 809 5,394 12 34 sept. 12 oct. 15 wyoming 38 100 10,000 311 81 sept. 01 nov. 20 montana 88 95 60,926 821 87 sept. 15 nov. 30 north dakota 7 47 36,514 51 101 sept. 04 dec. 13 colorado 65 130 1,540 571 25 sept. 12 oct. 14 minnesota 30 200 875 — season closed 2013 maine 28 1,424 5,320 25 24 sept. 28 nov. 28 vermont 21 639 2,036 161 6 oct. 1 oct. 26 new hampshire 22 391 1,365 20 9 oct. 18 oct. 26 1 offers early archery season. 12 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers 2003, fig. 2). in 2015 calf hunting was reduced to a 2-week period, and beginning in 2016, opening seasons will be delayed by 1 week (omnrf 2016). in addition, ontario hunters may apply in groups of up to 15 hunters for the chance of obtaining an adult tag that is area-specific, allowing a limited number of tags to be spread more evenly among hunter groups (omnrf 2015, 2016). sharing a moose is currently required in at least 2 jurisdictions including quebec (minimum 2 hunters/moose) and manitoba which provide an option of purchasing a conservation license (2 hunters using a single tag; s. lefort and v. crichton, pers. comm. 2015). timmermann (2003) detailed additional moose-sharing mechanisms including a “group hunt” and a “limited entry shared hunt” in british columbia, a “special antlered moose partner license” in alberta, and a “companion moose hunting stamp” in nova scotia. each successful permittee in maine, new hampshire, and vermont may select a sub-permittee to hunt together and harvest a single moose. to help maximize hunter opportunity, newfoundland gives preference to party applications (restricted to 2 individuals) over individuals, and to those unsuccessful in previous years (timmermann 2003). yukon has “special guide licenses” that annually allow 100 yukon residents to guide a non-resident hunter, in part to accommodate family members/friends from outside the territory (s. czetwertynski, pers. comm. 2016). harvest assessment all sources of mortality must be assessed to monitor the effectiveness of various harvest strategies. hunters, whether successful or not, are required to report their hunting activity in 11 of 23 jurisdictions, and harvest registration is compulsory in 13 (table 3). twelve jurisdictions apply a non-compliance penalty to hunters failing to report, although enforcement of these requirements varies among agencies. timmermann (2003) provided a more detailed description of this subject, including the use of interactive voice response technology, use of a telephone questionnaire, and modeling to predict population changes resulting from various harvest strategies. moose hunter education and engagement all first time hunters are required to successfully complete a hunter safety/education course in 23 jurisdictions that managed a moose hunt in 2014. most (19 of 23) charged fees ranging in canada from no fee in yukon to can $160 in ontario, and in the united states from no fee in washington and vermont, us $5 in new hampshire, and $10 in utah and wyoming. four states (idaho, utah, vermont, and new hampshire) incorporate a practical shooting test in their hunter safety/education course. all jurisdictions (22 of 23) except nunavut provide moose hunters with information on their official websites, and 9 use social media. printed hunting regulations were available in 22 of 23 jurisdictions, and 15 provided printed pamphlets, brochures, and/or fact sheets. television and/or radio was used to provide information in 4 jurisdictions, 10 used newspapers and/or magazines, and 17 used email and/or traditional mail. harvest by native and subsistence users currently, most north american moose management agencies give primary consideration to subsistence use by canadian indigenous peoples and native american peoples in recognition of obligations made under historical treaties signed by both federal governments (crichton et al. 1998, lynch 2006). currently, nine 9 of 24 jurisdictions (yukon, northwest territories, british columbia, manitoba, ontario, quebec, nova scotia, idaho, montana) report primary allocation of the moose resource to subsistence use by indigenous people under treaty or other legal agreements. in many areas unrestricted alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 13 access to moose exists year-round, and current regulations are considered liberal given the widespread use of modern technology (courtois and beaumont 1999). because conflicts often occur between licensed sport hunters and indigenous people, moose managers must consider the annual harvest by both groups in formulating hunting regulations (lynch 2006). the harvest by indigenous hunters is difficult to quantify and unfortunately, little effort has been made to measure the magnitude of this harvest which some managers believe approaches or exceeds the licensed harvest in certain jurisdictions. metis, who are considered people of mixed indian and white ancestry (swail 1996), are testing their perceived rights in court in alberta, manitoba, and nova scotia where they claim the right to hunt and fish on traditional territory both within and outside their current harvesting zone (v. crichton, pers. comm. 2016). the supreme court of canada table 3. moose harvest assessment strategies used in north america, 2014–2015. hunt activity report kill registration non-compliance penalty agency compulsory voluntary compulsory voluntary yukon territory x x1 fine cdn$100.00 northwest territories x x1 none nunavut x x fine cdn$200.00 british columbia x x fine cdn$230.00 alberta none x n/a saskatchewan x none n/a manitoba x none n/a ontario2 2 x 2 none1 n/a2 quebec none x finecdn$250-750.00 new brunswick x x finecdn $100-500.00 nova scotia x x n/a newfoundland x x none alaska x x loss of future eligibility washington x x fine us$25.00 idaho x x fine us$25.00 1,000 & jail utah x none ineligible to apply next year wyoming x x none montana x x n/a north dakota x x ineligible to apply next year colorado x x loss of future eligibility minnesota season closed 2013 maine x x fine us$100-1,000 / lic. loss vermont x x fineus$262 + & lic. loss new hampshire x x fineus$248-1,000 & loss of future eligibility 1 export permit/trophy fee. 2 compulsory hunt activity/harvest report in 5 wildlife management units; $150.00 fine for non-compliance and inability to receive a tag in subsequent year. 14 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers has refused to hear an appeal involving metis hunting and fishing rights in alberta following a supreme court ruling 10 years previous that granted hunting rights to ontario metis. however, in april 2016 the supreme court declared that the federal government has constitutional responsibility for métis and nonstatus indians, which could have important implications for their hunting and fishing rights. timmermann (2003) provided estimates of the annual moose harvest by indigenous and metis peoples in british columbia, alberta, saskatchewan, manitoba, ontario, quebec, new brunswick, nova scotia, northwest territories, and the yukon territory. a moose monitoring program was established in the northwest territories between the dehcho first nations and government biologists that yielded valuable biological information to identify changes in moose populations (larter 2009). in a similar effort to involve indigenous people in northwestern ontario, leblanc et al. (2011) suggested that provincial calculations may underestimate total harvests by up to 40%, and concluded that developing a working relationship with indigenous communities is necessary to effectively manage moose in ontario. a moose management plan was drafted by the nova scotia mi’kmaq in 2005 to aid in tripartite negotiations between the band and the provincial and federal governments (bridgland et al. 2007). the nova scotia dnr is currently working with the nova scotia mi'kmaq to develop a collaborative management plan for cape breton moose (p. macdonald, pers. comm. 2016). it is hoped such collaborative management will sustain moose populations long-term on cape breton island. cooperative management of moose between state and 3 tribal bands in northeastern minnesota has led to increased levels of trust since 1988 (edwards et al. 2004). in colorado, moose harvested by native and subsistence hunters is monitored by the brunot agreement with the southern ute and ute mountain tribes (a. holland, pers. comm. 2015). future sustainable harvests and population goals will largely remain elusive until the total harvest, including harvests by indigenous and metis peoples and subsistence users, are both agreed to and verifiable. under federal regulations in alaska, all rural residents are subsistence users which allows certain communities or households to harvest moose. in addition, special regulations allow moose harvesting outside of normal hunting seasons by alaskan natives for ceremonial and cultural purposes. all agencies suggested such harvests were “substantial” in specific local areas during the period 2000–2001. oregon has 2 native tribes which hunt ungulates in areas where moose occur; one tribe hunts moose and the other is considering a moose season (p. matthews, pers. comm., 2016). illegal hunting losses appear to be significant in some jurisdictions including colorado, utah, and ontario (timmermann 2003). most agencies encourage all hunters to report illegal infractions using a toll-free telephone number. ontario introduced a “moose watch” program in 2001 to help reduce poaching (todesco 2004). conservation officers in ontario’s northeast region found 1,741 illegally killed moose from 1997–2014 (todesco 2004, c. todesco, pers. comm. 2016). during this period, >238,000 hunters were contacted, 7,328 warnings issued, and 5,514 charges were laid while conducting moose hunt enforcement duties. managing unhunted populations parks, refuges, and special areas most north american jurisdictions where moose occur provide for areas where hunting is not a primary management objective. currently, 7 states have no open moose hunting season and 11 of 23 jurisdictions provide closed seasons in 2–33 management alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 15 areas (fig. 2). the assumed common objective of closed areas that include game or wildlife reserves, national, provincial, territorial, and state parks, and nature reserves is the preservation of moose in representative natural habitats for education and recreational enjoyment. further maintenance of biodiversity and ecosystem function is often a stated objective. a review of moose management objectives and programs in parks, refuges, and special areas was detailed by timmermann and buss (1995) and timmermann (2003). moose are native to at least 30 north american national parks in 18 jurisdictions with isle royale perhaps the most famous, boasting a 58-year continuous ecological study of wolves and moose beginning in 1959 (vucetich and peterson 2015). timmermann (2003) detailed moose-related studies in several national parks including isle royale in michigan, elk island in alberta, voyageurs in minnesota, and gros morne in newfoundland. west (2009) reported approximately 39,000 moose inhabited 35 national wildlife refuges in the united states, with ~38,000 in alaska alone; 9 refuges used management practices to specifically benefit moose (i.e., prescribed wildland fire). discussion telfer (1984) found a close correspondence between the southern limit of moose distribution worldwide and the 20 °c july isotherm. since then, several studies have suggested that moose numbers and the southern limit of their distribution may be affected by climate change (thompson et al. 1998, murray et al. 2006, lenarz et al. 2010) but others have documented increasing numbers and expanding moose populations at the southern edges of their jurisdictional boundary (wattles and destefano 2011, harris et al. 2015, laforge et al. 2016). our survey across north america further indicated that 10 of 15 moose populations at the southern limit of their distribution are stable or increasing. although we acknowledge that many of the reported population estimates and trends were based on professional opinion rather than systematic surveys, the weight of evidence suggests that moose are not immediately at risk of disappearing from southern regions of their distribution. the more interesting enquiry at this time would be to determine how and why moose continue to thrive and even expand their range southward in certain areas. moving forward in the face of climate change and ever-increasing human development, it will be vitally important to maintain systematic aerial surveys to monitor moose population trends, and to implement these wherever they are not in use. this will be a challenge not only because of the financial and human resources required, but also because climate change may hinder the collection of long-term data due to lack of snow required to efficiently conduct aerial surveys. thus, further research into new technologies that are not hindered as much by environmental conditions will be highly beneficial, such as forward-looking infrared radiometer systems (flir; millette et al. 2014) and sensor-equipped drones. our survey revealed variation in moose population trends across north america, and local variation expected within jurisdictions. as indicated by survey respondents, numerous factors can affect moose population trends both locally and regionally. although research might be undertaken to identify the most important factors in a particular area, the responses of local moose managers are limited because many cannot be controlled directly (e.g., parasites and disease, weather). instead, most moose managers can only use harvest and habitat management to mitigate declines in moose population numbers. these options will become ever more important as climate change and human development gradually increase their influence on moose numbers and distribution across north america. 16 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers acknowledgements appreciation is extended to the following individuals in the united states who provided unpublished information in response to a 9-page questionnaire survey: bruce dale and kim titus, alaska department of fish and game, juneau, alaska; rich harris, washington department of fish and wildlife, olympia, washington; pat matthews, oregon department of fish and wildlife, enterprise, oregon; steve nadeau, idaho fish and game, boise, idaho; kent hersey, division of wildlife resources, salt lake city, utah; andy holland, colorado parks and wildlife, fort collins, colorado; doug brimeyer, wyoming game and fish department, jackson, wyoming; nick decesare, montana fish, wildlife and parks, missoula, montana; jason smith, north dakota game and fish department, jamestown, north dakota; glenn delgiudice and leslie mcinenly, minnesota department of natural resources, forest lake, minnesota; dan storm and kevin wallenfang, department of natural resources, rhinelander, wisconsin; dean beyer, michigan department of natural resources and northern michigan university, marquette, michigan; lee kantar, maine department of inland fisheries and wildlife, bangor, maine; kristine rines, new hampshire fish and game department, new hampton, new hampshire; cedric alexander, vermont fish and wildlife department, st. johnsbury, vermont; andrew labonte, department of energy and environmental protection, north franklin, connecticut; david scarpitti, massachusetts division of fisheries and wildlife, west boylston, massachusetts; steven heerkens, department of environmental conservation, utica, new york. appreciation is extended to the following individuals in canada who provided unpublished information in response to a 9-page questionnaire survey: sophie czetwertynski and rob florkiewicz, environment yukon, whitehorse, yukon; angus smith and jan adamczewski, department of environment and natural resources, government of northwest territories, yellowknife, northwest territories; mathieu dumond, department of environment, arviat, nunavut; gerry kuzyk, ministry of forests, lands and natural resources operations, victoria, british columbia; rob corrigan, , environment and parks, edmonton, alberta; rob tether and mike gollop, saskatchewan ministry of environment, saskatoon, saskatchewan; hank hristienko and ken rebizant, manitoba conservation, winnipeg, manitoba; vince crichton, retired manitoba conservation, winnipeg, manitoba; greg lucking and patrick hubert, ontario ministry of natural resources and forestry, peterborough, ontario; charlie todesco, ontario ministry of natural resources and forestry, wawa, ontario; sébastien lefort, quebec ministère des forêts, de la faune et des parks, quebec city, quebec; dwayne sabine, new brunswick department of natural resources, fredericton, new brunswick; peter macdonald, nova scotia department of natural resources, kentville, nova scotia; john neville and conor edwards, newfoundland and labrador department of environment and conservation, corner brook, newfoundland. references andreozzi, h. a., p. j. pekins, and m. langlais. 2014. impacts of moose browsing on forest regeneration in northeastern vermont. alces 50: 67–79. armstrong, e. r., and r. simons. 1999. moose hunting opportunities for physically challenged hunters in ontario: a plot study. alces 35: 125–134. base, d. l., s. zender, and d. martorello. 2006. history, status, and hunter harvest of moose in washington state. alces 42: 111–114. beazley, k., h. kwan, and t. nette. 2008. an examination of the absence of established moose (alces alces) populations alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 17 in southeastern cape breton island and nova scotia, canada. alces 44: 81–100. bisset, r. a. 1996. standards and guidelines for moose population inventory in ontario. ontario ministry of natural resources, fish and wildlife branch, peterborough, ontario, canada. ———, and m. a. mclaren. 1999. moose population aerial inventory plan for ontario: 1999–2002. ontario ministry of natural resources, northwest science and technology. information report ir-004. thunder bay, ontario, canada. boertje, r. d., k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. ———, m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314–327. bontaites, k. m., k. gustafson and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69–75. bottan, b., d. euler, and r. rempel. 2002. adaptive management of moose in ontario. alces 38: 1–10. bridgland, j., t. nette, c. dennis, and d. quann. 2007. moose on cape breton island nova scotia: 20th century demographics and emerging issues in the 21st century. alces 43: 111–121. b.c. f&w (british columbia fish and wildlife) branch. 2015. provincial framework for moose management in british columbia. ministry of forests, lands and natural resources operations, fish and wildlife branch, victoria, british columbia, canada. brimeyer, d. g., and t. p. thomas. 2004. history of moose management in wyoming and recent trends in jackson hole. alces 40: 133–143. buss, m., r. gollat, and h. r. timmermann. 1989. moose hunter shooting proficiency in ontario. alces 25: 98–103. child, k. n., d. a. aitken, r. v. rea, and r. a. demarchi. 2010. potential vulnerability of bull moose in central british columbia to three antler-based hunting regulations. alces 46: 113–121. compton, b. b., and l. e. oldenburg. 1994. the status and management of moose in idaho. alces 30: 57–62. courtois, r., and a. beaumont. 1999. the influence of accessibility on moose hunting in northwestern québec. alces 35: 41–50. crichton, v. f. j., w. l. regelin, a. w. franzmann, and c. c. schwartz. 1998. the future of moose management and research. pages 655–663 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. ———, t. barker, and d. schindler. 2004. response of a wintering moose population to access management and no huntinga manitoba experiment. alces 40: 87–94. cumberland, r. e. 2012. potvin doublecount aerial surveys in new brunswick: are results reliable for moose? alces 48: 67–77. decesare, n. j., t. d. smucker, r. a. garrott, and j. a. guide. 2014. moose status in montana. alces 50: 35–51. delgiudice, g. d. 2013. minnesota moose aerial survey. minnesota department of natural resources, st. paul, minnesota, usa. dodge, w. b. jr., s. r. winterstein, d. e. beyer jr., and h. campa iii. 2004. survival, reproduction, and movements of moose in the western upper peninsula of michigan. alces 40: 71–85. 18 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers edwards, a. j., m. schrage, and m. lenarz. 2004. northeastern minnesota moose management a case study in cooperation. alces 40: 23–31. fieberg, j. r., and m. s. lenarz. 2012. comparing stratification schemes for aerial moose surveys. alces 48: 79–87. franzmann, a. w. 2000. moose. pages 578– 600 in s. demarais and p. r. krausman, editors. ecology and management of large mammals in north america. prentice hall, upper saddle river, new jersey, usa. gasway, w. c., s. d. dubois, d.j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. fairbanks, alaska, usa. ———, and s. d. dubois. 1987. estimating moose population parameters. swedish wildlife research (supplement) 1: 603–617. gosse, j., b. mclaren, and e. eberhardt. 2002. comparison of fixedwing and helicopter searches for moose in a midwinter habitat-based survey. alces 38: 47–53. harris, r. b., m. atamian, h. ferguson, and i. keren. 2015. estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty. alces 51: 57–69. heard, d. c., a. b. d. walker, j. b. ayotte, and g. s. watts. 2008. using gis to modify a stratified random block design for moose. alces 44: 111–116. hickey, l. 2008. assessing re-colonization of moose in new york with hsi models. alces 44: 117–126. hoffman, d., h. h. genoways, and j. r. choate. 2006. long-distance dispersal and population trends of moose in the central united states. alces 42: 115–131. judd, s. l. 1972. minnesota’s 1971 moose hunt. proceedings of the north american moose conference and workshop 8: 240–243. karns, p. d. 1998. population distribution, density and trends. pages 125–139 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d. c., usa. kantar, l. e. 2011. broccoli and moose, not always best served together: implementing a controlled moose hunt in maine. alces 47: 83–90. ———, and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose abundance in maine. alces 49: 29–37. kelsall, j. p. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research supplement 1: 1–10. ———, and e. s. telfer. 1974. biogeography of moose with particular reference to western north america. naturaliste canadien 101: 117–130. kilpatrick, h. j., r. riggs, a. labonte, and d.celotto. 2003. history and status of moose in connecticut. connecticut department of environmental protection, hartford, connecticut, usa. http://www. ct.gov/deep/lib/deep/wildlife/pdf_files/ game/moose02.pdf (accessed july 2016). kuzyk, g. 2016. provincial population and harvest estimates of moose in british columbia. alces 52: 1–11. ———, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. british columbia ministry of forestry, lands and natural resources operations, victoria, british columbia, canada. http://www2.gov.bc.ca/assets/gov/ environment/plants-animals-and-ecosys tems/wildlife-wildlife-habitat/wildlifehealth/wildlife-health-documents/2014_ bc_moose_research_design.pdf (accessed july 2016). alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 19 http://www.ct.gov/deep/lib/deep/wildlife/pdf_files/game/moose02.pdf http://www.ct.gov/deep/lib/deep/wildlife/pdf_files/game/moose02.pdf http://www.ct.gov/deep/lib/deep/wildlife/pdf_files/game/moose02.pdf http://www2.gov.bc.ca/assets/gov/environment/plants-animals-and-ecosystems/wildlife-wildlife-habitat/wildlife-health/wildlife-health-documents/2014_bc_moose_research_design.pdf http://www2.gov.bc.ca/assets/gov/environment/plants-animals-and-ecosystems/wildlife-wildlife-habitat/wildlife-health/wildlife-health-documents/2014_bc_moose_research_design.pdf http://www2.gov.bc.ca/assets/gov/environment/plants-animals-and-ecosystems/wildlife-wildlife-habitat/wildlife-health/wildlife-health-documents/2014_bc_moose_research_design.pdf http://www2.gov.bc.ca/assets/gov/environment/plants-animals-and-ecosystems/wildlife-wildlife-habitat/wildlife-health/wildlife-health-documents/2014_bc_moose_research_design.pdf http://www2.gov.bc.ca/assets/gov/environment/plants-animals-and-ecosystems/wildlife-wildlife-habitat/wildlife-health/wildlife-health-documents/2014_bc_moose_research_design.pdf labonte, a. m., h. j. kilpatrick, and j. s. barclay. 2013. opinions about moose and moose management at the southern extent of moose range in connecticut. alces 49: 83–98. laforge, m. p., n. l. michel, a. l. wheeler, and r. k. brook. 2016. habitat selection by female moose in the canadian prairie ecozone. journal of wildlife management 80: 1059–1068. larter, n. c. 2009. a program to monitor moose populations in the dehcho region, northwest territories, canada. alces 45: 89–99. leblanc. j. w., b. e. mclaren, c. pereira, m. bell, and s. atlookan. 2011. first nations moose hunt in ontario: a community’s perspectives and reflections. alces 47: 163–174. legg, d. 1995. the economic impact of moose hunting in ontario, 1993. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. lenarz, m. s. 1998. precision and bias of aerial moose surveys in northwestern minnesota. alces 34: 117–124. ———, j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. lichtenwalner, a., n. adhikari, l. kantar, e. jenkins, and j. schurer. 2014. echinococcus granulosus genotype g8 in maine moose (alces alces). alces 50: 27–33. lynch, g. m. 2006. does first nation’s hunting impact moose productivity in alberta? alces 42: 25–31. ———, and g. e. shumaker. 1995. gps and gis assisted moose surveys. alces 36: 145–151. maskey, j. j. 2011. giant liver fluke in north dakota moose. alces 47: 1–7. matthews, p. e. 2012. history and status of moose in oregon. alces 48: 63–66. mcgraw, a. m., r. moen, g. wilson, a. edwards, r. peterson, l. cornicelli, m. schrage, l. frelich, m. lenarz, and d. becker. 2010. an advisory committee process to plan moose management in minnesota. alces 46: 189–200. mclaren, b. e., b. a. roberts, n. djanchekar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 45–59. ———, and w. e. mercer. 2005. how management unit license quotas relate to population size, density, and hunter access in newfoundland. alces 41: 75–84. millette, t. l., e. marcano, and d. laflower. 2014. winter distribution of moose at landscape scale in northeastern vermont: a gis analysis. alces 50: 17–26. minnesota dnrc (department of natural resources moose advisory committee). 2011. minnesota moose research and management plan. minnesota department of natural resources, st. paul, minnesota, usa. moen, r., m. e. nelson, and a. edwards. 2011a. using cover type composition of home ranges and vhf telemetry locations of moose to interpret aerial survey results in minnesota. alces 47: 101–112. ———, r. o. peterson, s. k. windels, l. frelich, and m. johnson. 2011b. minnesota moose status: progress on moose advisory committee recommendations. nrri technical report no. nrri/ tr-2011–41. minnesota department of natural resources, st. paul, minnesota, usa. murray, d. l. e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t.w. custer, t. barnett, and t. d. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. 20 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers (omnrf) ontario ministry of natural resources and forestry. 2008. ontario moose program review 2008. information package. queen’s printer, toronto, ontario, canada. ———. 2009a. moose management policy. june 2009. queen’s printer, toronto, ontario, canada. ———. 2009b. moose. population objectives setting guidelines. june 2009. queen’s printer, toronto, ontario, canada. ———. 2015. hunting regulations 2015– 2016. ontario ministry of natural resources and forestry. queen’s printer, toronto, ontario, canada. ———. 2016. hunting regulations 2016– 2017. ontario ministry of natural resources and forestry. queen’s printer, toronto, ontario, canada. peek. j. m., and k. i. morris. 1998. status of moose in the contiguous united states. alces 34: 423–434. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. peterson, r. o, and r. e. page. 1993. detection of moose in midwinter from fixed-wing aircraft over dense forest cover. wildlife society bulletin 21: 80–86. rodgers, a. 2001. moose. world life library, voyageurs press, stillwater, minnesota, usa. seaton, k. a. k. 2014. evaluating options for improving gspe performance and developing a sightability correction factor. alaska department of fish and game, division of wildlife conservation, federal aid final research performance report 1 july 2007–30 june 2014, federal aid in wildlife restoration project 1.66, juneau, alaska, usa. smits, c. m., r. m. p. ward, and d. g. larsen. 1994. helicopter or fixed-wing aircraft; a cost-benefit analysis for moose surveys in yukon territory. alces 30: 45–50. smith, c., k. beazley, p. duinker, and k. a. harper. 2010. the impact of moose (alces alces andersoni) on forest regeneration following a severe spruce budworm outbreak in the cape breton highlands, nova scotia, canada. alces 45: 135–150. smith, m. a., s. kilpatrick, b. younkin, l. work, and d. wachob. 2011. assessment of crucial moose winter habitat in western wyoming. alces 47: 151–162. swail, p. j. 1996. blais vs the queen. a conviction under the criminal code of canada on august 22, 1996, for hunting big game out of season contrary to section 26 of the manitoba wildlife act. http://www.canlii.org/ca/cas/scc/2003/ 2003scc44.html. (accessed july 2016). tape, k. d., d. d. gustine, r. w. ruess, l. g. adams, and j. a. clark. 2016. range expansion of moose in arctic alaska linked to warming and increased shrub habitat. plos one | doi: 10.1371/ journal.pone.0152636. telfer, e. s. 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145–182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. thomas, t. p. 2008. moose population management recommendations. wyoming game and fish department, cheyenne, wyoming, usa. https://wgfd.wyo. gov/wgfd/media/content/pdf/habitat/ swap/mammals/moose.pdf (accessed july 2016). thompson, i. d., m. d. flannigan, b. m. wotton, and r. suffling. 1998. the effects of climate change on landscape diversity: an example in ontario forests. environmental monitoring and assessment 49: 213–233. timmermann, h. r. 1987. moose harvest strategies in north america. swedish alces vol. 53, 2017 timmermann and rodgers – status of moose in na circa 2015 21 http://www.canlii.org/ca/cas/scc/2003/2003scc44.html http://www.canlii.org/ca/cas/scc/2003/2003scc44.html https://wgfd.wyo.gov/wgfd/media/content/pdf/habitat/swap/mammals/moose.pdf https://wgfd.wyo.gov/wgfd/media/content/pdf/habitat/swap/mammals/moose.pdf https://wgfd.wyo.gov/wgfd/media/content/pdf/habitat/swap/mammals/moose.pdf wildlife research supplement 1: 565–579. ———. 1993. use of aerial surveys for estimating and monitoring moose populationsa review. alces 29: 35–46. ———. 2003. the status and management of moose in north america – circa 2000– 01. alces 39: 131–151. ———, and m.e. buss. 1995. the status and management of moose in north americaearly 1990’s. alces 31: 1–14. ———, and ———. 1998. population and harvest management. pages 559–615 in a. w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. ———, r. gollat, and h. a. whitlaw. 2002. reviewing ontario’s moose management policy 1980–2000: targets achieved, lessons learned. alces 38: 11–45. ———, and a. r. rodgers. 2005. moose: competing and complementary values. alces 41: 85–120. todesco, c. 2004. illegal moose kill in northeastern ontario: 1997–2002. alces 40: 145–159. toweill, d. e., and g. vecellio. 2004. shiras moose in idaho: status and management. alces 40: 33–43. tyers, d. b. 2006. moose population history on the northern yellowstone winter range. alces 42: 133–149. vucetich, j. a., and r. o. peterson. 2015. ecological studies of wolves on isle royale. annual report, michigan technological university, houghton, michigan, usa. ward, m. p., w. c. gasaway, and m. m. dehn. 2000. precision of moose density estimates derived from stratification survey data. alces 36: 197–203. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. ———, and ———. 2013. space use and movements of moose in massachusetts: implications for conservation of large mammals in a fragmented environment. alces 49: 65–81. west, r. l. 2009. moose conservation in the national wildlife refuge system, usa. alces 45: 59–65. wilton, m. l. 1995. the case against calling and hunting moose during the main rut period-a viewpoint. alces 31: 173–180. wolfe, m. l., k. r. hersey, and d. c. stoner. 2010. a history of moose management in utah. alces 46: 37–52. wünschmann, a., a. g. armien, e. butler, m. schrage, b. stromberg, j. b. bender, a. m. firshman, and m. carstensen. 2015. necropsy findings in 62 opportunistically collected freeranging moose (alces alces) from minnesota, usa (2003–2013). journal of wildlife diseases 51: 157–165 doi: 10.7589/2014–02-037. young, d. d. jr., and r. d. boertje. 2004. initial use of moose calf hunts to increase yield, alaska. alces 40: 1–6. ———, and ———. 2008. recovery of low bull:cow ratios of moose in interior alaska. alces 44: 65–71. ———, and ———. 2011. prudent and imprudent use of antlerless moose harvests in interior alaska. alces 47: 91–100. ———, ———, c. t. seaton, and k. a. kellie. 2006. intensive management of moose at high-density: impediments, achievements, and recommendations. alces 42: 41–48. xu, c., and m. s. boyce. 2010. optimal harvesting of moose in alberta. alces 46: 15–35. 22 status of moose in na circa 2015 – timmermann alces vol. 53, 2017 and rodgers the status and management of moose in north america -circa 2015 historical distribution and current status eastern north america western north america central north america factors affecting decreasing densities harvest management economic impact harvest control objectives allocation of hunting opportunities control concepts license qualifications and fees seasons management areas and harvest strategies harvest assessment moose hunter education and engagement harvest by native and subsistence users managing unhunted populations parks, refuges, and special areas discussion acknowledgements references alces 56 (2020) contents abundance of winter ticks (dermacentor albipictus) in two regenerating forest habitats in new hampshire, usa ...................................... brent i. powers and peter j. pekins 1 metrics of harvest for ungulate populations: misconceptions, lurking variables, and prudent management . .................... terry bowyer, kelley m. stewart, vernon c. bleich, jericho c. whiting, kevin l. monteith, marcus e. blum, and tayler n. lasharr 15 revisiting the recruitment-mortality equation to assess moose growth rates .......................................... ian w. hatter 39 tracking mooseand deer-vehicle collisions using gps and landmark inventory systems in british columbia ................................... caleb sample, roy v. rea, and gayle hesse 49 a review of circumpolar moose populations with emphasis on eurasian moose distributions and densities ................................................... william f. jensen, roy v. rea, colin e. penner, jason r. smith, eugenia v. bragina, elena razenkova, linas balciauskas, heng bao, stanislav bystiansky, sándor csányi, zuzana chovanova, gundega done, klaus hackländer, marco heurich, guangshun jiang, alexander kazarez, jyrki pusenius, erling j. solberg, rauno veeroja, and fredrik widemo 63 assessing moose hunter distribution to explore hunter competition .............................. tessa r. hasbrouck, todd j. brinkman, glenn stout, and knut kielland 79 assessing age of harvested moose prior to population declines in british columbia ............. gerald w. kuzyk, kaitlyn d. schurmann, shelley m. marshall, and chris procter 97 browse selection by moose in the adirondack park, new york ...........................................samuel peterson, david kramer, jeremy hurst, and jacqueline frair 107 estimation of moose parturition dates in colorado: incorporating imperfect detections .........................eric j. bergman, forest p. hayes, and kevin aagaard 127 53rd north american moose conference and workshop carrabassett valley (sugarloaf), maine june 10-14th, 2019 ............ 137 previous meeting sites of the north american moose conference and workshop and international moose symposia ............................................................................................................... 139 distinguished moose biologist – 2019 recipient .............................. 141 distinguished moose biologist – award criteria ........................... 143 editorial review committee ......................................................................... 144 additional copies additional copies and back issues of alces (issn 0835-5851 ‒ called proceedings of the north american moose conference and workshop up to vol. 16, 1980), most international moose symposia, and special symposia can be purchased online through the lakehead university alumni bookstore ‒ http://bookstore.lakeheadu.ca/esolution/course.php. all past publications are priced at cdn or us $49.99 each plus applicable tax (gst or hst). price includes mailing and handling costs. prices are subject to change. for more information contact the alces business editor. alces home pages further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our websites: http://flash.lakeheadu.ca/~arodgers/alces/alces.html. http://alcesjournal.org http://bookstore.lakeheadu.ca/esolution/course.php http://flash.lakeheadu.ca/~arodgers/alces/alces.html http://alcesjournal.org 141 lee kantar distinguished moose biologist – 2019 recipient the distinguished moose biologist award was presented to lee kantar at the 53rd north american moose conference and workshop held in carrabassett valley (sugarloaf), maine in june of 2019. this award is in recognition of his contributions to our understanding of moose ecology and management, managing the largest moose population in the lower 48 states, and building one of the most progressive and research-oriented management programs in north america as moose project leader for the maine department of inland fisheries and wildlife. lee grew up in new hampshire and spent his childhood freely roaming the woods around his home and exploring his dad’s woodlot when not being put to work. lee was fortunate enough to grow up in an era when a young kid could leave the house in the morning and roam free until called in for supper. the ability to explore the woods and streams of new hampshire set the stage for his future wildlife career. lee graduated from brandeis university with a degree in anthropology and after his last exam left for georgia to walk back to maine on the appalachian trail. five years of work in the outdoor recreation and education field led him back to the university of new hampshire to finish a 2nd undergraduate degree in wildlife management under the tutelage of dr. pete pekins, head of the wildlife program. lee spent 5 years working in various wildlife positions including natural resource manager for conservation organizations as well as the washington department of fish and wildlife. he worked as a technician on various projects including in fisheries and with black bears. later he decided to pursue his ms degree at new mexico state university within the new mexico cooperative fish and wildlife research unit where he investigated resource conflicts with a migratory elk herd along the new mexicocolorado border. in new mexico lee met his wife danielle who is also a wildlife biologist. he focused his work on applying research to management and began work for the us forest service as a wildlife biologist on the darrington ranger district in the north cascades of washington state. later he worked as district wildlife biologist for the washington department of fish and wildlife before danielle and he moved back to new england. in 2005 he became the deer specialist for the maine department of inland fisheries and wildlife (mdifw), and when kim morris retired in 2007, assumed the responsibilities for moose management as well. in 2009 lee was instrumental in establishing a controlled moose hunt to alleviate moose depredation of crops. working in partnership with multiple stakeholders, he laid the groundwork for a long-standing program that morphed into a disabled veteran’s hunt that relieves farmers of perennial crop damage. this unique program offers disabled veterans an opportunity to establish bonds with fellow veterans and help in addressing personal issues in a positive outdoor atmosphere. in 2010 lee attended his first north american moose conference in minnesota and quickly discovered a community of mentors that proved invaluable in understanding moose biology and management. he also began an aerial survey program that provided a new level of information for moose in maine. in 2014 he collaborated with the university of new hampshire (pete pekins) and the new hampshire fish and game department (kristine rines) to co-lead a regional study focused on adult cow and calf survival in light of increased winter tick impacts. expanding work into 3 separate study areas, mdifw has gps-collared and monitored over 600 moose since the beginning of the project. lee is considered a regional leader and authority of moose management and has published several papers on moose management and techniques. importantly, he mentors students and young biologists, serves on graduate student committees, and has given scores of public 142 presentations as the face of maine moose. his work has been featured in yankee magazine, northern woodlands, the boston globe, the new york times, as well as maine public, national geographic, and discovery channel-canada. lee organized the 53rd north american moose conference in carrabassett valley (sugarloaf), maine and continues as a regional leader in moose management. he and danielle, also a wildlife biologist for mdifw, live in orrington, maine where they raise their beloved daughters ella and wren. along with their two dogs they enjoy spending time outdoors throughout the year and exploring the wide variety of places that maine offers from the mountains to the ocean. both daughters have tagged along on many wildlife adventures from christmas bird counts with mom to relocating wayward moose with dad. the girls have learned to realize that a pair of binoculars is standard equipment even on a short trip. and although at times exasperated by the endless focus on wildlife, the girls have provided hints that perhaps a similar career is in their future! the effects of sex, terrain, wildfire, winter severity, and maternal status on habitat selection by moose in north-central alaska kyle joly1, mathew s. sorum1, tim craig2,4, and erin l. julianus3 1national park service, gates of the arctic national park and preserve, 4175 geist road, fairbanks, ak 99709; 2us fish and wildlife service, kanuti national wildlife refuge, 101 12th avenue, fairbanks, ak 99701; 3bureau of land management, central yukon field office, 1150 university avenue, fairbanks, ak 99709; 4retired abstract: habitat selection is a central component of the ecology of individual animals as it affects body condition, survivorship, and reproductive output. we instrumented male and female moose (alces alces) in north-central alaska with gps radio-collars to assess factors we hypothesized were important to their habitat selection. using synoptic modeling techniques, we found that models with more covariates were better predictors of moose habitat selection than more simplistic models. as expected, moose selected for habitats with high canopy cover and/or that typically have abundant forage such as 11-30 year old burned areas. however, we detected differences in habitat selection between sexes, seasons (i.e., winter versus summer), during winters of varying severity, and females with differing maternal status. during winter males moved to lower elevations areas, presumably to avoid greater snow depths, whereas females remained at relatively similar elevations. females selected burned habitat and areas that received higher amounts of solar radiation. we found that all moose selected for lower elevation habitats closer to rivers during moderate and severe winters, but elevation was not a strong influence during mild winters. we found that females with calves avoided riparian habitats and selected areas with more forested habitat than females without calves during both summer and winter. this suggests a trade-off between maximizing forage intake and reducing predation risk for their offspring. our and similar data are useful to improve moose management strategies and provide a benchmark against which the impacts of climate change and industrial development are assessed in this rapidly-changing region. alces vol. 52: 101–115 (2016) key words: alces alces, arctic, brooks range, maternal status, moose, synoptic modeling, winter severity an animal’s use of the landscape affects its body condition, reproductive output, survivorship, and fitness (gaillard et al. 2010). thus, studies of habitat selection are also informative to understanding the ecology of vagile species. although habitat selection by moose (alces alces) has been well documented in north american populations (see peek 1997), there is a paucity of habitat selection studies in northern alaska. patterns of selection by moose differ among and within populations, and between sexes and seasons. alaskan moose are sexually dimorphic in body size, and sexual segregation is well documented (miquelle et al. 1992, bowyer et al. 2001, oehlers et al. 2011). barboza and bowyer (2000) suggested that sexrelated differences in habitat selection patterns can be explained by differences in body size and annual changes in the physiology and morphology between sexes. large males are able to consume large quantities of low-quality forage, whereas smaller-bodied females are better adapted for smaller 101 quantities of high-quality forage. risk of predation is also thought to play a major role in the selection of habitats by moose, particularly females with calves (dussault et al. 2005, poole et al. 2007, oehlers et al. 2011). moose reduce risk of predation by avoiding travel routes used by predators (kunkel and pletscher 1999, dussault et al. 2005) and selecting habitats that provide greater concealment (oehlers et al. 2011). terrain features and snow conditions also influence patterns of distribution and selection by moose (poole and stuart-smith 2006). within interior mountain areas, moose tend to descend to lower elevation valley bottoms during winter (poole and stuart-smith 2006). thus, differences in habitat selection patterns among populations of moose are dependent on local conditions with respect to forage, predators, and weather. climate change is predicted to profoundly affect land mammals in the arctic (lawler et al. 2009, marcot et al. 2015). wildfire is already common in the region (joly et al. 2009) and is predicted to increase under warming scenarios (kasischke and turetsky 2006, johnstone et al. 2010, joly et al. 2012). early seral stage shrub communities, which follow wildfires, provide abundant high quality forage for moose (schwartz and franzmann 1989). moreover, moose populations have increased where these early seral habitats have expanded due to wildfire (spencer and hakala 1964, schwartz and franzmann 1989). increased shrub abundance has been documented around the arctic and is thought to be linked to warming (tape et al. 2006). thus, climate change may produce more moose habitat and more moose in this region if patterns of selection for early seral stage and shrubby habitats by moose are similar to other areas of the boreal forest (joly et al. 2012). understanding current patterns of habitat selection will aid in assessing the effects of climate change into the future. we analyzed habitat selection by moose on the southern flanks of the brooks range and the adjacent lowlands in north-central alaska (fig. 1), near the northern extent of moose range in this region. our goal was to provide information about habitat selection patterns in alaska’s arctic interior to improve moose management. we focused on selection within the home range across individuals during winter and summer seasons using variables we believed important to moose. we assessed whether patterns of habitat selection were driven primarily by spatial factors related to abundance of adequate forage, predator avoidance, or physiography. we hypothesized that habitat selection would be driven by a complex mix of factors, highlighting the trade-offs among access to forage, energy expenditure, and exposure to predation pressure. further, we hypothesized that maternal status and winter severity would influence patterns of habitat selection. we expected females with calves to select more forested areas further from rivers than females without calves, presumably to reduce predation risk, and that moose would select areas lower in elevation during more severe winters. methods study area this study took place in the upper reaches of the koyukuk river in northcentral alaska (fig. 1). the area supported a low density (~ 0.1 moose/km2) moose population (lawler et al. 2006), as well as the full complement of naturally occurring species including caribou (rangifer tarandus), dall’s sheep (ovis dalli), wolves (canis lupus), grizzly bears (ursus arctos), and black bears (u. americanus). the upper koyukuk river drainage had a strong continental climate with short, hot summers and long, cold winters. temperatures dropped below – 45 °c and snow persisted on the ground from october until may (western regional climate center, www.wrcc.dri.edu/). 102 habitat selection in alaska – joly et al. alces vol. 52, 2016 www.wrcc.dri.edu/ snow pack was typically >60 cm most winters and often >90 cm. summers were brief but temperatures can exceed 30 °c. large wildfires were common during warm dry summers, particularly south of the brooks range which consisted of boreal forest vegetation dominated by fire-prone communities such as black spruce (picea mariana) forests. fig. 1. moose habitat selection and use study area (white polygon) in north-central alaska, 2008– 2013. gps locations (dots) of individual moose are color-coded. alces vol. 52, 2016 joly et al. – habitat selection in alaska 103 the northern half of the study area consisted of the central brooks range rugged mountains that reach up to 2000 m in elevation that contain narrowly-confined glacial river valleys, and where wildfire is much less common. the valleys supported spruce and birch (betula papyrifera) forests, tussock tundra, shrub lands (alnus spp., salix spp.), and muskeg. tall and low shrub communities occurred on hillsides, but eventually gave way to alpine vegetation. this area included the southeastern portion of gates of the arctic national park and preserve (gaar) and lands managed by the bureau of land management (blm) and the state of alaska. the southern portion of the study area was much less rugged and lower in elevation; typically about 300 m above sea level with hills generally lower than 500 m. it had more wetland habitat, was extensively forested, and wildfires were prevalent. the southern portion of the study area primarily contained lands managed by the kanuti national wildlife refuge (knwr), the state of alaska, and the blm. the town of bettles, alaska was in the middle of the study area. moose capture, gps data, maternal status, winter severity we captured adult male and female moose between march 2008 and april 2011 via aerial darting. moose were fitted with gps radio-collars (telonics tgw-4780) that also had a very high frequency (vhf) radio beacon (joly et al. 2015a); collars were removed when the project ended in april 2013. collars collected 3 locations/ day except those deployed in march 2008. for our analyses, all location datasets began on 15 may, and we excluded all individualyears that were sampled <330 days. maternal status, as indicated by the presence or absence of a calf, was determined by tracking collared females in small, fixedwing aircraft (e.g., piper pa-18 supercub). we attempted to locate all collared females just after calving (late may–early june), in the fall (september–october), and during the following spring (march–april) to visually determine if the female was accompanied by a calf. if we could not make this determination, the individual was excluded from analyses related to maternal status. we classified each winter as mild, moderate, or severe based on the total number of days with snow and snow depth as recorded in bettles, alaska (joly et al. 2015a). the 3 classifications were: 1) mild winters had <135 days with ≥30 cm snow or <7 days with ≥60 cm snow, 2) moderate winters had >170 days with ≥30 cm snow, >50 days with ≥60 cm, or <14 days with ≥90 cm snow, and 3) severe winters had >170 days with ≥30 cm snow, >100 days with ≥60 cm, or >30 days with ≥90 cm snow. we used these non-continuous categories to highlight that the classifications were distinctive – all winters fell into a single category. two winters (2009–10, 2012–13) were categorized as mild, 3 (2007–08, 2010–11, and 2011–12) as moderate, and 1 (2008–09) as severe. we defined biological seasons as summer (1 july–24 august) and winter (16 december–14 may) based on regional weather patterns. study design scale is critical to understanding ecological processes (wiens 1989, wheatley and johnson 2009, decesare et al. 2012). habitat preferences modify with changes in the relative amount of available habitat (osko et al. 2004, herfindal et al. 2009). due to physiographic differences between the northern and southern portions of our study area, we estimated seasonal habitat selection by moose at the home range scale (3rd order; johnson 1980) using the synoptic model of space use (horne et al. 2008, slaght et al. 2013). this model uses a 104 habitat selection in alaska – joly et al. alces vol. 52, 2016 weighted distribution to simultaneously model an individual’s space use and habitat selection (johnson et al. 2008) within its home range, and is capable of estimating home range and resource selection simultaneously. thus, the probability of use at location x and time t was modeled using: f u x; tð þ ¼ f a xð þ � wðx; tþr f a xð þ � wðx; tþ ð1þ where f a xð þ is the null distribution of space use that models the probability of use in the absence of habitat selection (i.e., the availability distribution), and wðx; tþ is a selection function that transforms f a xð þ to f u x; tð þ by selectively weighting different areas based on habitat conditions (johnson et al. 2008). we defined f a xð þ ¼ bvn hð þ to be a stationary (i.e., time invariant) bivariate normal (bvn) distribution with parameters h describing the means and variances in the x and y dimensions and the covariance. by describing f a xð þ in this way, the areas considered available for selection can be thought of as a bvn distribution characterizing the entire home range of an individual. the bvn distribution characterizes the space use of an animal that biases movement towards a central place (horne et al. 2008, wilson et al. 2014b). we defined the selection function as: w x; tð þ ¼ exp h xð þ0bpðtþ h i ð2þ where h xð þ0 is a vector of covariate values describing the habitat or environmental conditions at location x, b is a vector of parameters (i.e., selection coefficients) to be estimated, and p(t) is an interaction term representing functions of time (i.e., winter, summer) to allow for temporal variation in habitat selection. others have used similar approach for modeling habitat selection through time (see ferguson et al. 2000, forester et al. 2009). we used maximum likelihood (via numerical optimization) to estimate the parameters governing the null model of home range (θ) and the selection coefficients (β) with a program written in r (r development core team 2013) with code developed by j. horne (see slaght et al. 2013 for example code). we used odds ratios to aid interpretation of the estimated coefficients βi. an odds ratio approximates the relative change in probability of event x occurring (e.g., a moose being present) given a 1-unit change in a given parameter (hosmer and lemeshow 1989). environmental variables based on previous research, we formulated 11 models to analyze seasonal habitat selection by moose (table 1). weixelman et al. (1998) and maier et al. (2005) suggested that moose select habitats that burned 11–30 years prior to usage because these areas tend to revegetate with deciduous shrubs. riparian zones often have abundant and high-quality forage that moose use in alaska (collins and helm 1997, maier et al. 2005, stephenson et al. 2006). areas with extensive vegetative cover typically have more deciduous trees (e.g., birch) and tall shrubs (e.g., willows) that are preferred moose forage than areas with low cover (e.g., tussock tundra). we expected moose to select areas that contained preferred forage, such as forested and burned habitat, and areas closer to rivers. the models that highlighted the importance of forage contained a mixture of covariates that included ‘fire’ (if a moose was in habitat that burned 11–30 years prior to use), ‘forest’ (if a moose used areas with extensive vegetative cover based on landcover type), and/or ‘dist_river’ (distance from a riparian area; table 1). we identified areas as ‘fire’ using the alaska fire service’s geodatabase which catalogs the extent, number, and location of large fires mapped from 1950–2014 (fig. 2; alces vol. 52, 2016 joly et al. – habitat selection in alaska 105 data at http://fire.ak.blm.gov/predsvcs/maps. php), and ‘forest’ using the national land cover database – alaska 2001 coverage (http://www.epa.gov/mrlc/nlcd-2001.html). major rivers were identified using the usgs 1:2,000,000 digital line graphs dataset (https://lta.cr.usgs.gov/dlgs). we expected moose, particularly females with calves, to select areas further from riparian areas and that were more forested to reduce predation pressure. riparian areas are often utilized by predators as travel corridors and forested areas provide more cover to hide from predators (peterson 1995, kunkel and pletscher 1999, mcphee et al. 2012). thus, we interpreted moose responses to riparian areas as a proxy for responding to areas of increased predation risk. the model highlighting the importance of predation pressure included the covariates of ‘dist_river’ and ‘forest’. moose select areas based on physiography and 4 models were used to assess the importance of terrain including a mixture of the covariates slope, elevation and their squared terms (to assess non-linear relations), and ‘sri’ (a solar radiation index, keating et al. 2007; table 1). these covariates were derived from our digital elevation model. higher solar radiation is correlated with reduced snow depth during winter and increased net primary productivity (i.e., forage) during summer (crabtree et al. 2009). we expected moose to select for terrain features that reduced snow depth, and subsequently increased forage availability, and that these patterns would be more prominent during more severe winters. we hypothesized that habitat selection by moose is influenced by a wide array of factors, rather than just forage abundance, predation pressure, or terrain acting alone. we used 4 models to assess this hypothesis, and covariates for these models included the entire suite used in the previous models. we used remotely sensed data to quantify the spatial distribution of habitat covariates and included interaction terms between resource selection coefficients and functions of time (ferguson et al. 2000, forester et al. 2009) to account for temporal variation in habitat selection. before modeling resource selection, we screened predictor variables for collinearity. we assumed that if │r│< 0.60, then correlation was not a concern between predictor covariates (sawyer et al. 2006, ciarniello et al. 2007). slope and elevation were considered positively correlated and were not included together in any model. model selection we used an information-theoretic approach for evaluating synoptic models of table 1. models and their structure used to analyze different hypotheses related to moose habitat selection in north-central alaska, usa, 2008– 2013. model covariates fire firea forage fire+dist_riverb+forestc predator dist_river+forest terrain1 srid+eleve terrain2 sri+slope terrain3 sri+elev+elev2 terrain4 sri+slope+slope2 complexity1 fire+forest+dist_river+sri+elev complexity2 fire+forest+dist_river+sri +slope complexity3 fire+forest+dist_river+sri+elev +elev2 complexity4 fire+forest+dist_river+sri +slope+slope2 a ‘fire’ denoted if a moose location was in habitat that burned 11-30 years prior to use b ‘dist_river’ is distance to a riparian area a moose was located c ‘forest’ denoted if a moose location was in habitat that was extensively vegetated (i.e. forest or tall shrubs) d ‘sri’ is a solar radiation index e ‘elev’ is elevation 106 habitat selection in alaska – joly et al. alces vol. 52, 2016 http://fire.ak.blm.gov/predsvcs/maps.php http://fire.ak.blm.gov/predsvcs/maps.php http://www.epa.gov/mrlc/nlcd-2001.html https://lta.cr.usgs.gov/dlgs habitat selection and determined a set of a priori candidate models that we deemed biologically relevant (burnham and anderson 2002). we fit models to location data for each individual and year. we ranked the models for each moose and year using the difference in akaike information criterion adjusted for small sample size (aicc) from the model with the smallest value (δaicc), and determined the relative likelihood of each model using akaike weights (burnham and anderson 2002). models, including the top model, which had an aicc score of <2 from the top model were designated as being in the top model set. we evaluated habitat selection by sex, winter severity, and maternal status. we averaged estimates of selection coefficients across models based on akaike weights for each individual and year. we scaled the weights to total 1 across models containing each variable (burnham and anderson 2002). for individuals that we observed during multiple years, we averaged the value of estimated-coefficients across years. to make class-level (i.e., sex, maternal status, and severity of the winter) inferences, we calculated the means and standard errors of univariate parameter estimates across all individuals for each parameter. if a parameter (e.g., fire) was not used by an individual, then no estimate was included for that individual for class-level inferences. for a conservative measure of precision at the class-level, we considered a coefficient to be significant if 2 times the standard error of the mean did not contain zero (boyce 2006, fieberg et al. 2010). complete separation of the data occurred where habitats were available but not used by a moose. for these individuals, we did not estimate a coefficient for the variable but simply noted avoidance (e.g., nielsen et al. 2002). we entered elevation and slope as quadratic terms to allow for selection, or avoidance, at intermediate values of elevation and slope. results we retrieved 71,675 gps locations from 37 moose between march 2008 and april 2013 via remote download and collar retrieval; 6 moose did not provide enough data to be included in our analyses. the remaining 31 moose (20 females and 11 males) produced 70 moose-years of data (range: 1–4 years per individual). for male moose, the complexity3 model (which included the covariates fire, forest, dist_river, sri, elev, and elev2) best described habitat selection within home ranges during both winter and summer (tables 1 and 2). complexity3 was in the top model set for 48% and 46% of the individual moose-years during the winter and summer, respectively. complexity1 was in the top model set for 29% and 38% of the individual moose-years during the winter and summer, respectively. both complexity4 and terrain3 were in the top model set for 10% of the individual moose-years during the winter, and complexity4 in the top model set for 8% of the individual mooseyears during the summer. the remaining models were in the top model set for ≤10% of the individual moose-years during either season (table 2). for female moose, the complexity3 model best described habitat selection within home ranges during both winter and summer (tables 1 and 3). complexity3 was in the top model set for 49% and 41% of the individual moose-years during the winter and summer, respectively. complexity2 was in the top model set for 14% and 24% of the individual moose-years during the winter and summer, respectively. terrain3 was in the top model set for 16% of the individual mooseyears during the winter, and complexity4 in the top model set for 16% of the individual moose-years during the summer. alces vol. 52, 2016 joly et al. – habitat selection in alaska 107 the remaining models were in the top model set for ≤10% of the individual moose-years during either season (table 3). nearly half (11 of 23) of the moose in the northern portion of the study area, where wildfire is less common than in the southern portion, did not use burned habitat during either winter or summer. all 8 moose in the southern portion of the study area used burned habitat, with 1 animal located only within burned habitat. seasonal selection patterns by moose patterns of selection by moose varied between season and sex (table 4). male moose consistently selected areas that were forested, lower in elevation, and with gentler slopes in winter; during summer they selected areas that were forested. during both seasons males were more variable in their selection of areas that received higher amounts of solar radiation, that were closer to riparian habitat, or that had been burned. across seasons, female moose consistently selected areas that were forested, burned, and lower in elevation. further, during winter females selected areas that received higher amounts of solar radiation, and during summer they avoided steeper slopes. distance to riparian habitat was not consistently selected or avoided by females during either season. winter severity the severity of the winter influenced habitat selection. as expected, moose selected areas lower in elevation with gentler slopes during more severe winters suggesting that snow depth influenced habitat selection. based on average probability ratios, moose were 56% less likely to select a location for every 100 m higher during severe winters, but only 6% less likely during mild winters. in addition, moose selected areas closer to rivers during more severe winters. during mild table 2. top models of habitat selection by male moose in north-central, alaska, usa, 2008–2013. the number of individual-years of data (n) for which each of the top 3 models of habitat selection received the most support, average and range of akaike weights, and percent of times (%) each model occurred in the top model set (<2 aicc of the top model) are presented by season. winter summer model n akaike weight % n akaike weight % complexity3 10 0.95 (0.62–1.00) 48 11 0.85 (0.22–1.00) 46 complexity1 6 0.92 (0.55–1.00) 29 9 0.77 (0.30–1.00) 38 complexity4 2 0.69 (0.38–1.00) 10 2 0.47 (0.24–0.70) 8 terrain3 2 1.00 (1.00–1.00) 10 table 3. top models of habitat selection by female moose in north-central, alaska, usa, 2008-2013. the number of individual-years of data (n) for which each of the top 3 models of habitat selection received the most support, average and range of akaike weights, and percent of times (%) each model occurred in the top model set (<2 aicc of the top model) are presented by season. winter summer model n akaike weight % n akaike weight % complexity3 31 0.88 (0.32–1.00) 49 26 0.79 (0.25–1.00) 41 complexity1 9 0.73 (0.27–1.00) 14 15 0.77 (0.27–1.00) 24 complexity4 10 0.66 (0.17–0.98) 16 terrain3 10 0.79 (0.23–1.00) 16 108 habitat selection in alaska – joly et al. alces vol. 52, 2016 winters moose were more variable in their selection of most land-cover classes and landscape features (table 4). maternal status six females successfully raised at least 1 calf through to the following spring, 11 lost calves by fall, and 11 either did not give birth or lost their calves during the first month post-birth. we were unable to determine the maternal status of 9 females. during both seasons, females with calves selected areas further from rivers, more forested, and with less burned habitat than females without calves (table 5). for example, based on average probability ratios, females with calves were 20% more likely to select a site 1000 m further from a river, whereas females without calves were 13% less likely to be found there. females with calves were 70% more likely to be in forested habitat, whereas females without calves were only 40% more likely to be there. discussion similar to studies in other northern regions, we found that moose in north-central alaska selected for habitats with extensive canopy cover. where habitat that burned 11–30 years previous was available, moose, particularly females, selectively used it (presumably) because habitats at this seral stage tend to have abundant forage (maccracken and viereck 1990, weixelman et al. 1998, maier et al. 2005). this appears to support the hypothesis that moose habitat selection is primarily driven by availability of forage abundance and quality (peek 1997). however, table 4. average parameter estimates (β) used to characterize selection by moose in north-central, alaska, usa, 2008–2013. bold values were significant at the class level (i.e., sex and population). values in parentheses represent n for each class. ‘f’ denotes female and ‘m’ male. winter summer winter male (11) female (20) male (11) female (20) mild (19 f, 8 m) mod/severe (18 f,11m) firea,x �1.12 1.37 0.02 3.57 1.15 0.10 srib �1.98 0.70 2.26 0.60 0.19 �0.64 non-linear elevc elev 5.08 35.89 65.27 52.71 50.68 3.23 elev2 �30.66 �86.05 �103.20 �114.05 �107.74 �33.61 non-linear slope slope 2.87 2.58 12.53 2.52 5.12 0.62 slope2 �11.24 �5.74 �18.94 �13.66 �8.72 �7.09 dist_riverd �3.08 �1.18 �2.17 0.50 �0.25 �2.29 foreste 0.34 0.39 0.76 0.36 0.46 0.33 elev �22.38 �4.63 �2.47 �6.94 �1.21 �17.19 slope �3.67 �0.89 0.41 �2.85 �0.32 �2.73 a ‘fire’ denoted if a moose location was in habitat that burned 11–30 years prior to use b ‘sri’ is a solar radiation index c ‘elev’ is elevation d ‘dist_river’ is distance to a river a moose was located e ‘forest’ denoted if a moose location was in habitat that was extensively vegetated (i.e. forest or tall shrubs) x not all moose utilized recently burned habitat so sample sizes were: winter male, n=4; winter female, n=12; summer male, n=7; summer female, n=10 alces vol. 52, 2016 joly et al. – habitat selection in alaska 109 this hypothesis was not supported by our top models that included the greatest number of variables. the majority (>66%) for both males and females included indices of forage abundance (time since last fire), extensive vegetative cover (forest), distance to river, in addition to elevation and solar radiation. while many of these covariates can be associated with forage abundance, our results suggest that a wide array of factors likely influence habitat selection by moose – supporting the hypothesis that habitat selection by moose is driven by a complex interaction of diverse factors. we found that patterns of habitat selection differed between sex and season. male and female moose exhibited similar patterns of selection for terrain features, particularly elevation during summer. during winter, however, sex-related differences were evident. as expected, males moved to lower elevations, but unexpectedly, females remained at similar elevations throughout the winter. this behavioral difference might provide smaller-bodied females some benefit of higher quality forage (i.e., in burned areas) and terrain features that reduced snow pack (i.e., higher sri). however, predation on moose is often greater in lowland areas (fuller and keith 1980). within the region, wolves often travel along riparian corridors (lake et al. 2013). we suspect that predators focus their hunting along riparian corridors, owing to the concentration of prey in areas of lower snow depth and that travel is probably easier for predators due to smooth, hard surfaces afforded by rivers (peterson 1995, kunkel and pletscher 1999, mcphee et al. 2012). moose reduce their vulnerability to wolf predation by avoiding areas used by wolves for table 5. average parameter estimates (β) used to characterize selection by maternal status of female moose in north-central, alaska, usa, 2008–2013. bold values were significant. values in parentheses represent n for each class. winter summer calf (6) no calf (22) calf (17) no calf (11) fireax �0.30 0.73 4.16 7.23 srib 0.14 0.39 0.41 0.58 non-linear elevc elev 43.02 39.93 31.59 47.95 elev2 �84.31 �79.63 �95.45 �129.86 non-linear slope slope 2.47 2.31 0.63 4.93 slope2 �1.79 �5.42 �10.49 �17.71 dist_riverd 1.62 �1.26 1.90 �0.59 foreste 0.51 0.34 0.38 0.24 elev �3.26 �0.60 �6.80 �5.79 slope 0.27 �0.52 �3.24 �1.13 a ‘fire’ denoted if a moose location was in habitat that burned 11–30 years prior to use b ‘sri’ is a solar radiation index c ‘elev’ is elevation d ‘dist_river’ is distance to a river a moose was located e ‘forest’ denoted if a moose location was in habitat that was extensively vegetated (i.e. forest or tall shrubs) x not all moose utilized recently burned habitat so sample sizes were: winter calf, n=1; winter no calf, n=10; summer calf, n=2; summer no calf, n=7 110 habitat selection in alaska – joly et al. alces vol. 52, 2016 travel (kunkel and pletscher 1999). our findings are consistent with the hypothesis that females try to minimize predation risk, whereas males adopt a strategy to maximize forage intake (fuller and keith 1980, oehlers et al. 2011). moose, with their large size and formidable strength, are well-adapted to snow (telfer and kelsall 1979, peek 1997). nevertheless, deep (65–70 cm) snow can affect moose movement, distribution, and home range size (van ballenberghe 1977, miquelle et al. 1992, peek 1997, ball et al. 2001, joly et al. 2015b). as expected, during severe winters moose selected habitats that were at lower elevations, with gentler slopes, and closer to rivers than during mild winters. deep snow at midto high elevations, or in early successional stages of burns, can cover preferred browse inducing moose to move to lower elevations and use areas where forage is more concentrated within riparian areas (weixelman et al. 1998). relatively higher moose densities in valley bottoms during severe winters may attract predators such as wolves, and thus increase localized predation risk (mcphee et al. 2012, lake et al. 2013). thus, harsh winters may have indirect, as well as direct, negative impacts on moose in our study area that may reduce their productivity and survivorship. we found that moose were more variable in their selection of land-cover classes and landscape features during mild winters, suggesting that tendencies for moose to select lower elevational areas closer to rivers during winter may be more related to snow depth, and subsequently, forage availability not forage quality. interestingly, we did not find that moose utilized habitats with higher canopy cover during severe winters which may reflect that many trees are diminutive in our highlatitude study area. even though forest stands can have relatively high canopy cover, there may be insufficient overhead foliage to intercept snow and reduce underlying depths, or provide thermal cover. we found that maternal status influenced patterns of habitat selection by females. as expected, females with calves avoided riparian habitats and selected areas with more forested habitat than females without calves during both summer and winter. in addition, females with calves selected areas with less burned habitat than females without calves. both riparian habitat and burned areas tend to provide more high quality moose forage than other habitat types (collins and helm 1997, maier et al. 2005, stephenson et al. 2006). these results suggest that maternal status-related differences in habitat selection patterns were likely more related to the specific needs of females with regard to protection of calves. however, due to the small sample size of females with calves (n = 6), our results should be considered preliminary. habitat selection is fundamental to the ecology of wildlife species. understanding patterns of habitat selection by moose can improve their management. an obvious example to use this information is to help guide where and when development occurs to minimize loss of critical moose habitat. our work is timely, given that a proposed industrial road would bisect the study area (wilson et al. 2014a, guettabi et al. 2016). further, enhanced knowledge of moose movements, distribution, and habitat selection should be useful to abate conflicts between subsistence and non-subsistence hunters by spatially or temporally separating users near high quality moose habitat. the arctic is undergoing rapid warming which will result in measurable ecological changes (hinzman et al. 2005, ipcc 2007). further, wildfire is predicted to increase in the region, potentially creating more productive foraging habitats for moose (joly et al. 2012). by collecting baseline data on habitat selection and use, future researchers will be better able to assess the impacts of alces vol. 52, 2016 joly et al. – habitat selection in alaska 111 climate change on moose at their northern extent of range in north america. acknowledgements funding for this project was provided by the national park service, us fish and wildlife service, blm, and the alaska department of fish and game. we thank pilots t. cambier, m. spindler, m. webb, p. zaczkowski, c. cebulski, p. christian, n. guldager, h. bartlett, a. greenblatt, l. dillard, d. sowards, and s. hamilton for efficient and safe flying. j. burch, t. hollis, j. lawler, n. pamperin, c. roberts, and g. stout provided expert assistance with captures. c. harwood, s. miller, and r. sarwas provided database and gis expertise. t. hollis and other biologists from the alaska department of fish and game conducted the radiotracking flights during calving. g. stout provided critical assistance with project management. all moose captures adhered to state of alaska animal care and use committee (acuc) guidelines (#07-11). we thank anonymous reviewers and n. decesare for constructive reviews that improved our manuscript. literature cited ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7: 39–47. barboza, p. s., and r. t. bowyer. 2000. sexual segregation in dimorphic deer: a new gastrocentric hypothesis. journal of mammalogy 81: 473–489. bowyer, r. t., b. m. pierce, l. k. duffy, and d. a. haggstrom. 2001. sexual segregation in moose: effects of habitat manipulation. alces 37: 109–122. boyce, m. s. 2006. scale for resource selection functions. diversity and distributions 12: 269–276. burnham, k. p., and d. r. anderson. 2002. model selection and multimodel inference: a practical informationtheoretic approach. springer-verlag, new york, new york, usa. ciarniello, l. m., m. s. boyce, d. r. seip, and d. c. heard. 2007. grizzly bear habitat selection is scale dependent. ecological applications 17: 1424–1440. collins, w. b., and d. j. helm. 1997. moose, alces alces, habitat relative to riparian succession in the boreal forest, susitna river, alaska. canadian fieldnaturalist 111: 567–74. crabtree, r., c. potter, r. mullen, j. sheldon, s. huang, j. harmsen, a. rodman, and c. jean. 2009. a modeling and spatio-temporal analysis framework for monitoring environmental change using npp as an ecosystem indicator. remote sensing of environment 113: 1486–1496. decesare, n. j., m. hebblewhite, f. k. a. schmiegelow, d. hervieux, g. mcdermid, l. neufeld, m. bradley, j. whittington, k. smith, l. e. morgantini, m. wheatley, and m. musiani. 2012. transcending scale-dependence in identifying habitat with resource selection functions. ecological applications 22: 1068–1083. dussault, c., j.-p. ouellet, r. courtois, j. huot, l. breton, h. jolicoeur, and d. kelt. 2005. linking moose habitat selection to limiting factors. ecography 28: 619–628. ferguson, s. h., m. k. taylor, and f. messier. 2000. influence of sea ice dynamics on habitat selection by polar bears. ecology 81: 761–772. fieberg, j., j. matthiopoulos, m. hebblewhite, m. s. boyce, and j. l. frair. 2010. correlation and studies of habitat selection: problem, red herring or opportunity? philosophical transactions of the royal society london b biological sciences 365: 2233–2244. 112 habitat selection in alaska – joly et al. alces vol. 52, 2016 forester, j. d., h. k. im, and p. j. rathouz. 2009. accounting for animal movement in estimation of resource selection functions: sampling and data analysis. ecology 90: 3554–3565. fuller, t. k., and l. b. keith. 1980. wolf population dynamics and prey relationships in northeastern alberta. journal of wildlife management 44: 583–602. gaillard, j.-m., m. hebblewhite, a. loison, m. fuller, r. powell, m. basille, and b. van moorter. 2010. habitat–performance relationships: finding the right metric at a given spatial scale. philosophical transactions of the royal society of london b: biological sciences 365: 2255–2265. guettabi, m., j. greenberg, j. little, and k. joly. 2016. evaluating potential economic effects of an industrial road on subsistence in north-central alaska. arctic 69: 305–317. herfindal, i., j.-p. tremblay, b. b. hansen, e. j. solberg, m. heim, and b.-e. sæther. 2009. scale dependency and functional response in moose habitat selection. ecography 32: 849–859. hinzman, l. d., et al. 2005. evidence and implications of recent climate change in northern alaska and other arctic regions. climatic change 72: 251–298. horne, j. s., e. o. garton, and j. l. rachlow. 2008. a synoptic model of animal space use: simultaneous estimation of home range, habitat selection, and inter/intra-specific relationships. ecological modeling 214: 338–348. hosmer, d. w., and s. lemeshow. 1989. applied logistic regression. wiley series in probability and statistics. john wiley & sons, inc., new york, new york, usa. ipcc (intergovernmental panel on climate change). 2007. climate change 2007: synthesis report. contribution of working groups i, ii and iii to the fourth assessment report of the intergovernmental panel on climate change. core writing team, r. k. pachauri, and a. reisinger, editors. ipcc, geneva, switzerland. johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61: 65–71. johnson, d. s., d. l. thomas, j. m. ver hoef, and a. christ. 2008. a general framework for the analysis of animal resource selection from telemetry data. biometrics 64: 968–976. johnstone, j. f., f. s. chapin, iii, t. n. hollingsworth, m. c. mack, v. romanovsky, and m. turetsky. 2010. fire, climate change, and forest resilience in interior alaska. canadian journal of forest research 40: 1302–1312. joly, k., t. craig, m. s. sorum, j. s. mcmillan, and m. a. spindler. 2015a. moose (alces alces) movement patterns in the upper koyukuk river drainage, northcentral alaska. alces 51: 87–96. ———, ———, ———, ———, ———. 2015b. variation in fine-scale movements of moose (alces alces) in the upper koyukuk river drainage, northcentral alaska. alces 51: 97–105. ———, p. a. duffy, and t. s. rupp. 2012. simulating the effects of climate change on fire regimes in arctic biomes: implications for caribou and moose habitat. ecosphere 3: 1–18. ———, t. s. rupp, r. r. jandt, and f. s. chapin. 2009. fire in the range of the western arctic caribou herd. alaska park science 8: 68–73. kasischke, e. s., and m. r. turetsky. 2006. recent changes in the fire regime across the north american boreal region – spatial and temporal patterns of burning across canada and alaska. geophysical research letters 33: 1–5. keating, k. a., p. j. p. gogan, j. m. vore, and l. r. irby. 2007. a simple solar radiation index for wildlife habitat studies. alces vol. 52, 2016 joly et al. – habitat selection in alaska 113 journal of wildlife management 71: 1344–1348. kunkel, k., and d. h. pletscher. 1999. species-specific population dynamics of cervids in a multipredator ecosystem. journal of wildlife management 63: 1082–1093. lake, b. c., m. r. bertram, n. guldager, j. r. caikoski, and r. o. stephenson. 2013. wolf kill rates across winter in a low‐density moose system in alaska. journal of wildlife management 77: 1512–1522. lawler, j. j., s. l. shafer, d. white, p. kareiva, e. maurer, a. r. blaustein, and p. j. bartlein. 2009. projected climate-induced faunal change in the western hemisphere. ecology 90: 588–597. ———, l. saperstein, t. craig, and g. stout. 2006. aerial moose survey in upper game management unit 24, alaska, fall 2004, including state land, and lands administered by the bureau of land management, gates of the arctic national park and preserve, and kanuti national wildlife refuge. national park service technical report nps/ar/nr/ tr-2006-55. fort collins, colorado, usa. maccracken, j. g., and l. a. viereck. 1990. browse regrowth and use by moose after fire in interior alaska. northwest science 64: 11–18. maier, j. a. k., j. m. ver hoef, a. d. mcguire, r. t. bowyer, l. saperstein, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics: effects of scale. canadian journal of forest research 35: 2233–2243. marcot, b. g., m. t. jorgenson, j. p. lawler, c. m. handel, and a. r. degange. 2015. projected changes in wildlife habitats in arctic natural areas of northwest alaska. climatic change 130: 145–154. mcphee, h. m., n. f. webb, and e. h. merrill. 2012. hierarchical predation: wolf (canis lupus) selection along hunt paths and at kill sites. canadian journal of zoology 90: 555–563. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monograph 122: 1–57. nielsen, s. e., m. s. boyce, g. b. stenhouse, and r. h. m. munro. 2002. modeling grizzly bear habitats in the yellowhead ecosystem of alberta: taking autocorrelation seriously. ursus 13: 45–56. oehlers, s. a., r. t. bowyer, f. huettmann, d. k. person, and w. b. kessler. 2011. sex and scale: implications for habitat selection by alaskan moose alces alces gigas. wildlife biology 17: 67–84. osko, t. j., m. n. hiltz, r. j. hudson, and s. m. wasel. 2004. moose habitat preferences in response to changing availability. journal of wildlife management 68: 576–584. peek, j. m. 1997. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., usa. peterson, r. o. 1995. the wolves of isle royale: a broken balance. willow creek press, minocqua, wisconsin, usa. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammalogy 88: 139–150. ———, and k. stuart-smith. 2006. winter habitat selection by female moose in western interior montane forests. canadian journal of zoology 84: 1823–1832. r core team. 2013. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. sawyer, h., r. m. nielson, f. lindzey, and l. l. mcdonald. 2006. winter habitat 114 habitat selection in alaska – joly et al. alces vol. 52, 2016 selection of mule deer before and during development of a natural gas field. journal of wildlife management 70: 396–403. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose, and forest succession: some management considerations on the kenai peninsula, alaska. alces 25: 1–10. slaght, j. c., j. s. horne, s. g. surmach, and r. j. gutiérrez. 2013. home range and resource selection by animals constrained by linear habitat features: an example of blakiston's fish owl. journal of applied ecology 50: 1350–1357. spencer, d. l., and j. b. hakala. 1964. moose and fire on the kenai. proceeding of the tall timbers fire ecology conference 3: 11–33. stephenson, t. r., v. van ballenberghe, j. m. peek, and j. g. maccracken. 2006. spatio-temporal constraints on moose habitat and carrying capacity in coastal alaska: vegetation succession and climate. rangeland ecology and management 59: 359–372. tape, k. d., m. sturm, and c. racine. 2006. the evidence for shrub expansion in northern alaska and the panarctic. global change biology 12: 686–702. telfer, e. s., and j. p. kelsall. 1979. studies of morphological parameters affecting ungulate locomotion in snow. canadian journal of zoology 57: 2153–2159. van ballenberghe, v. 1977. migratory behavior of moose in southcentral alaska. transactions of the 13th international congress of game biologists 13: 103–109. weixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces 34: 213–238. wheatley, m., and c. johnson. 2009. factors limiting our understanding of ecological scale. ecological complexity 6: 150–159. wiens, j. a. 1989. spatial scaling in ecology. functional ecology 3: 385–397. wilson, r. r., d. d. gustine, and k. joly. 2014a. evaluating potential effects of an industrial road on winter habitat of caribou in north-central alaska. arctic 67: 472–482. ———, j. s. horne, k. d. rode, e. v. regehr, and g. m. durner. 2014b. identifying polar bear resource selection patterns to inform offshore development in a dynamic and changing arctic. ecosphere 5 (10): 136. alces vol. 52, 2016 joly et al. – habitat selection in alaska 115 the effects of sex, terrain, wildfire, winter severity, and maternal status on habitat selection by moose in north-entral alaska methods study area moose capture, gps data, maternal status, winter severity study design environmental variables model selection results seasonal selection patterns by moose winter severity maternal status discussion acknowledgements literature cited alces vol. 45, 2009 sine et al. estimating tick abundance on moose 143 assessment of a line transect field method to determine winter tick abundance on moose meghan sine1, karen morris2, and david knupp3 1p.o. box 773 unity, maine 04988; 25510 bennoch road, lagrange, maine 04453; 3unity college, 90 quaker hill road, unity, maine 04988 abstract: high infestations of winter ticks (dermacentor albipictus) can exact high physiological costs on moose and are associated with high rates of juvenile mortality. quantifying tick abundance on moose may help managers calculate overall mortality rates for moose and make harvest recommendations. we compared winter tick counts along hair transects on samples of moose hides to tick counts obtained from chemical digestion of those same samples. winter tick counts from the two methods were strongly correlated (p <0.001, r2 = 0.88, n = 31). we field-tested the hair transect count method to determine its practicality at moose check stations. tick counts on 4 body areas per moose (n = 60) generally took ≤10 minutes and were rapid, non-destructive, inexpensive, and easily employed. this method has potential to serve as an effective method to index winter tick loads on moose. alces vol. 45: 143-146 (2009) key words: alces alces, dermacentor albipictus, index, monitoring, moose, winter ticks. winter ticks (dermacentor albipictus) are ectoparasites that grow from larval stage into engorged adults while feeding on a single ungulate host (addison and mclaughlin 1988). they were recognized as an important ectoparasite of moose as early as 1909 (seton 1909), and samuel (2004) has provided a summary of north american moose mortality related to winter ticks. moose are particularly suitable as hosts for winter ticks because of their ineffective grooming behavior and long hair (welch et al. 1991). winter tick loads on 183 moose in alberta averaged 30,683 ticks per moose, ranging from 2,774-149,916 (samuel et al. 2000). musante et al. (2007) estimated that blood loss during the 8-week engorgement period ranged from 64-112% of the normal blood volume of calf moose, and that this blood loss represented 50-100% of their daily protein requirement. calf survival and recruitment rates may be reduced through the combined effects of protein loss, thermal energy loss associated with alopecia, and energy loss associated with increased grooming (mooring and samuel 1999). given the potential negative impact that winter ticks may have on moose survival and recruitment, biologists have attempted to monitor winter tick infestations. however most field survey methods, such as late winter aerial surveys to assess hair loss, are both time consuming and costly. welch and samuel (1989) developed a laboratory technique to estimate winter tick numbers by digesting moose hide in potassium hydroxide and counting the undigested ticks; however, this method is restricted to the laboratory and is also time consuming and costly. our objective was to develop an efficient method to estimate winter tick abundance on moose by using easily accessed hunter-harvested moose. methods hides from hunter-killed moose were collected from an on-site meat processor at a hunter check station in greenville, maine, 2-3 october 2005. we collected hides from moose harvested within 1-23 h. pieces of hide measuring about 40 x 20 cm were cut from either the right shoulder or right rump; estimating tick abundance on moose – sine et al. alces vol. 45, 2009 144 we selected rump and shoulder areas based on a diagram of tick abundance presented by samuel (2004). after the moose was skinned, hide samples were placed in sealed plastic bags in <30 min, put on ice, and frozen (-17 c°) within 6 h. subsequently, we cut 10 cm x 10 cm samples from each 40 cm x 20 cm piece of hide in january 2006, and refroze the 10 cm x 10 cm samples for up to 2 weeks before counting ticks. each sample was systematically divided into 9 hair transects, 1 cm apart. four of the 9 were sampled randomly; the hair was parted and ticks that were visible along the lines were counted using an illuminated 10x magnifier. the average width in which ticks were visible along a single transect was about 0.5 cm. tick life-stages were not recorded. after counting ticks in the hair transects, each sample was immediately placed into a 1,000 ml beaker and digested following the procedures of welch and samuel (1989). ticks found loose in a sample bag were included in the digestion process; no sample had more than 4 loose ticks. the remaining solution was filtered through a 180 µm sieve and placed under an illuminated lens (10x magnification) to count the undigested tick exoskeletons. linear regression analysis was used to determine whether the numbers of ticks counted in the hair transects were correlated with the total number of tick exoskeletons counted after digestion. tick numbers on calves vs. adults were compared, and we analyzed tick numbers relative to median time since death. the generalizability theory, a method of partitioning measurement error, was borrowed from the field of educational measurement (brennan 2001) to determine the appropriate number of transects to count on a given square of hide. results hide samples were collected from 27 adult males, 2 adult females, and 2 calves (4-5 months). because of the small number of females and calves, we did not test for statistical difference in tick counts for age or sex of moose. the number of ticks estimated from the hair transect method and the number of ticks enumerated from our total counts of digested samples were positively correlated (p <0.001; r² = 0.88) (fig. 1). the proportion of ticks counted in the 4 transects averaged 42% of the total ticks per digested sample. in addition to determining whether the hair transect method could be used to predict the number of ticks determined by total count, we wanted to verify that 4 transects were sufficient in predicting tick numbers. using 4 lines yielded a dependability index of 0.95 (scale from 0-1), indicating that adding more lines would not greatly improve the reliability of the process. the difference in transect counts vs. total counts from the rump and shoulder area were compared to see if location may influence results. initially, the ticks and a dummy variable indicating body region were included in the regression equation. when including both predictors, the body region coefficient was not different from zero (t = 1.023, p = 0.317). therefore, the most statistically parsimonious and practical predictive equation included only the counted ticks as a predictor of the total number of ticks on the 10 cm2 piece of hide. calves averaged twice as many ticks per 100 cm2 than adults (261 vs. 113). the average tick count from moose dead <6.3 h (i.e., y = 3.6511x 3.3319 r² = 0.8791 p < 0.001 0 100 200 300 400 500 600 0 20 40 60 80 100 120 140 160 transect tick counts d ig es te d ti ck c ou nt s fig. 1. the relationship between the number of winter ticks counted on 4 transects within a 10 x 10 cm piece of moose hide to the number of exoskeletons of winter ticks counted in the related 100 cm² of digested moose hide. alces vol. 45, 2009 sine et al. estimating tick abundance on moose 145 median time of death) was higher than that from moose dead >6.3 h (~ 50%; t = 2.25, 22 df, p = 0.017). time spent counting ticks on each set of 4, 10 cm transect lines varied, but did not exceed 5 min. time required for chemical digestion and counting ticks was approximately 5 h per run of 4 samples. discussion the strong relationship between transect counts and total counts of winter ticks on a 10 cm2 plot indicated that the transect method may be useful for estimating winter tick numbers, and indexing trends in winter tick abundance on moose. tick counts on the rump and shoulder areas of moose were not different, and the best predictor of tick density on moose was the total number of ticks counted on all transects. counting ticks on hair transects is well suited to field applications because it requires little equipment and training, and does not require the moose to be skinned. during maine’s 2006 moose hunting season we field-tested the practicality of our method on >60 hunterharvested moose. we found that thin pointed objects such as knitting needles, pencils, and rat-tailed combs were excellent tools for parting moose hair; when marked with a 10 cm designation, the hair transect could be created and measured simultaneously. counting ticks on 4 transects on each of the rumps, posterior ribs, shoulder, and neck areas required about 5 min with a counter and a recorder; time was doubled if the same person did both tasks. we also found that ticks were plainly visible under bright light without magnification for people with good vision. welch and samuel (1989) found that approximately 15% of the hide had to be counted to estimate the total number of winter ticks on a moose. measuring transects at this level of intensity would be impractical in most field situations. however, fewer transects would be necessary to describe broad infestation categories (e.g., benign vs. pathogenic, or low, medium, and high) for monitoring or indexing annual winter tick abundance. calves are of primary concern when monitoring winter ticks because they are most susceptible to tick-induced mortality (addison et al. 1994). welch and samuel (1989) found that calves (<12 months) had higher tick densities than adults, and our limited sample (2 calves) supports this for maine moose as well. however, obtaining an adequate, annual sample size of calves could be problematic in maine because most hunters prefer to shoot larger moose. unless calves became more available at check stations, we suggest that adult moose are more practical for developing a tick index. tick emigration from dead moose was anticipated when collecting samples, and our data indicate that this was a legitimate concern. tick counts were reduced >50%, on average, after the median time since death (6.3 h). we recommend measuring only moose dead <6-8 h, or at least compensating for time of death. we believe that this transect method has good potential for objectively and rapidly assessing tick numbers on moose in the field, and will be most useful when used as an index. standardizing time after death and transect location and length will be required for meaningful comparisons between years and areas. acknowledgements we wish to thank unity college for providing support in the form of materials, equipment, and use of facilities. walter jakubas provided valuable insight in the composition process. rick gray and morgan reed volunteered field assistance in data collection and tawnya knupp contributed her assistance in statistical analysis. the help from the department of inland fisheries and wildlife biologists who tested the transect counts of ticks at various hunter registration stations was invaluable. windham butcher shop provided access to hides for sample collection. estimating tick abundance on moose – sine et al. alces vol. 45, 2009 146 references addison, e. m., r. f. mclaughlin, and j. d. broadfoot. 1994. growth of moose calves infested and uninfested with winter ticks. canadian journal of zoology 72: 1469-1476. _____, and _____. 1988. growth and development of winter ticks, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670-678. brennan, r. l. 2001. generalizability theory. springer-verlag, new york, new york, usa. mooring, m. s., and w. m. samuel. 1999. premature loss of hair in free-ranging moose (alces alces) infested with winter ticks (dermacentor albipictus) is correlated with grooming rate. canadian journal of zoology 77: 148-156. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces: 101-110. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1, federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m., m. s. mooring, and o. i. aalangdong. 2000. adaptations of winter ticks (dermacentor albipictus) to invade moose and moose to evade ticks. alces 36: 183-195. seton, e. t. 1909. lives of game animals. volume iii: hoofed animals. doubleday, doran and company, inc., garden city, new york, usa. welch, d. a., and w. m. samuel. 1989. evaluation of random sampling for estimating density of winter ticks on moose. international journal of parasitology 19: 691-693. _____, _____, and c. j. wilke. 1991. suitability of moose, elk, and white-tailed deer as hosts for winter ticks (dermacentor albipictus). canadian journal zoology 69: 2300-2304. persistent organic pollutants in the livers of moose harvested in the southern northwest territories, canada nicholas c. larter1, derek muir2, xiaowa wang2, danny g. allaire1, allicia kelly3, and karl cox3 1department of environment & natural resources, government of the northwest territories, po box 240, fort simpson, northwest territories, canada x0e 0n0; 2aquatic contaminants research division, environment and climate change canada, burlington, ontario, canada l7s 1a1; 3department of environment & natural resources, government of the northwest territories, po box 900, fort smith, northwest territories, canada x0e 0p0. abstract: moose (alces alces) are an important traditional and spiritual resource for residents of the southern northwest territories and local residents are concerned about contaminants that may be present in the country foods they consume. as part of a larger program looking at contaminants in moose organs, we collected liver samples from moose harvested in two separate but adjoining regions within the mackenzie river drainage area, the dehcho and south slave. we analyzed liver samples for a wide range of persistent organic pollutants (pops) including polychlorinated biphenyls (pcbs), ddt related compounds, toxaphene, brominated diphenyl ethers (pbdes) and perfluorinated alkyl substances (pfass). overall concentrations of major groups of pops (total (σ) pcbs, σpbdes, σpfass were consistently low (generally < 2 ng/g wet weight) in all samples and comparable to the limited data available from moose in scandinavia. pfass were the most prominent group with geometric means (range) of 1.3 (0.81–2.5) ng/g ww in the dehcho and 0.93 (0.63–1.2) ng/g ww in the south slave region. decabromodiphenyl ether (bde-209) was the most prominent pbde congener, similar to that found in other arctic/subarctic terrestrial herbivores. in general, bde-209 and pfass, which are particle-borne and relatively non-volatile, were the predominant organic contaminants. alces vol. 53: 65–83 (2017) key words: dehcho region, pops, polychlorinated biphenyls, perfluorooctane sulfonate, persistent organic pollutants, liver, moose, south slave region, northwest territories introduction moose (alces alces) are an important source of traditional food and of cultural significance for canada’s northern first nation communities, and is a frequently consumed food in the southwestern northwest territories (nt) (kuhnlein et al. 1995, receveur et al. 1997, berti et al. 1998). the long range transport and deposition of contaminants which often bio-magnify as they move through the food chain are of concern, especially as related to human exposure from consumed country foods (van oostdam et al. 2005, donaldson et al. 2010). recent studies have documented baseline levels of various heavy metals in the organs of moose from northern canada and alaska (o’hara et al. 2001, gamberg et al. 2005a, gamberg et al. 2005b, arnold et al. 2006, landers et al. 2008, larter et al. 2016); however, information on levels of persistent organic pollutants (pops) in the tissues of moose is scarce. analyses conducted in the early 1970s indicated that moose in idaho had very low corresponding author: nicholas c. larter, department of environment & natural resources, government of the northwest territories, po box 240, fort simpson nt x0e 0n0, canada, nic_larter@gov.nt.ca 65 mailto:nic_larter@gov.nt.ca levels of organochlorine pesticides (ocps) (benson et al. 1973). liver and muscle of moose from denali national park in central alaska were analysed for a suite of pops as part of the western airborne contaminants assessment project (wacap) (landers et al. 2008). low and variable con‐ centrations of pcbs and hexachlorobenzene (hcb), and ocps were detected in 3 moose liver and muscle samples; ddt related compounds (p,p′ddd + p,p′ddt, ∼34–340 ng/g lipid weight (lw)) and hcb predominated (∼0.1–0.72 ng/g lw). moose muscle, fat, and liver from communities in the mackenzie valley of nt that had been cooked/ baked had total (σ) pcb concentrations ranging from 3 to 23 ng/g (berti et al. 1998), but raw muscle and liver were not analysed. recent studies in scandinavia report low con‐ centrations of a wide range of pops in moose muscle and liver including pcbs, ocps, polybrominated diphenyl ethers (pbdes), and polychlorinated dibenzo-p-dioxins/dibenzofurans (pcdd/fs) (danielsson et al. 2008, mariussen et al. 2008, suutari et al. 2009, holma-suutari et al. 2016). a recent assessment of data for pops in canadian arctic food webs concluded that pbdes, particularly decabromodiphenyl ether (bde-209) and perfluorinated alkyl substances (pfass), were the predominant halogenated contaminants in caribou (rangifer tarandus) with concentrations typically higher than pcbs or ocps (muir et al. 2013); no results for moose were included. as part of a study assessing baseline concentrations of various contaminants in the organs of moose harvested for consumption by local residents, we analyzed livers to investigate the baseline levels of a wide range of pops and emerging contaminants of concern including pcbs, ocps, pfass, pbdes, and non-pbde brominated flame retardants (bfrs). knowledge of baseline levels of these contaminants in moose is important because comparable data are limited and moose consumption by local residents may increase in future due to the declining availability of caribou as an alternate country food resource. methods supporting information regarding more specific description of analytical methods, quality assurance, and raw data is pro‐ vided in “supporting information” at http:// alcesjournal.org/index.php/alces. reference to “supporting information” follows through‐ out, and nomenclature in tables beginning in “s” refers to tables in “supporting information.” study area the dehcho (ca. 154,000 km2) and south slave (214,000 km2) are administrative regions of the southern nt located substantially in the northern boreal forest where moose, boreal woodland caribou (r. t. caribou), and wood bison (bison bison athabascae) are the dominant ungulates. the samples for this study were collected by local harvesters from jean marie river, hay river, fort smith and fort resolution (fig. 1). moose samples first nation harvesters were requested to provide biological samples and general information from harvested moose. for the purpose of this study, we requested a minimum 5 cm x 5 cm piece of liver, an incisor tooth, and the following information: name of hunter, date and location of harvest, sex, estimated age (calf, yearling, adult), general body condition (excellent, good, fair, poor), and whether pregnant (yes, no) (table s1). liver samples (n = 7) from moose harvested in 2006 in the dehcho and in 2010 in the south slave (n = 7) were analyzed for 202 individual organohalogen compounds. a first incisor from each moose was forwarded to matson’s laboratory (manhattan, montana, usa) for aging by counting cementum 66 pops in moose livers – larter et al. alces vol. 53, 2017 http://alcesjournal.org/index.php/alces http://alcesjournal.org/index.php/alces annuli from the root of the first incisor; 1 june was used as the birthdate (matson 1981). liver tissue analysis liver samples were analyzed for pcbs, organochlorine pesticides (ocps), and other chlorinated organics (ocos) following us epa method 1699 (us epa 2007) by als global laboratories (burlington, ontario, canada), except for 4 samples from the dehcho which were analysed by environment canada (national laboratory for environmental testing [nelt]); both labs are accredited by the canadian association for laboratory accreditation and iso 17025 certified. the nlet used previously established methods (hoekstra et al. 2002, muir et al. 2006), and 3 samples from the dehcho were analyzed by both labs (see quality assurance section). for both methods, sample preparation was done in a clean room laboratory (positively pressurized with carbon and high-efficiency particulate arresting filters) at the canada centre for inland waters (cciw, burlington, ontario, canada). sample preparation, extraction, and cleanup/isolation of the ocps, ocos, and bfrs is described in detail in supporting information. clean extracts were concentrated to a final volume of 40 to 100 l in isooctane prior to analysis by gas chromatography (gc) using either elec‐ tron capture detection (gc-ecd), gc highresolution mass spectrometry (gc-hrms), or gc-low resolution ms (gc-lrms). a list of individual pcb/oco/ocp analytes is provided in supporting information (table s2). the gc-ecd analysis was conducted on 7 dehcho samples. final extracts of all samples from south slave and 3 of 7 from dehcho were analyzed by gc-hrms for 31 ocp related compounds (table s2) using gc-hrms at ≥10,000 resolution, and for 87 individual + co-eluting pcb congeners using gc-lrms. analyses of all pbdes and other brominated flame retardants (bfrs) toxaphene-related compounds including fig. 1. locations of moose collected from the eastern dehcho and south slave regions of the northwest territories showing roads and communities. moose were harvested in 2016 in dehcho (▲) and in 2010 in south slave (■). alces vol. 53, 2017 larter et al. – pops in moose livers 67 22 polychlorinated bornane congeners as well as αand β-endosulfan and endosulfan sulfate, were measured by gcelectron capture-negative ion mode (ecni) low resolution mass spectrometry. pfass were measured in liver samples (0.25–0.30 g) as described by lescord et al. (2015). further details of analytical methods for all pops are provided in supporting information. quality assurance and data analysis gc-ms and gc-ecd analysis of the 3 dehcho samples analysed by both methods indicated that gc-ms yielded 28–75% higher values for pcbs, hcb, and chlordane-related compounds (σchl) and 49% lower values for σhch (table s3); the discrepancies may reflect the low sample concentrations. recovery studies for ocps, pcbs, and bfrs spiked into moose tissue prior to extraction were very good and provided in tables s4 and s5. low levels of pcbs and pbdes were present in lab blanks and therefore all results were blank subtracted. method detection limits (mdls) for pcbs, ocp/ocos, and pbde/bfrs were calculated for all analytes based on results from 6 laboratory blanks that were analyzed in the same laboratory at approximately the same time, where mdl = 3x standard deviation of the blanks. for analytes with nondetectable blank values, the instrument detection limit (idl) based on a signal to noise ratio of approximately 10:1 was used for statistical calculations. further details on quality assurance are provided in supporting information. preliminary statistical analyses indicated that results for most individual analytes and total (σ) groups were not normally distributed based on the shapiro-wilk statistic <0.05 and coefficients of skewness and kurtosis > 2. log transformed data generally were normally distributed. correlations and comparison of means using the student’s t-test were conducted with log transformed wet weight data using systat version 13 (systat software inc., san jose, california, usa). results for males and females were pooled because preliminary analysis showed no significant differences in mean concentrations by sex. also, mean concentrations of all major analytes were compared between adult moose (n = 8) and calves (≤ 1 yr; n = 6) and no significant differences were found; therefore, only correlations with age were examined. significance were set at p ≤ 0.05 for all tests. results and discussion sample and data characteristics the 14 animals sampled were harvested over a wide area and based upon harvester reports, 13 of 14 were considered as excellent or good condition; one calf was rated as fair condition (table s1). all moose from the dehcho (n = 7) were harvested near the community of jean marie river and likely came from one localized population. south slave moose were harvested both east and west of the slave river and were possibly from two localized populations (fig. 1). concentrations of major groups of pops in moose liver are summarized in table 1, and results for 202 individual analytes are provided in tables s6, s7, and s8. a primary indication that low concentrations were common overall is that only 95 of the 202 individual target compounds were detectable (table s6). of these, 73 (dehcho) and 67 (south slave) were > mdl which is the 99% level of confidence that a given analyte is present (gomez-taylor et al. 2003). we report all results in order not to censor the data, but for those analytes 2.5 cm), whole leaves, and very small pieces of chewed food. the moisture content of the various intestinal tract samples fluctuated from 86 to 93% and was dependent on the type of diet consumed. the nitrogen intake from both dietary protein and nonprotein sources (including amino acid concentrations) varies significantly in moose by season. in spring, the branches of woody plants have the highest level of crude protein, which decreases during the summer due to the accumulation of structural compounds. in autumn, the concentration of crude protein and amino acids in deciduous twigs are 2 – 3 times lower than in spring samples (badlo and simakov 1990). concentrations of nitrogen and amino acids in dietary and intestinal digesta dry matter in autumn show that most amino acids in intestinal digesta are higher than in the diet (table 1). such autumn intestinal digesta levels demonstrate the transformation of dietary protein by microbes, although not as elevated as would be expected with summer diets. thus, protein and amino acid content in intestinal digesta of moose is dependent on the nutritional status of the animal, with more than likely a great part of the dietary protein being transformed by microbes. rumen and reticulum ingesta are transported through the reticulo–omasal orifice and strained (as if through a “screen”), allowing the passage of small particles of table 1. concentration of amino acids (g/kg dry matter) in the diet and digesta of moose forestomach. forestomach dietary amino acids ration rumen omasum abomasum aspartic acid 6.20 14.50 9.90 10.63 treonine 2.51 5.68 3.96 4.66 serine 2.91 5.19 4.20 4.86 glutamic acid 6.07 13.78 10.11 11.47 proline 2.33 4.57 3.13 4.36 cystine – 0.14 0.02 0.05 glycine 3.17 5.43 4.13 5.19 alanine 3.22 6.89 4.90 5.86 valine 3.01 6.22 4.34 5.23 methionine 0.06 0.57 0.20 0.41 isoluecine 0.05 0.57 0.20 0.41 isoleucine 2.48 5.81 4.48 4.49 leucine 4.66 8.42 6.17 7.43 tyrosine 1.72 3.45 2.07 2.70 phenylalanine 2.53 5.27 3.48 3.26 lysine 3.09 7.08 4.36 6.29 histidine 1.15 1.45 1.13 1.66 arginine 2.26 3.49 2.92 3.92 total 47.36 97.94 69.50 82.47 crude protein 75.00 113.12 82.50 111.87 alces suppl. 2, 2002 simakov – gastrointestinal tract of moose 121 food along with the soluble nutrients in ruminal fluids. larger particles of food that pass through the reticulo–omasal orifice are retained between the omasum’s folds, with fluid components being transported comparatively rapidly to the abomasum. omasum–retained materials have a higher percentage of dry matter (24.7%) but a lower concentration of nitrogen than ruminal fluids. the fluid component of samples from the abomasum is greater than in the samples from rumen and reticulum, due to “abomasum juice” secretion. the concentration of nitrogen compounds closely resembles the concentrations of amino acids in the rumen digesta. however, nonprotein nitrogen levels were higher, probably as a result of the partial hydrolysis of some proteins, and also urea that has passed into the rumen directly through the rumen wall from the circulating blood. the digesta flowing into the small intestine is characterized by its high proportion of fluid containing small particles of food and soluble nutrients. secretions from the main digestive glands (pancreatic juice and bile, which are continuously secreted) cause the digesta sample to have a higher concentration of total dry matter nitrogen (6.5%). nonprotein nitrogen is increased (60%) as a result of the proteolytic activity of intestinal juices and digestive tract enzymes. the concentration of amino acids is 9–10 times higher in the digesta dry matter of the small intestine than in the diet sample (table 2). glutamic acid, proline, and glycine are markedly higher in the small intestine table 2. concentration of amino acids (g/kg dry matter) in digesta from various segments of moose intestine. small large amino acids 4 m 7 m 13 m 20 m ceacum colon feces aspartic acid 26.94 22.81 12.20 11.41 6.78 7.65 6.21 treonine 26.63 17.10 9.85 4.99 2.69 3.36 2.90 serine 12.61 10.39 3.89 3.96 2.80 3.40 3.04 glutamic acid 66.24 46.12 30.40 18.62 6.30 7.47 6.01 proline 23.16 17.08 11.49 7.78 2.35 2.70 2.20 cystine 0.94 1.27 0.72 0.25 – – – glycine 37.24 26.86 21.10 7.85 2.97 3.39 2.95 alanine 28.77 18.16 13.10 8.09 3.52 3.71 3.61 valine 26.26 19.84 14.00 9.01 2.57 3.75 2.63 methionine 6.97 2.04 0.54 0.30 0.14 0.36 0.37 isoleucine 19.35 18.06 12.54 6.36 2.02 3.01 2.10 leucine 31.56 27.36 19.53 10.55 3.82 5.03 3.63 tyrosine 15.04 12.34 6.44 3.55 0.97 0.88 1.78 phenylalanine 24.10 15.64 8.28 5.23 2.40 2.61 2.61 lysine 21.56 12.07 10.49 6.77 2.15 3.39 2.80 histidine 8.41 5.08 2.76 1.71 0.66 0.95 0.84 arginine 19.46 13.68 11.65 3.69 1.48 2.25 1.60 total 392.24 285.90 188.98 82.20 43.62 53.91 45.28 crude protein 407.50 317.50 205.60 111.25 69.37 78.75 84.37 gastrointestinal tract of moose – simakov alces suppl. 2, 2002 122 samples. this may be due in part to the fact that, as has been observed by others for reindeer and sheep, glycine is secreted in the bile. in addition, there is evidence that glycine from plasma proteins may also be transferred through the intestinal wall of domestic ruminants. it is well documented that the small intestine is the main site of nutrient absorption to the blood stream. the concentrations of amino acids in small intestine digesta, determined at distances of 4, 7, 13, and 20 m from the pylorus, continuously decrease with passage along the small intestine, with the greatest absorption occurring in the first half of the small intestine. the digesta from the large intestine has dry matter contents and amino acid levels similar to that of consumed food. this is shown particularly in the concentrations of amino acids in the digesta dry matter of the caecum. in the hindgut, fermentation processes continue due to the enzymes of microorganisms inhabiting it. the fermentation intensity depends on the presence of appropriate substrata in the digesta flowing to the large intestine. in the digesta samples furthest along in the large intestines, degradation processes dominate. trends of increased concentrations in some amino acids indicate the presence of microbial protein synthesis in the colon. thus, the release of amino acids from dietary protein by digestion in moose is enriched as a result of the activity of ruminal microorganisms and protozoa. simultaneously, the proportion of nonprotein nitrogen is likewise higher. the concentration of amino acids is 9–10 times higher in dry matter of small intestine digesta than in the daily dietary ration. this elevation may indicate an additional impact on dietary protein from the protein–enzymes of digestive juices, which after a loss of activity are denatured and degraded to amino acids. the continuous decrease in the concentration of amino acids in digesta as it flows to the distal end of the small intestine indicates an absorption process. no significant changes in the concentration of amino acids in hindgut digesta were observed when compared to amino acid values of digesta from the distal end of the small intestine. we observed no other changes in concentration of amino acids in dry matter of hindgut digesta. references aliev, a. a., f. m. muchadov, x. x. rachmanov, and d. j. durdaev. 1981. seasonal dynamics of metabolism in karakul sheep. pages 86–107 in voprosy fyziologii i bioximii pitania ovec. kolos, russia. (in russian). badlo, l. p., and a. f. simakov. 1990. the processes of digestion in the forestomach and metabolism of nitrogen in moose. series preprint, “scientific report”. komi science center, ural division, russian academy of science, vol. 234. (in russian). cederlund, b. m. 1987. parturition and early development of moose (alces alces l.) calves. swedish wildlife research supplement 1: 399–422. stringham, s. f. 1974. mother–infant r e l a t i o n s i n m o o s e . n a t u r a l i s t e canadien 101:325–369. temporal effects of mechanical treatment on winter moose browse in south-central alaska sharon smythe1, dana sanchez1, and ricardo mata-gonzalez2 1fisheries and wildlife department, oregon state university, 104 nash hall, corvallis, oregon 97331, usa; 2animal and rangeland sciences, oregon state university, 120 withycombe, corvallis, oregon 97331, usa abstract: sites containing winter browse species utilized by moose on the copper river delta of south-central alaska were mechanically treated (hydraulic-axed) to counteract possible earthquakerelated increases in less-preferred forage species, and to measure treatment effects on biomass, height, nutritional quality (crude protein, lignin, and tannin), utilization, and snow burial on preferred (willow [salix spp.]) and less-preferred forage species (sweetgale [myrica gale], cottonwood [populus trichocarpa], and alder [alnus viridis sinuata]) within 3 winter scenarios (mild, moderate, and severe). sites were treated in 4 winters (1990–1992, 2008, 2010, and 2012) within 5 stand types in 20 sites varying from 0.9–63.4 ha. we found few significant differences in biomass, height, nutritional quality, utilization, and snow burial relative to controls. however, our ability to detect differences may have been limited by sample size (n = 1–9), as visual comparison suggests hydraulic-axing may be an effective method for increasing willow biomass while reducing alder biomass without influencing nutritional quality. however, because treated willows were shorter than untreated willows, treatment may result in less preferred forage for moose in severe winters with deep snow. our results have implications for habitat management of moose but further research is needed to determine incremental and longterm effects of treatment on willow growth and productivity. alces vol. 51: 135–147 (2015) key words: alaska, alces alces gigas, alnus viridis sinuata, copper river delta, forage biomass, hydraulic axing, myrica gale, nutrition, populous trichocarpa, salix spp. since many deer species in north america rely on early-successional forage, habitat management efforts commonly delay forest succession through mechanical treatment via shearing, crushing, or axing of overstory vegetation (scotter 1980, hundertmark et al. 1990, renecker and schwartz 1997, thompson and stewart 1997, suring and sterne 1998). mechanical treatment (hydraulicaxing) was applied on a limited scale to increase availability of preferred winter forage for an alaskan moose (alces alces gigas) population on the copper river delta (crd, stephenson et al. 1998), the location of this study (fig. 1). moose were introduced to the crd from 1949–1958 to establish a harvestable population, having likely been excluded by topography (maccracken et al. 1997). with a potential range encompassing >54,000 ha, the more managed and hunted western subpopulation has since grown to >600 animals (c. westing, alaska department of fish and game, unpublished data). however, intense winter winds through the copper river canyon, variable snow depths, and snow drifting can restrict winter range access to 4,800–12,900 ha (maccracken et al. 1997, stephenson et al. 2006). this seasonal effect constrains accessible browse and has sharon smythe, 104 nash hall. oregon state university, corvallis, oregon 97331, usa, sharonsmythe77@gmail.com 135 mailto:sharonsmythe77@gmail.com historically been thought to limit adult moose survival (regelin et al. 1985, schwartz et al. 1988, maccracken et al. 1997). furthermore, a 9.2 magnitude earthquake in 1964 uplifted the area by 1.0–4.0 m (grantz et al. 1964, ferrians 1966, plafker 1969, stover and coffman 1993), initiating changes in hydrology, soil salinity, and vegetation, including an acceleration of succession in some stands to stages with increased production of less-preferred browse (thilenius 1990, 2008). managers are concerned that the combined effects of winter range restrictions and earthquake-initiated vegetation changes might limit the performance or persistence of this locally important population (maccracken et al. 1997, stephenson et al. 2006). as a result, the usda forest service cordova ranger station initiated experimental treatments of moose habitat with hydraulic-axing machines (hereafter hydroaxing) which use rotary axes to cut down and splinter trees or shrubs up to 15 cm in diameter (stephenson et al. 1998). initial treatment plots were cut in 1990–1992, followed by additional plots in 2008, 2010, and 2012 (m. burcham, usda forest service cordova ranger district, personal communication, stephenson et al. 1998). because wintering crd moose depend on 5 willow species (feltleaf willow, barclays willow, undergreen willow, hookers willow, and sitka willow [s. alexensis, s. barclayi, fig. 1. sites mechanically-treated (hydraulic-axed) in 1990–1992, 2008, 2010, and 2012 on the west copper river delta of south-central alaska to improve the availability of willow forage for wintering moose. 136 mechanical treatment of moose browse – smythe et al. alces vol. 51, 2015 s. commutata, s. hookeri, s. sitchensis, respectively]), and only occasionally on black cottonwood (populus trichocarpa), sweetgale (myrica gale), and sitka alder (alnus viridis sinuata) (maccracken et al. 1997), treatments have focused on increasing the willow component of stands. in the kenai national forest willows re-sprouted following mechanical treatment whereas mature red alder (a. rubra) experienced high mortality (oldemeyer and regelin 1980, harrington 1984). thus, most treatments on the crd were sited on alder-dominated stands with remnant willow components, though spruce-cottonwood-, sweetgale-, and willow-dominated stands have also been treated (table 1). stephenson et al. (1998) evaluated the success of the initial (1990–1992) treatments 1–3 years post-treatment, and found that alder mass generally declined and sitka willow mass increased in treated sites. however, responses in biomass and utilization by other browse species varied by stand or were statistically precluded by sample size (stephenson et al. 1998). in addition, mean height of browse in treated stands was often less than in controls, and snow-buried browse varied by location, treatment, and stand type. it was hypothesized that sitka willow at full height (5 m) in alderand willow-dominated stands would be especially important in winters with deep snow and heavy drifting. therefore, it is possible that extensive treatment might increase the prevalence of shorter willows, coincidentally limiting browse available to moose in severe winters. however, hydro-axing effects in this system have not been studied beyond the first 3 years post-treatment. our objectives were to 1) evaluate species-specific and time-since-treatment responses of available biomass, height, nutritional quality, and moose utilization of winter browse species to hydro-axing 1, 3, 5, and 23 years post-treatment, and 2) estimate how biomass availability within treated sites varies with snow depth (winter severity). our results will assist managers in assessing the relative benefits of hydro-axing to maintain willow availability for moose in a dynamic ecosystem. table 1. characteristics of mechanically treated (hydraulic-axed) sites sampled (2012–2013) for moose browse species on the western region of the copper river delta, alaska, including site age (years since treatment), control stand type, soil type, area (ha), and sampling replicates. soil types include ast = alluvium and stream terrace deposits, opn = glacial outwash plains, nonforested, and gm = undifferentiated glacial moraines (davidson and harnish 1978). age (yr) winter treated control stand types soil type replicates (n) size (ha) 1 2012–2013 spruce-cottonwood ast 1 57.9 alder ast 2 23.9, 63.4 3 2010–2011 alder opn 1 3.4 sweetgale ast 3 8.0, 3.4, 5.7 5 2008–2009 spruce-cottonwood gm 2 10.7, 7.6 willow ast 2 11.8, 10.5 22–23 1990–1991 & 1991–1992 spruce-hemlock opn 2 0.9, 1.5 alder ast/opn 2 3.0, 2.2 alder-willow ast 2 0.9, 4.9 willow ast 1 1.5 sweetgale opn 2 2.6, 0.8 alces vol. 51, 2015 smythe et al. – mechanical treatment of moose browse 137 methods study area the crd lies within the chugach national forest and is bordered by 3 glaciers, the chugach mountain range, and the gulf of alaska (fig. 1). as the largest continuous wetland in the pacific northwest, it extends 120 km along the coast and supports abundant early-successional browse in a moist, relatively mild climate, lengthy growing season, and continuous channel changes by glacial streams and the copper river (christensen 2000, kesti et al. 2007, thilenius 2008). using a map derived from satellite pour l’observation de la terre (spot version 5 [spot5], 2011, red castle resources, inc.), we identified 7 stand types that produce winter moose forage: sprucehemlock, spruce-cottonwood, cottonwood, alder, alder-willow, willow, and sweetgale (viereck 1992). spruce-hemlock, sprucecottonwood, alder, and sweetgale can all form late-successional stands depending on hydrology, but alder-willow, willow, and sweetgale stands are generally considered early-successional (boggs 2000). drainage and desalination resulting from the 1964 earthquake increased the distribution of spruce-hemlock and alder stands, while accelerating succession or increasing the composition of willow, alder, sitka spruce (picea sitchensis), and western hemlock (tsuga heterophylla) within some stands (boggs 2000, stephenson et al. 2006, thilenius 2008). total winter snow depths range from 83.3–548.6 cm (1994– 2013; acrc 2014), and the area also receives substantial rainfall (annual mean of 236 cm), frequently interspersed within periods of snowfall (kesti et al. 2007). this phenomenon varies with winter severity, which can significantly affect snow accumulation, drifting, and compaction. thus, efforts to understand the complex interactions among snow depth, moose behavior, and browse availability are complicated and challenging. treatments and data collection prior to initial treatments, managers subjectively rated the suitability of potential treatment sites as high, medium, or low using factors of willow composition, encroachment by other woody species, and the level of understory organic matter (m. burcham, usda forest service cordova ranger district, unpublished data); only highly suitable sites were treated. due to the logistical difficulty of moving heavy equipment through wetlands, treatment occurred during winters with sufficiently frozen ground, and sites were partially determined by road access. managers refined their site selection techniques after the 1990–1992 treatments, selecting stands with the greatest potential for increased willow production. in total, the forest service treated approximately 300 ha from 1990–2012. treatments were applied to 32 sites in 5 stand types varying from 0.9–63.4 ha in the east-central, midcentral, and north-central regions of the west delta (table 1; fig. 1). all sites were unfenced and open (available) to moose. we sampled sites in august–september 2012–2013 and april–may 2013 to capture pre-winter available biomass and overwinter utilization and nutrition, respectively. because of logistical difficulties and differences in moose browsing pressure among sites, we selected 20 comparable sites treated in the east-central and mid-central region of the delta (table 1; fig. 1). we randomized sampling plots in treated sites and untreated adjacent controls, categorizing each site by the current control stand type. our study plots consisted of 3 random-start belt transects (1 × 10 m) separated by 5 m and running north, north, and east, respectively. we estimated the forage biomass available to moose (total biomass of twigs with diameters ≤8.3 mm; g/stem) with basal diameter-mass regression equations (table 2; maccracken and van ballenberghe 1993, stephenson et al.1998). at every 0.5 m along 138 mechanical treatment of moose browse – smythe et al. alces vol. 51, 2015 table 2. regression equations used to estimate species-specific available biomass (g/stem) and biomass consumed (g/twig) by moose wintering on the copper river delta, alaska, usa. time since treatment browse species 1 yeara 3 yearsa 5 yearsb 22–23 yearsb untreatedc consumptionc cottonwood = exp(−4.22) (bd2.85) = 0.64 (bd) = 0.15 (bd1.97) g– = 2.37 (bd) = 0.04 (bd2.6) alder = exp(−3.89) (bd2.77) = exp(−2.45) (bd1.8) = 0.03 (bd2.58) = 4.12 (bd) = 2.33 (bd) d= 0.03 + 0.06 (bd2.5) or = 0.34 (bd4) sitka willow = exp(−3.16) (bd2.52) = exp(−0.93) (bd1.46) = 0.13 (bd2.02) = 0.21 (bd1.8) = 11.07 (lnbd) = 0.03 + 0.06 (bd2.5) barclay willow e,f= exp(−3.50) (bd2.72) f= 0.98 (bd) = 1.74 (bd) = 2.56 (bd) e= 0.14 (bd1.93) = 0.05 + 0.03 (bd2.7) hooker’s willow e,f= exp(−3.50) (bd2.72) f= 0.98 (bd) = 0.11 (bd2.09) = 1.43 (bd) e= 0.18 (bd1.80) = 0.05 + 0.03 (bd2.7) undergreen willow = exp(−3.12) (bd2.48) = 0.56 (bd) = 1.51 (bd) = 1.40 (bd) = 0.55 (bd) = 0.05 +0.03 (bd2.7) sweetgale = 0.12 (bd) = 0.22 (bd) = 1.26 (bd) = 1.70 (bd) = exp(−3.33) (bd2.15) d= 0.05 + 0.03 (bd2.7) or = 0.12 (bd2) available biomass and biomass consumed equations are derived from measurements of basal diameters (bd, mm) and bite diameters (bd, mm), respectively. available biomass equations were developed in both mechanically-treated (hydraulic-axed) and untreated control sites. treated site equations are presented according to their site age (time since treatment, as of sampling in 2012 & 2013). adeveloped by stephenson et al. (1998). bdeveloped by smythe et al. (current). cdeveloped by maccracken and van ballenberghe (1993). drevised by stephenson et al. (1998). erevised by smythe et al. (current); negative added to coefficient. fseparate equations were not developed for hooker’s and barclay willows (smythe unpublished). gsample size was insufficient to develop a regression equation. a l c e s v o l . 5 1 , 2 0 1 5 s m y t h e e t a l . – m e c h a n ic a l t r e a t m e n t o f m o o s e b r o w s e 1 3 9 the belt transects, we measured basal diameters (mm; above the moss layer) of the 3 stems closest to the transect line. past research indicated that very large stem basal diameters (>60.0 mm) increased regression equation heteroskedasticity (maccracken and van ballenberghe 1993). thus, with such stems we instead measured a branch diameter and estimated how many equivalent branches were on the stem. within the belt transects, we calculated stem density (stems/belt; stems/ha), measured shrub height (m) on 3 replicates of every species, and estimated the available biomass (%) on each stem in 1-m vertical increments from 0–6 m to reflect the range of moose winter browsing heights, depending on crd snow pack conditions (t. joyce, usda forest service cordova ranger district, personal communication). we calculated the total available biomass (kg/ha; stem biomass × stem density) of every species in each plot. to calculate moose utilization, we measured every instance of browsing (bite diameters) on the closest 0.5 m stem. we estimated biomass consumed (g/twig) with bite diameter-mass regression equations (maccracken and van ballenberghe 1993) and summed the biomass removed per stem (g/stem). we collected nutritional samples of every browse species found at each plot, stored them fresh-frozen, removed all leaves, and sent them to the washington state university wildlife habitat and nutrition lab (pullman) for analysis. we developed 3 winter scenarios (mild, moderate, and severe) by summarizing data on mean winter snow depth (cm) from 1917–2012 collected by the alaska climate research center (acrc 2014) at cordova’s “mudhole smith” airport weather station. we could not accurately model the interaction between snow depth, snow compaction, and biomass available within the moose browsing window (0.5–3.0 m without snow). instead, we estimated the overall change in available biomass of browse in each plot according to our estimates of mean snow depth under each winter scenario, assuming that moose browsing height increased equally with snow depth. to evaluate differences between treated sites and their controls, we used t-tests to compare individual browse species and total plot available biomass, height, crude protein, lignin, tannin, and utilization, as well as the ratio of willow:alder biomass. individual willow species effects did not differ significantly and willow counts were pooled; feltleaf willow was not observed in any plot and was removed from analyses; the 1990– 1992 treatments were analyzed as a single treatment because they were not documented separately. furthermore, because we found few differences in time-since-treatment effects across stand types, we pooled all stand types for time-since-treatment analyses and used analysis of variance (anova) to compare treatments across time and winter scenarios. results treated willow, sweetgale, and total available biomass in 1990–1992 sites were higher than at control sites (p = 0.05, 0.003, and 0.001, respectively; table 3) no other differences were found between treated and control sites in available biomass of any browse species or treatment year (table 3). when weighted according to their untreated control (cut/control × 100), neither the relative total available biomass nor the relative total willow biomass differed significantly across times-since-treatment (fig. 2). treated alders in 2012 plots were shorter than in controls (p = 0.03). there was no significant effect on average willow height for timesince-treatment (fig. 2), but the average treated willow was shorter than the average control willow (p = 0.003). there were no significant differences in nutritional quality 140 mechanical treatment of moose browse – smythe et al. alces vol. 51, 2015 table 3. species-specific and total mean (±sd) available biomass (kg/ha), height (m), crude protein (%), lignin (%), tannin (mg/g), and use (%) of winter browse for moose in mechanically treated (cut, via hydraulic-ax) and untreated (control) sites on the copper river delta, alaska, usa. browse species age (yr) treatment biomass (kg/ha) height (m) crude protein (%) lignin (%) tannin (mg/g) use (%) black cottonwood 1 cut 10.89 (−) 1.0 (−) a– a– a– a– control 2343.00 (−) 6.0 (−) a– a– a– a– 3 cut b– b– b– b– b– b– control b– b– b– b– b– b– 5 cut 15.18 (11.05) 2.3 (1.2) 8.16 (2.56) 12.47 (0.65) 0.00 (0.00) 18.47 (0.32) control 573.53 (522.92) 4.0 (2.8) 5.45 (−) 13.28 (−) 0.00 (0.00) 0.00 (0.00) 23 cut b– b– b– b– b– – control 21.49 (51.19) 4.5 (2.1) 4.74 (−) 18.7 (−) 0.00 (−) 18.47 (−) sitka alder 1 cut 18.15 (13.43)* 1.0 (0.0)** a– a– a– a– control 605.42 (307.10)* 4.7 (1.2)** a– a– a– a– 3 cut 3.78 (4.99) 1.5 (0.7) c7.64 (−) c14.7 (−) c31.6 (−) 57.05 (15.20) control 138.59 (240.04) 6.0 (−) c7.64 (−) c14.7 (−) c31.6 (−) 0.40 (−) 5 cut b– b– b– b– b– b– control 125.48 (149.59) 4.0 (0.0) 7.64 (−) 14.7 (−) 31.6 (−) 0.00 (0.00) 23 cut 143.42 (430.25) 5.0 (−) 7.64 (−) 14.7 (−) 31.6 (−) 0.00 (−) control 257.49 (429.99) 4.5 (1.29) 7.64 (−) 14.7 (−) 31.6 (−) 7.17 (12.41) willow spp. 1 cut 78.13 (75.01) 1.3 (0.6) a– a– a– a– control 279.81 (253.04) 3.9 (1.9) a– a– a– a– 3 cut 386.19 (416.60) 1.4 (0.5) 7.04 (0.71) 11.87 (0.48)* 49.07 (17.64) 14.50 (10.87) control 405.41 (244.35) 2.3 (0.7) 6.91 (0.89) 15.47 (2.08)* 44.51 (19.55) 12.67 (5.03) 5 cut 550.79 (370.05) 1.6 (0.5) 7.91 (1.13) 15.53 (1.47) 32.28 (30.92) 3.25 (4.27) control 260.67 (112.35) 3.7 (1.5) 6.85 (1.18) 13.71 (1.29) 43.52 (3.03) 0.00 (0.00) 23 cut 1225.01 (614.71)** 2.0 (0.5) 7.06 (0.54) 15.60 (1.67) 48.26 (16.92) 16.17 (15.45) control 522.89 (408.90)** 2.5 (0.7) 7.07 (0.64) 15.61 (0.54) 37.48 (30.56) 11.83 (8.35) table 3 continued . . . . a l c e s v o l . 5 1 , 2 0 1 5 s m y t h e e t a l . – m e c h a n ic a l t r e a t m e n t o f m o o s e b r o w s e 1 4 1 table 3 continued browse species age (yr) treatment biomass (kg/ha) height (m) crude protein (%) lignin (%) tannin (mg/g) use (%) sweetgale 1 cut 21.06 (36.47) 1.0 (−) a– a– a– a– control 0.04 (0.08) 1.0 (−) a– a– a– a– 3 cut 76.63 (86.54) 1.0 (0.0) 8.50 (0.64) 22.42 (0.86) 44.53 (1.33) 53.00 (23.07) control 250.13 (221.49) 1.0 (0.0) 6.85 (−) 22.61 (−) 98.90 (−) 33.50 (21.92) 5 cut 403.28 (547.33) 1.0 (0.0) 6.75 (−) 17.00 (−) 41.00 (−) 10.80 (6.22) control b– b– b– b– b– b– 23 cut 503.02 (560.63)** 1.0 (0.0) 7.53 (0.46)* 21.73 (0.59)* 56.98 (27.95)* 7.75 (8.18) control 56.30 (103.77)** 1.0 (0.0) 6.91 (0.11)* 22.51 (0.17)* 95.37 (6.12)* 40.00 (40.15) total winter 1 cut 120.96 (80.93) 1.17 (0.29)* a– a– a– a– control 1666.26 (1292.39) 4.13 (1.63)* a– a– a– a– 3 cut 466.59 (476.44) 1.35 (0.47) 7.52 (0.27) 14.80 (1.27) 43.66 (6.70) 23.28 (12.00) control 794.12 (212.15) 2.38 (1.19) 6.94 (0.80) 16.40 (2.53) 51.06 (23.13) 16.00 (15.09) 5 cut 969.24 (852.05) 1.76 (0.77) 8.13 (1.50) 14.47 (0.23) 23.79 (23.96) 5.00 (5.83) control 959.69 (663.60) 3.44 (1.50) 6.81 (0.55) 14.13 (0.83) 30.43 (8.34) 0.00 (0.00) 23 cut 1871.44 (711.48)** 1.96 (0.56)* 7.19 (0.38) 16.77 (2.04) 46.76 (16.92) 10.67 (4.23) control 858.17 (454.79)** 2.86 (1.01)* 7.06 (0.60) 16.20 (1.30) 40.75 (29.50) 9.33 (7.58) treated sites were sampled 1, 3, 5, or 23 years post-treatment (age) in 2012 & 2013. are-growth of sites had not occurred by the time of spring nutritional sampling in one-year-old sites, but had occurred by the time of fall biomass sampling. bspecies did not occur in site. calder samples combined for nutritional analysis. *t-test, p = 0.06–0.10 between cut and control. **t-test, p ≤ 0.05 between cut and control. 1 4 2 m e c h a n ic a l t r e a t m e n t o f m o o s e b r o w s e – s m y t h e e t a l . a l c e s v o l . 5 1 , 2 0 1 5 fig. 2. total relative biomass (cut/uncut × 100, ±sd) of all winter browse species, relative biomass of willows (salix spp.), and mean heights of treated willows available to wintering moose within mechanically treated (via hydraulic-ax between 1990–2012) sites on the copper river delta, alaska as of 2012–2013 sampling. the dashed line represents the point at which treated sites have recovered pre-treatment biomass (100%) or the mean height of untreated willows (2.85 m). relative biomass across the 4 treatments was not significantly different (p = 0.15 and 0.13, respectively, 3 df). the average treated willow is significantly shorter than the average untreated willow (p = 0.003), but treated willow heights across treatment years are not significantly different (p = 0.13, 3 df). alces vol. 51, 2015 smythe et al. – mechanical treatment of moose browse 143 or utilization across any comparison. the ratio of willow:alder in treated sites was higher than in control sites at 23 years post-treatment (treated = 1163.37, control = 205.82, p = 0.004), though treated sites 1, 3, and 5 years post-treatment were not different (treated = 11.26, 323.63, 550.79, respectively; control = 0.77, 360.11, and 74.38, respectively). all treatment years were different (p = 0.02, 3 df). the 3 winter scenarios (mild, moderate, and severe) occurred 49, 29, and 11 times, respectively, with 6 winters uncategorized due to missing data. mean snow depth differed by scenario; 11.4 cm (±9.9–12.9), 25.8 cm (±23.3–28.3), and 63.9 cm (±47.4– 80.4), respectively. total available biomass across times-since-treatment varied significantly by scenario (p = 0.007–0.03, 4 df; fig. 3). total available biomass in treated 1990–1992 plots also differed across scenarios (p = 0.04, 3 df), declining 61% from mild to severe winters. further, available willow biomass across times-since-treatment varied significantly by scenario (p = 0.01– 0.05, 4 df; fig. 3). treated willow biomass in the 2008 plot differed across scenarios (p = 0.05, 3 df), declining 95% from mild to severe winters. discussion our data indicate that hydro-axing produces more total and willow biomass, with the effect increasing over time. given the observed variability, our a posteriori power analyses suggested sample sizes of 9–17 would be necessary to detect significance in comparisons of willow-only or all-species browse; however, treatment caused significant increase in the ratio of willow:alder over time. our results support those of harrington (1984), and further suggest that hydro-axing can be an effective method to increase willow biomass and counter ecologically-initiated (including earthquakeinfluenced hydrological or successional) increases in alder. hydro-axing did not influence the nutritional quality of the treated browse, as suggested by the lack of difference in crude protein, lignin, tannins, and utilization by moose. bowyer et al. (2001) reported similar findings for treated feltleaf willow in interior alaska, whereas rea and gillingham (2001) measured nutritional differences in scouler’s willow (salix scouleriana); however, both studies were short-term (≤3 years post-treatment). the high variability in height (m) of treated willows makes it difficult to determine if hydro-axing affects final regrowth height and the biomass available to moose across winter scenarios. because the average treated willow is shorter, yet more productive than the average untreated willow, hydro-axing may be causing a bushier growth form in treated willows, with more biomass concentrated in many smaller shoots on recovering stems. a changed architecture may explain the larger decrease in available biomass relative to controls in 1990–1992 treated sites as winter severity and snow depth increased. however, after 23 years of regrowth, mean available biomass in severe winters was similar to the mean available biomass provided by controls, suggesting that overall availability of treated biomass may compensate for losses due to snow burial. if so, hydro-axing would be an effective tool for increasing biomass available to moose in mild and moderate winters, while maintaining “normal” availability in severe winters, given sufficient time for regrowth. given the large gap between the 2008 and 1990–1992 treatments, we were unable to determine the regrowth asymptote or the minimal time required for winter browse species to recover sufficiently from treatment to provide equivalent (or potentially increased) biomass during severe winters. overall, our results indicate that mechanical treatment of moose winter browse 144 mechanical treatment of moose browse – smythe et al. alces vol. 51, 2015 species via hydro-axing has potential to reduce alder and increase willow biomass for wintering moose on the crd. however, extensive treatment could limit browse availability during extreme winter scenarios (deep snow) until regrowth occurs in a few decades. managers should be cautious in applying this technique across large areas concurrently. furthermore, monitoring at more frequent intervals should determine the temporal development and long-term effects of mechanical treatment on moose forage in the crd. this study provides a substantial summary of the effects of fig. 3. reductions in total and willow (salix spp.) biomass (kg/ha, ±ci) available to moose due to mean snow depths in 3 winter scenarios (mild, moderate, severe) in mechanically treated (via hydraulic-ax) sites cut over 4 years (1990–1992, 2008, 2010, and 2012) on the copper river delta, alaska. sites were sampled in 2012–2013. all biomass differences within winter scenarios are significant (p = 0.007–0.03 and 0.01–0.05, respectively, 4 df), and the 1990–1992 acrossscenario differences are significant (p = 0.04, 3 df). alces vol. 51, 2015 smythe et al. – mechanical treatment of moose browse 145 mechanical treatment on winter browse species, and should provide habitat managers of the crd and similar areas with a useful structure for current management decisions and further research. acknowledgements special thanks go to the managers at the cordova ranger district of the us forest service and the alaska department of fish and game, who provided critical financial, logistical, and personal support, including t. joyce, e. cooper, m. burcham, c. westing, d. crowley, and many seasonal employees. g. reeves, usfs-pnw research station, provided vital support in arranging and implementing this project. references alaska climate research center (acrc). 2014. applied climate information system daily data browser of university of alaska fairbanks. (accessed may 2014). boggs, k. 2000. classification of community types, successional sequences, and landscapes of the copper river delta, alaska. u.s. department of agriculture, forest service, pacific northwest research station, portland, oregon, usa. bowyer, t. r., b. m. pierce, l. k. duffy, and d. a. haggerstrom. 2001. sexual segregation in moose: effects of habitat manipulation. alces 37: 109–122. christensen, h. h. 2000. alaska’s copper river: humankind in a changing world. u.s. department of agriculture, forest service, pacific northwest research station, portland, oregon, usa. davidson, d., and c. harnish. 1978. soil and water resource inventory of the copper river delta. u.s. department of agriculture, forest service, chugach national forest, anchorage, alaska, usa. ferrians, o. j. jr. 1966. effects of the earthquake of march 27, 1964 in the copper river basin area, alaska. geological survey professional paper 543 e. u.s. government printing office, washington, d.c., usa. grantz,a.,g.plafker,andr.kachadoorian. 1964. alaska’s good friday earthquake, march 27, 1964: a preliminary geologic evaluation. u.s. department of the interior, geological survey, washington, d.c., usa. harrington, c. a. 1984. factors influencing initial sprouting of red alder. canadian journal of forest research 14: 357–361. hundertmark, k. j., w. l. eberhardt, and e. bail. 1990. winter habitat use by moose in southeastern alaska: implications for forest management. alces 26: 108–114. kesti, s., m. burcham, b. campbell, d. davidson, r. develice, c. huber, t. joyce, d. lang, b. macfarlane, d. sherman, and l. yarborough. 2007. west copper river delta landscape assessment. u.s. department of agriculture, forest service, chugach national forest, anchorage, alaska, usa. maccracken,j.g., andv. van ballenberghe. 1993. mass-diameter regressions for moose browse on the copper river delta, alaska. journal of range management 46: 302–308. ———, ———, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. wildlife monographs 136: 3–52. oldemeyer, j., and w. regelin. 1980. response of vegetation to tree crushing in alaska. alces 16: 429–443. plafker, g. 1969. tectonics of the march 27, 1964 alaska earthquake. u.s. geological survey professional paper 543-i. u.s. government printing office, washington, d.c., usa. rea, r. v., and m. p. gillingham. 2001. the impact of the timing of brush management on the nutritional value of woody browse for moose alces alces. journal of applied ecology 38: 710–719. 146 mechanical treatment of moose browse – smythe et al. alces vol. 51, 2015 http://climate.gi.alaska.edu/acis_data http://climate.gi.alaska.edu/acis_data regelin, w. l., c. c. schwartz, and a. w. franzmann. 1985. seasonal energy metabolism of adult moose. the journal of wildlife management 49: 388–393. renecker, l. a., and c. c. schwartz. 1997. food habits and feeding behavior. pages 403–440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. schwartz, c. c., m. e. hubbert, and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. the journal of wildlife management 52: 26–33. scotter, g. w. 1980. management of wild ungulate habitat in the western united states and canada: a review. journal of range management 33: 16–27. stephenson, t. r., v. van ballenberghe, and j. m. peek. 1998. response of moose forages to mechanical cutting on the copper river delta, alaska. alces 34: 479–494. ———, ———, ———, and j. g. maccracken. 2006. spatio-temporal constraints on moose habitat and carrying capacity in coastal alaska: vegetation succession and climate. rangeland ecology & management 59: 359–372. stover, c. w., j. l. coffman. 1993. seismicity of the united states, 1568-1989 (revised). u.s. geological survey professional paper 1527. suring, l., and c. sterne. 1998. winter habitat use by moose in south-central alaska. alces 34: 139–147. thilenius, j. f. 1990. woody plant succession on earthquake-uplifted coastal wetlands of the copper river delta, alaska. forest ecology and management 33: 439–462. ———. 2008. phytosociology and succession on earthquake-uplifted coastal wetlands, copper river delta, alaska. u.s. government printing office, washington, d.c., usa. thompson, i. d., and r. w. stewart. 1997. management of moose habitat. pages 377–402 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian intitution press, washington, d.c., usa. viereck, l. a. 1992. the alaska vegetation classification. u.s. department of agriculture, forest service, pacific northwest research station, portland, oregon, usa. alces vol. 51, 2015 smythe et al. – mechanical treatment of moose browse 147 temporal effects of mechanical treatment on winter moose browse in south-entral alaska methods study area treatments and data collection results discussion acknowledgements references alces_a_160168_o 141..152 hunter and tourist outfitter preferences for regulating moose hunting in northeastern ontario len m. hunt1 and peter davis2 1centre for northern forest ecosystem research, ontario ministry of natural resources and forestry, 103–421 james street south, thunder bay, ontario, canada p7c 2v6; 2northeast region, ontario ministry of natural resources and forestry, ontario government complex, 5520 hwy. #101e, p.o. bag 3020, south porcupine, ontario, canada p0n 1h0 abstract: it is important for managers to understand preferences of moose (alces alces) hunters and other stakeholders regarding options for harvest management. we determined harvest preferences of resident moose hunters and tourist outfitters in 2013 in northeastern ontario, canada through surveys that provided 5 management options. we tested 2 hypotheses: 1) that moose hunters will support options that are least impactful to them, and 2) that tourist outfitters will support restrictive calf harvest regulations more than resident hunters. we found little support for the first hypothesis as resident hunters and tourist outfitters ranked the status quo as the second least and least preferable option, respectively. resident hunters and tourist outfitters preferred shortened seasons for adult moose and less than a week long season for calves that would result in major departure from the status quo. we contend that this support arises because the hunters and outfitters are responding to the expectation of increased opportunities to hunt adult moose if they accept more restrictive regulations. consistent with the second hypothesis, tourist outfitters preferred options focused on restricting calf hunting opportunities more than resident hunters because clientele of tourist outfitters generally have low demand for calf hunts. resident hunters from areas where adult moose hunting opportunities were scarcer were surprisingly, less supportive than other hunters of change from an open to controlled hunt for calf moose. individuals in both groups that responded by mail, versus online, had stronger support for the status quo. alces vol. 52: 141–152 (2016) key words: alces alces, calf moose, hunting regulations, moose hunters, outfitters, preferences, selective harvest system, social surveys of the many stressors that impact moose (alces alces) populations, managers are often best capable of controlling mortality attributed to licensed hunting. this fact encourages managers to understand not only the biological consequences of management actions, but also the impact that such actions have on licensed hunters (ericsson 2003, hunt 2013). for example, while several different types of regulations can achieve the same level of harvest, some regulations are more preferable to hunters. thus, information about moose hunter preferences for regulations can help managers select more desirable (or less undesirable) paths for management while achieving biological objectives. harvest strategy has a prominent role in moose population management, with hunters managed through either direct or indirect controls on harvest. the selective harvest system (shs) that places hard limits on the maximum number of harvested bull and/or cow moose provides direct control. these limits often vary among management units and require hunters to obtain a special license or tag. regulations that provide indirect control over moose harvest typically alter harvest efficiency and/or hunter effort. for example, 141 the timing and season length can limit hunting effort and regulations governing permitted firearms, equipment, and party hunting can limit hunter effectiveness and harvest. here we report results from a 2013 survey of moose hunters and tourist outfitters in northeastern ontario, canada to identify support for different suites of direct and indirect controls on harvest. we focus on 12 wildlife management units (wmus) in this area that have been characterized by low calf recruitment since 2002 as documented through mid-winter aerial inventories, hunter post card surveys, and low calf hunter success rates (unpublished data, ontario ministry of natural resources [omnr]). nine wmus had a population density near or below the desired ecological minimum of 20 moose/100 km2 (omnr 2009a), and calves represented 43% of the resident moose harvest from 2008 to 2012 (unpublished data, omnr). it is suspected that the population decline was primarily caused by high calf harvest (sensu patterson et al. 2013) and cumulative effects of stressors such as morbidity and mortality from parasites including the winter tick (dermacentor albipictus) (rempel 2011). for these reasons, there was interest in understanding moose hunter and tourist outfitter support for different harvest strategy options. while many studies have evaluated trade-offs that moose hunters make when deciding where and how often to hunt (e.g., boxall and macnab 2000, bottan et al. 2003), trade-offs have not been identified when evaluating support for different options for managing moose hunting. we expected that providing supplemental information that makes such trade-offs evident to hunters would better allow hunters and outfitters to evaluate their preferences regarding regulations, contingent upon achieving expected future improvements in terms of licenses available to hunt adult moose. we developed 2 hypotheses to predict how moose hunters and tourist outfitters would prefer different sets of options. first, consistent with the view that individuals holding pro-hunting beliefs focus on wildlife for human use (fulton et al. 1996, teel and manfredo 2009), moose hunters will prefer options that are least impactful to their behaviours. thus, when evaluating more restrictive regulations than the status quo, moose hunters should be reluctant to support regulations other than the status quo (decker et al. 1996). for example, sigouin et al. (1999) noted that a greater share of quebec resident moose hunters rated the status quo (nonselective harvest hunt) as making an enjoyable hunting experience when compared to 4 different scenarios limiting or prohibiting cow moose harvest. status quo support, however, can be significantly lessened if hunters believe that a new management option will provide a positive future benefit to moose populations and, in turn, greater moose abundance for hunters (courtois and lamontange 1999). besides the status quo, hunters should prefer options requiring only slight modification to their behaviours such as changes to the length of the season. second, tourist outfitters will generally be more supportive of restrictive regulations on calf harvest than would resident moose hunters. given that most clients of tourist outfitters do not target calf moose when hunting 1 , outfitters should view regulations that restrict calf moose hunting more positively than resident moose hunters. we conducted a further set of analyses focused on the influence of context on the support for management options. in one instance, we compared support for management options among individuals hunting in 1from unpublished survey data collected between 2008 and 2012 by the omnrf within the wmus under study. of all moose harvested by tourist outfitters only 5.5% were calves, whereas for resident hunters, 43% were calves. 142 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 areas with large variability in the odds of obtaining a special license or tag required to hunt for adult moose. this comparison provides information about how the scarcity of tags can influence hunter support for different harvest strategy options. for example, as tags become scarcer, hunters should exhibit less support for the status quo than for options that increase the availability of tags. we also evaluated whether the response mode (i.e., online versus mail) influenced support for the options to help managers better understand and interpret public input. many wildlife management agencies are contemplating collecting survey data exclusively with online methods, yet collecting online data can be problematic (duda and nobile 2010). study area the study area included 12 wmus (28, 29, 31, 32, 35, 36, 37, 38, 39, 40, 41, and 42) in northeastern ontario, canada (fig. 1) that are part of cervid ecological zones (cez) c2 and d2 and represent a substantial portion of the core moose range in the province (see omnr 2009b for details on cezs). the 12 wmus were governed by the same set of moose hunting regulations in 2013 that included 1) a gun hunt conducted from the saturday nearest october 8 until november 15, and 2) a bow hunt conducted 3 weeks prior to the gun hunt. a 2013 resident license to hunt moose authorized the holder to hunt for calf moose in all wmus in the study area. a 2013 resident license to hunt moose and an adult validation tag (avt) fig. 1. wildlife management units (wmus) within the study area; study wmus were associated with cervid ecological zones (cez) c2 and d2 in the northeast region in northeastern ontario. alces vol. 52, 2016 hunt and davis – regulation preferences for moose hunting 143 authorized the holder to hunt for either a bull or cow moose, as specified on the avt in the wmu and under the conditions specified on the tag. party hunting for moose is legal as long as the applicable rules are adhered to under the fish and wildlife conservation act, 1997, ontario regulation 665/98, part iii, hunting in a party. applying to the resident moose draw for an avt requires a hunter to select a preferred (choice 1) wmu and season (i.e., bow or gun hunt). about 32% of all ontarians applying to this draw selected one of the 12 wmus as their choice 1 wmu. in 2012, 1,886 avts were available for the 31,449 individuals applying to the draw in these units (unpublished data, omnr). the ratios of applicants to avts were 15:1 and 37:1 in 2012 for cezs c2 and d2, respectively. from other research data about moose hunters in these wmus, it was estimated that hunters were predominantly 40-69 year-old males who had hunted for 20 years, and ~9 days in 2012. further, these hunters stated that the availability of an avt was the most important factor influencing where they hunted, and most hunted between october 6 and november 2 (hunt 2014). a total of 163 tourist outfitters catering to moose hunters were operating in 2012, and their allocation of avts was in addition to the 1,886 avts available for resident hunters. methods in 2013 surveys were developed to explore preferences of resident hunters and tourist outfitters for different harvest strategy options to hunt moose. resident moose hunters were defined as those individuals who applied as a choice 1 applicant in the 2012 resident draw for an avt in one of the 12 wmus in the study area (fig. 1). a total of 5,229 individuals were selected to receive the resident hunter survey. to permit analyses at the cez and wmu scales, respondents were stratified by their choice 1 wmu before the random selection (i.e., 1/6th of choice 1 applicants were randomly selected for each wmu). we mailed a survey to all 163 tourist outfitters. the survey process included up to 2 potential contacts for both groups. sampled individuals were contacted by mail in late march 2013 and were requested to complete the survey either online or by mail; no return postage was provided to encourage online submissions. unfortunately, some individuals experienced difficulty completing the survey online because they failed to use the correct web address and could not access the survey through a search engine. all mailed survey packages included a unique identification number for each potential respondent. individuals not completing the survey before 17 may received a second mail contact encouraging their submission by 7 june. we accepted completed surveys on-line or by mail until 12 july 2013. the survey contained a single question that asked individuals to rank 5 options by their preference for managing moose populations (table 1). the options included the status quo and 4 choices that altered the total length of the hunting season, a calf validation tag system, and a shortened calf hunting season. all options were plausible based on directives from the moose harvest management guidelines (omnr 2009c). the end early option reduced the gun season by 15 days, and the 14d gun option provided 14-day archery and gun seasons. the 6d calf option combined a 6-day calf hunt with a 42-43 day adult moose season. the calf tag option required hunters to obtain a calf validation tag through a draw system in a given wmu, with no change in the length of any season. supplemental information was provided in the survey that described the challenges of managing moose populations over the past decade and the possible outcomes from adopting each season option. these outcomes 144 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 were based on a moose population maintenance perspective for the wmus, and tradeoffs were communicated in terms of increased availability of avts for resident hunters and tourist outfitters given the anticipated reduction in calf and adult moose harvest for each option (table 2). the trade-offs were assessed by analyzing the temporal distribution of reported harvest by wmu and sector over the period 2000-2009. the temporal distribution of the calf harvest was 0.8% in september, 79.7% in october with the majority in the first 2 weeks of the gun season, and 19.5% in november. resident hunters accounted for 99.2% of calf harvest over this period. the supplemental information for the 5 options was aggregated for the 12 wmus and provided with the survey (table 2). to develop this information, several assumptions were invoked. first, the wmu-specific calf validation tag option had a planned cap of 25% of the total harvest. second, for the shortened season ending 31 october, we assumed that the past proportion of november-harvested calves would not be harvested in this shortened season. third, for the shortened season ending 31 october with the 6-day gun calf hunt (which would overlap the southern region gun moose hunt), we assumed that the proportion of november-harvested calves would be saved from harvest and that 2/3 of the current october calf harvest would be met in this 6-day gun calf hunt. fourth, for the 14-day archery and gun hunts, we assumed that the proportion of november-harvested calves would be saved from harvest and that 100% of the current october calf harvest would still occur. fifth, for the 3 options with shortened seasons, we assumed that the avt tag fill rates for adult moose would decline equal to the corresponding adult harvest that occurs in november, with the exception of wmu 28 where avt table 1. presented harvest strategy options for moose management in northeastern ontario, canada. label specific option status quo keep moose harvest management the same (manage by altering adult validation tags) calf tag same current seasons; calf harvest restricted through resident moose draw for wmu-specific calf validation tags end early reduce gun season by 15 days with last day october 31, same opening dates 6d calf 21-day archery season and 21-22 day gun season, same opening dates for adult moose; calf hunting allowed only for 6 days corresponding with the moose gun season south of frenchmattawa rivers 14d gun 14-day archery season and 14-day gun season table 2. supplemental information provided with the survey to identify potential increases in tag allocations for each harvest option. option total # of days (resident moose seasons) potential % increase to resident gun avts potential % increase to tourist outfitter moose tags status quo 57 to 63 d 0 0 calf tag 57 to 63 d 36 33 end early 42 to 47 d 21 17 6d calf 42 to 43 d; 6 d calf hunt 41 37 14d gun 28 d 21 17 alces vol. 52, 2016 hunt and davis – regulation preferences for moose hunting 145 tag fill rates would not change. finally, all other harvest planning parameters remained constant (e.g., harvest rates, bull:cow ratios, proportional sector allocations). individuals ranked all 5 options from 1 to 5 where 1 was the most preferred and 5 was the least preferred option. when an individual ranked all options but did not follow the instruction of using each rank only once, we allowed ties among options and recoded the original ranks such that ranks were compatible with surveys completed correctly. for example, if an individual gave 2 options a rank of 5, these 2 options were recoded with a rank of 4.5; this approach ensured that the sum of ranks equaled 15 in each survey. another tendency was that some supporters of the status quo option failed to rank any other option. to avoid underestimating support for status quo, we accepted these responses and ranked all other options of equal preference (i.e., 3.5). the rank data were analyzed with 2 approaches. first, the non-parametric friedman and wilcoxon-sign tests were conducted to assess whether differences in ranks (preference) existed among the options, and then to identify differences between options. a bonferroni correction for the number of pairwise corrections was made to the probabilities estimated from the wilcoxon-sign tests (i.e., a significant difference was based on p < 0.005 rather than p < 0.05 because of the 10 pairwise-combinations among the 5 options). second, for any single option and different groups of respondents, mean ranks between the groups were assessed with an independent samples t-test; where necessary, correction was made to the t-values for unequal variances between samples. while the rank data for the groups were discrete, it is common practice to analyze this type of survey data with parametric tests for assessing statistical inferences (vaske 2008). the hypotheses were tested by examining support among options and between respondents from resident hunters and tourist outfitters. hypothesis 1 focused on whether support for the status quo was greatest, and for the other options, whether support for end early was next most preferred. hypothesis 2 focused on the role of the calf tag and if it was less preferred by tourist outfitters than resident hunters owing to the lower harvest of calf moose by tourists. the effect of context was examined by 1) comparing responses of individuals completing the survey online versus by mail, and 2) for resident hunters, cez c2 and d2 as defined by the choice 1 wmu. all analyses were conducted with r (r core team 2015) with null hypothesis testing and significance at p < 0.05 except when adjusted for multiple comparisons. we assessed non-response bias by testing whether early and late responders to the survey differed in their support for options by assuming that late responders would be more like non-responders than would early responders (miller and smith 1983). late responders were defined as individuals responding 5 weeks after the initial distribution of the survey. results a total of 2,507 resident moose hunters and 108 tourist outfitters completed at least part of the survey. after accounting for undeliverable mail addresses, the response rate was 48.9% for resident hunters and 68.8% for tourist outfitters; late responders represented 27% and 29% of these groups, respectively. proportionally, online surveys represented 43% of resident hunter and 34% of tourist outfitter responses. differences (friedman test, p < 0.05) in the ranks of the 5 options were found in all wmus except #37 (table 3). rank preferences of the 5 options by resident moose hunters were different (χ2 = 375.3, 146 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 df = 4, p < 0.001). the 6d calf and end early options were ranked higher than the other options. the calf tag and status quo options were third and fourth ranked, and the 14d gun option was lowest ranked (table 4, fig. 2). the top 3 options (6d calf, end early, calf tag) garnered similar support when combining the top 2 ranks (most or second most preferred), although the calf tag option had a higher percentage indicating it was the least preferred option. similarly, despite ~20% ranking the status quo option as most preferred, a large percentage rated it least preferred. while the ranks also differed among the options for the tourist outfitters (χ2 = 78.9, df = 4, p < 0.001), fewer pairwise differences were identified (fig. 3). no significant differences were found among the comparisons of the 6d calf, end early, and calf tag options. the same negative response pattern by resident hunters for the calf tag option was also found for tourist outfitters. as with resident hunters, the status quo option was ranked the least preferred option by the majority (~60%) of tourist outfitters. certain significant differences existed between resident hunters and tourist outfitters in their mean ranks of the 5 options (table 5). tourist outfitters preferred the calf tag and 6d calf options more than resident table 3. mean ranks for 5 harvest strategy options by resident moose hunters’ choice 1 wildlife management unit (wmu), 2013. ranks range from 1 (most preferred) to 5 (least preferred). * signifies difference (friedman test, p < 0.05) among the ranks of options for a wmu. wmu status quo calf tag end early 6d calf 14d gun 28* 3.12 2.99 2.65 2.76 3.48 29* 3.48 3.00 2.67 2.69 3.17 31* 3.02 3.13 2.46 2.90 3.49 32* 3.50 2.57 2.89 2.48 3.56 35* 2.93 3.04 2.78 2.73 3.52 36* 2.98 3.26 2.08 3.17 3.52 37 3.17 3.24 2.77 2.67 3.15 38* 3.08 2.95 2.67 2.74 3.56 39* 2.84 3.20 2.71 2.73 3.53 40* 3.43 2.87 2.67 2.64 3.39 41* 3.20 3.31 2.58 2.78 3.14 42* 3.29 3.19 2.59 2.79 3.14 table 4. unadjusted significance probabilities from wilcoxon pairwise statistical tests of ranks for harvest strategy options by northeastern ontario resident licensed hunters and tourist outfitters in 2013 (c2, d2 – hunters with choice 1 wmu in cervid ecological zones c2 and d2, respectively). option 1 option 2 resident tourist resident c2 resident d2 end early 6d calf 0.021 0.115 0.097 0.033 end early calf tag <0.001**1 0.527 <0.001**1 <0.001**1 end early status quo <0.001**1 <0.001**1 <0.001**1 <0.001**1 end early 14d gun <0.001**1 <0.001**1 <0.001**1 <0.001**1 6d calf calf tag <0.001**1 0.019 <0.001**1 <0.001**1 6d calf status quo <0.001**1 <0.001**1 <0.001**1 <0.001**1 6d calf 14d gun <0.001**1 <0.001**1 <0.001**1 <0.001**1 calf tag status quo 0.006*1 <0.001**1 0.003**1 0.943 calf tag 14d gun <0.001**1 0.009*1 <0.001**1 0.924 status quo 14d gun <0.001**1 <0.001**2 0.004**1 0.896 1, 2 indicates the option with the higher mean rank * p < 0.10 (bonferonni-adjusted probability) ** p < 0.05 (bonferonni-adjusted probability) alces vol. 52, 2016 hunt and davis – regulation preferences for moose hunting 147 hunters, whereas resident hunters preferred the status quo option more than tourist outfitters, although it was the least preferred option of both groups. resident hunters in c2 and d2 ranked the options differently (χ2 = 340.2, df = 4, p < 0.001; χ2 = 45.0, df = 4, p < 0.001, respectively; table 5). those associated with c2 preferred the calf tag and were less supportive of the 14d gun option than those associated with d2 (table 5). no other differences between the groups were found in rankings of the status quo, end early, and 6d calf options. resident hunters responding by mail were more supportive than online respondents for the status quo, and less supportive of the end early and 6d calf options (table 6). similarly, the status quo option received higher support from tourist outfitters 0% 20% 40% 60% 80% 100% end early 6d calf calf tag status quo 14d gun least preferred 2nd least preferred 3rd preferred 2nd preferred most preferred fig. 2. rank preferences of resident hunters to 5 harvest management options in a moose harvest survey in northeastern ontario, 2013. 0% 20% 40% 60% 80% 100% end early 6d calf calf tag status quo 14d gun least preferred 2nd least preferred 3rd preferred 2nd preferred most preferred fig. 3. rank preferences of tourist outfitters to 5 harvest management options in a moose harvest survey in northeastern ontario, 2013. 148 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 responding by mail. no other significant differences were found for the other options. for all respondents, ~60% completed the questionnaire correctly (i.e., assigned each rank once and ranked all options). the remaining respondents either ranked each option but used the same rank value more than once (~27%), provided an incomplete set of ranks (~10%), or ranked the status quo option as most preferred without ranking the other options (~3%). by including these additional responses in the analysis, support increased for the status quo option for both resident hunters and tourist outfitters (t = �12.32, df = ~1771, p < 0.001 and t = �3.35, df = ~109, p = 0.002 for resident hunters and tourist outfitters, respectively). conversely, it led to reduced support for the 6d calf hunt option in both groups (t = 8.31, df = ~1951, p < 0.001 and t = 1.95, df = ~58, p = 0.056 for resident hunters and tourist outfitters, respectively), and less support for the end early option by resident hunters (t = 8.31, df = ~1951.4, p < 0.001). we report results from the full data set as we believe it best describes the relative support of the 5 options by both groups. the results were potentially compromised by non-response bias owing to the fact that many hunters and tourist outfitters did not complete the survey. the status quo table 5. results from pairwise comparisons of mean ranks between resident hunters and tourist outfitters and resident hunters applying to cervid ecological zones (cez) c2 and d2 in 2013 (note: ranks range from 1 to 5 with 1 being most preferred). option resident hunter tourist outfitter t-test (t, df, p) cez c2 cez d2 t-test (t, df, p) end early 2.65 2.61 (0.34, 108, 0.735) 2.66 2.58 (1.12, 369, 0.263) 6d calf 2.73 2.29 (2.73, 106, <0.001) 2.72 2.80 (�1.06, 371, 0.290) calf tag 3.05 2.73 (2.15, 105, 0.034) 3.03 3.21 (�2.08, 375, 0.039) status quo 3.20 4.01 (�5.90, 109, <0.001) 3.19 3.23 (�0.32, 374, 0.749) 14d gun 3.37 3.36 (0.04, 106, 0.971) 3.39 3.18 (2.80, 373, 0.005) table 6. results from pairwise comparisons of mean ranks between resident hunters and tourist outfitters in 2013 by mode of survey completion (note: ranks range from 1 to 5 with 1 being most preferred). option online mail t-value df p resident hunters end early 2.74 2.51 4.86 2005 <0.001 6d calf 2.81 2.63 3.54 1912 <0.001 calf tag 3.04 3.08 �0.63 1820 0.529 status quo 3.06 3.40 �5.17 1973 <0.001 14d gun 3.35 3.39 �0.69 1862 0.492 tourist outfitters end early 2.56 2.72 �0.74 66 0.459 6d calf 2.35 2.16 0.78 62 0.438 calf tag 2.89 2.41 1.52 60 0.134 status quo 3.77 4.50 �2.89 79 0.005 14d gun 3.43 3.22 0.87 63 0.389 alces vol. 52, 2016 hunt and davis – regulation preferences for moose hunting 149 option was more preferred by the late (mean rank = 3.71) than early (mean rank = 4.13) responding resident hunters (t = 3.85, df = ~1102, p < 0.001). no other pairwise difference among the options was found in either group. discussion the results provided little support for the first hypothesis that respondents should generally prefer the status quo and options that result in least impact to hunter behaviours. the weak support for the status quo option and strong support for the 6d calf hunt option suggested that hunters, and to a greater extent tourist outfitters, were generally willing to move to a system away from the status quo. the strong support for the 15-day reduction to the moose hunting season (end early) was consistent with this hypothesis as <20% of resident hunters in these wmus hunted after 1 november (hunt 2014). explaining why support for the status quo was less than expected is difficult. perhaps it was the expectation of future benefits in terms of increased avts by accepting other options (table 2). this would be consistent with choice model studies that illustrate how trade-offs influence hunter selection of hunting sites (e.g., boxall and macnab 2000, bottan et al. 2003). unexpectedly, however, the value of increased avts was not influenced by the scarcity of avts because applicants to cez c2, where tags were relatively more plentiful than in d2, were not more supportive of the status quo option than applicants to d2. this choice contrasts with previous information that indicated hunters placed greater value on tags in wmus where tag availability was scarce (hunt 2013). another possible explanation for the lack of support for the status quo option is that hunters were responding more in the interest of moose rather than themselves. individuals are characterized by value orientations ranging from mutualism to domination. mutualism value would focus more on respect for moose regardless of their value to people, whereas domination value would focus on the human benefit derived from moose (teel and manfredo 2009). pro-hunting beliefs are closely aligned with a domination orientation (fulton et al. 1996, teel and manfredo 2009). furthermore, satisfaction with wildlife management depends on hunting success (miller and graefe 2010) suggesting that management views of hunters are domination-oriented. finally, in a survey conducted simultaneously, strong support was found for the status quo option among moose hunters (hunt 2013). the primary difference in ranking options between these surveys was the provision of information (this study) relating to future avt benefits to hunters by adopting other options. it follows that the lack of support for the status quo option was related to future, expected benefits for hunting moose. many hunters, however, are concerned about moose regardless of hunting opportunities. for example, fulton and hundertmark (2004) found strong support for a selective harvest system among alaskan hunters; ~2/3 recognized benefits of the system for moose and <1/2 recognized benefits for hunters. our different conclusion is probably related to our maintaining, not increasing moose populations, suggesting that hunters might only respond to benefits to moose in terms of increased recruitment of calves. given this perspective, the hunter support and preference for harvest options were largely self-serving. the results generally supported the second hypothesis that tourist outfitters would prefer more restrictive options than resident hunters, especially with regard to hunting calves. tourist outfitters preferred the calf tag and 6d calf options more than resident hunters, whereas resident hunters were more supportive of the status quo option. given that most tourist outfitters cater to clients hunting adult moose, the benefits 150 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 of reduced calf harvest through shortened seasons and a tag draw system may have been more appealing to outfitters. this is consistent with the idea that hunters and tourist outfitters exhibit a domination value orientation that encourages support for options that increase hunting opportunity. context also affected support for the options as notable differences in ranks were found between online and mail respondents. mail respondents in both groups had higher support than online respondents for the status quo option, and online resident hunters ranked the end early and 6d calf options higher than mail respondents. that responses may differ between online and conventional mail surveys is certainly not novel (e.g., duda and nobile 2010), but our data indicate the potential for bias by using a single survey response mode. our non-response test revealed an important difference between early and late responding hunters in that late responding hunters were more supportive of the status quo option. consequently, our sample might underestimate support for this option, although this underestimation was qualitatively unimportant in affecting our conclusions. even if we adjusted for non-response bias in resident hunters, the status quo option remained the second least supported option. the potential underestimation of support for the status quo option strongly influenced our decision to use information from all surveys with ranked options, rather than only those surveys completed correctly. had we excluded the responses from the “incorrect surveys”, the status quo option would have been ranked last and, we believe not reflective of the moose hunter population. our study illustrates an effective approach to measure support and preferences of hunters and other stakeholders with respect to options for managing moose. importantly, the data were not supportive of one of our original hypotheses and point to the complexity of survey construction, interpretation, and potential bias. we hope that other researchers will build upon our survey methodology and embrace the importance of enabling respondents to consider trade-offs when assessing preference or support for management options (cornicelli et al. 2011). otherwise, support will probably be biased upwardly for status quo management programs despite options that provide positive future outcomes for moose and moose hunters. acknowldegements we thank l. dix-gibson for summarizing and providing data from the postcard moose hunter surveys. we thank p. hubert for reviewing an earlier version of the manuscript. we appreciate the support for the research by the resident moose hunters and tourist outfitters for taking the time to complete the questionnaires. finally, we are grateful to two anonymous reviewers and an associate editor for providing constructive criticism and suggestions to improve an earlier draft manuscript. references bottan, b., l. m. hunt, and w. haider. 2003. a choice modelling approach to moose management: a case study of thunder bay moose hunters. alces 39: 27–39. boxall, p. c., and b. macnab. 2000. exploring the preferences of wildlife recreationists for features of boreal forest management: a choice experiment approach. canadian journal of forest research 30: 1931–1941. courtois, r., and g. lamontagne. 1999. the protection of cows: its impact on moose hunting and moose populations. alces 35: 11–29. cornicelli, l., d. c. fulton, m. d. grund, and j. fieberg. 2011. hunter perceptions and acceptance of alternative deer management regulations. wildlife society bulletin 35: 323–329. alces vol. 52, 2016 hunt and davis – regulation preferences for moose hunting 151 decker, d. j., t. l. brown, and b. a. knuth. 1996. human dimensions research: its importance in natural resource management. pages 29–50 in a.w. ewert, editor. natural resource management: the human dimension. westview press, boulder, colorado, usa. duda, m. d., and j. l. nobile. 2010. the fallacy of online surveys: no data are better than bad data. human dimensions of wildlife. 15: 55–64. ericsson, g. 2003. of moose and man: the past, the present, and the future of human dimensions in moose research. alces 39: 11–26. fulton, d. c., and k. hundertmark. 2004. assessing the effects of a selective harvest system on moose hunters’ behaviors, beliefs, and satisfaction. human dimensions of wildlife 9: 1–16. ———, m. j. manfredo, and j. lipscomb. 1996. wildlife value orientations: a conceptual and measurement approach. human dimensions of wildlife. 1: 24–47. hunt, l. m. 2013. using human-dimensions research to reduce implementation uncertainty for wildlife management: a case of moose (alces alces) hunting in northern ontario, canada. wildlife research 40: 61. http://doi.org/10.1071/wr12185. ———. 2014. result from the 2012 ontario moose hunter questionnaire: comparisons among moose hunters grouped by hunting areas and residency. centre for northern forest ecosystem research technical report tr-013. queen’s printer for ontario, thunder bay, ontario, canada. miller, c. a., and a. r. graefe. 2010. effectofharvestsuccess onhunterattitudes toward white-tailed deer management in pennsylvania. human dimensions of wildlife 6: 189–203. miller, l. e., and k. l. smith. 1983. handling nonresponse issues. journal of extension 21: 45–50. ontario ministry of natural resources (omnr). 2009a. moose population objectives setting guidelines. queen’s printer for ontario, toronto, ontario, canada. ———. 2009b. cervid ecological framework. queen’s printer for ontario, toronto, ontario, canada. ———. 2009c. moose harvest management guidelines. queen’s printer for ontario, toronto, ontario, canada. patterson,b.r.,j.f.benson,k.r.middel, k. j. mills, a. silver, and m. e. obbard. 2013. moose calf mortality in central ontario, canada. wildlife society bulletin 77: 832–847. r core team. 2015. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. https://www.r-project. org/.r core (accessed may 2016). rempel, r. 2011. effects of climate change on moose populations: exploring the response horizon through biometric and systems models. ecological modelling 222: 3355–3365. sigouin, d., s. st-onge, r. courtois, and j.p. ouellet. 1999. change in hunting activity and hunters’ perceptions following the introduction of selective harvest in quebec. alces 35: 105–123. teel, t. l., and m. j. manfredo. 2009. understanding the diversity of public interests in wildlife conservation. conservation biology 24: 128–139. vaske, j. j. 2008. survey research and analysis: applications in parks, recreation, and human dimensions. venture, state college, pennsylvania, usa. 152 regulation preferences for moose hunting – hunt and davis alces vol. 52, 2016 http://doi.org/10.1071/wr12185 https://www.r-project.org/.r%20core https://www.r-project.org/.r%20core hunter and tourist outfitter preferences for regulating moose hunting in northeastern ontario study area methods results discussion acknowldegements references i in memoriam edmund (ed/eamon) s. telfer it is with a deep sense of sorrow and loss that we announce the passing of edmund (ed) stewart telfer on april 29, 2018 at the age of 87. ed was born on december 13, 1930 and raised in rural nova scotia, where he developed a love for the natural world. he acquired a b.sc. in forestry from the university of new brunswick in 1953. after working as a forester (timber cruiser) and land surveyor, ed returned to school, where he obtained a b.ed. specializing in teaching science from acadia university in 1962 and a m.sc. in wildlife biology, also from acadia university, in 1965. under the guidance of don dodds, ed’s thesis ‘studies of moose and white-tailed deer ecology in northern nova scotia’ set the tone for much of his career. ed was then hired by the canadian wildlife service, environment canada, first in new brunswick (4 years), then in edmonton alberta (27 years). ed was a tireless worker, applying his curiosity about the natural world to his job, first as a wildlife biologist and later as a research scientist. he led research projects that studied ungulates in particular moose and white-tailed deer and the impact of forest management on their habitat. this research included work on understanding the impact of timber harvesting on watersheds along the eastern slopes of the alberta rockies. he became a distinguished moose and habitat biologist and a loggerhead shrike expert. in the latter part of his career, he focused on the habitat requirements of forest birds. he advised on environmental impact statements for various development projects in the alberta rockies. ed was a long time member of the wildlife society and served on their publication awards committee. in 1994, the alberta chapter of the wildlife society awarded him the ‘william rowan distinguished service award’ for his outstanding contributions to the management and conservation of wildlife and their habitats. while a member of the canadian institute of forestry, he served as an associate editor for the forestry chronicle scientific journal. when he was a member of the international union of forest research organizations, he participated in many international conferences on fish and wildlife habitat and coordinated the group’s work in canada. he was a member of the north american moose conference and workshop, virtually from its inception, and was an associate editor with the research journal, alces. in 1984 ed received their ‘distinguished moose biologist’ award for “outstanding contribution to the field of moose management”. ed retired in 1996, and it is noteworthy that many of his accomplishments were after that. ed served as a scientist emeritus with the cws for several years, attending moose conferences, ii most recently with his beloved wife morag, such as in international falls, minnesota, jackson, wyoming and brandon, manitoba. ed loved to read and he had a tremendous memory for what he read. he also was a great storyteller and listener. he loved family and he loved visiting with people. ed’s character was exemplified by kindness, a compassionate heart, humble, meek, patient, self-controlled, thankful, and showing perfect courtesy towards all people. because of these characteristics, ed was well respected by his colleagues and associates within the scientific community, and, quite simply, by anyone who had the distinct pleasure of meeting and knowing him. there were many accolades from his moose colleagues. here are a few examples. ed “was much respected in the moose community for his knowledge and his willingness to share his thoughts with other colleagues over the years.... he made the effort to participate, as he had done since we first met in thunder bay at the 8th north american conference in 1972. over those 40+ years ed kept active and will be remembered by those who knew him. he left his mark” (tim timmermann). “ed telfer did some of the first and best work on moose, deer, and snow relationships, plus some of the shrub production methodology that i picked up on in days gone by.....he initiated a lot of work that i extended over the years, as did many others. so i consider ed to be among the first of the moose biologists who made initial contributions that have been built upon since. i’m very gratified to have known him, and his legacy in the moose world will quite obviously continue” (jim peek). ed was predeceased by his former wife, leona (gorman) telfer, who was the mother of their six children; and infant children, kenneth clifton and isabel anne. he is survived by his second wife, morag; his six children, dan (mary) telfer (okotoks, ab), angus (hazel harper) telfer (vancouver, bc), christine telfer (dennis livingstone) (riverview, nb), ellen (peter) paczkowski (kamloops, bc). jean (charles) ottosen (regina, sk), kathleen (dave) macnearney (montague, pei); by morag’s three children, scott (theresa) ozeroff, kimberly (mike) ozeroff, and christopher (sharla) ozeroff; and by his 15 grandchildren, 8 great-grandchildren, and 7 step-grandchildren. this gentle giant will be fondly missed. 1 in memoriam gerry lynch, 1941–2022 the moose management and biology world lost another pioneer on june 21, 2022 with the passing of gerry lynch at the age of 81. gerry died peacefully at his home near raleigh, north carolina with his family at bedside, having dealt valiantly with a serious heart condition for several years. for many years, gerry was a highly regarded regional wildlife biologist with alberta fish and wildlife in edson, alberta. eventually, he became the provincial moose manager based at the edmonton headquarters where his management skills flourished. as an “early” moose manager, gerry was progressive and innovative, designing management initiatives and developing long-term data sets that were uncommon at the time, yet are now the basis of sound moose management in alberta and elsewhere. his exemplary efforts and leadership were recognized by his receiving the distinguished moose biologist award (2000) from the international alces working group. the current cohort of moose biologists and managers in alberta continue to marvel about gerry’s dedication and productivity in the “early days” and the wisdom still evident in those ‘old programs’ such good data and ideas remain valid and ageless. gerry was born and raised near madison, wisconsin. after completing his bachelor of applied science degree in natural resources and conservation at university of wisconsin – stevens point (1963), gerry completed a ms in wildlife, fish and wildland science and management at the university of wisconsin – madison (1965–1967). his ms research involved aspects of skunk predation on waterfowl in manitoba, after which his studies, research experience, and motivation set the tone for a productive and highly respected career. gerry was the consummate professional – a hard-working, kind, and considerate man throughout his life. much of his work took place in the rugged and bog-filled boreal forest of the swan hills where he used innovative techniques to trap and track moose. he also built innovative allterrain vehicles to aid his efforts long before atvs became common in fieldwork. always willing to help others, he assisted dr. bill samuel in his pioneering work with winter ticks and moose. gerry loved outdoor recreation, was an avid hunter, and true friend of many. often described as “the best moose hunting partner ever”, he cherished the comradery of moose hunting with friends who considered him a great “mooser”. 2 beyond his professional dedication, gerry’s family and their spiritual lives were foremost in his daily thoughts and actions. he and janet were married 59 years and raised three children. unfortunately, yet again, the moose world loses another early leader in the short history of moose management. with deep admiration and respect for our friend, bill samuel and margo pybus alces16_137.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces15_245.pdf alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 alces vol. 15, 1979 201 44th north american moose conference and workshop idaho state university, pocatello, idaho 14-17 june 2009 could not consume all of the refreshments. wednesday started with a predator-prey roundtable discussion organized by matt kauffman, with participants terry bowyer, mark boyce, doug brimeyer, mike mitchell, mike wolf, and pete zager. the proposed delisting and management of gray wolves in the rocky mountains was a central theme of discussion, for which the audience provided important contributions. the roundtable was followed by the final three contributed talks, and then the capstone presentation by kris hundertmark (2007 recipient of the distinguished moose biologist award) who discussed genetics of moose worldwide, and why there is only one species of moose. this excellent presentation was followed by a tour of the exhibits and collections at the idaho museum of natural history and a discussion of ice age mammals lead by bill akersten. the banquet was held wednesday evening, with a further, and this time successful, attempt to reduce the available refreshments. lisa shipley and matt kaufman were given newcomer awards, and ken child received the senior’s travel award. the members of the organizing committee received the order of alces, and the contributions of the staff of the department of biological sciences at isu to the conference were praised. the highlight of the evening was ken child receiving the 2009 distinguished moose biologist award for his many contributions to the management of moose. these awards were followed by a lively auction by expert auctioneer, john nelson, where everyone did their best to keep the conference from going into the red. in total, there were 7 plenary presentations, 16 contributed presentations, 4 posters, a roundtable discussion, a home range workshop, a field trip, the establishment of a memorial award, and the capstone lecture. in all, there was a diversity of high-quality and productive activities for all to attend. we are grateful to our sponsors, including monsanto, the north american moose foundation, agrium, the us fish and wildlife service, southeast idaho national wildlife refuges, the idaho department of fish and game, stoller, inc., the 44th north american moose conference and workshop was held at idaho state university (isu) in pocatello, idaho, 14-17 june 2009. the theme for the conference was "population, behavioral, and landscape ecology of moose: implications for theory and management." the conference was hosted by the department of biological sciences at isu, with extensive help from the idaho department of fish and game and the us fish and wildlife service. there were 77 delegates registered for the meeting who came from throughout north america and sweden. the us was represented by delegates from 16 states, and canada by participants from 5 provinces; their broad experience resulted in vigorous discussions about moose research, management, and biology. all enjoyed beautiful weather, at least during some of the mornings, and the predicted plague of mosquitoes never emerged. following welcoming remarks by pamela crowell, the vice president for research at isu, the conference was initiated with a plenary session. topics of presentations were far ranging, and included applied and theoretical material of interest to moose biologists and managers. mark boyce, vince crichton, lisa shipley, mike wolfe, john fryxell, matt kauffman, and terry bowyer gave talks that generated considerable discussion for the remainder of the conference. the plenary session was followed by a poster session. contributed presentations filled the following monday, again ranging across a broad range of topics, including harvest, productivity of populations, parasites, and climate change. the presentations were followed by a business meeting where the albert w. franzmann and distinguished colleagues memorial award was established. on tuesday, delegates chose between a field trip to camas national wildlife refuge organized by mark gamblin and carl mitchell, and a home range workshop presented by john kie and art rodgers. both activities were great successes, and those on the field trip observed a shiras moose as well as adding numerous bird species to their life lists. the workshop and field trip were followed by a well-attended picnic, during which even moosers 202 and the department of biological sciences at idaho state university. we also thank those who provided items for the auction, including, but not limited to, us national park service, yellowstone center for resources, shoshone-bannock tribes, tribal enterprises, stackpole books, bruce becker, teton photo works, and midway usa. chair: r. terry bowyer host: department of biological sci ences, idaho state university location: 921 south 8th ave., stop 8007 pocatello, id 83209-8007 usa date: 14-17 june 2009 # of delegates/participants: 77 f:\alces\supp2\pagema~1\rus8s.pdf alces suppl. 2, 2002 bogomolova et al. study of moose behavior 37 the study of moose behavior on the kostroma moose farm ekaterina m. bogomolova1, yuriy a. kurochkin1, and alexander n. minaev2 1anokhin research institute of normal physiology, russian academy of medical science, moscow, russia; 2institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: we studied 35 adult farm moose-cows, about 50 wild bulls (in the breeding season), and more than 250 newborn and developing calves. we used the “los-2” radio-tracking system, the “los-3” radio-telemetry system, radio-communication between experimenters, photo, film, and video recording of behavior, and magnetic recordings of moose vocalizations. various aspects of moose behavior and corresponding changes of their heart rate and breathing rate were studied during 1977 to 1990 on the kostroma experimental moose farm. the results and advantages of farms for the study of animal behavior are listed and discussed. findings include work with specific time intervals, first behavioral reactions, bonding behaviors, and passive and active defensive mechanisms. the time interval from a calf’s birth to first standing varied from 12 to 58 minutes. the developmental sequence of the first functional behavior in a newborn calf is not preprogrammed genetically but is determined by the actual circumstances of the newborn. alces supplement 2: 37-40 (2002) key words: behavior, heart rate, moose, newborn, radio-tagging, telemetry experimental moose farms offer new opportunities for studying species-specific m o o s e b e h a v i o r ( b o g o m o l o v a a n d kurochkin 1987a). the animals on these farms do not sever connections with the natural environment because they inhabit the forest and do not essentially alter their behavior under the influence of the initial process of domestication. working with such tamed animals gives scientists the following advantages: 1. biologists can always find the radiotagged animal in the forest and investigate its behavior in its natural environment. tame moose get accustomed to human beings so much that they do not pay attention to their presence, even during such important events as parturition, suckling, protection of calves, and sexual and competitive interactions. 2. besides behavioral patterns, the corresponding emotional reactions of moose may be investigated by measuring heart rate and breathing dynamics because tame moose allow staff to equip them with detectors and radio-telemetry transmitters, which is convenient for investigators. 3. observing tame, radio-tagged moose allows the opportunity to simultaneously study the behavior of interacting wild animals such as wild moose-bulls during the rutting period or wild calves that grew up in the forest. methods postnatal development and physiological mechanisms of suckling, feeding, defense, breeding, and other activities, as well as mother-calf bond consolidation and subsequent separation were studied in 1977– 1 9 8 9 o n t h e k o s t r o m a m o o s e f a r m (bogomolova and kurochkin 1984, 1987b, 2002). heart rate and breathing dynamics study of moose behavior bogomolova et al. alces suppl. 2, 2002 38 were studied to estimate the emotional states of animals during various behavioral activities. we studied 35 adult farm moosecows, about 50 wild bulls (in the breeding season), and more than 250 newborn and d e v e l o p i n g c a l v e s . w e u s e d t h e radiotracking system “los-2,” radio-telemetry system “los-3” (minaev 2002), radiocommunication between experimenters, photo, film, and video recording of behavior, a n d m a g n e t i c r e c o r d i n g s o f m o o s e vocalizations. the gathered data were processed on a computer. results for the first time the entire parturition activity pattern of mixed-aged cows was investigated in detail (bogomolova and kurochkin 1984). we observed and studied parturition in more than 60 cows. in some cases, ecg and breathing patterns were recorded during the various stages of parturition. duration of parturition (time from the appearance of the amniotic sac to delivery) was extremely variable: from 4 to 136 min and on the average, 31–41 min. the time interval from a calf’s birth to its first standing varied from 12 to 58 minutes. there was no significant difference in the average duration of this interval between calves born single or the first of twins (29.2 min) and those born second of twins (25.2 min). the first standing up, the first achievement of a functional behavioral sequence, is a decisive moment in the integration of efferent and afferent activity of nervous, muscular, vestibular, visual, jointal, and other components in the whole, integrated system. once the calf stands up on 4 legs for the first time and keeps its feet, it is able to stand up the next time in only 1– 3 seconds. the average time interval from birth to the first milk sucking in calves born single or the first of twins was 65.5 ± 5.1 min; in those born second of twins, the interval was 69.8 ± 7.1 min. the difference is not significant. there was no correlation between the interval from birth to first standing and that from birth to the first milk sucking (r = 0.076, n = 32). the time from birth to the first sucking varied widely from 8 minutes to 3 hours. but, as soon as the calf found the mother’s nipple and obtained milk the first time, the calf found the nipple later practically at once. these results suggest that learning plays an important part in the development of such innate behavior (just as it does in standing up). the developmental sequence of the first functional behavior in a newborn calf is not preprogrammed genetically and is determined by the actual circumstances of the newborn. for example, as a rule, calves learn at first to stand up and walk on 4 feet, and only after that do they learn to find their mother’s nipples. but in 6 cases of 55, we observed that at first calves began to suck the milk and only after that did they learn to stand up. the analysis of a calf’s emotional behavior on the first and following milk sucking reveals that here, too, the decisive moment is the first achievement of a functional result. only after the first sucking do calves show marked, emotional reaction to obtaining milk. one of the most important behaviors of newborn calves is the “following response”, which we usually interpret as an inborn, genetically preprogrammed behavior that is essential for attachment to their mother. through this behavior, the calf learns step by step to recognize its mother and distinguish her from other objects. only afterward will the following response manifest itself in the closely formed “mother-infant” bond. during the first days of calves’ lives, our experiments show that newborn calves follow not only their mother but also any human being. moreover, calves can easily follow humans in the presence of their calm alces suppl. 2, 2002 bogomolova et al. study of moose behavior 39 mother. in our study we observed some cases in which newborn calves left their mothers for other moose-cows and cases in which the moose-cow accepted a calf of another cow and raised it. these data testify against widespread opinion about the existence of mutual mother-newborn imprinting in moose. we are convinced that the calf’s attachment to its mother is built essentially on the basis of its food behavior: when we succeeded in teaching wild calves captured at the age of 2–3 weeks to suck milk through a rubber nipple, later on they revealed the same attachment to human beings as the calves on the farm that were removed from their mother before first sucking and brought up by people. we found that the moose-cow and calf finally learned to recognize one another only after 7–8 days of permanent interactions at the place of parturition (bogomolova and kurochkin 1984, 2002). these features of moose behavior may be well explained from the point of view of anokhin’s concept of systemogenesis. according to this concept, every species has its own laws of ontogenetical development, which must exactly correspond to the ecological factors of the species-specific environment. actually, rapid establishment of the close mother-newborn bond is not ecologically necessary for moose because they are solitary forest animals. on the first days of life in a natural environment, the moose calf does not see anyone except its mother, and therefore cannot mistake her for anybody else. both creatures have enough time to learn little by little to recognize each other. working with reared calves, we had an opportunity to study their active and passive defensive behavior. the calves revealed both during the first hours of life. however, the first is formed gradually by active learning, while the second reveals itself in full form within the first hours of a calf’s life. the features of this specific behavior are the following: in response to a rapid approach of a frightening object from which the calf cannot run away, the calf’s muscular tonus weakens abruptly, the calf falls down to the ground and sprawls with stretched head and closed eyes. at the first moment of fright, for a few seconds (when the calf often tries to run away) the calf’s heart rate (hr) and breathing rate (br) increase abruptly (from 190–200 up to 230– 240 bpm and from 70–80 up to 130–140 per minute, respectively). after that, hr and br fall rapidly and become 1.5–2 times as low as baseline levels (up to 140–150 bpm and up to 40–45 per minute, respectively). the calf may remain in this state 10–15 minutes. the behavior described disappears in normally developing calves after 8–10 days of life, and they begin to react to danger by flight or by attack. but weak or sick calves continue this behavior even at the age of 1 month. only because the tamed animals allowed us to approach very near them could we record and describe practically the whole vocalization repertory of moose in various situations, including the sounds of wild males in the rutting period (bogomolova et al. 1984). having radio-tagged animals, we could investigate the complex behavioral organization of sexual interactions among moose living in the study area. we studied ethological and physiological features of male and female behavior at different stages of breeding, such as the onset of the rutting period, characteristics of forming, and subsequent disintegration of breeding pairs, the entire pattern of the breeding ritual up to copulation, ontogenesis of breeding ritual elements, characteristics of male interactions at various ages and dominance, the circumstances of the beginning, and entire pattern of tournament battles. moose sexual activity appeared to fit into the whole sysstudy of moose behavior bogomolova et al. alces suppl. 2, 2002 40 tem of social interactions of moose in the study area. it was found that moose-cows inhabit rather restricted and stable home range areas up to 60 km2 (bogomolova et al. 2002). references bogomolova, e. m., and y. a. kurochkin. 1984. parturition in a moose-cow. behavior of a moose-cow and a newborn calf. zoological journal 63:1713– 1724. (in russian). , and . 1987a. biological premises of rational moose domestication. pages 85-97 in problems of animal domestication. nauka, moscow, russia. (in russian). , and . 1987b. system organization of behavioral acts. pages 353-369 in functional systems of organisms. medicina, moscow, russia. (in russian). , and . 2002. parturition activity of moose. alces supplement 2:27-31. , , and a. n. minaev. 2002. home ranges and migrations of the kostroma farm moose. alces supplement 2:33-36. , , and a. a nikolslkiy. 1984. sound signals in communicative behavior of moose (alces alces). zoological journal 63:1878–1888. (in russian). minaev, a. n. 2002. use of telemetry to study behavior of domesticated moose. alces supplement 2:89-92. alces vol. 44, 2008 seip interactions among caribou, wolves, and moose 1 mountain caribou interactions with wolves and moose in central british columbia dale r. seip british columbia ministry of forests and range, 1011 fourth ave., prince george, b.c. v2l 3h9, canada, e-mail: dale.seip@gov.bc.ca abstract: mountain caribou (rangifer tarandus caribou) populations in south-eastern british columbia are declining over most of their range and are listed as threatened. predation has been documented as the major cause of declining caribou numbers. excessive predation by wolves (canis lupus) has been related to increased moose (alces alces) numbers. the increase in moose appears to be the result of a natural colonization process that has been enhanced by human-caused habitat change. options to reduce the rate of predation include reducing wolves, reducing moose, and reducing the amount of early seral habitat that supports moose. current management includes population control of moose and wolves. monitoring and assessment of these approaches will guide the future management strategy used to maintain mountain caribou in south-eastern british columbia. alces vol. 44: 1-5 (2008) key words: alces alces, british columbia, canis lupus, habitat, management, mountain caribou, moose, rangifer tarandus caribou, wolves mountain caribou are an ecotype of woodland caribou (rangifer tarandus caribou) that live in the mountains of south-eastern british columbia, extending into northern idaho (heard and vagt 1998). the mountains inhabited by these caribou experience extremely high snowfall during the winter which prevents the caribou from cratering through the snow to obtain terrestrial forage. consequently, mountain caribou winter in upper elevation subalpine forests (terry et al. 2000, apps et al. 2001, johnson et al. 2004) and feed almost exclusively on arboreal lichens (bryoria spp.). historically, mountain caribou were abundant and widely distributed throughout the mountains in south-eastern british columbia, but over the past century their numbers and distribution have greatly declined (spalding 2000). population declines and range contraction have continued in recent years (wittmer et al. 2005). for example, the george mountain herd just east of prince george was observed to be declining over the past few decades and was completely extirpated by 2004 (seip, unpublished data). as a result of these declines, mountain caribou are nationally listed as threatened and the subject of ongoing recovery efforts (mountain caribou technical advisory committee (mctac) 2002). initially, management concern for mountain caribou focused on protecting old-growth subalpine forests that provide the arboreal lichens that they rely on during winter. in response, a large amount of subalpine forest habitat has been protected from logging to provide caribou winter range (mctac 2002). however, research over the past few decades has indicated that the primary cause of declining mountain caribou herds is excessive levels of predation on both calves and adults during the summer (seip 1992, kinley and apps 2001, wittmer et al. 2005). grizzly bears (ursus arctos), black bears (u. americanus), wolves (canis lupus), and cougars (felis concolor) can all be major predators. bear predation is important throughout the range of mountain caribou, whereas cougar predainteractions among caribou, wolves, and moose seip alces vol. 44, 2008 2 tion is more significant in the southeastern parts of the province, and wolf predation is more significant in the central portion of the province (wittmer et al. 2005). winter survival of mountain caribou is quite high because they have minimal exposure to predators during this time. bears are hibernating, and wolves are found at lower elevations where they are sustained primarily by moose (alces alces) (seip 1992). conversely, moose and wolves often move to higher elevations in summer which increases both the spatial overlap of wolves and caribou and the predation of adult and calf caribou. the key question is why is predation by native predators resulting in decline of caribou now, given that these animals co-existed in the past? it is unlikely that grizzly bear populations have increased above historic levels, rather, they have probably been reduced in many areas due to hunting. in contrast, wolf populations have probably increased in response to increasing moose populations. wolves exhibit a strong numerical response to increasing prey availability (messier 1995), and the number of moose in central british columbia has greatly increased over the past century. today, the density of moose in central british columbia is greater than 1 moose/km2 (walker et al. 2006). there is some debate as to whether moose were completely absent from this region (hatter 1950, peterson 1955), or were sparse and scattered in the early 1900s (spalding 1990). regardless, moose began to increase in number in the early 1900s, and their distribution spread contiguously across central british columbia. for example, in the mountains immediately east of prince george, the colonization was recorded by local settlers: "in 1911 ernest jenson left his native denmark and came to canada. a few months later, he moved to the dome creek area … in 1912 and 1913, he was hired on as a hunter for the grand pacific railway…. although moose were present in the area since about 1900, ernest spent a whole year in the forests before seeing one. the moose were working their way south at about eight to ten miles per year. it wasn’t long before they were in great abundance… caribou were in abundance and it was common to see small herds in every valley" (boudreau 1998). early settlers also reported that the arrival of moose coincided with the decline in caribou, although the interaction among moose, wolves, and caribou was not obvious to them: "i asked him what he thought the reason was for the decline of the caribou populations… do you think the moose had any bearing on it? i asked. i can’t see how. they don’t even eat the same kind of food, and yet the caribou did start going down hill shortly after the moose arrived. probably the wolves were more responsible than anything else. i mean they sure killed a lot of them" (boudreau 1998). it was similarly reported in other parts of their range that decline in mountain caribou numbers coincided with an increase in moose numbers (edwards 1956, bergerud and elliot 1986). it appears that the decline in mountain caribou and their distribution in central british columbia are primarily due to excessive predation, and that this predation is related to high wolf numbers that are sustained by the increased numbers of moose (bergerud and elliot 1986, seip 1992). the cause of the increased number and distribution of moose is probably a combination of natural and human-caused factors. the initial increase and spread of moose during the early 1900s alces vol. 44, 2008 seip interactions among caribou, wolves, and moose 3 may have been related to a warming period following the end of the little ice age (luckman 2000). subsequent climatic warming in central british columbia over the past century (b.c. ministry of environment 2002) would have further encouraged increased moose abundance and distribution as severe winters became less common. consequently, at least part of the increase in moose numbers is probably related to climate change. however, there is also reason to believe that the increase in moose and the associated decline in mountain caribou are, in part, related to human-caused habitat change. the forest ecosystems in and adjacent to traditional range of mountain caribou had a very low frequency of natural disturbance resulting in a landscape dominated by mature and old forests (seip 1998, delong 2007). under the natural disturbance regime, these landscapes would provide little early seral habitat at low elevations that is preferred by moose as winter range. however, forest harvesting and human settlement has greatly increased the amount of early seral habitat, thereby increasing habitat suitability for moose within much of the range of mountain caribou. such habitat change appears to have benefited moose at the expense of mountain caribou. an analogous situation for mountain caribou exists in south-eastern british columbia where cougars (felis concolor) are a major predator of caribou (kinley and apps 2001). climatic and landscape changes that benefit deer (odocoileus spp.) and elk (cervus elaphus) habitat have increased these ungulate populations that, in turn, sustain high cougar numbers resulting in increased predation of caribou. management implications recently completed recovery plans for mountain caribou identify the need to reduce the rate of predation to achieve recovery (mctac 2002, hart and cariboo mountains recovery implementation group (hcmrig) 2005). these plans recognise that management actions could occur at different trophic levels to reduce predation, and include: 1. directly reduce wolf numbers by killing wolves or implementing reproductive control. this approach would have the most direct and immediate effect on reducing wolf numbers and wolf predation on caribou. however, wolf control is very controversial, and the management action would have to be used as an ongoing program because wolves would quickly recover following the cessation of any control program if moose habitat and moose numbers were not reduced. 2. reduce wolf numbers by reducing the number of moose which provide the primary prey of the wolf population. moose populations could be reduced by liberalizing hunting regulations, which is somewhat less controversial than predator control, but still a significant concern to hunters and first nations. also, moose reduction would have to be an ongoing program because the moose population will soon recover if good quality habitat is available. the potential for enhanced moose hunting to be a long-term solution is compromised by an ongoing decline in the number of hunters in british columbia, as well as the feasibility of hunters being able to adequately reduce moose to very low densities in inaccessible areas. there is concern that in the short term, wolves deprived of moose will temporarily increase their predation on caribou. a short-term wolf control program may be required to reduce wolf numbers to a level compatible with reduced moose numbers. 3. reduce the rate of forest harvesting, prescribed burning, and other practices which create early seral habitat. reducing early seral habitat on the landscape should reduce interactions among caribou, wolves, and moose seip alces vol. 44, 2008 4 moose numbers and lead to lower wolf numbers. this approach is favoured by environmental groups, but further reductions in forest harvesting concerns the forest industry (hcmrig 2005). although much of the upper elevation habitat that is used by caribou is already protected from forest harvesting, reducing the number of moose and wolves would require an additional reduction in the rate of logging in adjacent valley bottoms, even though caribou seldom use those low elevation areas in central british columbia. if the latter approach was implemented, there would be a substantial lag time before the existing cut blocks regenerated to a stage where they no longer provided good moose habitat. consequently, moose reduction and/or wolf control for several decades may be required until cut blocks mature beyond good moose habitat. some environmental groups accept use of interim predator management if it is used as a temporary tool while habitat recovers, but oppose predator management (reduction) if it is used as a permanent solution in place of habitat management (hcmrig 2005). the degree that forest harvesting would have to be reduced to adequately reduce wolf numbers is unclear. the hcmrig (2005) recommended that moose habitat suitability in areas adjacent to mountain caribou habitat should not exceed the amount that would occur under a natural disturbance regime in that area. however, given that moose are a recent colonizer to mountain caribou range, it may be that even under the natural disturbance regime moose and wolf numbers would be too great to allow caribou to co-exist. if so, there may be no possibility of using habitat management to maintain self-sustaining caribou herds in some or all areas, and ongoing predator-prey management would be required to maintain caribou populations. in practice, these three approaches are not mutually exclusive, and a management strategy could include some combination of reducing the amount of early seral habitat, increasing the moose harvest, and some direct wolf control or reduction. a similar set of options also applies to managing the predatorprey interactions among caribou, deer, and cougars in south-eastern british columbia. at this time, the final management approach to maintain mountain caribou in british columbia is still evolving. in the meantime, a variety of these management approaches have been implemented, including reproductive control of wolves in the quesnel highlands, moose reduction in the parsnip watershed, and moose reduction in the revelstoke area. limiting the creation of early seral habitat is an important management consideration in parks where there is no competing concern for timber supply and production. research projects are in place to monitor all these management programs, and the resulting information will be used to refine the future application of these management practices. references apps, c. d., b. n. mclellan, t. a. kinley, and j. p. flaa. 2001. scale-dependent habitat selection by mountain caribou, columbia mountains, british columbia. journal of wildlife management 65: 65-77. bergerud, a. t., and j. p. elliot. 1986. dynamics of caribou and wolves in northern british columbia. canadian journal of zoology 64: 1515-1529. boudreau, j. 1998. crazy man’s creek. caitlin press inc., prince george, british columbia, canada. british columbia ministry of environment. 2002. indicators of climate change for british columbia, 2002. http://www.env. gov.bc.ca/air/climate/indicat/pdf/indcc. pdf (accessed january 2008) delong, s. c. 2007. implementation of natural disturbance-based management in northern british columbia. forestry http://www.env.gov.bc.ca/air/climate/indicat/pdf/indcc.pdf http://www.env.gov.bc.ca/air/climate/indicat/pdf/indcc.pdf http://www.env.gov.bc.ca/air/climate/indicat/pdf/indcc.pdf alces vol. 44, 2008 seip interactions among caribou, wolves, and moose 5 chronicles 83: 338-346. edwards, r. y. 1956. snow depths and ungulate abundance in the mountains of western canada. journal of wildlife management 20: 159-168. (hcmrig) hart and cariboo mountains recovery implementation group. 2005. recovery implementation plan for threatened woodland caribou (rangifer tarandus caribou) in the hart and cariboo mountains recovery area, british columbia, canada. unpublished report. http:// www.centralbccaribou.ca/crg/24/rap (accessed january 2008) hatter, j. 1950. the moose of central british columbia. ph.d. thesis, state college of washington, pullman, washington, usa. heard, d. c., and k. l. vagt. 1998. caribou in british columbia: a 1996 status report. rangifer special issue no.10: 117-123. johnson, c. j., d. r. seip, and m. s. boyce. 2004. a quantitative approach to conservation planning: using resource selection functions to identify important habitats for mountain caribou. journal of applied ecology 41: 238-251. kinley, t. a., and c. d. apps. 2001. mortality patterns in a sub-population of endangered mountain caribou. wildlife society bulletin 29: 158-164. luckman, b. h. 2000. the little ice age in the canadian rockies. geomorphology 32:357-384. messier, f. 1995. on the functional and numerical response of wolves to changing prey density. in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. university of alberta press, edmonton, alberta, canada. (mctac) mountain caribou technical advisory committee. 2002. a strategy for the recovery of mountain caribou in british columbia. british columbia ministry of water, land and air protection, victoria, british columbia, canada. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. seip, d. r. 1992. factors limiting woodland caribou populations and their interrelationships with wolves and moose in southeastern british columbia. canadian journal of zoology 70: 1494-1503. _____. 1998. ecosystem management and the conservation of caribou habitat in british columbia. rangifer, special issue no. 10: 203-211. spalding, d. j. 1990. the early history of moose (alces alces) distribution and relative abundance in british columbia. contributions to natural science, royal british columbia museum, victoria, british columbia, canada. _____. 2000. the early history of woodland caribou (rangifer tarandus caribou) in british columbia. wildlife bulletin no. b-100. british columbia ministry of environment, lands and parks, wildlife branch, victoria, british columbia. terry, e. l., b. n. mclellan, and g. s. watts. 2000. winter habitat ecology of mountain caribou in relation to forest management. journal of applied ecology 37: 589-602. walker, a. b. d., d. c. heard, v. michelfelder, and g. s. watts. 2006. moose density and composition around prince george, british columbia. final report for forests for tomorrow. project no. 2914000. british columbia ministry of environment, prince george, british columbia, canada. wittmer, h. u., b. n. mclellan, d. r. seip, j. a. young, t. a. kinley, g. s. watts, and d. hamilton. 2005. population dynamics of the endangered mountain ecotype of woodland caribou (rangifer tarandus caribou) in british columbia, canada. canadian journal of zoology 83: 407-418. http://www.centralbccaribou.ca/crg/24/rap http://www.centralbccaribou.ca/crg/24/rap f:\alces\supp2\pagema~1\rus1s.pdf alces suppl. 2, 2002 ashihmina importance of moose in the sub-urals 11 the importance of moose to the people in the northern sub-urals during the bronze and early iron ages1 lidija i. ashihmina institute of language, literature, and history, komi scientific center, ural division of the russian academy of sciences, 167610, syktyvkar gsp, komi republic, russia abstract: moose (eurasian elk) are often found at ancient archaeological sites in the northern sub-urals. moose were an important component of the hunters’ lifestyle as a source of food. their skins were used for clothing and footwear and to make various household items. skins were also used to make tents. moose influenced the life and well being of these ancient hunters and were prolific in their myths and legends. hunters believed the sun was a gigantic moose “running” over the entire horizon during the day. this investigation is devoted to the interpretation of some moose images and objects, dating to the bronze and early iron age, collected from archaeological sites in the northern sub-urals. the analysis of these images and objects indicated that in ancient times the peoples of this region originally personified the universe as a gigantic moose “mother”. later, a tree-of-the-world concept was added to this image. this concept was the foundation for the organization and structure of people’s lifestyles in the northern sub-urals. alces supplement 2: 11-17 (2002) key words: archaeological sites, bronze age, eurasian elk, iron age, moose, sub-urals moose were given a special and important place in the culture of the people of the taiga and forest-taiga boundary because of its size and strength. it was commonly worshipped and as such its image was manifested in a variety of different forms. sculptural representations of moose were made out of stone, antler, bone, wood, and bronze. the image was also carved in silver, bronze, bone, on pendants, and in stone. the predominance of moose in these ancient representations indicates that many of the concepts of the universe that were held by the northern sub-urals’ peoples were connected with moose. buildings moose were ritually sacrificed during the building of homes and forts. archaeological excavations from the early bronze and early iron ages documented that moose heads were placed as “building sacrifices”. some of these can be seen at the bujskoje hill-fort on the vjatka river (author’s excavations during 1976, 19781981) and in majdanskoje at the mazarskoje i settlements on the middle volga river (nikitin 1980). moose heads were also placed above the entrance or on the roofs of ancient buildings (nikitin 1980). decorations on the roof ridges often took the form of moose heads or antlers. these were common in later excavations where the dwelling or rampart was shaped as if it was growing out of the moose’s head and crowned with a similar head or the image of another moose. the dwelling represented the tree-of-the-world concept in which moose played an important role. moose, however, were also depicted in another 1editor’s note: this manuscript was submitted without references and due to difficulties contacting the author the manuscript was published without them. importance of moose in the sub-urals ashihmina alces suppl. 2, 2002 12 major theme that of the role of adversary, or of life and death. “the building sacrifice” and roof ridge decorations could represent a sacrificial ritual (bajburin 1983). in this region, during the eneolith, the floors of the dwellings were powdered with red ochre. ochre was a symbol of blood and life that was used to sanctify the dwelling. it represented the life of the dwelling and corresponded to their vision of the world. traces of ochre were found within the walls of dwellings excavated at the vomyn’jag i site (ashihmina 1988). combs in the basin of the vjatka river, there were numerous objects made of metal, bone, and stone that were decorated by images of moose. of particular interest, are the bone combs found at the bujscoe hill-fort during excavations in 1978 and 1979 (personal observation). as combs are used in the hair, they are typically symbols of fertility, strength, and health. this symbolism is evident in the folktales of many different peoples in the northern sub-urals. for example, a thick forest appears at the place where a thrown comb touches the earth. in russian erotic short stories, a comb often corresponds to a phallus. the same idea is implied by the ritual of combing the bride’s hair and by presenting the bride with a comb. the comb itself is present during childbirth and during burial rituals. the ritual of combing the hair is also associated with water spirits, mermaids, and other representatives of the lower world (uspensky 1982). the combs from the bujscoe hillfort are decorated with a stylized moose head and the tree-of-the world (fig. 1a and fig. 1b). the moose heads are turned towards the edge and the ears are either oval or pointed with stylized, curling antlers. these curls probably represent the sun and reflect its movement across the sky. according to the direction of the curls depicted on the comb, it is possible to define the parts of the world on the comb. the curls that are turning counterclockwise symbolize the sun’s movement to the east. the curls that are turning clockwise symbolize the sun’s movement to the west. the comb can be divided into the four cardinal points, with the north being behind the tree in the center and the south in front of it. the east-west concept can also represent the parts of the day such as “morning-evening” and the seasons of the year “spring-autumn”, whereas the “northsouth” corresponds to “day-night” and “winter-summer”. the half-moon above the tree in fig. 1a indicates that the north personifies the night. the four moose are shown in one plane. it is unusual that the heads of the central pair face away from the tree. it is possible that this pair is a part of the ritual scene and, together with the second pair, represent the “keepers” or guards of the four parts of the world. in the most archaic classification systems, it is common that an animal is used to depict or correspond to a part of the world (semeka 1977). the fact that these combs are composed of three parts is also of interest. the upper part contains the ritual scene and is separated from the teeth by three straight lines. it is likely that this represents a ritual similar to that observed in priajan tungus. they worship the moose head and had a ritual that they observed when they cooked it. the head was divided into nine pieces in a definite order: first the lower part of the head was cut off, the tongue was removed, and the jaw was torn into pieces, each of which was again torn into two halves (pekarsky and tzvetcov 1912). in general, whole combs that depicted moose heads could represent the way that ancient peoples in the northern sub-urals thought of alces suppl. 2, 2002 ashihmina importance of moose in the sub-urals 13 fig. 1. findings of objects with moose images on archaeological monuments of the northern suburals. 1a, b, c – bujskoje hillfort, basin of the middle vjatka river; d – sejminsky burial place, the oka river (bader 1970, p. 116, fig. 50); e – find from the vicinity of perm’ city (studzittsskaja 1969, p. 227-229); f – woman’s headdress belonging to a woman on the upper volga river (maslova 1951, p. 44, fig. 2). materials: 1a, b – bone; c – stone; d, e, – bronze; f – wool, silk and gold embroidery on canvas. importance of moose in the sub-urals ashihmina alces suppl. 2, 2002 14 their world. later, this image evolved to become the image of a tree-of-the-world. spindle whorls in kama settlements, ornamented round spindle whorls are thought to reflect the calendar and certain mythological images. the circle is found in the tree-of-the-world image with the sun or its symbols often a stereotyped concept. since spindle whorls are connected with the circle, as well as the tree-of-the-world, both the time and seasonal cycles are also reflected in them such that the movement of the sun during a day and a year links motion in time and space in a circular pattern (toporov and mejlah 1982). the spindle whorls represent the vertical and horizontal spheres of the world and the directions of the four parts of the world can be represented by the crossed branches or the silhouette of moose heads (fig. 1c). at the center, where they cross, is a wooden core, which describes the vertical axis. the spindle whorl motion symbolizes the sun’s motion through the daily and yearly cycle. tree-of-the-world the tree-of-the-world determined the horizontal space of the circle in a year and the vertical space of the circle in a day (toporov and mejlah 1982). there are references to the sun’s circular motion within the language of the komi, for example, the expression “time kills” literally means “the sun pushes” (“shondice jetke”). the cycling concept of time is expressed in many images and is based on the sun’s movement from east to west. a bronze casting, excavated from the gljadenov settlement, pozhegdin ii, in the middle vychegda basin, and dated to the third and fourth centuries a.d. is an excellent example of the concept-of-the-world (vaskul 1989). in this casting, the mistress of the universe is a moose with twins on her back. a snake is at the base with water coming out of it. the moose head symbolizes both a newly born moon and the sun, and the onset of a new day. in this, the moose represents reincarnation dying and rising again. all the details are composed and located in such a way that as a whole they reflect a concept-of-the-world with four parts, each corresponding to a part of the world. all dualisms are represented: the sun and moon, top and bottom, right and left, positive and negative, and life and death. this casting is an excellent illustration of a space and time fusion and their indivisible unity. a bronze object from the hearth, in an early ananjin dwelling of borganjel, a settlement in the middle branch of the nivshera river, represents the tree-of-life concept (ashihmina, unpublished data 1985). this bronze object, created by flat casting, depicts the stem of the tree as curved. this curve also includes a crawling snake and is depicted as a deep furrow. the crown of the tree is bent and represents a new moon with its points facing upward. parallel to the moon are smaller branches in a crescent form. the points on these branches are pointing downwards. on the lower edge of the crescent, there are four small projections and four hollows on one side. on the tree stem, there are three projections and three hollows. the tree-of-life is one of the variations on the tree-of-the-world and is a mythological representation of life. in the concept, there are three zones: upper (the crown), middle (the stem), and lower (the crawling snake). each corresponds to the three vertical spheres of the universe: upper, middle, and lower worlds. sometimes, the image of the tree-of-life has a negative interpretation as the opposition to life (toporov 1980). the reason for this is evident in the myths of the broken moon or about being consumed by inhabitants of a alces suppl. 2, 2002 ashihmina importance of moose in the sub-urals 15 lower world. such myths are well known among the abkhasians (ivanov 1980), the hittites (ivanov 1980), and the selkups (prokofjeva 1976), wherein a cruel god, a representative of the lower world, tears or eats up the moon. myths surrounding the moon, which falls to earth, are also prominent among the khets (ivanov 1977). on the borganjel bronze, the role of a chaotic character is manifested in the snake. it has gnawed through part of the lower crescent moon, on the concave side where there are traces of teeth described by four projections and four holes on one side, and three projections and three holes on the other side of the tree stem. the old moon does not exist any longer and gives way to a young moon. it has the features of a man’s face and the snake is crawling up to it, indicating that the cycle will be repeated. in komi folklore, there are no legends about the broken moon. but, from the analogy presented, it is supposed that such a myth might have existed in the legends of the people of the northern sub-urals. in evenk mythology, among the spirits and masters of the lower world is a mythological creature “khalir” resembling a moose or reindeer, and having the antlers of a moose and the tail of a fish. its role is to guard a shaman’s mythological river, along which the shaman and his spirits travel to the lower or upper worlds of the universe. since the mythological river connects all three worlds of the universe, the “guard” of this river, “khalir”, possesses attributes of each. moose antlers are a cosmic symbol of the earth and the fish tail is a cosmic symbol of water (anisimov 1959). other myths explain that during this ritual, an evenk shaman called for his spirits creatures that were half human and half non-human. the people thought that when the shaman died, the animal counterpart went to the “lower world”. similarly, the creature was thought to have this duality. among shaman spirits of double nature, the image of the mythological mother-animal is also prevalent. in some cases, this was presented as the image of a mother-moose, but also mother-wild reindeer, mother-bear, mother-bird, etc. have been observed (anisimov 1958). because the creature observed in the object has a nature that is half human and half moose, it can be connected with a shaman cult. therefore, this could also be interpreted as an image of a shaman and his spirits travelling along some mythological river, where the snake represents the water. in the mythology of the majority of siberian peoples, there are close links between the cosmological images of the moose and the mammoth. the hunters believed that the mammoth originated from old, wild reindeer. as the old, wild reindeer ages, it lays down in a bog where it eventually turns into a mammoth. the mammoth disappears under the earth and wanders throughout the lower world, splitting the banks of the river (anisimov 1959). the mammoth is the creator of the earth shape in the folklore of the dolgans. the mammoth goes down to the lower world after the earth is created and takes all his offspring with him (anisimov 1959). komi legends also include mammoths as the creators of relief or contours and rivers. “in the old times there lived the mammoth. the earth could not carry it because it was very heavy. where he stomped there appeared a furrow and water began to flow giving birth to a stream. and where he went to and through there appeared a river” (rochev 1984). the nenetz god num made the earth smooth, but the mammoth went along it and spoiled it. all the places where he dug became mountains and the places where he pressed became lakes. num got very angry at the mammoth and sent him underground to the lower world. in evenk legends, representatives of importance of moose in the sub-urals ashihmina alces suppl. 2, 2002 16 the lower world also possessed characteristic features of the cosmological image of khely (shely, sely), a mammoth, and djabdar (dzjabdar), a mythological snake. the evenk connected these symbols with the creation of the mountains and rivers in the middle world and they were considered to be the creators of the world. in the beginning, the middle world did not exist and there was water all around. man had no land to live on. the mammoth, khely, decided to help man. with his “horns”, he dug out enough earth for all the people. the snake, djabdar, helped the mammoth to smooth the clods of the earth. in the places where the djabdar crawled, there appeared rivers and where the clods were left unsmoothed, mountains appeared, and where the mammoth stomped or lay down to create deep holes, they became lakes (anisimov 1959). one of the legends of evenksorochons speaks about the origins of mountains, rivers, and lakes. they were created as a result of the fight between the mammoth and snake (mazin 1984). in the myths of different peoples, the mammoth communicates not only with the representatives of the lower world (fish, pangolin, snake), but also with symbols of the middle world such as moose, reindeer, horse, bear, and the upper world, such as birds (toporov 1980). the image of a mammoth as a gigantic bird is known among the ob ugres, selkups, and evenks (toporov 1982). these images can be found in the materials ca. sixth century a.d. some items found in podcheremsky, a complex of burial mounds in the northern sub-urals (ashihmina 1988) and west siberia, include bronze pendants made of a combined figure of a duck and moose where the head is of a moose cow and the body of a duck. on the borganjel’ object, we distinguish two images of moose. one image depicts the snake cutting the lower half-moon and touching the upper one, which seems to divide them into two parts. but this can also be viewed as joining them into one unit: there is a short segment to the right of the snake’s head that runs parallel to it and connects both half moons. the moose heads are separated by a triangular hole located between the snake head and the antler of the moose. the twin cult is also present. daggers the creation of the middle world is reflected on the handle of the bronze dagger from the seiminsky burial ground in the oka river basin and dated to the bronze age (fig. 1d). the end of the handle depicts a moose head. one of its sides is decorated by the outline of an eight-legged snake crawling after an elk mammoth. the symbol of the eight-legged snake reflects the twin cult. one more reflection of the same myth is represented on a stray find of a bronze dagger found in the vicinity of perm’ and belonging to the same time (fig. 1e). the tip of the handle is molded in the form of two sculptures of cow moose heads, with their muzzles toward the handle, and joined by a crosspiece. along the edge of the blade, the wavy lines could represent snakes. the twin cult is evident in this casting too. the world analysis of these materials leads to the conclusion that the ancient inhabitants of the northern sub-urals believed that the universe was a gigantic animal. this concept of the world could be based on the fact that moose occupied such an important and central place in the lifestyles of the ancient hunters. this is tied to their beliefs regarding their existence and their customs. it is reflected in their artifacts and language. this in its turn is well supported by the presence in the komi language of two names for the moose: “yera”, which means strength alces suppl. 2, 2002 ashihmina importance of moose in the sub-urals 17 and might and “lola”, which means soul and life. the people recognized their dependence on nature so strongly that the image of the world included many features that made it difficult to separate man from nature (gurevich 1984). due to the changes that took place in the economy of the population of the north sub-urals, the image of the moose gradually gives way to the image of the horse, but it is not substituted completely. the ritual of decorating roof ridges with moose heads and antlers remained prominent in the north for a long time. in embroidery and knitting, the “moosereindeer” near the tree-of-the-world is evident (fig. 1f). the komi people have a p r o v e r b , “ y e r a y d y d d z u d , n o i s i j e kon’jas’le”, which translates to be “moose is great, but also stumbles”. the role of an animal in the mythology of the inhabitants was defined by the importance those animals had during the early stages of community development. ethnographical materials from siberia show that the universe itself was thought of as a living being and was associated with different images of animals. orochi imagined the universe as an eight-legged moose cow often depicted in the well-known picture by shaman savely khatunka. the earth was also thought by nganassan to be a cow moose or wild deer with its head turned to the west it follows the sun. for the evenk people of podkamennaja tunguska river basin, the mistress of the universe is the mother of animals and people simultaneously. it is quite common for hunters to imagine the sun as a gigantic moose, covering the whole of the horizon during the day. language f o r t h e e v e n k p e o p l e o f t h e podkamennaja tunguska river basin, the mistress of the universe, “enintyn” which means belonging to the universe, their mother, is at the same time the mother of animals and people. this mistress is pictured as a woman and an animal. the evenk word “enin” has two meanings: mother and cow moose. in the evenk language, there are a number of words with the root “en”, such as: “enin” mother, cow moose; “enike” grandmother; “eniken” she-bear, “enty vazhenka” with a calf; and “entyl” parents (anisimov 1959). evidently, these words with similar roots have the meaning of mother, giving birth, and giving life. in the komi language, “yera” elk; “en’yera” cow moose; “en” can be translated as mother; and, “en’a-nyla” mother and daughter; “en’a-pia” mother and son; “en’osh” sow (timushev and kolegova 1961). here we observe the same phenomenon and the usage of “en’” in the meaning of mother, giving birth, and giving life. gradually the new concept of the world as the tree-of-the-world organizing and determining all the life cycles of man is based on the concept of the universe represented as an animal. alces vol. 45, 2009 carson et al.– effects of simulated browsing on aspen 101 compensatory shoot growth in trembling aspen (populus tremuloides michx.) in response to simulated browsing allan w. carson1, roy v. rea2, and arthur l. fredeen2 1undergraduate student, university of northern british columbia, prince george, b.c., v2n 4z9 canada ; 2ecosystem science and management program, university of northern british columbia, prince george, bc, v2n 4z9 canada abstract: moose (alces alces) browsing influences plant growth and architecture. we sought to determine the impact of the timing of moose browsing on bud development and growth in aspen shoots in the subsequent spring through simulation by clipping aspen (populus tremuloides) stems in the field in june, july, and august 2005 at the university of northern british columbia, prince george, bc. to observe new leaf+shoot development in aspen over a 60-day period, the top meristems of both simulated browse treatments and unbrowsed controls were harvested in january 2006, and incubated in a growth chamber that simulated local springtime conditions. total leaf+shoot biomass produced from stems was higher for juneand august-’browsed’ stems relative to unbrowsed controls. mean stem diameter was significantly higher and number of total buds significantly lower on clipped relative to unclipped stems. the number of buds that broke winter dormancy and became active in the growth chamber remained relatively constant for both clipped and unclipped aspen, but with fewer dormant buds on clipped stems than controls. overall, our findings suggest that the mechanical effects of moose browsing on aspen stimulate the production of compensatory leaf+shoot biomass, and therefore potential browse. alces vol. 45: 101-108 (2009) key words: alces alces, populus tremuloides, browsing, herbivory, plant-animal interaction. the nature and level of plant response to browsing by moose can vary (bergström and danell 1995). response may be species dependent or may vary individually within a species as a result of differences in time of year or the amount of tissue removed (rea and gillingham 2001). the compensatory growth response of many plants browsed in winter (danell et al. 1985) and the growing season (bergström and danell 1995, gadd et al. 2001) is equal to the level of annual growth in unbrowsed plants of the same species. however, the degree of compensatory growth (e.g., location of meristems, number of dormant buds activated, shoot size, and length) varies in response to the degree of browsing damage; such variance can affect both plant productivity and quality of forage. for example, birch (betula pendula and b. pubescens) produced larger shoots with larger and more chlorophyll-rich leaves following browsing (danell et al. 1985). almost all studies of plant response to herbivory have documented the overall effects of browsing damage to individual plant health and morphology, but few have investigated specifically how individual “plant units” respond. honkanen and haukioja (1994) speculated that individual plant units, such as branches or ramets, can act as semiautonomous units in that response to damage as an isolated unit would be similar to its response when attached to the parent tree. in order to examine the compensatory response of aspen meristem units, we observed isolated meristems under incubation that were clipped in simulated browsing treatments during the previous growing season. we believed that the response to clipping damage would result in greater allocation of new biomass to stems as compared to undamaged branches, as found in a similar study by stevens et effects of simulated browsing on aspen – carson et al. alces vol. 45, 2009 102 al. (2008). prior to the simulated browsing treatments, we experimented by incubating different stem cuttings of different woody shrubs and trees at different times of the year to observe their growth response. we determined that branches of aspen that were clipped at different times during the previous growing season altered their tissue repair physiology in response to clipping (carson et al. 2007). here, we sought to determine whether the timing of simulated browsing would influence the compensatory growth response of aspen in the next growing season. study area we conducted our study on an approximately 20 ha area located adjacent to the university of northern british columbia (unbc) endowment lands near prince george, b.c., canada (lat 53º 53’ n, long 122º 40’ w). the topography was rolling at an elevation of 780 m above sea level. the climate is continental and characterized by seasonal extremes with cold winters and warm, moist summers. mean annual precipitation is approximately 460 mm; snow fall averages approximately 200 cm and the mean annual temperature ranges 1.7-5 °c (atmospheric environment service 1993). the study area was clear-cut approximately 15 years prior to the study. young trembling aspen was the dominant tree species on site, while pioneering species such as shrub willows (salix spp.), paper birch (betula papyrifera), and alder (alnus spp.) were also present. moose and deer (odocoileus spp.) are both native and foraged within the study area. our observations indicated that most browsing of aspen was by moose (~1.5 moose/km2; walker et al. 2006) at the time of this study. methods the simulated browsing (clipping) treatments imposed on aspen saplings (approximately 1-5m height) within the aspendominated stand (14,240 ± 5696 s. d. stems/ ha) were described in carson et al. (2007). four simulated browsing treatments (no-browse control and three growing-season clipping dates: 1 june, 16 july, and 30 august 2005) were imposed on 200 naturally growing aspen saplings. to approximate the mechanical effects of browsing, apical stems were clipped at 4.0 mm stem diameter proximal to the apical meristem, which is the average bite diameter of shoots browsed by moose in the study area (carson et al. 2007). the top 50 cm of winter-dormant stems from the aspen sapling crowns of control and simulated browsing aspens were harvested 7-14 january 2006. approximately 5 aspen stems from within each treatment and control were collected on each of the 7 harvest dates for a total of 160 stems from the original 200. forty of the individuals were damaged or killed by moose between the time of treatment and harvest (carson et al. 2007). immediately after removal, stems were placed in water buckets with their cut stem ends immersed in water to a depth of approximately 10 cm to reduce the effects of cavitation (williamson and millburn 1995). harvested stems were then transported to the enhanced forestry laboratory (efl) at unbc to record the extent of stem dieback resulting from the simulated browsing treatments imposed during the previous summer (carson et al. 2007), and prepared for sprouting in an environmental growth chamber (egc; model gcw 30, chagrin falls, ohio, usa). the necrotic (dieback) region below the point of summer clipping of each harvested stem segment was cut off at the terminus to eliminate unproductive and potentially phytopathogenic stem tissue. harvested stems were reduced to a set mass of 12.0 ± 3.0 g by cutting from the stem bottom (harvest point) and were defined as “set weight stems.” set weight stems were incubated in water baths within the controlled growth chamber for 45 days at a light and temperature regime that approximated the mean local climate in may, followed by 15 days at the mean climate conditions in alces vol. 45, 2009 carson et al.– effects of simulated browsing on aspen 103 june (meteonorm 4.0; fig. 1). during the first 4 weeks, the daytime photosynthetically active radiation (par), air temperature, and relative humidity (rh) were set at 600 watts m-2, 15 °c, and 44% rh, respectively, over a 16-h photoperiod; a 17-h photoperiod at 650 watts m-2, 19 °c, and 48% rh was used in the last 2 weeks. conditions at night were held constant during the full incubation period (0 watts m-2, 6 °c, and 87% rh). water baths only contained plants from the same treatment to avoid possible effects due to water-mediated hormone transport between stems of different treatments. baths were covered with white plastic and trays were painted white to prevent any light-induced temperature change to the medium (fig. 1a). stems were incubated in the growth chamber for 60 days (fig. 1b). during incubation, stems were monitored for the time of bud burst and maximum growth time prior to leaf desiccation as a result of stem embolism and/or cavitation (williamson and millburn 1995). a data logger (hobo temp/external channel data logger, onset computer corporation, h08-002-02, massachusetts, usa) was used to monitor light intensity, temperature, and rh throughout the incubation period. after the 60-day growing period, set weight stems were harvested and separated into new growth (new leaf+shoot) and preexisting stem. the number of active and dormant buds was recorded for each stem. fresh weights for new growth and pre-existing stem were recorded, and then dried at 60 °c for 2 (leaf+shoot) or 6 (old stem) days to measure oven-dry weight. statistical analyses we used one-way analysis of variance for unequal sample sizes (anova; zar 1999) to compare differences between clipping treatments and controls; new growth and pre-existing stem mass, mean stem diameter normalized to set weight stem mass, and dormant and active buds normalized to set weight stem mass were compared. tukey’s honestly significant difference (hsd) test for unequal sample sizes (zar 1999) was used for post-hoc comparisons among treatments. all anovas were performed using statistica (version 6.0, statsoft 2005, tulsa, ok). we used linear regressions to determine the relationship between the number of active buds and dry leaf mass per stem unit. regression equations were computed using excel (microsoft office 2003). results overall, significant differences in the ratio of leaf+shoot mass:total branch mass (new growth + pre-existing stem) were observed between the treatments and controls when examining the fresh weight of incubated stems (table 1). specifically, june and august fig. 1. harvested stem tops of aspen (12 ± 3.0 g) within an environmental growth chamber at; a) initial and b) final stages of a 60-day incubation period to assess regrowth potential. effects of simulated browsing on aspen – carson et al. alces vol. 45, 2009 104 clipping trials had higher leaf+shoot mass to total branch mass when compared to controls. also, the ratio of leaf+shoot mass to total branch mass for august-clipped stems was higher than that of july-clipped stems. no differences were found relative to dry weight of incubated stems, although june and august clipped stems were about 10% heavier than controls and approached statistical significance (p = 0.092). the ratio of mean diameter normalized to the set weight stem mass was higher for treatments (~0.44 mm/g for all treatments) than controls (0.33 mm/g; f(1,3) = 24.5, p = < 0.001). however, tukey’s hsd indicated that only controls were different from treatments (p = < 0.001). the ratio of dormant buds (f (1,3) = 9.599, p < 0.001) and total buds (f (1,3) = 5.5015, p = 0.001) normalized to set weight stem mass was not different among clipping treatments, but was higher for controls than for any clipping treatment (fig. 2). we found no differences (f (1,3) = 0.4436, p = 0.722) in the ratio of active buds normalized to set weight stem mass between any clipping treatment or the control (fig. 2). weak relationships were detected between the number of active buds and dry leaf mass (fig. 3); as the number of active buds increased, the dry leaf mass increased for all treatments 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 june july august controln um be r o f b ud s/ h ar ve st m as s (g -1 ) active buds dormant buds total buds n=31 n=39n=34n=38 fig. 2. the number of active buds and dormant buds normalized by the set weight stem mass (12 ± 3.0 g) prior to incubation for treatments after 60 days of incubation in a growth chamber. the numbers of total and dormant buds for all treatments were significantly different from the control (p < 0.001). month of simulated browsing f p june july august control n = 31 n = 38 n = 34 n = 39 mean s.e. mean s.e. mean s.e. mean s.e. new growth:total branch ratio fresh weight 0.088ab 0.004 0.077cb 0.004 0.090a 0.002 0.075cd 0.004 4.385 0.006 dry weight 0.058 0.003 0.053 0.003 0.060 0.002 0.051 0.003 2.192 0.092 table 1. mean ratio of new growth (leaf+shoot) mass:total branch mass in clipped aspen stems and unclipped control stems after simulated browsing (clipping) at 3 different times during summer. means in a row not sharing a common superscript indicate significant differences as determined by tukey’s hsd post-hoc tests. alces vol. 45, 2009 carson et al.– effects of simulated browsing on aspen 105 and the control (june: y = 32.349x – 0.6691, r2 = 0.6339; july: y = 20x + 3.2593, r2 = 0.2317; august: y = 28.507x – 0.9541, r2 = 0.2533; and control: y = 17.704x + 4.8627, r2 = 0.2206). discussion clipping stems to simulate browsing generally produces the same responses as natural browsing (haukioja and huss-danell 1997), but the effects of clipping and natural browsing on plant morphology and productivity have not been adequately examined in aspen. indeed, the question of whether browsing animals such as moose positively ‘cultivate’ their browse species is an open one. we found no evidence that season of simulated browsing on meristems affected the overall production of leaf and stem mass or influenced the proportion of active vs. dormant buds in the spring following clipping. however, we were able to demonstrate a significant effect of simulated browsing on these quantitative aspects of regrowth in aspen stem units when compared with unclipped controls. given that young aspen is important browse for moose, and that aspen can rapidly grow beyond browsing height of moose, a positive feedback from aspen browsing on forage availability is of more than academic interest. overall, our findings suggest that moose browsing can stimulate the production of more compensatory leaf+shoot biomass (potential browse) than is produced by unbrowsed stems. although aspen is not a preferred browse species in our area, it is consumed frequently by moose in areas of northern bc and elsewhere (renecker and schwartz 1998), especially in the absence or low abundance of other preferred browse. aspen has a high juvenile growth rate and productivity that combined with its ability to tolerate stress better than other tree species (lieffers et al. 2001), may explain the compensatory response we observed in response to clipping. stevens et al. (2008) examined herbivory tolerance in aspen and found a positive correlation between tolerance and increased allocation of new biomass to stems under high nutrient conditions. because we clipped aspens on the main stem, a loss of apical meristem dominance may help explain the compensatory response we observed. according to the sink-source hypothesis, a change in the ability of meristems to compete with other plants and even other branches of the same plant for resources is the r2 = 0.2206 r2 = 0.6339 r2 = 0.2317 r2 = 0.2533 0 5 10 15 20 25 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 dry leaf mass (g) n um be r o f a ct iv e bu ds june july august control july june august control fig. 3. relationship between the number of active buds and the dry leaf mass per set weight stems (12 ± 3.0 g) for treatments and controls. effects of simulated browsing on aspen – carson et al. alces vol. 45, 2009 106 primary way in which damage affects plants (honkanen and haukioja 1994). in this way, plant tissues (such as our aspen meristems) that have been damaged or removed by browsing (or clipping) are no longer available to photosynthesize and “sink” resources. this results in a reallocation of plant root resources to shoot production and plant compensation derived from axillary bud development (pratt et al. 2005). simulated browsing treatments also had an effect on the mean diameter of winter-dormant stems (normalized to set weight stem mass), increasing mean diameter of such stems over unclipped controls. clipping was conducted at a diameter pre-determined from bite marks of moose within the study area, so it was not surprising that unclipped stems with their intact leaders would have a lower mean diameter than stems damaged from browsing or clipping. although this difference between the mean stem diameter was an artifact of the clipping treatment, the change in architecture (either by clipping or browsing) can have a direct effect on a tree’s ability to compensate for tissue loss from browsing over time. plants with larger mean diameters had a lower number of total buds, presumably affecting the plants capability for shoot production relative to smaller diameter shoots. like our aspens, the mean shoot diameter of birch (betula spp.) was shown to be higher on stems previously browsed by moose than on unbrowsed trees of the same age (danell 1983). while the number of active buds per gram of stem tissue was similar between treatment and control stems, the number of dormant buds was significantly less on clipped stems (fig. 2). the reduction of dormant buds is likely related to the availability of total buds on clipped stems and their capacity to activate in response to tissue loss. for example, active buds represented 76.8% of total buds on stems clipped in june and only 53.6% of total buds on controls. thus, stems clipped in june had approximately the same number of active buds as controls despite a reduction in the total number of buds available. therefore, it appears that aspen can compensate from a single summer browsing event during the following spring through the activation of dormant buds. if we relate the number of active buds to the production of new leaf+shoot mass for both treatment and control individuals we find some correlation (fig. 3; we did not test differences between clippings, but illustrate individual trends for the sake of interest). our results indicated a somewhat positive relationship between the number of active buds and production of leaf mass. for single browsing events, a stem’s ability to maintain the required number of active buds to maximize growth does not seem to reduce plant productivity. it is possible that repeated browsing events on the same stems could eventually hamper the tree’s ability to compensate for tissue losses and decrease new shoot production by reducing the availability of meristems. while not evaluated, this negative feedback on vertical growth could have other beneficial effects for the browser (e.g., shoots and leaves produced in the following year might remain within reach of moose). when we compared the response of plant units and individual plants to damage from simulated browsing, we found similar responses. clipped stems had significantly fewer mean buds per stem than the controls; similarly, bergstrom and danell (1987) found an overall reduction in the mean number of buds per tree on clipped individuals. as well, clipped individual stems in our experiment produced the same leaf+shoot biomass as unclipped stems. defoliation of long shoots on individual birch (betula pendula) during the summer resulted in lower leaf biomass on defoliated trees; however, total leaf biomass produced during the season was about the same on both treated and untreated individuals (bergstrom and danell 1995). although we did not see a difference in the production alces vol. 45, 2009 carson et al.– effects of simulated browsing on aspen 107 of new leaf+shoot biomass between clipping treatments, clipping at different times of the growing season can produce variable levels of biomass production as compared to unclipped stems. thorne et al. (2005) found that the frequency of clipping alone had no significant effect on biomass, rather, it was specific combinations of seasonal clipping that produced the highest variation. we suggest further investigation into the relationships among meristem availability, height-specific browse production, and aspen’s ability to compensate for tissue loss, specifically with respect to the influence of varying intensity and frequency of browsing events. related research has identified activation from bud dormancy as a basic component of compensatory response within plants (tuomi et al. 1994), but as with our study, has been tested only within the scope of a single browsing event. stevens et al. (2008) found that the response of aspen to herbivory was dependent on soil nutrient conditions; we presumed that soil conditions were reasonably consistent within our relatively small study site. a more detailed approach may be required to observe aspen response to repeated and variable levels of browsing intensity. furthermore, the relationship between stem volume and number of buds should be studied over a variety of branch sizes to better understand the general characteristics governing morphometric responses and browse production in aspen stems, as well as stems of other browse species. palatability and nutritional differences between compensatory growth of clipped aspen stems versus unclipped stems is also of interest. moose are known to select for compensatory shoots that grow from plants that have been browsed or cut (danell et al.1985), and appear to select for shoots based on the season of cutting (alpe et al. 1999). presumably, nutritive quality varies depending upon the season of browsing (rea and gillingham 2001), however, such responses are unmeasured in aspen. we did not find distinct differences in shoot/leaf production between clipping treatments as we did between controls and clipped stems. however, we did not assess whether our clipped samples included only new (current year) or a combination of new and old growth. in retrospect, accounting for whether we clipped new or old growth might have helped us discern any effects associated with new and older growth, and possible interactions with time of clipping. we recommend that similar research account for the age of clipped growth as opposed to clipping indiscriminately at the diameter of an average bite. acknowledgments the authors would like to thank a. kantakis and d. hoekstra for their assistance with the clipping trials, and a. skoblenick for volunteering his time during the harvest period. j. orlowsky and s.storch were of great assistance with the experimental design, and maintenance and operation of the growth chamber in the enhanced forestry lab at unbc. we are grateful to dr. p. jackson for providing modeled weather data specific to the study area. references alpe, m. j., j. l. kingery, and j. c. mosley. 1999. effects of summer sheep grazing on browse nutritive quality in autumn and winter. journal of wildlife management 63: 346-354. atmospheric environment service. 1993. canadian climate normals 1961-1990 british columbia. canadian climate program, atmospheric environment service, environment canada, downsview, ontario, canada. bergström, r., and k. danell. 1987. effects of simulated winter browsing by moose on morphology and biomass of two birch species. journal of ecology 75: 533-544. _____, and _____. 1995. effects of simulated browsing by moose on leaf and shoot effects of simulated browsing on aspen – carson et al. alces vol. 45, 2009 108 biomass of birch, betula pendula. oikos 72: 132-138. carson, a. c., r. v. rea, and a. l. fredeen. 2007. extent of stem dieback in trembling aspen (populus tremuloides) as an indicator of time-since simulated browsing. journal of rangeland ecology and management 60: 543-547 danell. k. 1983. shoot growth of betula pendula and b. pubescens in relation to moose browsing. alces 18: 197-209. _____, k. huss-danell, and r. bergstrom. 1985. interactions between browsing moose and two species of birch in sweden. journal of ecology 66: 1867-1878. gadd, m. e., t. p. truman, and t. m. palmer. 2001. effects of simulated shoot and leaf herbivory on vegetative growth and plant defense in acacia drepanolobium. oikos 92: 515-521. haukioja, k. e., and k. huss-danell. 1997. morphological and chemical responses of mountain birch leaves and shoots to winter browsing along a gradient of plant productivity. ecoscience 4: 296-303. honkanen, t., and e. haukioja. 1994. why does a branch suffer more after branchwide than after tree-wide defoliation? oikos 71: 441-450. lieffers, v. j., s. m. landhausser, and e. h. hogg. 2001. is the wide distribution of aspen a result of its stress tolerance? june 13-15, 2000. rocky mountain research station publications, grand junction, colorado. fort collins, colorado, usa. meteonorm 4.0 – global meteorological database for solar energy and applied climatology. meteotest, 2000. pratt, p. d., m. b. rayamajhi, t. k. van, t. d. center, and p. w. tipping. 2005. herbivory alters resource allocation and compensation in the invasive tree melaleuca quinquenervia. ecological entomology 30: 316-326. rea, r. v., and m. p. gillingham. 2001. the impact of the timing of brush management on the nutritional value of woody browse for moose alces alces. journal of applied ecology 38: 710-719. renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behaviour. pages 403-439 in a.w. franzmann and c.c. schwartz, editors. the ecology and management of the north american moose. smithsonian institution press, washington d.c., usa. statsoft inc. 2005. statistica [data analysis software system] version 6.0. statsoft inc., tulsa, ok, usa. stevens, m. t., e. l. kruger, and r. l. lindroth. 2008. variation in tolerance to herbivory is mediated by differences in biomass allocation in aspen. functional ecology 22: 40-47. thorne, m. s., p. j. meiman, q. d. skinner, m. a. smith, and j. l. dodd. 2005. clipping frequency affects canopy volume and biomass production of planeleaf willow (salix planifolia var. planifolia prush). journal of rangeland ecology and management 58: 41-50. tuomi, j., p. nilsson, and m. astrom. 1994. plant compensatory responses: bud dormancy as an adaptation to herbivory. ecology 75: 1429-1436. walker, a. d. b., d. c. heard, v. michelfelder, and g. s. watts. 2006. moose density and composition around prince george, british columbia. british columbia ministry of environment. final report for forests for tomorrow 2914000. williamson, v. g., and j. a. millburn. 1995. cavitation events in cut stems kept in water: implications for cut flower senescence. scientia horticulturae 64: 219-232. zar, j. h. 1999. biostatistical analysis. 4th ed. prentice hall, saddle river, new jersey, usa. alces vol. 34 (1), (1998) i patrick d. karns, 73, son of the late patrick and ruby karns, of salem, oregon died on 27 december, 2009. he attended northwestern college and michigan state earning a b.s., worked early on as a wildlife biologist for the michigan department of natural resources, and spent most of his career with the minnesota department of natural resources, first in ely and then grand rapids where he was group leader of the research unit. he was married to ruth meyer and they made their retirement home in salem, oregon. in addition to his many years of moose and deer research, pat enjoyed trains, flying kites, and travel. he is survived by his wife, ruth, daughters, kathleen lederle, mary elizabeth brennan, and bridget rossman, son patrick b. karns, and a number of grandchildren. pat was an annual attendee of both north american and international moose conferences until his last years when severe health issues finally prevented his active participation. he chaired the founders a tribute to pat karns and al elsey by vince chricton conferences, participated on and organized ad hoc committees, and was a regular presenter until his last meeting in whitefish, montana, 2007. pat will be remembered for his dry sense of humor even in failing health, his love of family, life, dancing, dedication to wildlife, and his consummate professionalism to help anyone within or outside his jurisdiction. charles allan (al) elsey, former regional biologist for the ontario ministry of natural resources, passed away on 28 november, 2009 at the age of 92 in burnaby, british columbia. he is survived by his wife of 62 years, margaret, their three daughters, eight grandchildren, and two great grandchildren; he was a quiet, thoughtful, and much loved father. his sense of humor delighted all, even in his years of failing health. al completed normal school (teacher’s training) and briefly taught in one-room schoolhouses in rural saskatchewan. he subsequently enrolled at the university of saskatchewan where he alces vol. 34 (1), (1998) ii completed a m.s. in fisheries biology, and then worked 37 years for the ontario ministry of natural resources (formerly department of lands and forests), both in small communities in northern ontario and for 17 years as a regional biologist in thunder bay. al was also a recognized leader in the figure skating world where he volunteered countless hours, both as a member and president of the port arthur figure skating club and as a competition judge. he chaired the local organizing committee for the 1988 national championships, effectively demonstrating that local support could make such events financially viable in small communities. pat karns was well recognized in the field of wildlife management and research during his 34-year career with the minnesota department of natural resources. he made an indelible mark on deer and moose population management, playing a major role in recovering minnesota’s deer herd in the 1970s, establishing minnesota’s first moose hunt in 1971, and aiding in management of manitoba’s caribou. his research skills and leadership resulted in the reorganization of research in the minnesota conservation department, and pat was appointed group leader of the newly formed forest wildlife populations and research group in 1968. pat moved to the carlos avery wildlife area near forest lake in 1971, where his research focused on deer and moose physiology. pat’s interest in physiology and wildlife diseases lead to the collection of age, sex, morphometric, reproductive, and physiological data during the first and subsequent moose hunts in minnesota. recognizing the need for effective planning, he focused his late professional years on developing “long range management plans” for fish and wildlife populations and their habitats. pat authored and co-authored dozens of professional and popular articles receiving numerous accolades. al elsey was known internationally for his leadership in wildlife conservation. though he was highly valued for his management and leadership skills, his first love was fieldwork. some of his early research involved the aurora trout in numerous lakes in the swastika area of ontario. in the early 1960s he moved from swastika to be a district biologist in fort francis; subsequent “professionalizing” within the ontario department of lands and forests dictated that all biologists be elevated to fish and wildlife supervisor, thus, al moved to thunder bay in 1963 as supervisor of the port arthur district. like many true leaders, he spent much of his career as a mentor, guiding both young biologists and programs. in his waning years, some of his favorite memories were of his work with first nation communities in the far north. al was from a professional generation with few women, however, as a true pioneer and leader, he recognized the value of women in the workplace, hiring and supporting many female biologists and conservation officers, and likewise encouraging his daughters to pursue careers of their choice. pat karns and al elsey were truly the founders and impetus of the north american moose conference, which was spawned from a series of their phone calls in the early 1960s. al was big on interagency cooperation and arranged their initial meeting to discuss moose, after which our annual conferences began. their first formal meeting was in fort francis, ontario, and with their interest sparked by randolph peterson’s book the north american moose, early discussions were pretty basic – what did north american biologists know about moose? thus, the first official moose meeting was arranged and held in st. paul, minnesota on march 7, 1963 and again in march 1964. to ensure pat’s continued participation and interest, al went so far as to grant pat a free ontario moose hunting license! the initial agendas changed little: how to count/estimate a moose population, the impact of hunting on populations, and “moose sickness” or brainworm. subsequent meetings alces vol. 34 (1), (1998) iii were in winnipeg, edmonton, alaska, kamloops, saskatoon, and the 8th in thunder bay in 1972. early on it was unclear whether to call it a workshop or conference. pat insisted it be a workshop as such terminology would make it easier for more to attend; henceforth, it was the “north american moose conference and workshop.” not surprisingly, both were the first recipients of the distinguished moose biologist award given at the annual conference in 1981 in thunder bay, ontario. i urge you to read pat’s 2000 paper “reminiscences and a bit of moose conference history” in alces in which he documents the history of the moose group that began as a special interest group within the great lakes deer group and northeastern chapter of the wildlife society in the early 1960s. pat ended his 2000 paper with this note he received from al on may 3, 1999: “it is wonderful to know that moose still get the associated scientists together. such a magnificent animal deserves it. congratulations, keep up the good work.” the north american and international moose conferences, the scientific journal alces, the “moose bible” ecology and management of the north american moose, the distinguished moose biologist award, the order of alces, student awards, senior travel grant, and the camaraderie and professionalism found at the annual conferences are all part of the rich history and heritage of pat karns and al elsey. our work reflects their lives as scientists, pioneers, consummate professionals, and gentlemen. their legacy is the foundation of modern-day moose management in north america. all moose biologists share in the responsibility of protecting their legacy. f:\alces\supp2\pagema~1\rus 28s alces suppl. 2, 2002 volokitin and kosinskaya – holocene ungulates 127 forest ungulates found in holocene archaeological materials from the european northeast1 alexander v. volokitin2 and lubov l. kosinskaya3 2institute of history and archaeology, ural division of the russian academy of sciences, 620026, ecaterinburg, russia; 3ural state university, 620083, ecaterinburg, lenin prospect, 51, russia abstract: the ungulate fauna present in mesolithic, neolithic, and eneolithic age sites of the pechora and vychegda basins were considerable. these sites differed in archaeological culture and in the type of settlement they represented. established sites included summer and winter camps, permanent settlements, and temporary campsites. in spite of many changes in natural conditions, moose (alces alces) were utilized consistently throughout the early and middle holocene. the significance of this trend for the economy was that it affected the ideology of the human population. alces supplement 2: 127-130 (2002) key words: archaeology, beaver, moose, holocene, hunting, osteology, pechora river, vychegda river the early and middle holocene (from mezolithic to eneolithic ages) was a time when moose meat prevailed in the diets of primitive people of the forest (taiga) zone, including northeastern europeans. the subsistence strategy of ancient hunters was often associated with moose. the great importance of moose to these ancient people was reflected in their art and tool craftsmanship, including drawings and engravings on rocks, and articles made of bone, antler, and wood. moose and the ideas associated with this species took a leading role in the spiritual culture of these ancient people (stolar 1983). to confirm this we may refer to the works of investigators and ethnographers. good analyses of such previous work were conducted by okladnicov (1950) and studzitskaya (1981). in the paleolithic age the depictions of moose were very few. the known depictions were located at sites dated to the final s t a g e o f t h e u p p e r p a l e o l i t h i c a g e (melnichuc and pavlov 1987). in a number of cases the interpretations of these depictions differed, and the artwork may have represented either moose or saiga antelope (arts and deeben 1987). moose were not the preferred catch for paleolithic hunters, and the bones of moose found at paleolithic sites were very few in number (tromnau 1987). paleogeographic and archaeological data may provide some explanation of a change in the primary species hunted by people. changes in the material culture of the people may have broadened their adaptive possibilities. materials and methods most holocene sites of the european northeast (vychegda and pechora river basins) were located on the sandy edges of pine forests. the soil conditions at these sites did not provide for adequate preservation of organic remains. bones obtained from these sites were in extremely bad condition and were represented as small charred fragments. in spite of the poor editor's note: this manuscript was submitted without references and due to difficulties contacting the authors the manuscript was published without them. holocene ungulates – volokitin and kosinskaya alces suppl. 2, 2002 128 condition of most bones, we managed to accumulate some osteological material (osteological analyses were conducted by p. a . k o i n t s e v , i n s t i t u t e o f e c o l o g y , ecaterinburg). this allowed us to draw some conclusions about regional specificity of information concerning the role of moose in the lifestyle and culture of ancient people. in addition to osteological materials, we used topographical data and the composition of implements found at the sites (table 1). archaeological description (chronologically and culturally) of these sites, individually and as groups, is as follows. archaeological site descriptions in collections from the parch 2 and 3 sites (vychegda river), a series of arrowheads on blades and inserts were of special interest. there were also different burins and side–scrapers present. the difference between the material culture of parch 2 and 3 sites and other known mesolithic sites of the region should be pointed out. the ages of the sites were derived from pollen analyses, radiocarbon dating, and from the complex of paleogeographic data (volokitin and kovalenko 1988, volokitin and kosinskaya 1989). u s t – u k h t a 1 a n d t h e l e c k – l e s a mesolithic sites (izhma river, pechora river tributary) referred to another archaeological culture based on their stone implements. at the ust–ukhta 1 site, a compact concentration of material was discovered that may be the remains of a ground dwelling. tools, blades, and inserts were represented in collections from this site. among numerous findings at the leck–lesa site, there were knife kits with large blades, double truncated bladeletes, points, and side–scrapers. the burins were few. several sites in the basins of the lower vychegda and izhma rivers were joined into 3 cultural groups and dated to the early neolithic age based on the character of stone implements and ceramics. the first group included the sites kochmas a and chernaya vadya on the lower vychegda river. findings from the kochmas a site included only flint implements. the chipped stone technique left blade–flakes. knives and scrapers prevailed in tool kits. many tools were used for bone and antler processing, although no bone articles were found. stone projectile points were not discovered, but through the presence of blade–inserts, we inferred the existence of manufactured points. flint implements of the chernaya vadya site were made according to blade technology. tools prevalent at this site were knives, scrapers, and inserts from short rectangular segments of blades. there were lateral points that may have served as edge inserts for composed heads of a hunter’s weapon. fragments from 2–3 vessels were also found in the dwellings. kochmas a and chernaya vadya are similar to the upper volga (early neolithic) sites and date back to the fourth and fifth millennia bc. the second group from the early neolithic age was represented by the revyu i site. flint manufacture included a combination of blade and flake production. side– scrapers and cutting tools were common at this site. small trapezium inserts from chips were prevalent in kits at this site, but large symmetrical trapeziums of average height were few. ceramics were not discovered. hypothetical dating was to the second half of the fourth millennium bc. t h e t h i r d g r o u p i n c l u d e d t h e chernoborskaya iii (izhma river) and niremka iii (vym river) sites. the flint implements of chernoborskaya iii were manufactured using the large blade and flake technique. narrow triple–edged arrowheads, knives on blades, and side–scrapers were numerous. several ceramic vessels were found. at niremka iii, flint alces suppl. 2, 2002 volokitin and kosinskaya – holocene ungulates 129 table 1. archaeological sites in the komi republic with faunal remains of ungulates. age sites type of topography fauna dating1 settlement mesolithic parch 2 2 ground flood–plain, beaver 32/13 bo–1 dwellings 5 m moose 3/? parch 3 site–work– flood–plain, beaver 7/1 bo–1 shop 5 m ust–ukhta 1 ii terrace, moose 1/1 at–1 19 m leck–lesa ground ii terrace beaver 3/3 at–1 excavation–1 15–16 m moose 1/1 dwelling bear 1/1 leck–lesa ground ii terrace, moose/? at–1 excavation–2 15–16 m dwellings early kochmas a winter ii terrace, moose 26/2 at–2 neolithic dwelling 8 m beaver 7/2 at–3 (with work– fox 1/1 (middle– shop) bird 1/1 southern taiga) chernaya dwelling ii terrace moose 1/1 at–2 vadya round–the– beaver 2/2 at–3 excavation–2 (middle– remains southern taiga) early revyu 1 winter ii terrace moose 42/? at–3 neolithic excavation 2 dwelling? beaver 1/1 (southern reindeer 6/? taiga) bear 4/? wolf 2/? chernoborskaya site ii terrace, moose 1/? at–3 iii 14 m reindeer 6/? neolithic niremka iii hunter’s ii terrace, moose 10/2 at–3 camp 11 m sb–1 (middle– southern taiga eneolithic niremka i dwellings and ii terrace, reindeer 1/1 sb–2 household 7 m (southern constructions taiga) niremka i dwellings– ii terrace, reindeer 1/1 sb–2 winter? 7 m beaver 1/1 niremka i dwelling– ii terrance, beaver 1/1 sb–2 winter? 7 m 1 according to holocene subdivisions by hotinksy (1977). 2 number of identified bones. 3 number of individuals. holocene ungulates – volokitin and kosinskaya alces suppl. 2, 2002 130 manufacture was practically nonexistent. the sites of this group hypothetically date back to the third and fourth millennia bc. three dwellings at niremka i were indicative of the eneolithic age. ceramics and stone implements were concentrated inside the dwellings; an indication of the winter character of the dwellings. the ceramics were not numerous (3–8 vessels per dwelling). flint implements were made using the flake technique and each dwelling contained indications of mass flint manufacture. side–scrapers prevailed in tool kit knives. small arrowheads of leaf–like shape and dart–heads were found. all 3 dwellings were different in age, were associated with the choinovtinskaya culture, and dated back to the latter half of the third and the first half of the second millennia bc. in summary, the stone implements present at all the sites mentioned above show that hunting was the occupational trend of ancient populations in this region during the mesolithic through eneolithic ages. conclusions in spite of differences in culture and natural conditions, the household activities of ancient people of the european northeast were based on hunting forest ungulates and beaver throughout a time period of several thousand years. this trend was characteristic of other taiga regions. the regional role of moose in the household activities and culture (ideology) of the population in the holocene age became evident based on depictions of this animal in the vis peat bog of sindor lake (vychegda river basin). the most realistic depiction is on a hunter’s ski. evidently, this was a case of the manifestation of hunting magic (burov 1968). the household (seasonal) activities of the ancient people of the vychegda and pechora basins were based on hunting large forest ungulates and beaver, in addition to fishing and gathering. based on ethnographical data, we conclude that evolution lead to the development of a population of taiga hunters and fishermen characterized by relative mobility throughout the annual seasonal cycle and by economical use of natural resources. this provided for stable development of the population. the methods and means of hunting evidently changed and improved over time. we note that investigators usually assume the use of passive hunting methods for large taiga ungulates during the mezolithic through eneolithic ages (savvateev and vereschagin 1983). the real evidence of the use of such implements in the european northeast were the findings at the vis peat bog. gigantic self–shooting bows made of small tree trunks were present in the collection of wooden articles found there. these bows might have been used for hunting large ungulates (burov 1973). rock carvings have depicted scenes of collective winter moose hunting by humans using skis (savvateev 1973). from the analysis of rock carvings, it was also presumed that moose hunting was done from boats at river crossings using horn axes. however, this method was considered to be more characteristic of reindeer hunting. it should be pointed out that in the region of the parch 1– 3 sites, moose still cross the vychegda river today. 39 revisiting the recruitment-mortality equation to assess moose growth rates ian w. hatter nature wise consulting, 308 uganda avenue, victoria, british columbia v9a 5x7, canada abstract: hatter and bergerud (1991) developed a recruitment-mortality (r-m) equation to estimate the annual finite rate of change (λ) in a moose (alces alces) population from a single estimate of calf recruitment and adult mortality. i present and assess an alternative formulation of the r-m equation and compare it with the original. a modification to the r-m equations is provided to accommodate early to mid-winter composition surveys where recruitment is measured when calves are less than 1 year-of-age. an example with the modified r-m equation illustrates estimation of λ for the female component of two moose populations under recent study in british columbia, canada. due to potential biases with estimating recruitment and mortality rates, the calculation of λ with the r-m equation should be verified with periodic density surveys whenever possible. the r-m equation is most useful for estimating λ when moose density surveys are not feasible or an estimate of the adult survival rate is available. alces vol. 56:39–47 (2020) key words: alces alces, finite rate of change, mortality rate, population growth rate, recruitment mortality equation, recruitment rate, survival rate hatter and bergerud (1991) developed a recruitment-mortality (r-m) equation to estimate the finite rate of change or growth rate (λ) of an ungulate population from estimates of juvenile recruitment (r) and adult mortality (m) rates where λ = (1−m)/(1−r). this equation has been used commonly with populations of moose (alces alces) (e.g., gasaway et al. 1992, boertje et al. 1996, kunkel and pletscher 1999, hayes et al. 2000, kuzyk et al. 2019b, severud et al. 2019), caribou (rangifer tarandus) (e.g., bergerud and elliott 1986, seip and cichowski 1996, mcloughlin et al. 2003, hebblewhite et al. 2007, sorensen et al. 2008, latham et al. 2011, decesare et al. 2012, hervieux et al. 2013, 2014, serrouya et al. 2017), elk (cervus elaphus) (e.g., kunkel and pletscher 1999, devore et al. 2018), black-tailed deer (odocoileus hemionus columbianus) (e.g., hatter and janz 1994), and white-tailed deer (odocoileus virginianus) (e.g., kunkel and pletscher 1999, patterson et al. 2002). refinements have been made by several authors to improve its utility as an ungulate population assessment tool (decesare et al. 2011, hervieux et al. 2013). the primary utility of the r-m equation is that it calculates λ from a single estimate of recruitment provided that the adult mortality rate is known. it is particularly appropriate when there are few other cost-effective alternatives. one drawback of using the r-m equation with moose is that recruitment is commonly measured from composition surveys conducted during early or mid-winter before calves are recruited into the adult population at 1 year-of-age (hatter and bergerud 1991). this is because aerial spring surveys are generally impractical due to lack of snow-cover, poor visibility, revisiting the r-m equation. – ian w. hatter alces vol. 56, 2020 40 and dispersal of moose from winter ranges (gasaway et al. 1986). the purposes of this paper were to review the original r-m equation (hatter and bergerud 1991), compare it with an alternative formulation, identify a modification that accommodates moose composition surveys before calves are 1 year-of-age, and provide an example of its use. the r-m equation the r-m equation enables the calculation of λ from annual recruitment and adult mortality rates. this equation expressed as a difference equation (serrouya et al. 2017) is: nt+1 = nt + rnt+1 ‒ mnt (1) where r is the proportion of juveniles at the end of their first year of life (i.e., year t1) and m is the adult (1+ year-old males and females) mortality rate during the year (i.e., from year t0 to t1). rearranging eq.1 and solving for λ = nt+1/nt yields: λ = (1 ‒ m)/(1 ‒ r). (2) as 1–m is the adult survival rate, eq. 2 also equals: λ = s/(1 ‒ r) (3) where s is the annual probability of adult survival. here, i refer to eq. 3 as the type 1 r-m equation. if hunting occurs, then s must account for both non-hunting (mn) and hunting (mh) mortality rates: s = (1‒ mh) × (1‒ mn) (4) alternative formulations of the r-m equation where r is the juvenile:adult ratio and adults refer to both sexes combined (guthery and shaw 2013, devore et al. 2018) are: nt+1 = nts(1+r) (5) and λ = s(1+r) (6) i refer to eq. 6 as the type 2 r-m equation. the r-m equation has also been used to estimate λ for the female segment of the population as growth rates are largely determined by females (caughley 1977) and because many studies focus on adult female mortality rates (hervieux et al. 2013, 2014, kuzyk et al. 2019a, 2019b). equations 7–12 apply specifically to females, although they are easily modified for males or both sexes combined (hatter and bergerud 1991). recruitment in the type 1 model is estimated as: = × × + r j j c /f c /f 1 (7) and in the type 2 model as r = c × j/f (8) where c is the proportion of recruited juveniles that are female and j/f is the ratio of juveniles/1+ year-old females. calf:cow ratios (j/f) for moose are usually measured in early or mid-winter when aerial survey conditions are optimal for determining herd composition and abundance (gasaway et al. 1986). however, these surveys do not provide an accurate measure of recruitment at 1 yearof-age since calves die at a higher rate than cows during winter (ballard et al. 1991, kuzyk et al. 2019b). failing to account for the differential winter mortality between calves and cows results in a biased estimate of r, which by definition is measured when calves are 1-year-ofage. calf:cow ratios from these surveys alces vol. 56, 2020 revisiting the recruitment-mortality. – ian w. hatter 41 must be adjusted to account for this differential mortality. the adjustment of r (r' ) for the type 1 model is: = × × × × + r' c j f sf c j f sj sj / ( / ) w w w (9) and the adjustment for the type 2 model is: = × ×r' c j f sj sf/ /w w (10) where sjw is the winter calf survival rate and sfw is the winter cow survival rate. these survival rates must be measured from the end of the winter survey to just before calves become 1 year-of-age, and is typically accomplished through telemetry studies of radio-collared animals (pollock et al. 1989). the finite rate of change for the type 1 model is then estimated as: λ = sf/(1 ‒ r' ) (11) and for the type 2 model as: λ = sf(1 + r' ) (12) where sf is the annual cow survival rate. figure 1 illustrates estimates of change in λ when there is differential overwinter mortality between cows and calves (sjw /sfw). for example, λ ranged from 0.98 (sjw /sfw = 0.50) to 1.06 (sjw /sfw = 1.00) when assuming j/f = 0.35, c = 0.5, and sf = 0.90. the above equations do not account for uncertainty in the parameters used to estimate λ. in order to determine the 95% ci for λ in eq. 11 or 12, the se must be measured for j/f, sf, sfw, sjw, and c. a number of researchers including caughley (1977), gasaway et al. (1986), skalski et al. (2005), and pollock et al. (1989) provide methods and examples for making these calculations. fig. 1. contour plot depicting the range of population growth rates (λ) based on plausible ranges in annual cow survival (sf) and differential winter mortality between calves and cows (sjw/sfw). the midwinter calf:cow ratio (j/f) was 0.35, and the calf sex ratio was 50:50. revisiting the r-m equation. – ian w. hatter alces vol. 56, 2020 42 following latham et al. (2011) and hervieux et al. (2013), one method of estimating the 95% ci for λ is to randomly draw from each year’s annual survival and recruitment distributions (i.e., x̅ and se) a large number of times (e.g., 10,000) using monte carlo simulation. survival rates should be drawn from a beta distribution (values range from 0–1) and calf:cow (j/f) ratios from a lognormal distribution (values > 0). for a quick and simple comparison, the normal approximation may be used to determine if two estimates (est) (e.g., j/f, sf, sfw, sjw, or λ) are significantly different using = − + >z est est var est var est ( ) ( ( ) ( )) 1.961 2 1 2 where var = se2 (sinclair et al. 2006). if estimates are not significantly different, they may be merged to produce a more precise estimate of λ using the procedure outlined by sinclair et al. (2006:234). an example factors affecting moose population declines in british columbia, canada are currently being investigated with population dynamics studied intensively (kuzyk et al. 2019a, 2019b). estimates of j/f, sf, sfw, and sjw were available for 3 consecutive years in the bonaparte study area (2016–17 to 2018–19) and 2 consecutive years in the prince george south study area (2017–18 to 2018–19) (table 1), and the proportion of 8-month-old female calves (c) was documented from the sex ratio of radio-collared calves from 2016–17 to 2019–20. calf sex ratios were not significantly different from 50:50 in either study area (bonaparte: χ2 = 0.45, p = 0.50, n = 80; prince george south: χ2 = 0.24, p = 0.62, n = 66), so c was set equal to 0.5. differential winter survival (sjw/sfw) varied from 0.47–0.89 in the bonaparte study area and from 0.78–0.87 in the prince george south area. growth rates based on unadjusted recruitment rates from the hatter and table 1. moose population parameters including calf:cow ratio (j/f), annual cow survival rate (sf), winter cow survival rate (sfw), and winter calf survival rate (sjw) within the bonaparte and prince george south study areas of british columbia, canada; data is from kuzyk et al. (2019). merged refers to combined estimates for 2017–18 and 2018–19. a. bonaparte study area year calf:cow ratio annual cow survival winter cow survival winter calf survival j/f se n sf se n sfw se n sjw se n 2016–17 0.13 0.028 208 0.91 0.057 79 0.96 0.028 52 0.45 0.051 20 2017–18 0.32 0.047 256 0.98 0.022 53 0.96 0.022 52 0.85 0.069 20 2018–19 0.28 0.056 148 0.95 0.026 70 0.95 0.028 63 0.80 0.073 20 merged 0.30 0.036 0.97 0.017 0.96 0.017 0.83 0.050 b. prince george south study area year calf:cow ratio annual cow survival winter cow survival winter calf survival j/f se n sf se n sfw se n sjw se n 2017–18 0.34 0.040 375 0.79 0.049 54 0.90 0.041 40 0.70 0.071 20 2018–19 0.31 0.056 168 0.79 0.052 55 0.85 0.052 47 0.74 0.076 19 merged 0.33 0.033 0.79 0.036 0.88 0.032 0.72 0.052 alces vol. 56, 2020 revisiting the recruitment-mortality. – ian w. hatter 43 bergerud (1991) r-m equation were higher than those based on adjusted recruitment rates (table 2). estimates of λ based on adjusted recruitment rates were identical between the type 1 and type 2 r-m equations. the 95% ci for λ, based on adjusted recruitment rates, ranged from <1.0 to >1.0 for each year in both populations, except for 2017–18 in the bonaparte study area. in this area estimates of λ were lower in 2016–17 than in 2017–18 (z = 2.64, p = 0.008) and also between 2016–17 and 2018–19 (z = 1.82, p = 0.068). growth rates were not different between 2017–18 and 2018–19 (z = 1.00, p = 0.32). merging these years produced a more precise estimate of λ (table 2). growth rates were similar in the prince george south area between 2017– 2018 and 2018–19 (z = 0.034, p = 0.97); merging these years also produced a more precise estimate of λ. estimates of growth rates from density surveys in the bonaparte study area during 2013–18 (λ = 0.97) were within the range calculated from recruitment and mortality (λ = 0.93–1.12), and estimates of population growth from density surveys in the prince george south study area during 2012–17 (λ = 0.91) were comparable to the r-m equation (λ = 0.90). discussion the original formulation of the r-m equation developed by hatter and bergerud (1991) requires an adjustment to the recruitment rate when calf:cow ratios are measured in early-to-mid winter. this adjustment accounts for differential overwinter survival of calves and adults which may be made using either the type 1 or type 2 r-m equations. failure to account for differential survival results in an overestimate of λ. users of the r-m equations may prefer the type 2 model since most moose biologists define calf recruitment as the calf:cow ratio rather than % calves and because calculations are simpler. table 2. moose recruitment (r), adjusted moose recruitment (r’), and population growth rates (λ) with monte carlo simulated confidence intervals (95% ci) for the type 1 and type 2 r-m equations within the bonaparte and prince george south study areas of british columbia, canada. merged refers to combined estimates for 2017–18 and 2018–19. a. bonaparte study area year r-m equation1 type 1 r-m equation type 2 r-m equation type 1 & 2 r (eq. 7) λ (eq. 3) r’ (eq. 9) λ (eq. 11) r’ (eq. 10) λ (eq. 12) 95% ci 2016–17 0.061 0.97 0.030 0.93 0.030 0.93 0.79–1.02 2017–18 0.138 1.13 0.124 1.12 0.142 1.12 1.04–1.18 2018–19 0.123 1.08 0.105 1.06 0.118 1.06 0.98–1.14 merged 0.132 1.11 0.116 1.09 0.131 1.09 1.04–1.14 1estimated using the original r-m equation from hatter and bergerud (1991). b. prince george south study area year r-m equation1 type 1 r-m equation type 2 r-m equation type 1 & 2 r (eq. 7) λ (eq. 3) r’ (eq. 9) λ (eq. 11) r’ (eq. 10) λ (eq. 12) 95% ci 2017–18 0.145 0.93 0.117 0.90 0.133 0.90 0.77–1.01 2018–19 0.134 0.91 0.119 0.90 0.135 0.90 0.76–1.02 merged 0.142 0.92 0.119 0.90 0.135 0.90 0.81–0.98 1 estimated using the original r-m equation from hatter and bergerud (1991). revisiting the r-m equation. – ian w. hatter alces vol. 56, 2020 44 there are 4 primary methods currently used to estimate λ for moose populations. the growth rate may be measured from 2 or more population surveys over time (van ballenberghe 1983), from calf recruitment and adult mortality rates (hatter and bergerud 1991), from survival and fecundity schedules (van ballenberghe 1983), or from fitting population models to multiple sources of observed data (kuzyk et al. 2018). the primary advantage of the r-m model is that it provides an estimate of λ from a single, late winter herd composition survey when the annual adult mortality rate is known. it may also be preferred where moose density surveys are not feasible, such as in densely forested areas or where moose occur at very low density (severud et al. 2019). guthery and shaw (2013) note that the r-m equation is a tautology and thus inevitably true and does not require empirical verification. however, survey timing and survey bias can affect the accuracy of λ calculations. serrouya et al. (2016) compared population growth rates for caribou between abundance surveys and the r-m equation and found that the r-m equation overestimated λ compared to survey-based λ. they proposed 3 possible reasons including: 1) measurement of recruitment at 10-months of age rather than as 1 year-old, 2) adult survival estimates that are biased towards more mature animals, and 3) errors in herd composition surveys (e.g., sightability differences between barren females and those with offspring). they found that the r-m equation explained 60% of the variation in survey-based λ and was a better predictor of this parameter compared to other approaches. users of the r-m equation should ensure that herd composition surveys are representative of the population and measure recruitment just before calves become 1 year-of-age. alternatively, the modified r-m equation may be used to account for differential winter mortality between calves and adults when survival estimates are available. in addition, users need to be careful to ensure that estimates of adult survival from radio-collared animals include a representative sample of their standing age distribution. this may require estimating survival rates over multiple years. in alberta, the population trend for boreal caribou was monitored using the r-m equation based on annual surveys of recruitment and an ongoing, intensive ratio telemetry program with adult females (hervieux et al. 2013). annual changes in λ were used to calculate realized population change, which were the successive product of λ calculated from the first year of monitoring up to and including the most recent year’s λ calculation. because this approach may compound errors in relative abundance over time (serrouya et al. 2017), periodic surveys of absolute abundance should be performed to help validate long-term population trends based on the r-m equation. the r-m equation may provide relatively large confidence intervals for λ. in order to improve the accuracy and precision of λ, it is important to ensure that samples of recruitment and mortality rates are representative and that sample sizes are sufficiently large. decesare et al. (2012) employed elasticity and life-stage simulation analysis for a woodland caribou population in alberta and found that adult female survival and recruitment rates were nearly equivalent drivers of population growth. this suggests that increased sampling to improve precision of λ should be directly towards both population parameters. alternatively, annual estimates of j/f, sf, sfw, and sjw may be merged over multiple years to improve precision, providing these parameters are not significantly different between years. biologists have been tasked with ensuring moose objectives for both conservation alces vol. 56, 2020 revisiting the recruitment-mortality. – ian w. hatter 45 and sustainable use are met. these challenges have recently intensified with broad scale moose declines occurring in parts of north america (timmermann and rodgers 2017, kuzyk et al. 2018). the r-m equation provides a simple and relatively inexpensive method to rapidly assess moose population trends when adult survival rates are available. however, numerous biases may exist in estimating recruitment and mortality rates that can substantially affect estimates of λ (serrouya et al. 2017, severud et al. 2019). thus, it is important to assess the accuracy and precision of both parameters when calculating λ using the r-m equation and to ensure that over-winter differential mortality between cows and calves is incorporated. if the r-m equation is used to monitor trends over multiple years, then estimates of λ should be validated periodically with surveys of absolute abundance. acknowledgements g. kuzyk, c. proctor, and s. marshall provided access to the british columbia moose data sets used in the example. m. scheideman kindly produced the estimates of cow winter survival rates and provided the winter calf sex ratios. i appreciate the constructive reviews from m. carstensen and 2 anonymous reviewers which greatly improved an earlier draft of this manuscript. this study is dedicated to the late w. a. bergerud who developed the proof for the original r-m equation. references ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. bergerud, a. t., and j. p. elliott. 1986. dynamics of caribou and wolves in northern british columbia. canadian journal of zoology 64: 1515–1529. doi: 10.1139/z86-226 boertje, r. d., p. valkenburg, and m. e. mcnay 1996. increases in moose, caribou and wolves following wolf control in alaska. journal of wildlife management 80: 474–479. doi: 10.2307/3802065 caughley, g. 1977. analysis of vertebrate populations. john wiley & sons, london, united kingdom. decesare, n. j., m. hebblewhite, m. bradley, k. smith, d. hervieux, and l. neufeld. 2012. estimating ungulate recruitment and growth rates using age ratios. journal of wildlife management 76: 144–153. doi: 10.1002/jwmg.244 devore, r. m., m. j. butler, m. c. wallace, and s. g. liley. 2018. population dynamics model to inform harvest management of a small elk herd in central new mexico. journal of fish and wildlife management 9: 531–544. doi: 10.3996/012018-jfwm-008 gasaway, w. c, r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larson. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120: 1–59. _____, s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska. number 22. institute of arctic biology, fairbanks, alaska, usa. guthery, f. s., and j. h. shaw. 2013. density dependence: applications in wildlife management. journal of wildlife management 77: 33–38. doi: 10.1002/jwmg.450 hatter, i. w., and w. a. bergerud. 1991. moose recruitment, adult mortality and rate of change. alces 27: 65–73. _____, and d. w. janz. 1994. apparent demographic changes in black-tailed deer associated with wolf control on northern vancouver island. canadian revisiting the r-m equation. – ian w. hatter alces vol. 56, 2020 46 journal of zoology 72: 878–884. doi: 10.1139/z94-119 hayes, r. d., a. m. baer, u. wotschikowsky, and a. s. harestad. 2000. kill rate by wolves on moose in the yukon. canadian journal of zoology 78: 49–59. doi: 10.1139/z99-187 hebblewhite, m., j. whittington, m. bradley, g. skinner, a. dibb, and c. a. white. 2007. conditions for caribou persistence in the wolf-elk caribou systems of the canadian rockies. rangifer 17(special issue): 79–91. doi: 10.7557/2.27.4.322 hervieux, d., m. hebblewhite, n. j. decesare, m. russell, k. smith, s. robertson, and s. boutin. 2013. widespread declines in woodland caribou (rangifer tarandus caribou) continue in alberta. canadian journal of zoology 91: 872–882. doi: 10.1139/ cjz-2013-0123 _____, _____, d. stepnisky, m. bacon, and s. boutin. 2014. managing wolves (canis lupus) to recover threatened woodland caribou (rangifer tarandus caribou) in alberta. canadian journal of zoology 92: 1029–1037. doi: 10.1139/cjz-2014-0142 kunkel, k., and d. h. pletscher. 1999. species-specific population dynamics of cervids in a multi-predator ecosystem. journal of wildlife management 63: 1082–1093. doi: 10.2307/3802827 kuzyk, g, i. hatter, s. marshall, c. procter, b. cadsand, d. lirette, h. schindler, m. bridger, p. stent, a. walker, and m. klaczek. 2018. moose population dynamics during 20 years of declining harvest in british columbia. alces 54: 101–119. _____, c. procter, s. marshall, and d. hodder. 2019a. factors affecting moose population declines in british columbia: updated research design. wildlife bulletin no. b-128. british columbia ministry of forests, lands and natural research operations and rural development, victoria, british columbia, canada. _____, _____, _____, h. schindler, h. schwantje, m. scheideman, and d. hodder. 2019b. determining factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife working report. no. wr-127. progress report: february 2012–april 2019. british columbia ministry of forests, lands and natural research operations and rural development, victoria, british columbia, canada. latham, a. d., m. c. latham, n. a. mccutchen and s. boutin. 2011. invading white-tailed deer change wolf-caribou dynamics in northeastern alberta. journal of wildlife management 75: 204–212. doi: 10.1002/jwmg.28 mcloughlin, p. d., e. dzus, b. wynes, and s. boutin. 2003. declines in populations of woodland caribou. journal of wildlife management 67: 755–761. doi: 10.2307/3802682 patterson, b. r., b. a. macdonald, b. a. lock, d. g. anderson, and l. k. benjamin. 2002. proximate factors limiting population growth of white-tailed deer in nova scotia. journal of wildlife management 66: 511–521. doi: 10.2307/ 3803184 pollock, k. h., s. r. winterstein, c. m. bunck, and p. d. curtis. 1989. survival analysis in telemetry studies: the staggered entry design. journal of wildlife management 53: 7–15. doi: 10.2307/ 3801296 seip, d. r., and d. b. cichowski. 1996. population ecology of caribou in british columbia. rangifer 9: 73–80. doi: 10.7557/2.16.4.1223 serrouya, r., s. gilbert, r. s. mcnay, b. n. mclellan, d. c. heard, d. r. seip, and s. boutin. 2017. comparing population growth rates between census and recruitment-mortality models. journal of wildlife management 81: 297–305. doi: 10.1002/jwmg.21185 alces vol. 56, 2020 revisiting the recruitment-mortality. – ian w. hatter 47 severud, w. j., g. d. delgiudice, and j. k. bump. 2019. comparing survey and multiple recruitment-mortality models to assess growth rates and population projections. ecology and evolution 9: 12613–12622. doi: 10.1002/ece3.5725 sinclair, a. r. e., j. m. fryxell, and g. caughley. 2006. wildlife ecology, conservation, and management. second edition. blackwell publishing, hoboken, new jersey, usa. skalski, j. r., k. e. ryding, and j. j. millspaugh. 2005. wildlife demography: analysis of sex, age, and count data. elsevier academic press, new york, new york, usa. sorensen,t., p. d. mcloughlin, d. hervieux, e. dzus, j. nolan, b. wynes, and s. boutin. 2008. determining sustainable levels of cumulative effects for boreal caribou. journal of wildlife management 72: 900–905. doi: 10.2193/2007-079 timmermann, h. r., and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. van ballenberghe, v. 1983. rate of increase in moose populations. alces 19: 98–117. alces vol. 45, 2009 laaksonen and oksanen – vector-borne nematode in finnish cervids 81 status and review of the vector-borne nematode setaria tundra in finnish cervids sauli laaksonen and antti oksanen finnish food safety authority evira, fish and wildlife health research unit, p.o. box 517, fi-90101 oulu, finland abstract: the filarioid nematode setaria tundra caused an outbreak of peritonitis in finnish semi-domesticated reindeer in 2003-2006. our research group studied the invasion and reservoirs of s. tundra in finnish cervid populations and this paper provides an overview of that research. the outbreak had detrimental effects on reindeer health and may, in part, explain the observed decline of the population of wild forest reindeer (rangifer tarandus fennicus). both range expansion by roe deer, and high summer temperatures that increased vector populations of mosquitoes and gnats and influenced habitat use by reindeer were implicated in the outbreak. we suggest that vector borne parasites will increase in the arctic owing to the effect of global climate change and have consequences for all cervid populations. alces vol. 45: 81-84 (2009) key words: cervid, climate change, filarioidea, population dynamics, reindeer. there is a growing body of literature documenting the expansion of emerging parasites in sub-arctic areas. the potential impact of global warming on shifts in the spatio-temporal distribution and transmission dynamics of vector-borne diseases in domesticated and wild ungulates may be remarkable (hoberg et al. 2008). contemporary finnish studies have revealed an array of filarioid nematodes and associated diseases that appear to be emerging in northern ungulates (laaksonen et al. 2007, nikander et al. 2007, solismaa et al. 2008). for example, members of the genus setaria (filarioidea: onchocercidae) are found in the abdominal cavity of artiodactyls (especially bovidae), equids, and hyracoids. all species produce microfilariae that are present in host blood, and known vectors are haematophagous mosquitoes (culicidae spp.; anderson 2000) and horn flies (haematobia spp.; shol and drobischenko 1973). the filarioid nematode setaria tundra was first described in semi-domesticated reindeer (rangifer tarandus tarandus) in the arkhangelsk area of russia by rajevsky (1928). setaria sp. infections appear to first emerge in scandinavian reindeer in the 1960s. s. tundra was observed initially in northern norway in 1973 where there was an outbreak of peritonitis in reindeer. in the same year, tens of thousands of reindeer died in the northern part of the herding area of finland. severe peritonitis and large numbers of setaria sp. worms were common. however, the incidence of setaria sp. in scandinavian reindeer diminished afterward (laaksonen et al. 2007). according to meat inspection data and clinical reports from practicing veterinarians in finland, the latest outbreak of peritonitis in reindeer started in 2003 in the southern and middle parts of the reindeer herding area. in the province of oulu, the proportion of reindeer viscera condemned due to parasitic lesions identified during meat inspections increased dramatically from 4.9% in 2001 to 47% in 2004; in lapland the increase was from 1.4% in 2001 to 43% in 2006. these increases caused substantial economic loss and increased workload associated with meat processing. the focus of the outbreak moved northward approximately 100 km/yr, and by 2005 only those reindeer in the small, northernmost part vector-borne nematode in finnish cervids – laaksonen and oksanen alces vol. 45, 2009 82 of finland (upper lapland) were free of lesions. during the same period, the peritonitis outbreak was apparently concentrating in the southern area (laaksonen et al. 2007). the causative agent was s. tundra based on morphologic and molecular data. samples of dna sequences of s. tundra parasitising reindeer in northern finland were deposited in genbank under accession number dq097309 (laaksonen et al. 2007, nikander et al. 2007) the prevalence and intensity of s. tundra microfilariae (smf) were higher in reindeer calves than adults; overall prevalence was 42%. the overall smf-prevalences for moose (alces alces), wild forest reindeer (rangifer tarandus fennicus), and roe deer (capreolus capreolus) were 1.4-1.8%, 23%, and 39%, respectively. the focus of microfilaremia in reindeer moved north as it declined simultaneously in the south as the observed peritonitis outbreak lessened. experimentally, reindeer calves infected in their first summer of life had peak microfilaremia in their second summer. captive reindeer were smf positive throughout the year, but smf disappeared from the blood after 2 years. the prepatent period of s. tundra was estimated to be about 4 months, with a life span of at least 14 months (laaksonen et al. 2008a) reindeer calves with heavy s. tundra infection expressed decreased thriftiness, poor body condition, and undeveloped winter coat. in kuusamo, 4603 slaughtered reindeer were examined clinically in 2003-04; meat inspections of diseased reindeer carcasses revealed ascites fluid, green fibrin deposits, adhesions, and live and dead s. tundra nematodes. histopathology indicated granulomatous peritonitis with lymphoplasmacytic and eosinophilic infiltration. no specific bacterial growth was found. no significant impact on ph values of meat or on organoleptic evaluation of meat was found. there was a significant positive correlation between worm counts and the degree of peritonitis, and a negative correlation between the degree of peritonitis and back-fat layer (laaksonen et al. 2007). setaria yehi has been associated with low grade chronic peritonitis in alaskan reindeer (dieterich and luick 1971). s. tundra, in combination with corynebacterium sp., has been associated with mild to severe peritonitis in swedish reindeer (rehbinder et al. 1975). based on the evidence in both ante and post-mortem inspections and histological examinations, our studies (laaksonen et al. 2007, laaksonen et al. 2008b) and historical data indicate that s. tundra can act as a significant pathogen in reindeer. we collected parasite samples from wild cervids in order to monitor the dynamics of s. tundra in nature. about 300 moose, the most abundant wild cervid in the reindeer herding area, were inspected and only a few cases of pre-adult encapsulated s. tundra nematodes were found on the surface of livers. however, no peritonitis was identified (laaksonen et al. 2007), and the prevalence and intensity of smf in 324 moose blood samples within and outside the reindeer herding area were low (1.4% and 1.8%, 1-3 smf/ml blood; laaksonen et al. 2008a). because the moose population in northern finland peaked in 2004-2005, moose are apparently not a suitable reservoir host for the s. tundra haplotype occurring in reindeer. there has been one previous report of a peritonitis outbreak in moose associated with setaria sp. nematodes in finnish lapland in 1989 (nygren 1990). although this earlier outbreak took place within the reindeer herding area, there was no concurrent report of any associated, increased morbidity in reindeer. it is possible that the high percentage (62%; 21 of 34) of wild forest reindeer with signs of peritonitis caused by s. tundra (laaksonen et al. 2007) may be related to its substantial population decline (1700 to 1000) in 2001-2005 (kojola 2007). two roe deer examined fresh in the field had s. tundra nematodes in their abdomen and smf in circulating blood, but no peritonitis. according to our studies, roe deer seem to be a capable host and asymptomatic carrier of s. alces vol. 45, 2009 laaksonen and oksanen – vector-borne nematode in finnish cervids 83 tundra. this conclusion is supported by the simultaneous appearances in the 1960-1970s of s. tundra in scandinavia (laaksonen et al. 2007) and roe deer in north scandinavia (haugerud 1989). considering the reservoir host capacity of roe deer and the dynamics of s. tundra, we suggest that young male roe deer that can disperse many hundred kilometers from their birthplace (cederlund and liberg 1995) could be efficient long-distance vectors for s. tundra. further support for this theory is that only minor nucleotide differences exist between the reindeer s. tundra sequence (648 bp) and that of specimens from roe deer in italy (genbank aj544874, casiraghi et al. 2004), indicating that they are the same haplotype. mosquitoes, particularly aedes spp. and to a lesser extent anopheles spp., play an important role in the transmission of s. tundra in reindeer herding areas in finland. the prevalence of filariod larvae in finnish mosquitoes naturally infected with s. tundra varied from 0.5-2.5%. however, the rate of development in mosquitoes is temperature dependent; infective larvae were present approximately 14 d after a blood meal in mosquitoes maintained at room temperature (mean 21° c), but did not develop in mosquitoes maintained outdoors for 22 days at a mean temperature of 14.1° c. the third-stage (infective) larvae had a mean length of 1411 μm (sd 207) and width of 28 μm (sd 2) (laaksonen et al. 2009). the 1973 s. tundra outbreak in sweden was associated with unusually warm weather and abnormally high numbers of mosquitoes and gnats (rehbinder et al. 1975). the summers of 1972 and 1973 in finland were also very warm, as were those in 2002 and 2003 (finnish meteorological institute data, pers. comm., s. nikander 2004). warm summers apparently promote transmission and genesis of disease outbreaks by favoring the development of s. tundra in its mosquito vectors, by improving the rate of mosquito development and reducing their mortality from frost, and finally, by forcing reindeer to stay in herds on mosquito-rich wetlands (laaksonen et al. 2009). mosquito-borne diseases are among those most sensitive to weather and predictably will be influenced by climate change. climate change can directly affect disease transmission by shifting the vector’s geographic range, increasing reproductive and biting rates, and shortening the incubation period of the pathogen (patz et al. 1996). thus, we predict that global climate change has the potential to promote the further emergence of filarioid nematodes and diseases caused by them in the subarctic ecosystem. this study indicated that s. tundra likely has an important impact on boreal ecosystems. it also revealed the absence of baseline knowledge concerning temporal parasitic biodiversity in cervids at high latitudes. therefore it is important to gain knowledge about these parasites, their ecology, transmission dynamics, and their impact on human and animal health. the potential relationship between climate change and a vector-borne disease identified in this paper indicates the potential and obvious threats to the individual and population health of arctic ungulates. acknowledgements these studies were partly funded by ministry of agriculture and forestry (makera). references anderson, r. c. 2000. the superfamily filarioidea in: nematode parasites of vertebrates; their development and transmission. second edition, cabi publishing, new york, usa. casiraghi, m., o. bain, r. guerro, c. martin, v. pocacqua, s. l. gardner, a. franceschi, and c. bandi. 2004. mapping the presence of wolbachia pipientis on the phylogeny of filarial nematodes: evidence for symbiont loss during evolution. international journal of parasitology vector-borne nematode in finnish cervids – laaksonen and oksanen alces vol. 45, 2009 84 34: 191-203. cederlund, g., and o. liber. 1995. rådjuret, viltet, ekologin och jakten (roe deer, game species, ecology and hunting). almqvist and wiksell tryckeri, uppsala. 113-117. (in swedish.) dietrich, r. a., and j. r. luick. 1971. the occurance of setaria in reindeer. journal of wildlife diseases. 7: 242-245. haugerud, r. e. 1989. rådyret vandrer mot nord (roe deer are expanding northward). ottar 5: 31-36. (in norwegian.) hoberg, e. p., l. polley, e. j. jenkins, s. j. kutz, a. m. veitch, and b. t. elkin. 2008. integrated approaches and empirical models for investigation of parasitic diseases in northern wildlife. emerging infectious diseases 14: 10-17. kojola, i. 2007. petojen vaikutus metsäpeurakannoissa (effect of large carnivores on populations of wild forest reindeer). suomen riista 53: 42-48. (in finnish.) laaksonen, s., j. kuusela, s. nikander, m. nylund, and a. oksanen. 2007. parasitic peritonitis outbreak in reindeer (rangifer tarandus tarandus) in finland. the ve-the veterinary record 160: 835–841. _____, a. oksanen, t. orro, h. norberg, m. nieminen, and a. sukura. 2008b. efficacy of different treatment regimes against setariosis (setaria tundra, nematoda: filarioidea) and associated peritonitis in reindeer. acta veterinaria scandinavica 50: 49. _____, m. solismaa, r. kortet, j. kuusela, and a. oksanen. 2009. vectors and transmission dynamics for setaria tundra (filarioidea; onchocercidae), a parasite of reindeer in finland. parasites & vectors 2: 3. _____, m. solismaa, t. orro, j. kuusela, s. saari, s. nikander, a. oksanen, and a. sukura. 2008a. setaria tundra microfilariae in finnish semi-domesticated reindeer (rangifer tarandus tarandus) and wild cervids. parasitology research 104: 257-65. nikander, s., s. laaksonen, s. saari, and a. oksanen. 2007. the morphology of the filarioid nematode setaria tundra, the cause of peritonitis in reindeer rangifer tarandus. journal of helminthology 81: 49-55. nygren, t. 1990. hirvikannan tila ja hirvitutkimutsen vaiche lapissa (the status of reindeer population and research in lapland). riistantutkimusosaston tiedote. bulletin of finnish game and fisheries institute, no. 104: 3-21. (in finnish.) patz, j. a., p. r. epstein, t. a. burke, and j. m. balbus. 1996. global climate change and emerging infectious diseases. journal of the american medical association 275: 217–23. rajevsky, s. a. 1928. zwei bisher unbekannten nematoden (setarien) von rangifer tarandus und von cervus canadensis asiaticus. (two hitherto unknown nematodes (setaria species) from rangifer tarandus and from cervus canadensis asiaticus) z. infektious krankheit und hygene des haustiere 35: 40-52. (in german.) rehbinder, c., d. christensson, and v. glatthard. 1975. parasitic granulomas in reindeer. a histopathological, parasitological and bacteriological study. nordisk veterinaermedicin 27: 499-507. shol` v. a., and n. i. drobischenko. 1973. development of setaria cervi (rudolphi, 1819) in cervus elaphus maral. helminthologia (bratislava) 14: 214-246. (in russian with english abstract.) solismaa, m., s. laaksonen, m. nylund, e. pikanen, r. airakorpi, and a. oksanen. 2008. filarioid nematodes in cattle, sheep and horses in finland. acta veterinaria scandinavica 50: 20. blood profiles and associated birth characteristics of free-ranging moose (alces alces) neonates in a declining population in northeastern minnesota glenn d. delgiudice1,2 and william j. severud2 1forest wildlife populations and research group, minnesota department of natural resources, 5463-c west broadway avenue, forest lake, minnesota 55025, usa; 2department of fisheries, wildlife, and conservation biology, university of minnesota, 2003 upper buford circle, suite 135, saint paul, minnesota 55108, usa abstract: sources of natural variability of blood analytes related to physiological development pose both challenges and opportunities to deriving and interpreting the most useful nutritional and health-related information from blood profiles of free-ranging animals. preliminary evidence suggests accurate interpretation of blood profiles may be particularly important relative to newborns given their high probability of death. our goal was to establish hematological and serum reference values for freeranging moose (alces alces) neonates. sixteen neonates (8 females, 8 males) were captured and blood was sampled during 8–12 may 2013. mean age was 2.9 days old (range = 1.4–6.0); mean body mass and hind foot length were 16.8 kg (13.8–20.5) and 46.8 cm (45.0–49.0). we present mean, 95% confidence interval and range of values for 15 hematological and 24 serum characteristics, including metabolites, chemistries, electrolytes, enzymes, and metabolic and stress hormones. we observed significant (r2 = 0.423–0.747, p ≤ 0.016) positive relationships between body mass and red blood cell and white blood cell counts, hemoglobin, and packed cell volume. hind foot length was positively related (r2 = 0.369, p = 0.028) only to red blood cell counts. no serum constituents were affected by body size metrics, but sex influenced (p ≤ 0.052) several whole blood and serum characteristics. at the individual level, blood profiles facilitated discrimination of one individual neonate in poor nutritional condition that was not evident in the original physical examination at capture. as wildlife researchers and veterinarians increasingly assess the nutritional and health status of free-ranging moose and other species by clinical biochemistry and laboratory methods, cumulative banks of blood reference values will aid in data interpretation. alces vol. 52: 85–99 (2016) key words: alces alces, blood profiles, blood reference values, hematology, moose neonates, serum profiles the effective use of blood profiles to assess the health of domestic animals and their metabolic, nutritional, and reproductive status has relied on a long history of biochemical and physiological research and the establishment of reference values of quantifiable constituents (davidsohn and henry 1969, cole 1980, benjamin 1981, swenson 1984, kaneko 1989). this work supports similar, more recent efforts of wildlife researchers and veterinarians investigating the influence of intrinsic and extrinsic factors on hematological and serum characteristics in captive and free-ranging wild animals, most commonly for adult cervids (kitchen and pritchard 1962, johnson et al. 1968, seal and erickson 1969, thurley and mcnatty 1973, white and cook 1974, seal et al. 1981, delgiudice et al. 85 1990a, 1990b, 1990c, 1994); however, blood reference values for juveniles are limited for most cervidae species (tumbleson et al. 1970, franzmann and leresche 1978, rawson et al. 1992, kunkel and mech 1994, sams et al. 1995, 1996, carstensen powell and delgiudice 2005, rostal et al. 2012). sources of natural variability of measured blood analytes related to physiological development must be understood to best distinguish that variability from the influence of altered nutrition or health issues on blood profiles of free-ranging animals. this understanding is important beginning with rapidly developing newborns. for example, quantifiable relationships between age and rapid growth, increasing energy requirements, and physiological development reflected by changes in hematology have been reported for white-tailed deer from birth to 90 days of age (rawson et al. 1992). additionally, chronic maternal nutritional restriction during gestation may subsequently compromise immunocompetence of neonates, affect hematological and serum profiles, and predispose young to mortality by various agents (sams et al. 1995, 1996). the probability of mortality is greatest for northern free-ranging ungulates within the first weeks of life (delgiudice et al. 2006, carstensen et al. 2009, lenarz et al. 2010, keech et al. 2011, severud et al. 2015a, 2015b). with the assistance of global positioning system (gps) collar technology, we recently had the opportunity to capture and handle 49 free-ranging moose (alces alces) neonates (≤ 6 days old) in northeastern minnesota to study survival and cause-specific mortality (severud et al. 2015a). the long-term persistence of this population is in jeopardy; moose numbers have declined an estimated 55% from 8,840 to 4,020 from 2006 to 2016 (delgiudice 2016). however, pregnancy rates indicate that fertility is comparable to moose across north america (84%, boer 1992, severud and delgiudice, unpublished data). our goal in sampling blood from a portion of these neonates was to establish reference values and increase our understanding of the potential of hematology and serum profiles in assessing the nutritional status and overall health of newborns. specifically, our objectives were to quantify the relationship and potential influence of 1) age at capture, 2) metrics of body size (i.e., mass, hind foot length), and 3) sex of moose neonates on values of hematological and serum characteristics, and 4) to highlight any particularly informative findings (i.e., values) at the individual level. study area we captured calves within a 6,068-km2 study area located between 47° 06’n and 47° 58’n latitude and 90° 04’w and 92° 17’w longitude in northeastern minnesota. this region is described as the northern superior upland (minnesota department of natural resources [mndnr] 2015) and includes bogs, swamps, lakes, and streams, with lowland stands of northern white cedar (thuja occidentalis), black spruce (picea mariana), and tamarack (larix laricina), and upland balsam fir (abies balsamea), jack pine (pinus banksiana), white pine (p. strobus), and red pine (p. resinosa). conifers are frequently intermixed with trembling aspen (populus tremuloides) and white birch (betula papyrifera). wolves (canis lupus) and american black bears (ursus americanus) are predators of moose (fritts and mech 1981, patterson et al. 2013, severud et al. 2015a) with recent densities estimated at 3.4 wolves and 23 bears/100 km2 (erb and sampson 2013, garshelis and noyce 2015). white-tailed deer (odocoileus virginianus) are managed at pre-fawning densities of < 4 deer/km2, and are primary prey of wolves in most of northern minnesota (nelson and mech 1986, kunkel and mech 1994, delgiudice et al. 2006, 86 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 carstensen et al. 2009, grund 2014). maximum daily temperatures have been generally increasing since at least 1960 (lenarz et al. 2010). mean daily minimum and maximum temperatures ranged from �5.2 °c to 13.3 °c and 3.3 °c to 24.6 °c, respectively, during april to july 2013 at ely, minnesota (midwestern regional climate center 2015). methods neonate capture and handling we began monitoring 73 gps-collared cows on 1 may and captured neonates during 8–17 may 2013; cow collars were programmed to obtain hourly fixes in may and to transmit 4 times daily (severud et al. 2015a). we used 3 different and complementary approaches for computer-monitoring the hourly locations and movements of dams and their gps-collared neonates: a base station computer, a web-mapping service, and automated reports (severud et al. 2015a). our primary monitoring objective was to record when and where individual pregnant females increased activity reflected by a “calving movement,” a variable atypical, long distance move that ends with localization for 1–15 days (bogomolova et al. 1992, poole et al. 2007, demars et al. 2013, severud et al. 2015a). we assumed that once a female localized, the birthing process had begun, and birth occurred within 12 h (hydbring et al. 1999, bogomolova et al. 1992, asher et al. 2014). we then allowed an additional 24 h for dam-calf bonding, whereupon calves were designated “eligible” for capture. actual bonding time was calculated as that 24 h plus the elapsed time prior to capture, which depended on the daily schedule and logistical constraints. additional details of our computer-monitoring approaches are provided in severud et al. (2015a). the capture team (quicksilver air, inc., fairbanks, alaska, usa) located designated (eligible) dams from their most recent gps coordinates, and captured and collared calves as time and conditions allowed on a daily basis. they located the target dam from the air and then landed some distance (≥100 m) away to allow handlers to approach the calves on foot. neonates were not netted or chemically immobilized; handlers could simply walk up to them with most moving <10 m from where first observed and subsequently captured and handled (delgiudice et al. 2015); twins were captured, handled, and released together. the handling protocol included fitting a 420 g-gps collar (gps plus vertex survey-1 globalstar with expandable belt, vectronic aerospace gmbh, berlin, germany); fixing ear tags; collecting 25 ml of blood by syringe from the jugular vein into a 5-ml ethylenediamine tetraacetic acid (edta) tube for hematology and into two 10-ml serum tubes for chemistry and hormone assays; measuring body mass (bm, ± 0.5 kg) by spring-scale, hind foot length (± 1 cm; hfl), and rectal temperature (± 0.1 °f) by digital thermometer; and a physical examination to record injuries or abnormalities. blood samples were stored on ice and allowed to clot for 1.7 ± 0.2 h before separation by a portable centrifuge. overall handling time averaged 12.9 minutes (± 1.14 [se], range = 7–18 min, n = 9). by 12 may 2013, 2 of 11 (18.2%) dams had abandoned 2 of 17 (11.8%) neonates in apparent response to capture and handling (delgiudice et al. 2015). consequently, we removed blood-sampling, presumably the most invasive handling technique, from our protocol. ultimately, despite excluding blood-sampling and other components of the protocol, capture-induced abandonment continued intermittently, during 13–17 may 2013. our analyses indicated that abandonments reflected more of a disturbance to the dams than the neonates (delgiudice et al. 2015). all captures and handling protocols adhered to requirements of the institutional animal care and use alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 87 committee for the university of minnesota (protocol 1302-30328a) and followed guidelines of the american society of mammalogists (sikes et al. 2011). laboratory and statistical analyses hematological analyses were performed using an advia 2120 hematology analyzer (siemens healthcare diagnostics, inc., tarrytown, new york, usa) at the university of minnesota’s clinical pathology laboratory (st. paul, minnesota, usa). additionally, a peripheral blood film evaluation was conducted to determine a manual 5-part leukocyte differential and to assess cell morphology. total plasma protein (tpp) and fibrinogen concentrations were determined by refractometry and by the heat precipitation method (stockham and scott 2008: 369–413). serum biochemical profiles were determined on an au480 chemistry analyzer (beckman coulter, inc., brea, california, usa). serum total thyroxine (tt4, clinical assays tm m total t4 125i ria kit), free thyroxine (ft4, gammacoattm free t4 [two-step] 125i, ria kit), and free triiodothyronine (ft3, clinical assaystm gammacoattm free t3 125i ria kit) were assayed with kits from diasorin inc. (stillwater, minnesota, usa) at the diagnostic center for population and animal health (dcpah, michigan state university, east lansing, michigan, usa). serum total triiodothyronine (tt3) concentrations were determined at dcpah; the assay procedure and a subsequent modification are described by refsal et al. (1984) and panciera et al. (1990), respectively. dcpah also conducted serum cortisol assays (coat-a-count cortisol radioimmunoassay, siemens medical solutions diagnostics, los angeles, california, usa). we examined potential relationships between physical development (bm, hfl) and values of hematological and serum characteristics with linear regression (ott 1984: 245–250, microsoft excel 2010). we examined the influence of sex on physical and blood characteristics of neonates with t-tests assuming equal variances (ott 1984: 140–142, microsoft excel 2010). we report data as means and 95% confidence intervals (ci) or ± se. results we collected blood from 16 moose neonates; edta tubes for hematology from 13 (6 males, 7 females) and serum tubes to examine chemistries, metabolites, electrolytes, and hormones from all 16 (8 males, 8 females). mean age was the same for both groups (tables 1 and 2). there was no significant (p ≥ 0.401) relationship between age at capture and bm or hfl; however, hfl was related (r2 = 0.530, p = 0.001) to bm (fig. 1). we also observed significant (r2 = 0.423–0.747, p ≤ 0.016) positive relationships between bm and red blood cell (rbc) and white blood cell (wbc) counts, hemoglobin (hgb), and packed cell volume (pcv, fig. 2). hfl was positively related (r2 = 0.369, p = 0.028) to rbc counts (fig. 3), but only marginally to wbc counts (r2 = 0.248, p = 0.083) and hgb (r2 = 0.245, p = 0.085). neither bm nor hfl was related to mean corpuscular volume (mcv), mean corpuscular hemoglobin (mch), mean corpuscular hemoglobin concentration (mchc), platelets, differential wbcs, tpp, or fibrinogen (table 1). there were no significant relationships between bm or hfl at capture and any serum characteristics presented in table 2. however, age at capture was positively related to ft3 (r2 = 0.260, p = 0.044) and ft4 (r 2 = 0.381, p = 0.011, fig. 4). blood characteristics that differed by sex included wbc (t11 = 2.20, p = 0.040), monocytes (t11 = 2.20, p < 0.001), fibrinogen (t11 = 2.20, p = 0.052), and serum calcium (ca, t14 = 2.14, p = 0.024), total protein (tp, t14 = 2.14, p = 0.021), globulin (t14 = 88 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 2.14, p = 0.011), alkaline phosphatase (alp, t14 = 2.14, p = 0.010), and gammaglutamyl transferase (ggt, t14 = 2.14, p = 0.002; table 3). discussion a total of 49 moose neonates of 31 dams were captured and handled during 8–17 may 2013 (severud et al. 2015a). based on physical examination, measurements and their behavior, all appeared to be calm and in good overall health. mean body size of the 16 blood-sampled neonates was comparable to that of neonates of the typically larger tundra moose (a. americanus gigas) in south-central and western interior alaska (ballard et al. 1996, keech et al. 2011). the range of neonatal body size captured during our 5-day operation indicated that skeletal development accounted for just over half of the variability in bm, suggesting the remaining 50% was associated with water, protein, and fat (i.e., brown fat) content (schoonderwoerd et al. 1986, delgiudice et al. 1990c, watkins et al. 1991). a similar association (r2 = 0.85) was observed for free-ranging deer neonates (carstensen powellanddelgiudice2005).thisrelationship has value for interpreting blood profiles, particularly hematology, given our findings (figs. 2 and 3). rawson et al. (1992) showed that increases in age, bm, and resting metabolic rate contributed to a concomitant steady table 1. mean, 95% confidence interval (ci), coefficient of variation, and range of age, physical characteristics, and values of hematological characteristics of moose neonates (n = 13) at capture, northeastern minnesota, 8–12 may 2013.a characteristicb mean 95% ci coefficient of variation range age at capture (days) 2.9 2.1–3.7 0.52 1.4–6.0 body mass (kg) 16.8 15.5–18.1 0.14 13.8–20.5 hind foot length (cm) 46.8 46.0–47.5 0.03 45.0–49.0 rbc (106/µl) 6.1 5.7–6.4 0.10 5.1–6.8 wbc (103/µl) 5.4 4.7–6.2 0.25 3.9–7.8 neutrophils segs (% [103/µl]) 74.4 69.4–79.2 0.12 52.1–84.1 lymphocytes (% [103/µl]) 20.0 15.5–24.6 0.42 6.9–36.0 monocytes (% [103/µl]) 3.4 1.9–5.0 0.85 0.0–9.1 eosinophils (% [103/µl]) 2.0 1.0–2.9 0.89 0.0–5.9 basophils (% [103/µl]) 0.1 -0.05–0.34 2.45 0.0–1.0 hgb (g/dl) 9.6 9.0–10.3 0.12 8.2–11.8 pcv (%) 31.4 29.3–33.5 0.12 25.7–38.9 mcv (fl) 51.8 49.5–54.1 0.08 45.9–57.2 mch (pg) 15.9 15.2–16.5 0.08 14.1–17.6 mchc (g/dl) 30.7 30.2–31.2 0.03 28.7–31.8 platelets (103/µl) 550 436–665 0.38 357–1,065 tpp (g/dl) 5.8 5.5–6.2 0.11 4.8–7.5 fibrinogen (g/dl) 0.5 0.39–0.53 0.27 0.3–0.7 ahandlers were able to approach and handle neonates with minimal excitement, no nets or chemicals (severud et al. 2015a). bcharacteristics include rbc = red blood cells, wbc = white blood cells, hgb = hemoglobin, pcv = packed cell volume, mcv = mean corpuscular volume, mch = mean corpuscular hemoglobin, mchc = mean corpuscular hemoglobin concentration, and tpp = total plasma protein. alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 89 increase in the total daily metabolic energy requirementfor developingdeerfawns.the positive relationships between bm and rbc, pcv, and hgb, also observed in deer fawns, may maximize the o2-carrying capacity of blood of growing juveniles to meet their increasing o2 requirements, and might have positive implications for escaping predators (rawson et al. 1992). hematologic characteristics may reflect changes in nutrition, hydration, and physical exertion, and frequently respond to table 2. mean, 95% confidence interval (ci), coefficient of variation, and range of age, physical characteristics, and values of serum constituents of moose neonates (n = 16) at capture, northeastern minnesota, 8–12 may 2013.a characteristicb mean 95% ci coefficient of variation range age at capture (days) 2.9 2.2–3.5 0.47 1.4–6.0 body mass (kg) 16.8 15.8–17.9 0.13 13.8–20.5 hind foot length (cm) 46.9 46.3–47.5 0.03 45.0–49.0 ca (mg/dl) 10.1 9.8–10.5 0.07 8.8–11.4 p (mg/dl) 8.9 8.3–9.5 0.13 5.9–11.0 na (meq/l) 141 139–142 0.02 137–150 k (meq/l) 5.1 4.8–5.3 0.08 4.4–5.9 cl (meq/l) 95.3 94–96 0.02 92–100 mg (mg/dl) 1.7 1.7–1.8 0.08 1.5–2.0 bicarbonate (meq/l) 22.3 20.3–24.3 0.19 15.2–32.6 sun (mg/dl) 17.9 14.1–21.8 0.44 10.0–44.0 creatinine (mg/dl) 0.7 0.6–0.8 0.27 0.5–1.2 total bilirubin (mg/dl) 0.4 0.34–0.44 0.25 0.3–0.7 glucose (mg/dl) 108 90.5–124.7 0.32 8.0–145 sun:c (mg:mg) 24.2 20.8–27.6 0.29 14.0–36.0 albumin (g/dl) 2.1 2.0–2.3 0.12 1.7–2.6 tp (g/dl) 4.2 3.9–4.5 0.16 3.3–5.9 globulin (g/dl) 2.1 1.7–2.4 0.32 1.0–3.7 tt4 (µg/dl) 7.5 6.5–8.5 0.26 5.0–12.1 free t4 (ng/dl) 1.5 1.3–1.7 0.24 0.9–2.1 tt3 (ng/dl) 272 251–293 0.16 195–352 free t3 (pg/ml) 6.9 5.4–8.3 0.42 3.5–13.8 cortisol (µg/dl) 6.2 1.9–10.5 1.41 1.5–38.5 alp (u/l) 294 252–335 0.29 170–436 ggt (u/l) 52.0 39.6–64.4 0.49 21.0–101 ast (u/l) 63.0 53.7–72.3 0.30 42.0–110 ck (u/l) 135 91–178 0.66 52.0–399 ahandlers were able to approach and handle neonates with minimal excitement, no nets or chemicals (severud et al. 2015a). bcharacteristics include sun = serum urea nitrogen, sun:c = serum urea nitrogen:creatinine, tp = total protein, ca = calcium, p = phosphorous, na = sodium, k = potassium, cl = chloride, mg = magnesium, alp = alkaline phosphatase, ggt = gamma-glutamyl transferase, ast = aspartate aminotransferase, ck = creatine kinase, tt4 = total thyroxine, free t4 = free thyroxine, tt3 = total triiodothyronine, and free t3 = free triiodothyronine. 90 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 infectious disease (seal et al. 1978, coles 1980, benjamin 1981, delgiudice et al. 1990b, sams et al. 1995). consequently, though hematological findings have inherent value in assessing the overall well-being of neonates, interpretation of findings is most accurate when the influence of physical and natural physiological development are understood. unfortunately, established reference values for blood constituents of free-ranging, north american ungulate neonates are rare (kunkel and mech 1994, ballard et al. 1996, carstensen powell and delgiudice 2005). the earliest published study of blood profiles of free-ranging moose that included juveniles (newborns to 5 months old pooled), and attempted to assess condition of the group, did not account for changes in hematology associated with natural physiological development and growth (franzmann and leresche 1978), which potentially confounds interpretation of condition assessments. not surprisingly, mean values of pcv and hgb from that study are more reflective of the physiological development that occurs by 3 months of age (rawson et al. 1992), 36 and 78% higher, respectively, than those observed in our moose neonates (table 1). these elevated values may also reflect greater capture and handling stress in the older calves (franzmann and leresche 1978, seal et al. 1981). ballard et al. (1996) reported fig. 1. relationship of hind foot length (cm) to body mass (kg) at capture of moose neonates (1.4–6.0 days old), northeastern minnesota, 8–12 may 2013. fig. 2. relationships (top to bottom) of body mass (kg) at capture to red blood cell (rbc) and white blood cell (wbc) counts, hemoglobin (hgb) concentrations, and packed cell volume (pcv) of moose neonates (1.4–6.0 days old), northeastern minnesota, 8–12 may 2013. alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 91 reference values for blood characteristics of free-ranging moose neonates, but the hematological profiles were limited to hgb and pcv. hematological profiles of similarly aged, free-ranging white-tailed deer neonates were generally comparable to those of our much larger moose neonates (carstensen powell and delgiudice 2005), emphasizing the importance of physiological development associated with bm within species, not related to size alone. mean wbc counts of our free-ranging moose neonates fall within the range of variability of leukocyte counts of free-ranging and captive white-tailed deer neonates, and young calves of domestic cattle (benjamin 1981: 77, rawson et al. 1992, sams et al. 1995, 1996, carstensen powell and delgiudice 2005). differences between values of differential leukocyte counts of free-ranging moose neonates (table 1) and older calves in norway (rostal et al. 2012) are difficult to interpret due to the potential for physiologic leukocytosis from capture pursuit and chemical immobilization of the older calves. the absence of relationships between body size (i.e., bm or hfl) of neonates and values of serum constituents simplifies data interpretation relative to assessments of health and nutritional status. as expected in generally healthy newborns, the variability of homeostatically-maintained serum electrolytes (cations and anions) was notably more limited (mean coefficient of variation [cv] = 8.6 ± 2.22%) than for proteins (20.4 ± 6.3%), metabolites (32.0 ± 4.2%), enzymes (43.5 ± 8.7%), and metabolic and stress hormones (49.7 ± 23.0%). we included these other serum analytes in the neonate serum profile primarily for their diagnostic value in assessing nutritional intake (e.g., crude protein, digestible energy) and status (i.e., condition), hydration, responses to excitement or exertion, or a combination of these measures (benjamin 1981, seal et al. 1981, warren et al. 1982, kaneko 1989, delgiudice et al. 1990b, 1990c, 1994, watkins et al. 1991). whereas the values of some of these constituents exhibit wide ranges (e.g., sun, glucose, enzymes, thyroid hormones), most of them exhibit relatively narrow cis (table 2). specifically, characteristics such as sun, glucose, tp, albumin, and thyroid hormones (e.g., tt4 and tt3) may respond (e.g., increasing, fig. 3. relationship of hind foot length (cm) to red blood cell (rbc) counts at capture of moose neonates (1.4–6.0 days old), northeastern minnesota, 8–12 may 2013. fig. 4. relationships of age at capture to serum free triiodothyronine (t3, top) and free thyroxine (t4, bottom) of moose neonates (1.4–6.0 days old), northeastern minnesota, 8–12 may 2013. 92 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 decreasing) readily but somewhat moderately to changes in recent nutritional intake, but more extreme values may reflect dehydration, low energy status, or poor condition (e.g., accelerated net catabolism of endogenous protein) relative to nutritional restriction (bahnak et al. 1981, benjamin 1981, warren et al. 1982, delgiudice et al. 1990c, 1994, watkins et al. 1991). at the study-cohort level, concentrations of most serum constituents (table 2) were similar to values reported for captive and free-ranging white-tailed deer neonates (sams et al. 1995, carstensen powell and delgiudice 2005), as well as elsewhere for free-ranging moose neonates and various domestic livestock (benjamin 1981, kaneko 1989: 886–891, ballard et al. 1996). however, serum ca and phosphorous (p) concentrations tended to be higher than in adult white-tailed and desert mule deer (odocoileus hemionus) (delgiudice et al. 1990b, 1990c), and highly variable cortisol values, on average, were higher than in most domestic livestock species and adult white-tailed deer, but were similar to values of white-tailed deer neonates (kaneko 1989: 886–891, delgiudice et al. 1990c, sams et al. 1995, carstensen powell and delgiudice 2005). elevated ca and p concentrations, associated with increased alp values, are normal in young animals and reflective of the increased osteoblastic activity of early skeletal development and postprandial effects of periodic nursing (jacobson and mcgilliard 1984). the variable cortisol concentrations are suggestive of varied excitement levels associated with handling, but also may be influenced by reduced energy intake and time since nursing (thurley and mcnatty 1973, seal et al. 1981, delgiudice et al. 1990b). serum creatinine concentrations are directly related to muscle mass, and along with albumin, tp, and globulin, are typically lower in the young of domestic livestock and table 3. mean, 95% confidence interval (ci), and range of age, physical characteristics, and values of blood constituents of moose neonates at capture that differ between males and females, northeastern minnesota, 8–12 may 2013.a males females characteristicb n mean 95% ci range n mean 95% ci range body mass (kg) 8 16.9 15.2–18.7 14.0–20.5 8 16.8 5.4–18.1 13.8–19.5 hfl (cm) 8 47.1 46.3–47.8 45.5–48.5 8 46.8 45.8–47.7 45.0–49.0 wbc (x103/µl) 6 6.2 5.0–7.5 3.9–7.8 7 4.8 4.3–5.2 3.9–5.6 monocytes (% [x103/µl]) 6 4.8 2.5–7.1 0.9–9.1 7 2.3 0.34–4.19 0.0–7.9 fibrinogen (g/dl) 6 0.53 0.45–0.62 0.4–0.7 7 0.4 0.31–0.49 0.3–0.6 ca (mg/dl) 8 9.7 9.3–10.2 8.8–11.1 8 10.5 10.1–10.9 9.8–11.4 tp (g/dl) 8 3.8 3.6–4.0 3.3–4.1 8 4.6 4.0–5.1 3.7–5.9 globulin (g/dl) 8 1.64 1.37–1.90 1.0–2.1 8 2.5 1.98–2.95 1.3–3.7 bicarbonate (meq/l) 8 24.4 21.7–27.1 20.4–32.6 8 20.3 17.9–22.6 15.2–24.0 alp (u/l) 8 243 204–282 170–350 8 345 290–399 211–436 ggt (u/l) 8 34.8 26.9–42.6 21–51 8 69.3 52.8–85.7 35–101 ahandlers were able to approach and handle neonates with minimal excitement, no nets or chemicals (severud et al. 2015a). bcharacteristics include hfl = hind foot length, wbc = white blood cells, ca = calcium, tp = total protein, alp = alkaline phosphatase, and ggt = gamma-glutamyl transferase. alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 93 wild deer species (summarized by benjamin 1981: 111–112, finco 1989, delgiudice et al. 1990c). among the serum enzymes, alp, ggt, and ck in particular were higher than in several domestic livestock species (kaneko 1989: 886–891). concentrations of serum enzymes can be highly variable. because these enzymes are synthesized intracellularly, relative to their primary source(s), moderate to extreme elevations in serum may be indicative of physiological functions or events ranging from normal bone turnover (e.g., alp), particularly in juveniles, to cellular damage associated with the heart and skeletal muscle (e.g., ck), liver, kidneys, and other organs from a variety of pathologies or extreme physical exertion (zimmerman and henry 1969, benjamin 1981: 229– 232, seal et al. 1981, kramer 1989). our ability to observe and recognize most of the “… circumstantial, behavioral, and physical conditions …” influencing the vulnerability of prey species relative to the forces of nature is limited (mech 1970: 247–248). the unique value of blood profiles at the individual level is demonstrated clearly in our assessment of the nutritional and health status of neonate number 520. this 2.6-day old, gps-collared neonate died within 4.5 h of capture and release, and was one of the largest calves at 20.5 kg (hfl = 48.0 cm, table 1). it appeared healthy and calm during handling, and the apparently attentive dam remained nearby during capture and handling, when the team departed, and when they returned to recover the dead calf. initially, the mortality was classified as “capture-related.” however, a detailed necropsy at the university of minnesota’s veterinary diagnostic laboratory, including extensive macroscopic and microscopic examinations, was inconclusive, except for an empty abomasum (devoid of curdled milk), suggestive of hypoglycemia. subsequent analyses of the neonate’s blood samples generated a hematological profile with the highest rbc, hgb, and pcv values, and the lowest tpp of the sampled cohort (table 1), indicative of hemoconcentration and dehydration concomitant with prolonged nutritional deprivation (coles 1980: 116– 117, benjamin 1981: 73, 146–147, delgiudice et al. 1994). this neonate also exhibited the cohort’s highest sun, creatinine, and sun:creatinine values, indicative of accelerated net catabolism of endogenous protein, and the cohort’s lowest ca, tp, globulin, glucose, and ft3 concentrations, reflecting nutritional deprivation and condition deterioration (table 2; coles 1980: 246, benjamin 1981: 175, 178, delgiudice et al. 1990c, 1994, watkins et al. 1991). additionally, this neonate exhibited the highest ast and ck concentrations, indicative of cellular damage associated with skeletal muscle (benjamin 1981: 231), as well as the most elevated serum cortisol concentration (38.4 µg/dl, table 2). elevated serum cortisol often accompanies severe nutritional restriction and increased protein catabolism, and, associated with stress, induces a physiologic leukocytosis (coles 1980: 50– 51, 280–281, delgiudice et al. 1990c), indicated in this neonate by the sampled cohort’s highest wbc count (7.8 x 103/µl) and neutrophilia (84.1% [x 103/µl], table 1). the blood profile of neonate 520 provided a plethora of evidence that this individual was in very poor condition, likely moribund prior to capture, an assessment and prognosis not possible simply from physical examination in the field or during laboratory necropsy. it is unclear whether this neonate’s poor condition was due to a problem with its ability to nurse or the dam’s inability to provide nutrition. however, many of its hematological and serum analyte measurements were starkly different from those of 2 other similarly large neonates, numbers 519 (19.5 kg) and 503 (18.5 kg). their rbc counts and hgb concentrations were similar to those of neonate 520, but not 94 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 due to hemoconcentration and dehydration. their rbc and hgb values, associated with a near average pcv, tpp, sun, creatinine, ca, and tp values, higher globulin, glucose, and ft3 concentrations, and markedly lower cortisol were reflective of calves adequately nourished and in good nutritional condition. indeed, neonate 519 was the only calf that yielded several analyte values (tpp, tp, globulin, and free t3) higher than 2 standard deviations (sd) above the sampled cohort’s respective mean values (95% portion of the distributions). this calf slipped its collar at 13 days old, whereas neonate 503 survived until at least 279 days old, when we removed its collar. reproductive success, defined as producing a calf that survives to 1 year of age (i.e., recruitment), is an important driver of population performance (gaillard et al. 2000, raithel et al. 2007). since 2012, the study population has experienced a 5-year interval of relative stability at about 4,000 moose (delgiudice 2016). recruitment has been low, but pregnancy rates indicate that fertility is not limiting reproductive success. neonatal blood profiles and morphological characteristics indicate that physical and physiological development were relatively robust and did not reflect any specific vulnerabilities, except in the case of neonate 520. comparing this case to the overall data set of reference values demonstrates the key role that blood profiles can play in improving understanding of the nutritional and health status of cervid neonates when mortality pressures are greatest. taking advantage of every research means possible becomes increasingly important in attempting to identify and best comprehend factors impacting a steadily declining moose population. acknowledgements thanks to t. enright, k. foshay, t. obermoller, j. lodel, b. betterly, a. jones, k. miedtke, b. smith, e. hildebrand, d. pauly, m. carstensen, m. dexter, l. cornicelli, m. larson, j. forester, a. wünschmann, a. armien, b. patterson, r. moen, a. mcgraw, j. terry, and m. schrage for varying combinations of technical and cooperative support during the planning and operational phases of this study. we are appreciative for the flying, capture, and calf-handling skills of r. swisher and m. keech (quicksilver air, inc., fairbanks, alaska) and the fixed-wing aircraft assistance of a. buchert and l. ettl. this project was funded in part by the minnesota department of natural resources section of wildlife and the wildlife restoration (pittman-robertson) program. we thank the minnesota deer hunters association for supplemental funding. references asher, g. w., a. j. wall, k. t. o’neil, r. p. littlejohn, a. bryant, and n. cox. 2014. the use of gps data to identify calving behaviour of farmed red deer hinds: proof of concept for intensively managed hinds. applied animal behavioral science 154: 93–103. bahnak, b. r., j. c. holland, l. j. verme, and j. j. ozoga. 1981. seasonal and nutritional influences on growth, hormone, and thyroid activity in white-tailed deer. journal of wildlife management 45: 140–147. ballard, w. b., p. j. macquarrie, a. w. franzmann, and p. r. krausman. 1996. effects of winters on physical condition of moose in south-central alaska. alces 32: 51–59. benjamin, m. m. 1981. outline of veterinary clinical pathology. the iowa state university press, ames, iowa, usa. boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces (suppl.) 1: 1–10. bogomolova, e. m., j. a. kurochkinja, and p. k. anokhin. 1992. observations alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 95 of moose behavior on a moose farm. alces (suppl.) 1: 216. carstensen powell, m., and g. d. delgiudice. 2005. birth, morphologic, and blood characteristics of free-ranging white-tailed deer neonates. journal of wildlife diseases 41: 171–183. ——, ——, b. a. sampson, and d. w. kuehn. 2009. survival, birth characteristics, and cause-specific mortality of white-tailed deer neonates. journal of wildlife management 73: 175–183. coles, e. h. 1980. veterinary clinical pathology. w. b. saunders company, philadelphia, pennsylvania, usa. davidsohn, i., and j. b. henry (editors). 1969. todd-sanford clinical diagnosis by laboratory methods. 14th edition. w. b. saunders company, philadelphia, pennsylvania, usa. delgiudice, g. d. 2016. 2016 aerial moose survey. technical report, minnesota department of natural resources, st. paul, minnesota, usa. http://files.dnr.state. mn.us/wildlife/moose/2016_moosesurvey. pdf (accessed march 2016). ——, j. fieberg, m. r. riggs, m. powell carstensen, and w. pan. 2006. a long-term age-specific survival analysis of female white-tailed deer. journal of wildlife management 70: 1556–1568. ——, p. r. krausman, e. s. bellantoni, m. c. wallace, r. c. etchberger, and u. s. seal. 1990a. blood and urinary profiles of free-ranging desert mule deer in arizona. journal of wildlife diseases 26: 83–89. ——, k. e. kunkel, l. d. mech, and u. s. seal. 1990b. minimizing capturerelated stress on white-tailed deer with a capture collar. journal of wildlife management 54: 299–303. ——, l. d. mech, and u. s. seal. 1990c. effects of winter undernutrition on body composition and physiological profiles of white-tailed deer. journal of wildlife management 54: 539–550. ——, ——, and ——. 1994. winter undernutrition and serum and urinary urea nitrogen of white-tailed deer. journal of wildlife management 58: 430–436. ——, w. j. severud, t. r. obermoller, r. g. wright, t. a. enright, and v. stlouis. 2015. monitoring movement behavior enhances recognition and understanding of capture-induced abandonment of moose neonates. journal of mammalogy 96: 1005–1016. demars, c. a., m. auger-méthé, u. e. schlägel, and s. boutin. 2013. inferring parturition and neonate survival from movement patterns of female ungulates: a case study using woodland caribou. ecology and evolution 3: 4149–4160. erb, j., and b. a. sampson. 2013. distribution and abundance of wolves in minnesota, 2012–2013. technical report, minnesota department of natural resources, st. paul, minnesota, usa. http://files.dnr.state.mn.us/fish_wildlife/ wildlife/wolves/2013/wolfsurvey_2013. pdf (accessed march 2016). finco, d. r. 1989. kidney function. pages 496–542 in j. j. kaneko, editor. clinical biochemistry of domestic animals. fourth edition. academic press, inc., new york, new york, usa. franzmann, a. w., and r. e. leresche. 1978. alaskan moose blood studies with emphasis on condition evaluation. journal of wildlife management 42: 334–351. fritts, s. h., and l. d. mech. 1981. dynamics, movements, and feeding ecology of a newly protected wolf population in northwestern minnesota. wildlife monograph no. 80. gaillard, j. m., m. festa-bianchet, n. g. yoccoz, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual review of ecology and systematics 31: 367–393. garshelis, d. l., and k. v. noyce. 2015. status of minnesota black bears, 2014. 96 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 http://files.dnr.state.mn.us/wildlife/moose/2016_moosesurvey.pdf http://files.dnr.state.mn.us/wildlife/moose/2016_moosesurvey.pdf http://files.dnr.state.mn.us/wildlife/moose/2016_moosesurvey.pdf http://files.dnr.state.mn.us/fish_wildlife/wildlife/wolves/2013/wolfsurvey_2013.pdf http://files.dnr.state.mn.us/fish_wildlife/wildlife/wolves/2013/wolfsurvey_2013.pdf http://files.dnr.state.mn.us/fish_wildlife/wildlife/wolves/2013/wolfsurvey_2013.pdf final report, minnesota department of natural resources, st. paul, minnesota, usa. http://files.dnr.state.mn.us/recreation/ hunting/bear/2014_bearharvest.pdf (accessed march 2016). grund, m. 2014. monitoring population trends of white-tailed deer in minnesota-2014. status of wildlife populations. minnesota department of natural resources, st. paul, minnesota, usa. hydring, a. m., e. macdonald, g. drugge-boholm, b. berglund, and d. k. olsson. 1999. hormonal changes during parturition in heifers and goats are related to the phases and severity of labour. journal of endocrinology 160: 75–85. jacobsen, n. l., and a. d. mcgilliard. 1984. the mammary gland and lactation. pages 863–880 in m. j. swenson, editor. duke’s physiology of domestic animals. 10th edition. comstock publishing associates, cornell university press, ithaca, new york, usa. johnson, h. e., w. g. youatt, l. d. fay, h. d. harte, and d. e. ullrey. 1968. hematological values of michigan white-tailed deer. journal of mammalogy 49: 749–754. kaneko, j. j. (editor). 1989. clinical biochemistry of domestic animals. fourth edition. academic press, inc., new york, new york, usa. keech, m. a., m. s. lindberg, r. d. boertje, p. valkenburg, b. d. taras, t. a. boudreau, and k. b. beckmen. 2011. effects of predator treatments, individual traits, and environment on moose survival in alaska. journal of wildlife management 75: 1361–1380. kitchen, h., and w. r. pritchard. 1962. physiology of blood. pages 109–114 in proceedings of first national whitetailed deer symposium. university of georgia, athens, georgia, usa. kramer, j. w. 1989. clinical enzymology. pages 338–363 in j. j. kaneko, editor. clinical biochemistry of domestic animals. fourth edition. academic press, inc. new york, new york, usa. kunkel, k. e., and l. d. mech. 1994. wolf and bear predation on white-tailed deer fawns in northeastern minnesota. canadian journal of zoology 72: 1557–1565. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. mech, l. d. 1970. the wolf: the ecology and behavior of an endangered species. university of minnesota press, minneapolis, minnesota, usa. microsoft excel. 2010. version 14.0. 7166.5000 (32-bit). microsoft corporation 2010, redmond, washington, usa. midwestern regional climate center. 2015. cli-mate, mrcc application tools environment. http://mrcc.isws.illinois. edu/climate/ (accessed march 2016). minnesota department of natural resources (mndnr). 2015. ecological classification system. minnesota department of natural resources. st. paul, usa. http://www.dnr.state.mn.us/ ecs/index.html (accessed march 2016). nelson, m. e., and l. d. mech. 1986. mortality of white-tailed deer in northeastern minnesota. journal of wildlife management 50: 691–698. ott, l. 1984. an introduction to statistical methods and data analysis. pws publishers, boston, massachusetts, usa. panciera, d. l., e. g. macewen, c. e. atkins, w. t. k. bosu, k. r. refsal, and r. f. nachreiner. 1990. thyroid function test in euthyroid dogs treated with i-thyroxine. american journal of veterinary research 51: 22–26. patterson, b. r., j. f. benson, k. r. middel, k. j. mills, a. silver, and m. e. obbard. 2013. moose calf mortality in central ontario, canada. journal of wildlife management 77: 832–841. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 97 http://files.dnr.state.mn.us/recreation/hunting/bear/2014_bearharvest.pdf http://files.dnr.state.mn.us/recreation/hunting/bear/2014_bearharvest.pdf http://mrcc.isws.illinois.edu/climate/ http://mrcc.isws.illinois.edu/climate/ http://www.dnr.state.mn.us/ecs/index.html http://www.dnr.state.mn.us/ecs/index.html interior montane ecosystems. journal of mammalogy 88: 139–150. raithel, j. d., m. j. kauffman, and d. h. pletcher. 2007. impact of spatial and temporal variation in calf survival on the growth of elk populations. journal of wildlife management 71: 795–803. rawson, r. e., g. d. delgiudice, h. e. dziuk, and l. d. mech. 1992. energy metabolism and hematology of whitetailed deer fawns. journal of wildlife diseases 28: 91–94. refsal, k. r., r. f. nachreiner, and c. r. anderson. 1984. relationship of season, herd, lactation, and pregnancy with serum thyroxine and triiodothyronine in holstein cows. domestic animal endocrinology 3: 225–234. rostal, m. k., a. l. evans, e. j. solberg, and j. m. arnemo. 2012. hematology and serum chemistry reference ranges of free-ranging moose (alces alces) in norway. journal of wildlife diseases 48: 548–559. sams, m. g., r. l. lochmiller, e. c. hellgren, m. e. payton, and l. w. varner. 1995. physiological responses of neonatal white-tailed deer reflective of maternal dietary protein intake. canadian journal of zoology 73: 1928–1936. ——, ——, ——, w. d. warde, and l. w. varner. 1996. morphometric predictors of neonatal age for white-tailed deer. wildlife society bulletin 24: 53–57. schooderwoerd, m., c. e. doige, g. a. wobeser, and j. m. naylor. 1986. protein and energy malnutrition and fat mobilization in neonatal calves. canadian veterinary journal 27: 365–371. seal, u. s., and a. w. erickson. 1969. hematology, blood chemistry, and proteinpolymorphisms in the white-tailed deer (odocoileus virginianus). comparative biochemistry and physiology 30: 695–713. ——, l. j. verme, and j. j. ozoga. 1978. dietary protein and energy effects on deer fawn metabolic patterns. journal of wildlife management 42: 776–790. ——, ——, ——. 1981. physiologic values. pages 17–34 in w. r. davidson, editor. diseases and parasites of white-tailed deer. tall timbers research station, tallahassee, florida, usa. severud, w. j., g. d. delgiudice, t. r. obermoller, t. a. enright, r. g. wright, and j. d. forester. 2015a. using gps collars to determine parturition and cause-specific mortality of moose calves. wildlife society bulletin 39: 616–625. ——, ——, ——, r. j. ryan, and b. d. smith. 2015b. an alternate method to determine moose calving and cause-specific mortality of calves in northeastern in minnesota. pages 93–108 in l. cornicelli, m. carstensen, m. d. grund, m. a. larson, and j. s. lawrence, editors. summaries of wildlife research findings 2014, minnesota department of natural resources, st. paul, minnesota, usa. http://files. dnr.state.mn.us/publications/wildlife/ research2014/binder.pdf (accessed march 2016). sikes, r. s., w. l. gannon, the animal care and use committee of the american society of mammalogists. 2011. guidelines of the american society of mammalogists for the use of wild mammals in research. journal of mammalogy 92: 235–253. stockham, s. l., and m. a. scott. 2008. fundamentals of veterinary clinical pathology. 2nd ed. wiley-blackwell, ames, iowa. swenson, m. j. (editor). 1984. duke’s physiology of domestic animals. 10th edition. comstock publishing associates, cornell university press, ithaca, new york, usa. thurley, d. c., and k. p. mcnatty. 1973. factors affecting peripheral cortisol levels in unrestricted ewes. acta endocrinologica 74: 331–337. tumbleson, m. e., j. d. cuneio, and d. a. murphy. 1970. serum biochemical and hematological parameters of captive 98 blood characteristics of moose neonates – delgiudice and severud alces vol. 52, 2016 http://files.dnr.state.mn.us/publications/wildlife/research2014/binder.pdf http://files.dnr.state.mn.us/publications/wildlife/research2014/binder.pdf http://files.dnr.state.mn.us/publications/wildlife/research2014/binder.pdf white-tailed fawns. canadian journal of comparative medicine 34: 66–71. warren, r. j., r. l. kirkpatrick, a. oelschlaeger, p. f. scanlon, k. e. webb, jr., and j. b. whelan. 1982. energy, protein, and seasonal influences on white-tailed deer fawn nutritional indices. journal of wildlife management 46: 302–312. watkins, b. e., j. h. witham, d. e. ullrey, d. j. watkins, and j. m. jones. 1991. body composition and condition evaluation of white-tailed deer fawns. journal of wildlife management 55: 39–51. white, m., and r. s. cook. 1974. blood characteristics of free-ranging whitetailed deer in southern texas. journal of wildlife diseases 10: 18–24. zimmerman, h. j., and j. b. henry. 1969. serum enzyme determinations as an aid to diagnosis. pages 710–748 in i. israel, and j. b. henry, editors. todd–sanford clinical diagnosis by laboratory methods. 14th edition. w. b. saunders company, philadelphia, pennsylvania, usa. alces vol. 52, 2016 delgiudice and severud – blood characteristics of moose neonates 99 blood profiles and associated birth characteristics of free-anging moose (alces alces) neonates in a declining population in northeastern minnesota study area methods neonate capture and handling laboratory and statistical analyses results discussion acknowledgements references fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics henry jones1, peter j. pekins1, lee e. kantar2, matt o’neil, and daniel ellingwood1 1department of natural resources and the environment: wildlife program, university of new hampshire, durham, nh, 03824, usa; 2maine department of inland fisheries and wildlife, research and assessment section: bangor, maine, 04401, usa abstract: moose (alces alces) populations in northern new hampshire and western maine experienced 3 successive years of high winter tick infestations (epizootics) in 2014–2016 that resulted in late-winter calf mortality rates >70%. to assess productivity in these populations, we measured fecundity rates of yearling and adult cow moose, and neonatal and summer calf survival. parturition, fecundity, and survival were measured via direct observation by stalking vhf and gps radio-collared cows (n = 177) in may-august, 2014–2016. calving rates for yearlings and adults averaged 0 and 57%, respectively; there was no twinning documented. summer calf survival to august was high overall (83%), with 85% of the mortality occurring in the first week of life. calving and twinning rates declined since last measured in new hampshire in 2002–2005 and were below the north american average; conversely, summer survival of calves was considered normal. given that optimal habitat has increased in the past 15 years in the study area that is dominated by commercial forestry, lower productivity is presumably related to the additive impacts of successive winter tick epizootics on year-round condition of cows. alces vol. 53: 85–98 (2017) key words: alces alces, calving, fecundity, maine, neonate survival, new hampshire, pregnancy, winter ticks many moose (alces alces) populations along the southern edge of their north american range are declining, including in minnesota, manitoba, nova scotia, vermont, new york, and new hampshire (murray et al. 2006, broders 2012). the cause of decline varies regionally, but it is generally associated with the warming climate which likely has an indirect influence through increasing incidence of parasites and disease (samuel 2004, murray et al. 2006, lankester 2010). in northern new england, winter ticks (dermacentor albipictus) are suspected to influence the population through periodic widespread mortality of calves during epizootics (musante et al. 2010) which have been occurring at an increasing frequency in the last 15 years (bergeron et al. 2013). an epizootic event (>50% calf mortality) was documented in northern new hampshire in 2002 (musante et al. 2010), and acknowledged to occur in northern new hampshire and western maine in 2008 and 2011 by the new hampshire fish and game department [nhfg] and the maine department of inland fisheries and wildlife [mifw] (bergeron et al. 2013). in these same areas, epizootics occurred in 3 consecutive years from 2014–2016. mortality of radio-marked 10–12 month-old calf moose was 60–80% between march and may from blood loss to winter tick parasitism; the average winter 85 tick infestation on dead calves was 46,800 ticks (range = 34,800–63,600) (jones 2016, l. kantar, mifw, pers. comm.). the population implications of successive or frequent epizootics with high calf mortality rates are cause for concern among regional moose managers. however, the status and trajectory of a moose population is also dictated by the number of calves recruited into the population, termed here as productivity, which is dependent on fecundity, the number of calves born per cow, and neonate and calf survival. pregnancy, calving, and twinning rates are parameters that contribute to the fecundity rate of a population (van ballenberghe and ballard 2007). productivity is influenced by the nutritional condition of cows and calves, and as such, allows for comparison with other populations and overall assessment of health (schwartz 2007). where productivity is high, a population is more resilient to mortality factors (franzmann 2000), hence, an important consideration given the recent decline in the northern new england moose population. winter ticks are known to cause population decline in moose through widespread mortality of calves during an epizootic year, and suspected long-term effects due to reduction of adult cow fitness and productivity (musante et al. 2010, bergeron et al. 2013). high winter tick infestations presumably exacerbate the negative energy balance of adult cows in late winter and early spring due to the compounding effects of substantial protein deficit from blood loss (musante et al. 2007) and the nutritional deficiency of late winter browse (schwartz and renecker 2007). optimal condition is relative to season, as moose experience a negative energy deficit during winter resulting in weight loss even on the best range (schwartz and renecker 2007). yet, because gestational and early lactational costs are met prior to spring green-up, availability of tissue energy, or minimizing weight loss, is paramount to production. poorer cow condition from the additive effect of blood loss to winter ticks may result in reduced fertility, low yearling productivity, increased age of first reproduction, and low twinning rates (musante et al. 2010), all of which have been documented in new hampshire (bergeron et al. 2013). given the increasing frequency of regional epizootics and concurrently declining estimates of productivity (bergeron et al. 2013), it is critical to measure yearling and adult female fecundity and survival rates of calves to accurately assess the status and trajectory of the regional population. this study was designed to investigate the productivity of two moose populations: one in northern new hampshire and the second in western maine. similar harvest strategies and ecological conditions between these areas provided the opportunity to compare and combine data sets. the specific objectives of this research were to measure: 1) adult cow pregnancy rates and parturition dates, 2) calving and twinning rates of yearling and adult cow moose, and 3) summer calf survival. study area the 2 study sites were in northern new hampshire and western maine, separated by approximately 120 km. the new hampshire site was located in the eastern portion of coos county centered on the town of milan (fig. 1). this site encompassed ∼1,250 km2 in nhfg wildlife management unit (wmu) c2 and portions of wmus a2, b, and c1, and replicated the study area of a comprehensive population dynamics study that occurred from 2002–2005 (musante et al. 2010). moose density was estimated as 0.46–0.87 moose/km2. the number of moose hunting permits issued in 2013–2015 averaged 28 either-sex and 10 antlerless-only permits. the western maine site extended north and west of the town of greenville to the 86 nh & me productivity – jones et al. alces vol. 53, 2017 quebec border, including parts of somerset, franklin, and piscataquis counties (fig. 1). this site was ∼5,620 km2 and encompassed mifw wildlife management district 8. moose density was estimated as 0.97–1.35 moose/km2. the number of moose hunting permits issued in 2013–2015 averaged 175 bull-only and 25 antlerless-only permits. potential predators of moose calves were black bears (ursus americanus) and coyotes (canis latrans); bear density was estimated as 0.38–0.58 bear/km2 in new hampshire (a. timmins, nhfg, pers. comm.) and 0.65 bear/km2 in maine (r. cross, mifw, pers. comm.). white-tailed deer (odocoileus virginianus) were sympatric with moose throughout the region and at an estimated density of 6.34 deer/km2 in new hampshire (d. bergeron, nhfg, pers. comm.) and 0.60 deer/km2 in maine (k. ravanna, mifw, pers. comm.). both study sites were privately-owned, managed for commercial timber, and considered high quality moose habitat (15–20% of the landscape in 4–16 year-old regenerating forest; ball 2017). the area was mountainous (max elevation 1220 m) and geographically diverse with lowland valleys, rolling hills, fig. 1. study areas located in northern new hampshire and western maine, 2014–2016. alces vol. 53, 2017 jones et al. – nh & me productivity 87 smaller mountains, and numerous lakes, ponds, and rivers scattered throughout. the dominant cover type was northern hardwood forest consisting of american beech (fagus grandifolia), sugar maple (acer saccharum), and paper birch (betula papyrifera). conifer stands of mostly red spruce (picea rubens) and balsam fir (abies balsamea) were common at high elevation, with white cedar (thuja occidentalis) and black spruce (picea mariana) common in wet lowland sites (degraaf et al. 1992). year-round access was available on logging roads, off highway recreational vehicle (ohrv) trails, and snowmobile trails. private landowners permitted access during the calving period (may and early june) which typically coincided with road-closures due to mud. climate data were available from the national climatic data center weather stations at york pond, berlin, new hampshire (id: ghcnd:usc00279966, lat/long: 44.5002, �71.333) and jackman, maine (id: ghcn: usc00174086, lat/long: 45.626, �70.246). annual ambient temperature ranged from 32 to �32 °c in both areas. for new hampshire and maine, respectively, annual precipitation ranged from 114.0 to 121.6 cm and 91.0 to 106.0 cm, and maximum snow depth ranged from 17.8 to 66.0 cm and 30.5 to 106.7 cm. the mean annual snowfall of 215.6 cm and maximum recorded snow depths in 2014– 2016 (61.0, 66.0, and 35.6 cm) in new hamp‐ shire were generally similar to those in maine (240.5 cm and 106.7, 53.3, and 43.2 cm). the average weekly snow depth measured at open sites in december–april 2013–2016 ranged from 2.5–26.4 cm in new hampshire and 5.8–48.7 cm in maine; average weekly snow depth did not exceed 70 cm at either site. methods capture and marking animal capture and handling protocols were approved by the institutional animal care and use committee at the university of new hampshire (iacuc #130805). cow and calf (both sexes) moose were captured by helicopter net-gunning and helicopter darting (aero tech inc., clovis, new mexico, usa in 2014 and 2015; native range capture services, elko, nevada, usa in 2016) with a 4–6 person crew: the pilot, animal handlers, and a veterinarian. captures occurred in january and were competed in ≤ 7 days at each study site. concentrations of non-collared moose were identified from helicopter and fixed-wing flights prior to captures. moose captured via darting were immobilized with 3 mg of carfentanil and reversed with 300 mg of naltrexone. additional captures in maine included 6 adult cows collared in water using small boats and a noose in august 2014 (crossley 1987), and 2 adult cows darted roadside using 2.2 ml of ketamine and reversed with 0.8 ml of medetomidine in december 2014. moose captured via net gunning were quickly removed from the net and restrained with leg hobbles and blindfolded; the handling process typically lasted < 15 min. age classes were categorized as calves (<1 year old), yearlings (>1 year but <2 years old), and adults (>2 years old) at the time of the fall breeding season. as it can be difficult to differentiate between yearling and adult cow moose without observing tooth wear, all cows were considered adults; the relative size of each was checked with the capture crew in an attempt to identify any obvious yearling. the yearling age class consisted of radio-marked female calves that survived their first winter. a 30 ml blood sample was taken from the jugular vein to be used for subsequent blood tests including pregnancy. each moose was fitted with numbered ear tags colorcoded by year (allflex usa, dallas, texas, usa) and a very high frequency (vhf) or global positioning system (gps) radio-collar. moose in new hampshire were fitted with either a vhf (n = 76; m2610b, advanced 88 nh & me productivity – jones et al. alces vol. 53, 2017 telemetry systems, isanti, minnesota, usa; mod-600, telonics, mesa, arizona, usa) or gps radio-collar (n = 54, gps plus vertex survey collar, vectronic aerospace gmbh, berlin, germany); all moose in maine were fitted with gps radio-collars (n = 142, gps plus vertex survey collar, vectronic aerospace gmbh, berlin, germany). vhf radio-collars had a motion sensor switch to indicate a 4-h period without movement; collars were continuously monitored with a r4500s datalogger (ats, isanti, mn) connected to a large omnidirectional antenna (cushcraft crx 150) mounted centrally in the study area. the gps radio-collars collected 2 gps fixes daily (0000 and 1200 hr est) and had a vhf beacon that was active at 0700–1900 hr est; after 5 h of non-movement, a motion sensor switch triggered a “mortality message” via e-mail and the pulse rate of the vhf signal increased. adults received collars sized to a standard fit; calves received retrofitted collars that allowed future expansion (see musante et al. 2010). this research was part of a larger study that also assessed cause-specific mortality of the radio-marked moose. fecundity pregnancy status of each adult female was determined from the blood samples collected at capture using the pregnancy specific protein-b test (biopryn, moscow, idaho, usa). calving and twinning rates were measured principally through direct observation by stalking adult and yearling cows within sighting distance (i.e., walk-ins; mech 1983, musante et al. 2010). calving rates included the annual calving rate or the proportion of cows documented as having a calf each year, and the successive calving rate or the proportion of cows that birthed in consecutive years. in new hampshire, walk-ins were conducted 2–3x weekly from 1 may– 1 july, and weekly thereafter until 1 august. in maine, movement was monitored through daily gps locations and walk-ins were initiated when daily locations became highly localized, indicative of birthing behavior (testa et al. 2000a, mcgraw et al. 2014). if localization did not occur, cows were checked once weekly from mid-may to 1 july. the age of neonatal calves was estimated from their wet or dry appearance, mobility, and coordination (larsen et al. 1989, musante et al. 2010). the time span between the initial calf sighting and the last observation of a cow without its calf was also used to estimate age and birth date. in the absence of direct observation, calves could be identified from tracks, vocalizations, and behavior (mobility and grunting) of cows leading their young. parturition date and age of the calf at first observation was estimated in new hampshire because walk-ins occurred multiple times weekly; however, parturition date and calf age at first observation in maine were assigned to weekly periods because walk-ins occurred ∼1x weekly which increased the probability of missing some early calf mortality (< 7 days). direct observation was considered the best method to document births and early survival because it was minimally invasive and reasonable access was available at both locations. further, because local spring green-up typically occurs ∼2 weeks after the median birth date (18 may; musante et al. 2010), observations were ideal during the immediate post-birth period when most calf mortality occurs. monitoring of unmarked calves calf survival was measured intensively for 60 d post-birth (summer survival) from direct observation or sign (e.g., tracks, beds, fecal matter, vocalizations) observed during walk-ins; we assumed that surviving calves would be near their radio-marked mother. survival of calves-at-heel was checked 2x and 1x weekly in new hampshire and maine, respectively. if a calf was not ob‐ served ≥3 consecutive times over 2 weeks, alces vol. 53, 2017 jones et al. – nh & me productivity 89 it was considered a mortality on the day midway between the last confirmed observation and the first missing date; specific cause of death was never identified. an unmarked calf was considered a mortality if the cow died during this time period. analysis fecundity and calf survival rates were compared between years at each study site using a chi-square independence test and between study sites using fisher’s exact test. calf survival was plotted on a kaplan-meier survival curve with 95% confidence limits and a chi-square goodnessof-fit test was used to compare the timing of unmarked calf losses. analyses were performed using program r (v 3.2.2, vienna, austria). results fecundity a total of 76 adult cows were monitored for at least one calving season: 46 in new hampshire and 30 in maine. the annual pregnancy rate in new hampshire ranged from 76 to 88%, averaging 78% (n = 45); there was no difference by year (χ2 = 0.54, p = 0.76, table 1). the annual pregnancy rate in maine ranged from 75 to 90%, averaging 88% (n = 24); there was no difference by year (χ2 = 0.28, p = 0.60, table 1). pregnancy rate did not differ between the study areas (p = 0.32), and the combined average was 81% (n = 69). parturition occurred from 10–31 may (n = 83) in both study areas and the median parturition date was 19 may (n = 45, new hampshire only). birthing was highly synchronous all years with 90% occurring from 14–25 may. the mean age at first observation (n = 29) was 1.8 d (sd = 1.4) and only 2 calves were first detected at ≥5 days old in new hampshire; calf age at first observation was less specific in maine due to differing methodology (∼90% of calves were observed during their first week of life). calving rates were similar in new hampshire (56%, n = 86) and maine (58%, n = 62) (p = 0.87), and among years in both states (new hampshire: χ2 = 2.60, p = 0.27; maine: χ2 = 0.43, p = 0.80) (table 1). no births by yearling cows (n = 17) or twinning by adults (n = 148) was documented in either state; table 1. annual and total observed fecundity rates of radio-collared adult cow moose in northern new hampshire and western maine, during 2014–2016. sample sizes are given in parentheses. rates did not differ by year or study area (p > 0.05). adult cow productivity % pregnancy rate calving rate successive calving rate new hampshire 2014 76 (21) 67 (21) na 2015 75 (16) 45 (33) 29 (17) 2016 88 (8) 59 (32) 18 (28) all years 78 (45) 56 (86) 22 (45) maine 2014 90 (20) 55 (20) na 2015 75 (4) 55 (20) 55 (11) 2016 na 64 (22) 25 (12) all years 88 (24) 58 (62) 39 (33) combined 81 (69) 57 (148) 24 (78) 90 nh & me productivity – jones et al. alces vol. 53, 2017 thus, the fecundity rate was equivalent to the calving rate. the successive calving rate was 22% (n = 45, range = 18–29%) in new hampshire and 39% (n = 33, range = 25–55%) in maine; the successive calving rate did not differ between years (new hampshire: χ2 = 0.29, p = 0.59; maine: χ2 = 1.04, p = 0.31) or study area (p = 0.60), and overall was 24% (n = 78) (table 1). unmarked calf survival calf survival to 60 days averaged 77% (n = 47) in new hampshire and 94% (n = 36) in maine (fig. 2). there was no difference in annual survival at either site, ranging from 64–87% in new hampshire (χ2 = 2.04, p = 0.36) and 91–100% in maine (χ2 = 1.35, p = 0.51). although survival was not different (p = 0.06) between study sites and the combined rate was 86%, survival in maine was 17% higher. nearly all mortality in new hampshire (82%) and maine (100%) occurred within 7 d post-birth when it was higher than in the remainder of the 60-d period (nh: χ2 = 9.0, p = 0.01; me: analysis precluded). discussion calf survival, pregnancy, and successive calving rates were higher in maine and could be considered biologically different resulting in a population trajectory ∼5% higher than that in new hampshire. however, we do not believe that the populations are measurably different because 1) the method for detecting calves in maine probably resulted in maine calves being a few days older (with higher summer survival) at first detection than in new hampshire, 2) the predominance of pregnancy data originated from only a single year in maine, and 3) substantial variation in the 2 years of successive calving data reduces the potential accuracy of these parameters to indicate any difference between the study sites. importantly, the new hampshire and maine study sites demonstrated similar trends of low adult fecundity and high summer calf survival, suggesting that a combined dataset is representative of the larger study area, and that the moderate variation in specific parameters was due to inherent variance associated with methodology and sample size. fig. 2. kaplan-meier survival curve for unmarked calf moose from birth to 60 days old collected at study sites in northern new hampshire and western maine, 2014–2016. alces vol. 53, 2017 jones et al. – nh & me productivity 91 most calf mortality in new hampshire (82%) and all in maine occurred in the first few days of life (≤7 days post-partum) when calves are most susceptible to predation by black bears (franzmann and schwartz 1985). black bear density within the study areas was high enough to incur predation (ballard 1992), but calf survival was much higher than in populations held at low density by predation (76% loss in weeks 0–8; gasaway et al. 1992). calf survival in maine (94%) was higher than in populations in scandinavia with minimal predation (87%; ericsson et al. 2001), further suggesting that early calf survival estimates in maine were probably artificially high. calf survival in new hampshire was similar to that measured previously (71%; musante et al. 2010) and the combined new hampshire and maine calf survival rate (88%) was similar to that in scandinavia (ericsson et al. 2001). again, this survival estimate is probably biased high due to the later detection date of certain maine calves; albeit, if so, the maine calving rate was underestimated. the median parturition date (19 may) and predominant calving season (14–25 may) were similar to those identified previously in new hampshire (musante et al. 2010) and central ontario (addison and mclaughlin 1993), but earlier than in alaska (testa et al. 2000b, bertram and vivion 2002). the timing and synchronous birthing pulse (90% in 10 d) occurs across north america (keech et al. 2000, testa et al. 2000b) and has been hypothesized as a response to maximize use of high quality forage in summer (bowyer et al. 1998) and to minimize the influence of predation (testa et al. 2000b). forage availability is probably not as important in the northeastern united states given the longer growing season relative to alaska. our productivity measures were simi‐ lar to those associated with a declining population, but these parameters often vary regionally because they reflect unique and multiple conditions in moose populations (table 2). the average pregnancy rate was similar to the north american average (84%; boer 1992), but the yearling pregnancy and adult twinning rates were low compared to other declining populations (boer 1992, murray et al. 2006, lenarz et al. 2010). yearling pregnancy and adult twinning rate are directly related and considered indicators of the relative nutritional status in moose populations (franzmann and schwartz 1985, boer 1992), and reflect habitat quality and body weight (adams and pekins 1995). the decline in yearling pregnancy from 20% in 2002–2005 (musante et al. 2010) to 0% in 2014–2016 corresponds with declines in ovulation rate (proximate measure for yearling pregnancy) and dressed body weight measured from 1998 to 2009 in new hampshire (adams and pekins 1995, bergeron et al. 2013). the lack of twinning in 2014–2016 also corresponds with a decline in the corpora lutea count since 2002–2005 in new hampshire, suggesting a similar, but more subtle decline in phy‐ sical condition of adult cows (bergeron et al. 2013). it is hypothesized that the lack of reproduction by yearling cows is caused by high annual infestations of winter ticks on calf moose, and consequently, reduced fit‐ ness and fecundity of surviving yearlings (musante et al. 2010, bergeron et al. 2013). calf moose that survive an epizootic event (>50% mortality) are likely in poor condition (mclaughlin and addison 1986) which will be reflected in lower body weight, ovulation rate, and productivity as yearlings (peterson 1977, saether and heim 1993, keech et al. 1999). adult cows with high winter tick infestations are also presumed to be in poorer condition due to the compounding ener‐ getic costs associated with winter ticks, ges‐ tation, and lactation while consuming a protein-deficient diet until spring green-up 92 nh & me productivity – jones et al. alces vol. 53, 2017 occurs 2–3 weeks post-birth (musante et al. 2007, schwartz and renecker 2007). reduced productivity in adult cows is consistent with the low rates of calving and successive calving measured in 2014–2016. decline in physical condition during late winter and early spring that stems from parasites is analogous to decline in physical condition that influences productivity following years of deep snow (mech et al. 1987) or harsh winter conditions (colder temperatures, greater maximum depth and duration of snow pack and shorter growing season; sand 1996). reduced productivity from a decline in physical condition of adult cows is expressed by low twinning rates (mech et al. 1987, sand 1996), lower calf: cow ratios in autumn (rolley and keith 1980), and reduced weight and viability of 9-month old moose calves (peterson et al. 1982). because prenatal mortality is low in deer and moose (verme and ulrey 1984), these parameters likely reflect still births or undersized and behaviorally abnormal calves predisposed to mortality (keech et al. 2000). the potential effect of winter on productivity during the subsequent year is attributed to the inability of cows to fully recover or compensate in one year (mech et al. 1987). the energetic costs of prior gestation and lactation are evident in maternal adult cows that have less fat, lower pregnancy rates, and smaller embryos in autumn, and body condition in autumn is positively correlated with pregnancy and calving rate, and negatively with reproductive losses and neonatal mortality the following spring (testa and adams 1998). these relationships illustrate that adult cows must compensate for the demands of pregnancy and lactation, and may require a year to recover which reduces productivity in the population (mech et al. 1987). poor table 2. productivity measures from moose populations across the southern range of moose in north america. pregnancy rates were from serum progesterone levels, calving rates and twining rates from direct observation, except for north american averages where pregnancy rates were from intrauterine counts, and twinning rates from direct observation and intrauterine counts. twinning rates are for yearling and adult cows. location population change adult pregnancy rate yearling pregnancy calving rate twinning rate source new hampshire and maine decreasing 82 0 57 0 this study northwestern minnesota decreasing 48 <20 45 19 murray et al. 2006 north america average decreasing 84 18 – 5 boer 1992 northeastern minnesota decreasing – – 78 – lenarz et al. 2010 norway n/a – – 77 – stubsjøen et al. 2000 new hampshire 2002-2005 stable 85 20 75 11 musante et al. 2010 upper peninsula michigan stable/ increasing 74 – 65 19 dodge et al. 2004 southern ontario increasing 87 2 – 17 murray et al. 2012 north america average 84 49 – 33 boer 1992 alces vol. 53, 2017 jones et al. – nh & me productivity 93 body condition in late winter-early spring may affect productivity 3 years later because calves born to cows in poor condition are smaller and remain so (peterson 1977, keech et al. 2000), reducing their survival and increasing age to sexual maturity (mech et al. 1987, saether and heim 1993, keech et al. 1999). the large difference between our measured pregnancy rate (82%) and calving rate (57%) can be attributed to a variety of outcomes including resorption, still birth, or undersized and behaviorally abnormal calves predisposed to mortality; however, the specific outcome for any individual could not be determined. adult cows in poor condition produce smaller neonates that experience higher mortality and slower development (peterson 1977, keech et al. 2000). certainly a portion of neonatal mortality could be attributed to the compromised condition of calves as a consequence of marginalized adult cows, given the clear association of frequent, high winter tick infestations and the declining condition and productivity in this population. similarly, the low successive calving rate (24%) suggests that maternal costs and inadequate compensatory growth prevented subsequent reproduction. the successive calving rate was 3x higher (75%) a decade earlier when a single epizootic occurred in a 4-year period in new hampshire (musante et al. 2010). the effect of high winter tick infestations on body condition and productivity is analogous to malnutrition associated with poor habitat (albon et al. 1983, albright and keith 1987). but an important distinction is that such malnutrition affects physical nutrition in all age classes (peterson 1977, skogland 1983, messier and crête 1984), induces starvation before old age (bergerud et al. 1983), and high browsing rates are evident (albright and keith 1987). field-dressed body weight, antler dimensions, and the onset of sexual maturity are correlated with the physical condition of moose (schwartz and hundertmark 1993, schmidt et al. 2007). since 1988 these measurements have declined measurably in yearling moose throughout the region, yet only slight downward trends have occur‐ red in adult age classes (bergeron et al. 2013, andreozzi et al. 2015). yearling condition reflects, in part, condition as a calf. the marked decline in body weight and productivity of yearling cows, yet subtle decline in other age classes, reflects the negative effects of frequent epizootics on productivity (bergeron et al. 2013). neither this or the previous study (musante et al. 2010) in the same area attributed any mortality to starvation (>350 radio-collared animals), and browsing intensity on the landscape is considered low-moderate overall (bergeron et al. 2011). further, the rate of annual forest harvest in the region has been relatively stable at 1–3% of the landscape since 1990, providing continuous availability of optimal foraging habitat (15–20% of the landscape in 4–16 year-old regenerating forest) through both growth and decline of the moose population (ball 2017). moose are at the southern edge of their continental range in this region and snow depth of 70 cm that impedes moose mobility and >90 cm that confines movement and increases mortality (coady 1974) occur infrequently. the winter of 2016 had nearly snowless conditions and temperatures slightly above normal (0.3°c, ncdc weather data), yet an epizootic occurred and the trend of low productivity continued. given the low adult productivity measured here, and the lack of productivity and deterred growth of yearlings, yet the lack of starvation and the constant production of optimal foraging habitat, it is evident that the winter tick, not habitat, is the predominant influence on this regional moose population. con‐ tinued slow decline in this population is predicted if the frequency of epizootics remains high. 94 nh & me productivity – jones et al. alces vol. 53, 2017 acknowledgements funding for this research was provided through nhfg and mifw in cooperation with the united states fish and wildlife service wildlife and sport fish restoration program, the university of new hampshire, and safari club international foundation. lighthawk llc and the maine forest service were key partners for completing aerial censuses. this research would not have been possible without access to property owned by american forest management, the conservation fund, e. j. carrier inc., gmo renewable resources, hilton timberlands llc, maine bureau of parks and lands, plum creek, moosehead wildlands inc., t. r. dillon inc., and wagner forest management ltd. the efforts of m. o’neal and s. mclellan in maine, and numerous student technicians in both states were critical for the collection of these data. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53–59. addison, e. m., and r. f. mclaughlin. 1993. seasonal variation and effects of winter ticks (dermacentor albipictus) on consumption of food by captive moose (alces alces) calves. alces 29: 219–224. albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of newfoundland. canadian field naturalist 101: 373–387. albon, s. d., b. mitchell, and b. w. staines. 1983. fertility and body weight in female red deer : a density-dependent relationship. journal of animal ecology 52: 969–980. andreozzi, h. a., p. j. pekins, and l. e. kantar. 2015. analysis of age, body weight and antler spread of bull moose harvested in maine, 1980–2009. alces 51: 45–55. ball, k. r. 2017. moose and winter tick epizootics in northern new england’s changing climate. m.s. thesis, university of new hampshire, durham, new hampshire, usa. ballard, w. b. 1992. bear predation on moose: a review of recent american studies and their management implications. alces supplement 1: 162–276. bergeron, d. h., p. j. pekins, h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study. alces 47: 39–51. ———, ———, and k. rines. 2013. temporal assessment of physical characteristics and reproductive status of moose in new hampshire. alces 49: 39–48. bergerud, a. t., w. wyett, and b. snider. 1983. the role of wolf predation in limiting a moose population. the journal of wildlife management 47: 977–988. bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior alaska. the jorunal of wildlife management 66: 747–756. boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1: 1–10. bowyer, r. t., v. van ballenberghe, and j. g. kie. 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79: 1332–1344. broders, h. g., a. b. coombs, and j. r. mccarron. 2012. ectothermic res‐ ponses of moose (alces alces) to thermoregulatory stress on mainland nova scotia. alces 48: 53–61. coady, j. w. 1974. influence of snow on behavior of moose. le naturaliste canadian 101: 417–436. crossley, a. 1987. summer pond use by moose in northern maine. m. s. thesis, university of maine, orono, maine, usa. alces vol. 53, 2017 jones et al. – nh & me productivity 95 degraaf, r. m., m. yamasaki, w. b. leak, and j. w. lanier. 1992. new england wildlife: management of forested habitats. general technical report ne-144. radnor, pa: united states department of agriculture, u. s. forest service, northeast forest experiment station, radnor, pennsylvania, usa. dodge, w. b. jr., s. r. winterstein, d. e. beyer, jr., and h. campa iii. 2004. survival, reproduction and movements of moose in the western upper peninsula of michigan. alces 40: 71–85. ericsson, g., k. wallin, j. p. ball, and m. broberg. 2001. age-related reproductive effort and senescence in freeranging moose, alces alces. ecology 82: 1613–1620. franzmann, a. w., and c. c. schwartz. 1985. moose twinning rates : a possible population condition assessment. journal of wildlife management 49: 394–396. ———. 2000. moose. pages 578–600 in s. desmarais and p. r. krausman, editors. ecology and management of large mammals of north america. prentice hall, upper saddle river, new jersey, usa. gasaway, w. c., r. d. boertje, d. v grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. jones, h. f. 2016. assessment of health, mortality, and population dynamics of moose in northern new hampshire during successive years of winter tick epizootics. m. s. thesis, university of new hampshire, durham, new hampshire, usa. keech, m. a., r. d. boertje, r. t. bowyer, and b. w. dale. 1999. effects of birth weight on growth of young moose: do low-weight neonates compensate? alces 35: 51–57. ———, t. r. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. the journal of wildlife management 64: 450–462. lankester, m. w. 2010. understand‐ ing the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53–70. larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rates of moose mortality in southwest yukon. journal of wildlife management 53: 548–557. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. mcgraw, a. m., j. terry, and r. moen. 2014. pre-parturition movement patterns and birth-site characterisitics of moose in northeast minnesota. alces 50: 93–103. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. mech, d. l. 1983. handbook of animal radio tracking. university of minnesota press, minneapolis, minnesota, usa. ———, r. e. mcroberts, r. o. peterson, and r. e. page. 1987. relationship of deer and moose populations to previous winter's snow. journal of animal ecology 56: 615–627. messier, f., and m. crête. 1984. body condition and population regulation by food resources in moose. oecologia 65: 44–50. murray, d. l., e. w. cox, w. b. ballard, h.a. whitlaw, m.s. lenarz, t.w. custer, t. barnett, and t.k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining 96 nh & me productivity – jones et al. alces vol. 53, 2017 moose population. wildlife monographs 166: 1–30. ———, k. f. hussey, l. a. finnegan, s. j. lowe, g. n. price, j. benson, k. m. loveless, k. r. middel, k. mills, d. potter, a. silver, m-j. fortin, b. r. patterson, and p. j. wilson. 2012. assessment of the status and viability of a population of moose (alces alces) at its southern range limit in ontario. canadian journal of zoology 90: 422–434. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. ———, ———, ———. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. national climate data center (ncdc). 2016. climate data online daily summaries january-december, 2014–2016. (accessed january 2017). peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. national park service scientific monograph series no. 11. u. s. government printing office, washington, d. c., usa. ———, j. m. schelder, and p. w. stephen. 1982. selected skeletal morphology and pathology of moose from the kenai peninsula, alaska and isle royale, michigan. canadian journal of zoology 60: 2812–2817. rolley, r. e., and l. b. keith. 1980. moose population dynamics and winter habitat use at rochester, alberta, 1965–1979. canadian field-naturalist 94: 1–9. saether, b-e., and m. heim. 1993. ecological correlates of individual variation in age at maturity in female moose (alces alces): the effects of environmental variability. journal of animal ecology 62: 482–489. samuel, w. m. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. sand, h. 1996. life history patterns in female moose (alces alces): the relationship between age, body size, fecundity and environmental conditions. oecologia 106: 212–220. schmidt, j. i., j. m. ver hoef, and r. t. bowyer. 2007. antler size of alaskan moose alces alces gigas: effects of population density, hunter harvest and use of guides. wildlife biology 13: 53–65. schwartz, c. c.2007. reproduction, natality, and growth. pages 141–145 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. ———, and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. the jorunal of wildlife management 57: 454–468. ———, and l. a. renecker. 2007. nutrition and energetics. pages 441–478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. skogland, t. 1983. the effects of den‐ sity dependent resource limitation on size of wild reindeer. oecologia 60: 156–168. stubsjøen, t., b. e. saether, e. j. solberg, m. heim, and c. m rolandsen. 2000. moose (alces alces) survival in three populations in northern norway. canadian journal of zoology 78: 1822–1830. testa, j. w., and g. p. adams. 1998. body condition and adjustments to reproductive effort in female moose (alces alces). journal of mammalogy 79: 1345–1354. ———, e. f. becker, and g. r. lee. 2000a. movements of female moose in relation to birth and death of calves. alces 36: 155–162. alces vol. 53, 2017 jones et al. – nh & me productivity 97 https://www.ncdc.noaa.gov/cdo-web/datasets#ghcndms� https://www.ncdc.noaa.gov/cdo-web/datasets#ghcndms� ———, ———, ———. 2000b. temporal patterns in the survival of twin and single moose (alces alces) calves in southcentral alaska. journal of mammalogy 81: 162–168. van ballenberghe, v., and w. b. ballard. 2007. population dynamics. pages 223–246 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. verme, l. j., and d. e. ullrey. 1984. physiology and nutrition. pages 91–118 in l. k. halls, editor. white-tailed deer ecology and management. stackpole, harrisburg, pennsylvania, usa. 98 nh & me productivity – jones et al. alces vol. 53, 2017 fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics study area methods capture and marking fecundity monitoring of unmarked calves analysis results fecundity unmarked calf survival discussion acknowledgements references 4006.p65 alces vol. 40, 2004 young and boertje moose calf hunts 1 initial use of moose calf hunts to increase yield, alaska donald d. young jr. and rodney d. boertje alaska department of fish and game, 1300 college road, fairbanks, ak 99701-1599, usa abstract: in 2002 the board of game authorized alaska’s first permit hunts specifically for calf moose (alces alces). we promoted these calf hunts to help stabilize a high-density, food-stressed moose population and to compensate for declining harvests of bulls. low harvest rates of cows (= 1% of the prehunt cow population, 1996–2001) were tightly controlled by the public. high harvest rates of bulls (21–26% of the prehunt bull population, 1995–1999) resulted in bull:cow ratios declining below the management objective of 30:100. to conserve bulls, the previous bag limit of any bull was changed to bulls with specific antler configurations. simultaneously, 300 calf drawing permits were made available in 7 different hunt areas with the allocation of permits based on estimated moose densities within individual hunt areas. we issued 274 permits, but 61% of the permittees did not participate, in part to protest the hunt. of 108 hunters, 33 reported taking a calf. the harvest accounted for about 1.3% (33/2,500) of the estimated prehunt calf population and 7% (33/471) of the total reported harvest. the calf harvest contributed only marginally to meeting the game management unit 20a harvest mandate of 500–720 moose. we observed decreasing acceptance of calf hunts and increasing acceptance of cow hunts during 2002 and 2003. in 2004 we expect to substantially increase the harvest of cows and calves using registration and late season hunts and continuing education programs. we deem gaining public acceptance of cow and calf hunts in increasing, food-stressed alaska moose populations to be a long-term, challenging, yet worthwhile endeavor. alces vol. 40: 1-6 (2004) key words: alaska, alces alces, calf, harvest, yield calf moose (alces alces) hunts have been used as a management tool to achieve a wide range of objectives in the united states, canada, and scandinavia. for example, alberta, british columbia, newfoundland, idaho, montana, north dakota, utah, and wyoming have all used limited antlerless permits to harvest cows and calves in selected management units to maintain balanced sex ratios and provide additional hunting opportunities (timmermann and buss 1998). in ontario, where there is no harvest restriction on calves and every licensed hunter is eligible to take a calf (hooper and wilton 1995), increased calf harvests, concurrent with reduced bull and cow harvests, were used as a tool aimed at doubling the provincial moose population (timmermann and whitlaw 1992, timmermann and rempel 1998). in scandinavia, where calves comprise 40% of the total harvest each year (h.r. timmermann, ontario ministry of natural resources, personal communication) maximizing harvest is paramount. in 2002 we promoted the first alaskan drawing hunts for calf moose. moose density was high and slowly increasing in this area, yet twinning rates had declined to low levels indicating the population should be stabilized (gasaway et al. 1992:24). also, harvest of bulls needed to become more restrictive, because bull:cow ratios were declining. reductions in total moose harvest in this area have legal ramifications under an intensive management law. ramifications include consideration of predator moose calf hunts young and boertje alces vol. 40, 2004 2 control programs and habitat improvement projects. while promoting calf hunts, we continued our decade-long encouragement for prescribed burns to rejuvenate habitat and improve moose productivity. prior to 2002 and after 2003 in alaska, bulls were defined as any male moose, so male calves were legal bulls, and calves were also legal in the relatively few antlerless hunts. in discussions with area biologists statewide we found calf harvest in alaska to be a small portion (< 5%) of the annual harvest. calves were taken incidentally each year but demand was very low. the drawing hunts we proposed for calves posed a different problem because state law requires that the majority of citizens’ advisory committees, residing in or adjacent to the management area, approve of antlerless hunts prior to the board of game’s vote. we successfully argued for calf hunts in 2002 and 2003, despite a board initiated 2year ban on calf hunts statewide (except approved permit hunts) in 2003. we also successfully argued to rescind this statewide ban in 2004, because there was no biological justification for the ban and it was an impediment to effective management. our objective here is to document alaska’s initial use of permit hunts specifically for calf moose. we discuss the problems encountered and make recommendations regarding the use of calf hunts to increase yield and hunting opportunity. study area our study area encompassed game management unit (gmu) 20a immediately south of fairbanks and across the tanana river. the study area is in interior alaska and is centered around 64°10'n latitude and 147°45'w longitude. game management unit 20a encompasses 17,000 km2, but only 13,044 km2 is below the upper limits of vegetation characteristically used by moose. gasaway et al. (1983) and boertje et al. (1996) described the physiography, habitat, climate, major predator and prey species, and moose population status and harvest from 1963 through 1994. the moose population peaked at an estimated 23,000 in 1965, likely due to largescale burns in the early 1940s and extensive predator control in the 1950s. the population declined to approximately 2,800 in 1975 because of a series of bad winters, accompanying high predation, and overharvest. the population increased to 11,000–13,000 by 1995 due to hunting restrictions, periodic wolf control, and sustained wolf harvest. moose numbers increased more slowly through 2003 as twinning rates declined. in november 2003, gmu 20a had the highest moose density in alaska for any equivalentsized area. we estimated 16,446 moose ±2,365 (90% ci) in 13,044 km2 of moose habitat. methods for estimating moose numbers included the use of spatial statistics and a sightability correction factor of 1.12 (gasaway et al. 1986, ver hoef 2001). we also documented the lowest moose twinning rates in alaska (0-18%) during 1993–2003 (boertje et al. 1996, 2000). twinning rates were higher (32–40%) during 1979–1983, when moose density was relatively low (gasaway et al. 1983). during most years, we used transect surveys a few days after the median calving date to e s t i m a t e t w i n n i n g r a t e s . w e u s e d radiocollared moose to determine the median calving date, and, in a few years, to estimate twinning rates. moose seasons and bag limits in gmu 20a varied markedly in recent history. long seasons and hunts for both antlered and antlerless moose were common through the 1960s and early 1970s when moose numbers were high. following the low point in the population in 1975, hunting seasons were shortened to 10 days and bag limits limited to bulls-only. as moose numbers increased from the late 1970s through the alces vol. 40, 2004 young and boertje moose calf hunts 3 mid-1990s, seasons were progressively lengthened to as many as 25 days. antlerless hunts were resumed again in 1996, primarily to maximize harvest, but harvest of antlerless moose remained very low (60–75 cow moose, 1% of the prehunt cow population) through 2001 (except 1999 when 0 were harvested). high harvest rates (21– 26%) of the prehunt bull population from 1995 to 1999 resulted in bull:cow ratios declining below the management objective of 30:100. in 2000, the hunting season was shortened 5 days to reduce bull harvests. additionally, in 2002, antler restrictions were instituted to further reduce the harvest of bulls to a sustainable level. m o s t s u c c e s s f u l m o o s e h u n t e r s accessed gmu 20a by airplane, propeller/ jet boat, or atv/off-road vehicles and to a lesser extent via horses, airboats, and highway vehicles. less than 5% of gmu 20a was accessible by road, but seasonal military and mining trails provided access to the foothills in autumn and winter. the only significant human settlements occurred along the perimeter of the game management unit although 1 subdivision was near the center of the unit and remote cabins and airstrips were scattered throughout much of the unit. methods in 2002, 300 calf moose permits were available by lottery. applications were accepted only in may, as for most drawing hunts in alaska. the hunt period was 1–25 september. game management unit 20a was divided into 7 different hunt areas with 1 hunt area divided temporally into 2 hunts (1–13 and 14–25 september) for a total of 8 different hunts.the allocation of permits was based on estimates of calf moose numbers within individual hunt areas from population surveys (gasaway et al. 1986, ver hoef 2001) conducted the previous november (2001). these estimates were adjusted slightly upwards based on the 2002 trend in parturition rates from radiocollared moose in central gmu 20a. we also had 175 cow moose permits available by lottery in may for 2 hunt areas (3 different hunts) within gmu 20a. we compared the proportional use of cow permits with those of calf permits to assess whether holders of calf permits protested the calf hunt by not hunting. our goal in the first year was to have hunters harvest up to 5% of the prehunt calf population. based on our experience with drawing permit hunts for antlerless moose, we assumed harvest success rates would not exceed 50%. therefore, the number of permits made available for each hunt area equaled 10% of the estimated prehunt calf population. for example, in hunt area dm755, we estimated a prehunt population of 300 calves and issued 30 permits for an estimated harvest of up to 15 calves. successful applicants to the calf and cow lotteries were notified in early july. successful applicants for the calf hunts received a letter with their permit explaining: (1) how to distinguish a calf from a yearling; (2) estimated weight range of calves in september; and (3) safety tips regarding cows that may be aggressive after their calves were shot. successful applicants to both calf and cow lotteries were notified that they were prohibited from hunting for bulls in gmu 20a, which was intended to reduce hunting pressure on the bull segment of the population. all hunters were required to report to the alaska department of fish and game if they successfully harvested a moose, were unsuccessful, or did not hunt. results of the 300 calf permits in 2002, 274 were issued to hunters and 61% failed to hunt (table 1). this failure to hunt was significantly higher than the failure of cow moose calf hunts young and boertje alces vol. 40, 2004 4 permittees to hunt (z-value = 6.5, p < 0.0001). of the 175 permittees for cow hunts, 54 (31%) did not hunt. reported harvest of calves totaled 33 (14 male and 19 female). three of the 8 hunts were undersubscribed; remoteness of these hunt areas was likely a contributing factor. the calf harvest accounted for about 1.3% (33/2,500) of the estimated prehunt calf population and 7% (33/471) of total reported harvest. in contrast, the harvest of bulls accounted for about 14% (344/ 2,500) of the prehunt bull population and 73% (344/471) of the total harvest. the harvest of cows accounted for about 1.2% (94/7,600) of the prehunt cow population and 20% (94/471) of the total harvest. discussion the calf harvest contributed only marginally to meeting the harvest objective of 500–720 moose. this was primarily due to the poor participation in the calf hunts. based on conversations with hunters, letters to the editor of local newspapers, and comments on harvest report cards, it was apparent that a large number of hunters applied for the permits with no intention of using them. secondarily, the success rate of those that did hunt was lower than expected, likely because harvesting a calf was more difficult than hunters anticipated. cows with calves cannot be lured to a call, tend to be more wary and alert, and typically utilize heavier cover than bulls and barren table 1. calf moose harvest data by permit hunt for game management unit 20a, central tanana river valley, alaska, 2002. cows. the calf hunts were contentious, particularly among local citizens’ advisory committees and hunters. interestingly, there was no opposition by anti-hunting or animal rights organizations. rather, one group commented that they favored the hunts because they emulated natural mortality more closely than bulls-only hunts. primary arguments against the calf hunts were philosophical, cultural, and biological in nature. many individuals made ethical statements such as “you just don’t shoot calves,” or “it’s just not right to shoot calves.” a woman testified that after her son had shot a calf she felt “embarrassed”. some comments were more anthropomorphic in nature. for example, a note on a harvest report card in which the permit holder did not hunt stated simply “saved a calf”. some individuals implied that shooting a calf was cruel to the mother. significant opposition to the calf hunts also came from 2 native athabascan communities. at a public meeting in which the reauthorization of the calf hunts was being discussed, several elders stated that it was not their custom to hunt calves and that they disliked the taste of calf meat. another common argument was that calves provided little or no meat. however, we had provided information to the hunting public regarding estimated september weights of calves (135–190 kg) from studies conducted in central gmu 20a during hunt permits issued did not hunt % unsuccessful hunters % successful hunters % males % females % unk % harvest dm750 65 39 60 20 77 6 23 2 33 4 67 0 0 6 dm752 65 44 68 13 62 8 38 3 38 5 63 0 0 8 dm754 37 23 62 9 64 5 36 2 40 3 60 0 0 5 dm755 30 6 20 16 67 8 33 5 63 3 38 0 0 8 dm756 5 1 20 2 50 2 50 0 0 2 100 0 0 2 dm757 20 10 50 9 90 1 10 1 100 0 0 0 0 1 dm758 33 27 82 4 67 2 33 0 0 2 100 0 0 2 dm759 19 16 84 2 67 1 33 1 100 0 0 0 0 1 totals 274 166 61 75 69 33 31 14 42 19 58 0 0 33 alces vol. 40, 2004 young and boertje moose calf hunts 5 1997–2001. a male, twin calf shot on 16 september had a gutted weight of 102 kg, a dressed weight of 76 kg, and yielded 43 kg of meat (bones removed). the estimated live weight of that calf based on a standard formula applied to cattle (live weight = 2 × dressed weight) was 152 kg. the yield in meat from calf moose is similar to that of adult barren-ground caribou (rangifer tarandus), which, like moose, are hunted primarily for human consumption by alaskan residents. the potential danger associated with cows in defense of their downed calves was also a concern of hunters. many individuals were concerned that hunters would be injured or even killed by defensive cows when approaching a downed calf. others felt that a large number of aggressive cows would be shot, the meat left to rot, and that these incidents would go unreported. to the contrary, we received no reports, official or otherwise, of overly aggressive cows, cows being shot, or of any hunter being injured by an aggressive cow. in manitoba, v. crichton (manitoba department of natural resources, personal communication) reported that, to his knowledge, there have been no documented cases of injury to hunters or aggressive cows being shot during calf hunts. the main biological argument from the public against calf hunts was that the harvest of calves would eliminate future breeding stock and particularly bulls, which would lead to decimation of the moose population and hunting opportunity for bulls. this concern persisted despite our repeated explanations that: (1) only about 5% of the prehunt calf population would be harvested annually; (2) the moose population was food-stressed at the current density and was not declining; and (3) the harvest of calves was partially compensatory. the scenario often heard from the public was that predation is largely additive to harvest (gasaway et al. 1983; i.e., predators are killing healthy calves). merit exists in this argument, but we presented data that calf mortality was high (47%) compared with annual cow mortality of about 2% between the ages of 2 and 7 (boertje et al. 2000). therefore, cow harvest would more likely constitute additive mortality, whereas calf harvest may not, particularly at high density (euler 1983, timmermann and rempel 1998). in retrospect, more public education was needed to harvest significant numbers of calves, because no prior hunts in alaska had targeted calves and because shooting calves became highly controversial. similar public resistance occurred in the 1960s when attempts were made to introduce antlerless hunts (rausch et al. 1974). furthermore, antlerless hunts reinstated in gmu 20a in 1996 were also highly controversial. however, by 2002, antlerless hunts had gained popularity with the hunting public. thus, time may be central to gaining public acceptance of calf hunts. in 2004 we expect to test whether cow and calf harvests can be substantially increased with registration and late season hunts for antlerless moose hunts not specific to calves. registration hunts, unlike drawing permit hunts, are much less restrictive and would allow greater latitude in terms of the number of permits issued and season length. we envision issuing several thousand registration permits, requiring a short reporting period for successful hunters, and closing the hunt by emergency order once the desired harvest has been reached. we established quotas for each of the 7 hunt areas. this approach is much more costly but will allocate permits to those hunters willing to take a cow or calf rather than to those individuals desiring to “save a calf”. in addition, the antlerless season will potentially remain open into december, well beyond the 1–25 septemmoose calf hunts young and boertje alces vol. 40, 2004 6 ber season for bulls. thus, additional hunting opportunity will be provided when most big game hunts are closed. recommendations to reduce the moose population to the management objective of 10,000–12,000 moose, we are recommending registration and late season antlerless hunts to increase harvest and hunting opportunity while maintaining 30 bulls:100 cows. ultimately, we hope these hunts will improve public acceptance of calf hunts. once the population is reduced below 12,000 moose, we will attempt to sustain a harvest ratio of approximately 60 bulls:20 cows:20 calves. acknowledgements we wish to thank the many pilots that assisted with capture and survey flights and numerous wildlife technicians and volunteers that helped with moose survey flights. this study was funded by the alaska department of fish and game and federal aid in wildlife restoration. references boertje, r. d., c. t. seaton, d. d. young jr., m. a. keech, and b. w. dale. 2000. factors limiting moose at high densities in unit 20a. alaska department of fish and game. federal aid in wildlife restoration. grant w-27-3. study 1.51. juneau, alaska, usa. _____, p. valkenburg, and m. e. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal of wildlife management 60:474– 489. euler, d. 1983. selective harvest, compensatory mortality and moose in ontario. alces 19:148–161. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r.o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 92. _____, s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological paper 22, university of alaska. fairbanks, alaska, usa. _____, r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. hooper, c. a., and m. l. wilton. 1995. a selective moose hunt in south central ontario. alces 31:139–143. rausch, r. a., r. j. sommerville, and r. h. bishop. 1974. moose management in alaska. naturaliste canadien 101:705– 721. timmermann, h. r., and m. e. buss. 1998. population and harvest management. pages 559–615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, and r. s. rempel. 1998. age and sex structure of hunter harvested moose u n d e r t w o h a r v e s t s t r a t e g i e s i n northcentral ontario. alces 34:21–30. _____, and h. whitlaw. 1992. selective moose harvest in north central ontario – a progress report. alces 29:137– 163. ver hoef, j. m. 2001. predicting finite populations from spatially correlated data. pages 93–98 in proceedings of the section on statistics and the environment of the american statistical association, august 13 – 17, 2000, indianapolis, indiana, usa. ≥ 4207(41-48).pdf alces vol. 42, 2006 young et al. elevating moose harvests 41 intensive management of moose at high density: impediments, achievements, and recommendations donald d. young jr., rodney d. boertje, c. tom seaton, and kalin a. kellie alaska department of fish and game, 1300 college road, fairbanks, ak 99701-1599, usa abstract: in 1994, the alaska legislature passed legislation directing the board of game to identify big game prey populations where “intensive management” (im) would be used to attain and sustain high but fails to mention that antlerless hunts are key to achieving high levels of harvest. we discuss im for moose in game management unit (gmu) 20a through 2005, because gmu 20a has a unique history of predator management and currently supports the highest moose density for any equivalent-sized area in alaska. moose numbers in gmu 20a exceeded the im population objectives beginning in 1999, but to achieving im harvest objectives in gmu 20a: (1) negative public attitude toward antlerless moose hunts; (2) local citizen advisory committees have veto power over antlerless hunts; (3) bull:cow ratios management activities, and public education. despite these impediments, liberal antlerless harvests annual harvests reached the highest levels recorded for gmu 20a. to facilitate the management of high-density moose for high levels of harvest, we recommend: (1) elimination of advisory committee ment activities, research programs, and public education. alces vol. 42: 41-48 (2006) key words: alaska, alces alces, harvest, impediments, intensive management, moose, recommendations in 1994, the alaska legislature mandated that the board of game (board) establish population and harvest objectives for intensive populations with the purpose of achieving high levels of harvest (alaska statutes 2005). ing “intensive management,” “harvestable big game prey populations,” and “sustained yield.” the intent of the legislation was to direct the board to choose areas where predator and habitat management would be used to attain and sustain high levels of harvest. in other areas, moose would be managed less intensively and for other purposes. hundertmark and schwartz (1996) provided a critical review of the concept of im for moose (alces alces) in alaska. they interpreted that the im legislation directed management for maximum sustained yield and they recommended managing at densities above maximum sustained yield. they also discussed the problems and expense involved with implementing cow harvests and managexamples. this paper differs in that we discuss im elevating moose harvests young et al. alces vol. 42, 2006 42 moose where we have ultimately been successful in managing for the highest levels of harvest compared to any equivalent-sized area in alaska today. thus, we discuss achievements in managing for high levels of harvest rather than simply impediments. this case history focuses on moose in game management unit (gmu) 20a in interior alaska, primarily since 1998 (when im was authorized in gmu 20a), but we also review regulatory and biological events leading up to im. in most im areas, moose populations are below established population objectives and the immediate challenge is to raise moose numbers to higher levels using predator control and habitat management. however, in gmu 20a the moose population surpassed the population objective in 1999, yet the harvest objectives were not reached (using reported harvest) during 2002-2005. gmu 20a is an important case history for 3 additional reasons including: (1) the highest moose density for any equivalent-sized area in alaska; (2) a history of periodic state wolf (canis lupus) control to elevate moose numbers; and (3) a history of high predator (wolves, black bears [ursus americanus], and grizzly bears [ursus arctos]) harvests, particularly of wolves by trapping. study area our study area encompassed gmu 20a encompasses 17,000 km2, but only 13,044 km2 contains topography and vegetation characteristically used by moose. gasaway et al. (1983), boertje et al. (1996), and keech et al. (2000) described the physiography, habitat, climate, major predator and prey species, and moose population status from 1963 to 1997. the only the perimeter of the game management unit, although one subdivision is near the center of the unit, and remote cabins and airstrips are scattered throughout much of the unit. less than 5% of gmu 20a is accessible by road, but seasonal military and mining trails provide access to most of the area in winter after the rivers freeze; usually november in recent years. regulatory and biological history since its passage in 1994, the im law has been the primary force behind increasgmu 20a, the board set the population and annual harvest objectives at 10,000-12,000 and 300-500 moose, respectively, in 1998 based on recommendations from the alaska the board increased the harvest objectives to 500-720 moose in 2001 and to 1,400-1,600 moose in 2004 based on recommendations the following history should help place these objectives in perspective. the moose population in gmu 20a increased to an estimated 23,000 moose in the early 1960s 1940s, federal predator control in the 1950s, and low bull-only harvests (rausch et al. 1974, gasaway et al. 1983). a dramatic population decline to an estimated 2,800 moose occurred by early winter 1975. causes for the decline included at least 5 harsh winters between 1961-1962 and 1974-1975, accompanying high predation, and excessive cow harvests during 1971-1974 (gasaway et al. 1983). managers had underestimated the effects of predation and the severity of the decline and mistakenly advocated cow hunts to improve birth rates. these ill-timed and misguided cow hunts led to legislation that authorized local citizen advisory committees to hold veto power over antlerless hunts. lation growth ensued from 1976 through 2003. causes for the increase included state wolf control (1976-1982, 1993-1994), public alces vol. 42, 2006 young et al. elevating moose harvests 43 harvest of predators, mostly conservative bullonly harvests, and nearly 3 decades of mostly mild winters (boertje et al. 1996, national weather service 1974-2004). by november 2004, gmu 20a had the highest moose density in alaska for any equivalent-sized area. we estimated 16,800 moose (14,980-18,650; 90% ci) in 13,044 km2 of moose habitat. methods for estimating moose numbers included the use of spatial statistics (ver hoef 2001) and a sightability correction factor of 1.20 (gasaway et al. 1986). estimates during parametric empirical bayes estimates (ver hoef 1996:1048). we suggest that the combined harvests of wolves, grizzly bears, and black bears likely contributed to higher survival rates (keech et al. 2000) and high densities of moose in gmu 20a. average annual reported harvest of wolves during 1976-2001 (not including was 42, and percent of the autumn population killed ranged from 12% (1980 and 1985) to 50% (2000; boertje et al. 1996; young 2000, 2003). average annual reported harvest of grizzly bears during 1976-2001 was 15 bears. the increase in average annual harvests from 10 grizzly bears, 1976-1979, to 17 bears, 1980-1991, reportedly led to a population decline by 1992 (reynolds 1999). average annual reported harvest of black bears 19762001 was 41, but harvest was highly variable among years (range 14-64). moose seasons and bag limits in gmu 20a have varied markedly in recent history. harvests of both antlered and antlerless moose were common through the 1960s and early 1970s when moose numbers were high, but total harvests were conservative (1 4% of 2005. elevating moose harvests young et al. alces vol. 42, 2006 44 prehunt numbers) except during 1971-1974 (6 19% of prehunt numbers; gasaway et al. 1983:25). in 1975, only bulls could be hunted until 1996 and, initially, seasons were shortened to 10 days. as moose numbers increased from the late 1970s through the mid-1990s, seasons were progressively lengthened to as many as 25 days. high harvest rates (21 26%) of the prehunt bull population from 1995 to 1999 resulted in bull:cow ratios declining below the management objective of 30:100 in 1999. experience has shown that with a bull:cow with the low opportunity to encounter a bull and particularly a mature bull. in 2000, the bull-only hunting season was shortened from 25 to 20 days to reduce harvests. in 2002, antler restrictions (i.e., harvest limited to bulls having spike or forked antlers, antlers having a width equal to or greater than 50 inches, or, at least one brow palm; schwartz et al. 1992) were instituted to further reduce the harvest of bulls to a sustainable level. during 1996-2001 (except 1999) antlerless hunts resumed but at very low levels (61-76 cow moose, 1% of the prehunt cow proval from local citizen advisory committees allowed antlerless hunts to be liberalized from a drawing hunt (300 permits) with a 25-day season to a registration hunt (unlimited permits) with a 101 day season. to more effectively distribute the antlerless harvest across the unit, gmu 20a was divided into 7 different hunt impediments to elevating moose harvests antlerless hunts: negative public attitudes and advisory committee veto power ing conservative antlerless hunts be resumed in gmu 20a to help slow moose population growth and to increase harvest, but public opposition remained strong, based on experiences from the early 1970s. the affected advisory committees were not supportive. by law, antlerless hunts require majority support annually from affected advisory committees and 4 advisory committees have jurisdiction over gmu 20a. advisory committees several times a year for nearly a decade and discussing the opportunities for increasing harvest, antlerless hunts were resumed on a limited drawing basis (300 permits) in 1996. the change occurred after initial disapproval by the 4 advisory committees in 1996. we then wrote an editorial in the local newspaper strongly advocating antlerless harvests and we thoroughly informed the board of this reoccurring dilemma. the advisory committees subsequently reversed their decisions in time for the 1996 hunt. still, the hunts were not popular with the hunting example, the average annual antlerless harvest was only 68 (1996-1998 and 2000-2001). also, the antlerless hunt was not held in 1999 because the advisory committees desired that show unequivocally that the moose population was increasing. harvest of cows was low because most permittees used the permits only if they did not shoot a bull. therefore, in 2002 the board changed the moose hunting regulations such that hunters with gmu 20a antlerless permits were prohibited from shooting a bull moose in that unit. as a result, the antlerless harvest increased to 94. also in 2002, a calf hunt was initiated, but interest was limited (young and boertje 2004). although 300 permits were available, the calf hunt was undersubscribed and only 275 permits were issued in 2002 and only 217 in 2003. harvest was also low, with only 32 and 24 calves harvested in 2002 and 2003, respectively. in addition, in 2002 a limited registration hunt with a quota of 20 alces vol. 42, 2006 young et al. elevating moose harvests 45 antlerless moose was initiated in the western to meet subsistence needs. these changes in the antlerless permit hunt conditions, along with the new calf and registration hunts in the 165 antlerless moose in 2003. the antlerless harvest, however, was still well below that estimated to curtail population growth. and increasing advisory committee support, in 2004 and 2005. the hunt changed from a limited drawing with 300 permits being issued to open registration with over 5,400 permits issued. as a result, reported harvest increased to 600 antlerless moose in 2004 and 690 in 2005. however, even with overwhelming biological evidence (boertje et al. 2007) and support for antlerless hunts from 3 advisory committees in 2004, one affected advisory committee rejected the gmu 20a hunt. if one other affected advisory committee had failed to support the hunt, the hunt would have been cancelled. maintaining bull:cow ratios most hunters prefer to harvest a bull, but meeting the im harvest objectives with of bulls by about 300 in 2002 and 2003 to recover declining bull:cow ratios, the im harvest objectives of 500-720 moose could not be met with bull-only hunting. also, it is not possible to attain current im harvest objectives (1,4001,600 moose) with bull-only hunting. access issues differential access across the unit affected the spatial and temporal distribution of the vehicle trail densities in zones 1 and 3, areas with high moose densities were accessed easily and desired harvests were quickly reached and, at times, exceeded. in contrast, given that access was essentially limited to aircraft in zone 5, the harvest objective of 120 moose was unmet with only 22 antlerless moose harvested. zone 2 presents unique challenges in that some high-density areas received heavy remote portions of the zone, harvest was low because of poor boat access. related to differential access was the temporal distribution of the harvest. in most portions of the unit, access limitations, hunter densities, and harvest quotas worked in hunt zones in a timely fashion to prevent overharvest. in contrast, in zones 1 and 3 where access was excellent and hunter densities were extremely high, harvests occurred the hunt in time to prevent surpassing the established quota. the harvest objective for zones 1 and 3 combined was 100 antlerless moose, yet nearly 200 moose were taken. this excessive harvest was the result of an early season opening that coincided with a long holiday weekend, a harvest reporting period that was too long (3 days), and a policy of 2areas with excellent access and high hunter densities is a challenge. social issues — in general, local hunters dislike non-local hunters hunting moose in their “backyard”. of the 3,008 hunters that reported hunting moose in gmu 20a in 2004, 42% were non-local hunters (i.e., resided outside of interior alaska). as a result, local residents were crowded by non-local hunters, which destabilized long-term hunting patterns (i.e., traditional hunting camps). the result was a high level of dissatisfaction by local hunters, who can — hunter densities were extremely high in the more elevating moose harvests young et al. alces vol. 42, 2006 46 accessible portions of the unit, particularly zones 1 and 3. although season lengths differed between years, the number of hunters that reported hunting moose in gmu 20a increased from 1,189 in 2003 to 3,008 in 2004, an increase of 153%. this resulted in considerable congestion at roadside pullouts, camping areas, trailheads, trails, and other accessible areas. trespass and garbage to human waste complaints. these complaints occurred on most private lands in gmu 20a, but particularly western zone 3. parking and camping spots were overrun. landowners in alaska typically do not post their lands, so hunters were usually unaware they were parking or camping on private property. landowners in the more remote gold king subdivision, located in the central portion of the unit, had additional complaints about moose gut-piles. gut-piles were potential attractants to black and grizzly bears and therefore posed a safety concern to subdivision residents. hunter-landowner to maintain support for intensive harvests of moose. higher incidence of illegal kills — although we do not have reliable numbers to compare illegal take among years, we hypothesize that illegal take was higher in 2004 with the registration permit hunt than it had been with limited drawing permit hunts. in addition, local hunters were convinced that illegal take had increased. according to the alaska bureau of wildlife enforcement, an inordinate amount of illegal activity was reported for one area in the southwestern portion of zone 3 along healy creek. the known illegal take of several antlerless moose in that area nearly resulted in loss of advisory committee support for the hunt in 2005. support was maintained when season and boundary changes were proposed for the 2005 season. lack of public support for habitat — preim to maintain and enhance moose habitat. the public is generally opposed to prescribed and possible damage to private property. in addition, the general public and even most moose hunters do not understand the value natural resources is the agency authorized to conduct prescribed burns in alaska. since and offering to fund a large-scale prescribed but without success. funding issues managing intensively requires infordeaths in moose populations (hundertmark and schwartz 1996). data on body mass, birth rates, survival rates, browse utilization, and population estimation were critical to convincing a skeptical hunting public that antlerless hunts were both timely and prudent in gmu 20a (boertje et al. 2007). during on annual population surveys. without that would have been successful in obtaining liberal hunting and harvest opportunities are likely being lost in adjacent im areas because of inadequate funding. achievements cant progress in elevating moose harvests to help meet im mandates. harvest strategies in gmu 20a in 2004 and 2005 provided the greatest moose hunting opportunities and alces vol. 42, 2006 young et al. elevating moose harvests 47 harvest in recorded history in gmu 20a (> 3,000 moose hunters). reported harvest totaled approximately 1,000 moose (about 390 bulls, 540 cows, and 60 calves) in 2004 and 1,100 (about 430 bulls, 620 cows, and 70 calves) in 2005. although these harvests did not meet the recent im harvest objectives of 1,400-1,600 moose, modeling indicated that the harvest of cows was likely high enough to halt population growth (a management oban additional drawing hunt (300 permits) for any-bull moose because increased recruitment of bulls has occurred since antler restrictions were initiated in 2002. harvests from this new hunt, in combination with the hunts of 2004 and 2005, should allow us to approach the im harvest objective in 2006. recommendations we recommend that legislation granting advisory committees veto power over antlerless hunts be rescinded. this legislation was enacted in the mid-1970s, long before the im law was passed. this 1970s legislation conmoose in im areas (e.g., variable seasons, multiple bag limits, calf hunts, and greater authority to regulate access) because strategies to intensively manage harvest often run counter to prevailing public opinion (hundertmark and schwartz 1996). regulations levels of harvest and to guard against potential overharvest. in addition, we recommend that managers closely monitor the myriad of hunting-related social issues associated with im of moose populations. social issues can be easily overlooked but are an integral part of securing and maintaining public support for hunts, especially those with high hunter densities and intensive harvests. be given greater authority and funding to an integral component of im to maintain and increase moose numbers. ing to determine population parameters and trends and to educate the public with this information. we concur with hundertmark and schwartz (1996) that implementing im programs without reliable information will lead to mismanagement, including undesired population declines. acknowledgements ration. we wish to thank laura mccarthy for her help preparing the manuscript for publication. references alaska statutes. 2005. section 16.05.255, regulations of the board of game; management requirements. pages 32-35 in lations annotated, 2005-2006 edition. lexisnexis , charlottesville, virginia, usa. (www.legis.state.ak.us [accessed april 2006]). boertje, r. d., k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin antlerless harvests. journal of wildlife management 71: in press. _____, p. valkenburg, and m. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal of wildlife management 60: 474-489. gasaway w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological paper number 22, university _____, r.o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. elevating moose harvests young et al. alces vol. 42, 2006 48 interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. hundertmark, k. j., and c. c. schwartz. 1996. considerations for intensive management of moose in alaska. alces 32: 15-24. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450-462. national weather service. 1974-2004. climatological data, alaska national climatic data center, asheville, north carolina, usa. rausch, r. a., r. j. somerville, and r. h. bishop. 1974. moose management in alaska. naturaliste canadien 101: 705712. reynolds, h. v., iii. 1999. effects of harvest on grizzly bear population dynamics in the northcentral alaska range. alaska in wildlife restoration. research progress report. grants w-24-5 and w-27-1. study 4.28. juneau, alaska, usa. schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula. alces 28: 1-13. ver hoef, j. m. 1996. parametric empirical bayes methods for ecological applications. ecological applications 6: 10471055. _____. 2001. from spatially correlated data. pages 9398 in proceedings of the section on statistics and the environment of the american statistical association, august 13-17, 2000. indianapolis, indiana, usa. young, d. d., jr. 2000. units 20 and 25 wolf management report. pages 151-167 in m. v. hicks, editor. wolf management report of survey and inventory activities 1 july 1996-30 june 1999. alaska department of _____. 2003. units 20 and 25 wolf management report. pages 154-166 in c. healy, editor. wolf management report of survey and inventory activities 1 july 1999-30 and game, juneau, alaska, usa. _____, and r. d. boertje. 2004. initial use of moose calf hunts to increase yield, alaska. alces 40: 1-6. f:\alces\supp2\pagema~1\rus 21s alces suppl. 2, 2002 panichev et al. – importance of salt licks 99 the importance of salt licks and other sources of sodium in the ecology of the ussuri moose (alces alces cameloides) alexander m. panichev1, olga y. u. zaumyslova2, and v. v. aramilev1 1pacific institute of geography, far eastern branch, russian academy of science, vladivostok, russia; 2sikhote–alin state reserve, terney, primoriye territory, russia abstract: the most important sources of sodium for moose (alces alces) in sikhote–alin are: (1) freshwater aquatic vegetation (river, lake, and bog); (2) marine water and algae; and (3) sodium– saturated ground waters and soils at salt licks. the distribution of local sources of sodium essentially determines the spatial and temporal structure of moose populations. salt licks play an important role in the ecology of moose as a factor promoting their regular distribution under conditions of the mountain–taiga landscape and also affecting breeding activity; i.e., increasing the probability of encounters of mating partners. the latter is of particular importance where population density is low. alces supplement 2: 99-103 (2002) key words: salt licks, sodium–deficient ecosystems, sodium needs the access of herbivorous animals to sodium sources is of particular importance in sodium–deficient ecosystems. the latter are known to include the majority of mountain, mountain–forest, and also some forest and tundra landscapes in temperate and high–latitude zones on the earth. the mountain sikhote–alin, the ancient place of origin of the ussuri moose (alces alces cameloides), is no exception in this respect. the present paper is an attempt to reveal the primary sources of sodium used by moose in sikhote–alin, with special reference to the role of natural salt outcropping in the ecology of these animals. study area and methods the investigations were conducted in the sikhote–alin state biosphere reserve, mainly in the upper reaches of the kolumbe river, and also in the basin of the upper reaches of the bikin river (the zeva bolshaya and malaya svetlovodnaya rivers) territory, which is thought to be the southern edge of moose range. the average population density there is hardly 1 individual per 1,000 ha. the work was based on collection and investigation of the chemical composition of the essential forage plants and detailed observations of the seasonal movements of moose and their behavior at salt outcrops. among the forage plants, chemical composition was studied in salix rosida, the conifer larix komarouvii, and also the graminoids carex p h y c h o p h y s a , a n d c a l a m a g r o s t i s landsdorfii. the following aquatic plants were studied: potamogeton perfoliatus, r a n u n c u l u s e r a d i c a t u s , s p a r g a n i u m stenofillum, and also some species of blue– green algae. the plant samples were taken in early summer (may–june) and during early autumn (september). preparation and analysis of the samples was performed in the laboratory of geochemistry, pacific institute of geography. the total number of hours of observations at salt licks over 1980–1990 was 2,774. visual observations importance of salt licks – panichev et al. alces suppl. 2, 2002 100 at the salt licks recorded the sex, approximate age, and the time spent at the lick, as well as the duration of the consumption of salt lick substances and the forms and duration of exploratory and social behavior. individual recognition of the majority of individuals was possible from differences in the shape of antlers, fur coloration, molt patterns, etc. annual observations of 7 animals (5 males and 2 females) were conducted in the region of the kaplanovsky salt licks. all the salt licks under investigation were studied for mineral and chemical composition of the substances used by the animals. some of the results of these studies have already been published (panichev 1987). results and discussion the concentration of sodium and potassium in the summer diet of ussuri moose differs only slightly from that of moose dwelling in northern european russia, as demonstrated by the chemical composition of terrestrial forage (table 1). the experience of giving such foods to moose in captivity indicates that these foods are unable to provide a positive sodium equilibrium in the body (kochanev et al. 1981). in boreal ecosystems, the sodium deficiency in terrestrial plants can be compensated for by moose through consumption of sodium–rich aquatic plants, which was demonstrated in a number of studies by north–american table 1. concentration of important chemical macronutrients in summer forage of the ussuri, european, and american moose (g/kg dry matter). forage species1 collection site na+ k+ ca2+ mg 2 + terrestrial plants calamagrostis landsdorffii (5) kolumbe river 0.19 13.40 1.75 1.02 carex phychophysa (2) flood plain and 0.13 8.43 1.44 1.11 leaves of willow (2) slopes of sikhote–alin 0.11 7.69 3.76 1.63 fir conifers (2) 0.13 6.93 2.16 1.18 herb mixture komi republic 0.14 23.74 16.24 4.99 birch leaves (kochanov et al. 1981) 0.29 6.86 8.50 3.43 willow and birch leaves 0.17 10.72 17.04 2.85 bird cherry leaves shore of lake superior, 0.051 16.0 9.90 2.80 poplar leaves canada (fraser et al. 1984) 0.059 11.0 9.70 2.00 willow leaves 0.050 13.6 10.0 2.10 aquatic plants potamogeton perfoliatus (1) creeks and lakes of 3.80 3.41 1.62 3.40 ranunculus eradicatus (11) kolumbe and bikin rivers 1.70 4.62 1.44 1.92 sparganium stenofillum (2) (sikhote–alin) 4.42 2.97 1.35 2.51 algae (4) 2.85 4.92 1.58 3.25 potamogeton perfoliatus (9) shore of lake superior, 4.10– 14.5– 10.9– 3.00– canada (fraser et al. 1984) 10.7 38.2 19.8 6.30 1 number of samples analyzed in parentheses. alces suppl. 2, 2002 panichev et al. – importance of salt licks 101 investigators (botking et al. 1973, belovsky and jordan 1981, fraser et al. 1984). data collected from aquatic plants in numerous creeks and crescent lakes of the rivers bikin and ussurka, and also in freshwater lakes and bogs of the zeva mountain plateau, demonstrate that in the sikhote–alin, moose dwelling where there is sufficient aquatic forage available do not utilize salt licks. at any rate, in terms of the quantity of sodium, the aquatic diet of the ussuri moose is comparable to that in their north american counterparts. along with freshwater plants, moose living in the coastal sikhote–alin rely on marine water and algae for sodium. according to interview data, mass migrations of moose to the sea during summer are common north of terney, particularly in the region of the edinka village. in the vast areas of mountain sikhote– alin, in particular, in rough country and forested areas, the sources of aquatic vegetation are normally small. under these conditions, the main sources of sodium for the animals are sodium–saturated ground water or soil. active visitation of such sources develops natural salt lick complexes. in the sikhote–alin mountains they are numerous and have been previously investigated (panichev 1987). as known from our detailed observations in the upper reaches of the kolumbe and svetlovodnaya rivers (territories where moose dwell in summer and where salt licks are the only sodium source), moose travel dozens of kilometers from the salt licks in winter, migrating downhill to the sparse fir or deciduous forests with undergrowth of willow, birch, and rhododendron. in this case, the autumn migration of moose from densely forested mountain areas, where spruce (picea sp.) and fir (abies sp.) predominate, is especially well–defined. in winter and spring, before green forage appears and the onset of the molting season, moose do not show particular interest in salt licks, although they occasionally come out to the salt licks closest to their winter grounds. with the onset of molting (mid–to late may), moose come close to the salt licks. late may–early july marks the beginning of salt lick visitation by moose in the mountains of sikhote–alin. in late june– early july there is a peak of salt–licking activity, after which the salt–licks are visited less often until early september. in september there is another peak of salt– licking activity that coincides with the rutting season. in some years (e.g., 1986) the autumn peak of visitation may be higher than in summer. during periods of salt– licking activity, the density of moose in the salt–lick area increases dozens of times. for example, daily visitations of the bolshoi kaplanovsky salt–lick exceeded 50 individuals. the density of animals in salt–lick regions can for some time remain at 10 individuals per 10,000 ha or more. at some salt licks of sikhote–alin, for instance, in the upper reaches of the kolumbe river, moose prefer drinking mineralized water from ground sources. at others, for example, in the upper reaches of the rivers pescherka, losevka, and maaka, moose willingly consume hard salt–lick substances (normally these are clay or clay– ceolith mineral substances). the preference for particular salt–lick substances is determined by the closeness of particular salt licks to the summer feeding grounds and also the content of sodium in these substances. because summer habitats of moose occur on flood plains of creeks and rivers, moose can most often be seen at water salt licks, normally formed on the flood plains. the latter are mostly weakly mineralized in sikhote–alin, the most important chemical mixtures of sodium and hydrocarbons predominating (panichev 1987). an example is found in the preference importance of salt licks – panichev et al. alces suppl. 2, 2002 102 of moose for salt licks depending on their chemical composition and water flow (table 2). the most visited within the salt– licking site kaplanovsky are the salt licks with the highest content of sodium and abundant water. the poor visitation by moose of the “dry” salt lick is explained by the fact that the source water flow is very low there; additionally, there is very little assimilable sodium in the salt–lick clays – only 2 g/kg of the clay, while in the salt licks of the pescherka river, where moose prefer consuming the clay, the quantity of sodium consumed is at least 10 times higher (panichev 1990). the sex and age differences in the visitation of salt licks are associated with the physiological conditions of the animals and the remoteness of the sites used for calving and feeding. in fact, at the kaplanovsky salt licks in june, the main visitors to the salt licks were adult bulls (75%). they begin molting earlier than cows and their antlers develop actively. in addition, the cows at that time are busy with young. during the first half of june, adult females and young under 3 years of age predominate (66% and 30%, respectively). in late summer it is mostly lactating cows (94%) that come to the salt licks, which is accounted for by the increasing loss of sodium with milk production. in this case, we recorded that cows during that period consume the salt–lick substances more than any other group. during the rutting season, in which the peak of autumn salt–licking activity occurs, it is males who most often come to the salt licks (69%). in this case the relative duration of drinking behavior by them is 26 + 10% of the time they stay at the lick, and in females 39 + 10%. in early summer this index in males and females was roughly similar (58 + 13% and 64 + 11%, respectively). the duration of social interactions in summer and in autumn in different sexes at t a b le 2 . c h e m ic a l c o m p o s it io n o f s p ri n g w a te r, c h a ra c te ri s ti c s o f th e s iz e o f s a lt l ic k s , a n d s it e i n d ic e s o f v is it a ti o n b y m o o s e a t s a lt l ic k s o f th e k a p la n o v s k y a re a a s o f j u ly 1 9 8 7 ( m a c ro n u tr ie n ts i n m g /l , m ic ro n u tr ie n ts – s i a n d t h e re a ft e r in m g /l ). v is it a ti o n w a s d e te rm in e d b y a v e ra g in g t h e to ta l n u m b e r o f v is it s o v e r th e p e ri o d o f c o n c u rr e n t v is it a ti o n a t s a lt l ic k s f ro m s p ri n g t h ro u g h t h e s u m m e r p e a k o f “ s a lt l ic k a c ti v it y ” . c o ll e c ti o n o p e n v is it s w a te r s it e a re a (p e r f lo w (m 2 ) h o u r) (l /h r) p h n a + k + c a 2 + m g 2 + h c o 3 – c l– s o 4 2 – s i f e m n c u p b z n a l a g c d s a lt l ic k 1 3 7 5 0 .1 1 7 .2 7 3 2 .1 2 .9 4 .5 1 .6 1 4 4 .9 0 .6 1 4 4 .0 1 2 .1 6 .0 2 .2 1 .0 2 .5 32 20 1 .0 0 .2 b o ls h o i k a p la n o v s k y 1 3 ,8 0 0 0 .9 10 7 .5 8 1 2 0 .8 4 .9 1 .6 0 .7 3 6 3 .6 0 .4 4 2 .0 7 .6 1 0 .2 0 .7 72 3 3 .5 7 5 4 < 1 0 0 .8 1 .3 m a ly k a p la n o v s k y 2 ,2 4 0 0 .7 10 7 .7 3 8 0 .4 3 .4 0 .4 0 .4 1 6 3 .0 1 .6 4 9 .6 1 7 .0 3 1 .9 3 .1 50 1 0 0 6 3 4 < 1 0 0 .8 1 .3 s a lt l ic k “ d ry ” 7 0 0 0 0 .1 7 .6 1 8 1 .2 1 0 .2 0 .4 0 .4 1 6 8 .4 4 .0 2 8 .8 1 2 .0 1 9 .4 4 .2 25 1 3 6 3 6 3 60 0 .8 1 .2 alces suppl. 2, 2002 panichev et al. – importance of salt licks 103 salt licks was as follows: in bulls it changed from 4% in summer to 50% in autumn, and in cows from 8% to 30%, respectively. the characteristic feature of the behavior of moose at salt licks in autumn was the fact that the animals showed active interest in their mating partners. the time spent on sexual behavior in all adult animals at salt licks was clearly in excess of that spent on the use of salt–lick substances. that fact leads us to conclude that under conditions of the mountain taiga area, in addition to their role as a source of mineral substances, salt licks are also centers for the breeding activity of moose. conclusions sources of extra sodium for moose in the sikhote–alin are: (1) freshwater plants of crescent lakes and creeks, lakes, and high and low bogs; (2) marine algae and oceanic water; and (3) mineralized sources of water and soil enriched by exchange sodium. the distribution of local sources of sodium over the territory essentially determines the spatial and temporal structure of the populations of the ussuri moose. salt licks, serving as sources of important mineral substances, primarily sodium, promote a regular settlement by moose of mountain areas. besides, they largely play the role of centers for breeding activity, which undoubtedly is of importance where moose populations are sparse. the great number of salt licks and attachment of moose to them in the sikhote– alin is indicative of a long period of adaptation by these animals to living in that landscape’s geochemical and climatic conditions. ussuri moose evolved in a mountain landscape with a humid climate wherein sodium is quickly leached from the soils. the bedrock in the region is formed from volcanic rock with a high content of silica and limited occurrence of calcium and phosphorus (major bone–forming elements), largely promoting the formation of the ussuri strain of moose characterized by small body size and weak deer–like antlers. acknowledgements t. v. lutsenko performed technical laboratory analyses. references belovsky, g. e., and p. a. jordan. 1981. sodium dynamics and adaptation of a m o o s e p o p u l a t i o n . j o u r n a l o f mammalogy 62:613–621. botking, d. v., p. a. jordan, a. s. dominski, h. s. lowedorf, and g. e. hutchinson. 1973. sodium dynamics in a northern ecosystem. proceedings of the national academy of science, usa 70:2745–2748. fr a s e r , d., e. r. ch a v e z, and j. e. paloheimo. 1984. aquatic feeding by moose: selection of plant species and feeding areas in relation to plant chemical composition and characteristics of lakes. canadian journal of zoology 62:80–87. kochanov, n. e., t. m. ivanova, a. e. weber. 1981. metabolism in wild ungulates (reindeer and moose). nauka, leningrad, russia. (in russian). panichev, a. m. 1987. salt licks of the sikhote–alin. far east scientific cent r e , u s s r a c a d e m y o f s c i e n c e , vladivostok, russia. (in russian). . 1990. lithophagy in animal life. nauka, moscow, russia. (in russian). alces vol. 44, 2008 härkönen et al.moose and aspen regeneration in finland 31 does moose browsing threaten european aspen regeneration in koli national park, finland? sauli härkönen1, kalle eerikäinen1, riikka lähteenmäki2, and risto heikkilä3 1finnish forest research institute, joensuu research unit, p.o. box 68, fi-80101 joensuu, finland, email: sauli.harkonen@metla.fi; 2university of joensuu, faculty of forest sciences, p.o. box 111, fi-80101 joensuu, finland; 3finnish forest research institute, vantaa research unit, p.o. box 18, fi-01301 vantaa, finland abstract: large european aspen (populus tremula) trees host hundreds of species of which many are threatened species of conifer-dominated, old-growth boreal forests. aspen is also one of the deciduous tree species most intensively used by moose (alces alces) in finland. in conservation areas aspen regeneration is facilitated by large-scale disturbances, especially fires and windstorms, and also by mortality of individual trees and small-scale disturbances that create small openings. these aggregated patches of young aspens provide high quality feeding sites for moose. in finland, it has been hypothesized that intense browsing pressure by moose on aspen may prevent new aspen cohorts from emerging, and thus endanger the spatio-temporal continuum of aspen occurrence in the long term. the aim of this study was to analyze the influence of moose browsing on the regeneration of aspen in koli national park in eastern finland at 2 different spatial scales, the landscape level and stand level. our results indicated that moose browsing on aspen has been very intense in the area. at the landscape level, moose damaged (twig-browsing, stem breakage, or bark stripping) 96% of aspens in the southern area and 62% in the northern area of the park. in addition, 23% of the damaged aspens (all <5 m) were dead in the southern area. according to counts of fecal pellet groups, moose activity was higher in the southern area than the northern area. at the stand level, on average, 79% of the aspens in the southern area and 73% in the northern area were damaged. the proportion of dead aspens (35%) was highest in stands in height category of 5-15 m. aspen density declined from young to old stands in both areas. bark stripping was relatively common in the height category of 5-15 m over the whole area. we concluded that the current browsing pressure retards the height development of young aspens because of the repeated break-off of main stems and leader shoots. although occurrence of aspen may decline due to high browsing pressure by moose, the majority of aspens have excellent tolerance to heavy and repeated browsing. hence, a high proportion of aspens may reach maturity, and thus maintain the spatio-temporal continuum of aspen occurrence at a level that contributes to its role in community dynamics and local and regional biodiversity. alces vol. 44: 31-40 (2008) key words: alces alces, biodiversity, browsing, european aspen, moose, populus tremula, regeneration browsing by cervids on tree seedlings and saplings has been regarded as a pivotal issue affecting the success of forest regeneration in managed commercial forests and in nature conservation areas across europe (e.g., gill 1992, angelstam et al. 2000, heikkilä et al. 2003, edenius and ericsson 2007) and in northern america (e.g., baker et al. 1997, kay 1997, rooney and waller 2003, mclaren et al. 2004). in finland, moose (alces alces) density has increased continuously over the last few decades (finnish forest research institute 2006), which has led to intense public debate on the need to search for balance between moose and forest management, as well as nature conservation. large european aspen (populus tremula) trees (i.e., living and decaying dead aspens) moose and aspen regeneration in finland härkönen et al. alces vol. 44, 2008 32 host hundreds of herbivorous and saproxylic invertebrates, polypore fungi, and epiphytic lichens of which many are threatened species of conifer-dominated, old-growth boreal forests (kouki et al. 2004). it has been estimated that there are at least 150 specialist species that are entirely dependent on aspen in fennoscandia (siitonen 1999, kouki et al. 2004). aspen is also one of the deciduous tree species most intensively used by large herbivores, and ranks among the highly preferred woody plants in the diet of moose (bergström and hjeljord 1987). due to increasing awareness about explicit ecological values of aspen on forest biodiversity, aspen is receiving more consideration in management planning of commercial forests, specifically in the restoration planning of formerly managed forests within conservation areas. it has been estimated that only 0.3% of the forest land in finland is covered by aspendominated stands (finnish forest research institute 2006). however, young mixed species stands with aspen trees are relatively common, whereas densities of large aspens are very low in older forests. according to the national forest inventory (nfi) of finland, there are only 0.35 aspens/ha with diameter at breast height >30 cm (finnish forest research institute 2006). aspen can regenerate sexually by seed and asexually by root suckers. however, in the finnish landscape asexual type of reproduction is more common and in most regeneration sites practically all seedlings originate from root suckers (tikka 1955). some disturbance is needed to start aspen regeneration. in natural forests, aspen regeneration is facilitated by large-scale disturbances, especially fires and windstorms, but also by the death of individual trees and small-scale disturbances that create small openings (syrjänen et al. 1994, cumming et al. 2000). these aggregated patches of young aspens provide high quality feeding sites for moose. in finland, it has been hypothesized that intense browsing pressure by moose on preferred aspens may prevent new aspen cohorts from emerging (kouki et al. 2004), thus, endangering the spatio-temporal continuum of aspen occurrence. even in conservation areas this may result in a situation where aspen-associated species may disappear both locally and regionally. the aim of this study was to analyze the influence of moose browsing on the regeneration of aspen in a conservation area. specifically, we assessed past and present moose browsing pressure on aspen at the landscape and stand levels within koli national park in eastern finland. the results are related to the regeneration success of aspen in forests under high browsing pressure by moose. study area the study occurred in the koli national park in eastern finland (fig. 1). the central hill sites of koli became state-owned in 1907, and from 1924-2007 the area was administered by the finnish forest research institute. koli national park was established by law in 1991 and covered a total area of 1,135 ha; it has since been enlarged to about 3,000 ha. as of 2008, the finnish forest and park service is responsible for administration of koli national park. the purpose of conserving the area by law was to ensure the preservation of koli’s heritage landscape and the old-growth forests of the koli highlands, as well as maintain the plant communities created in the past by swidden cultivation (i.e., clearing of land for cultivation by slashing and burning the forest vegetation cover). other aims were the promotion of environmental research, education, and nature recreation in the area. the park area is characterized by a highly variable landscape and topography where the altitude varies from 95-347 m above sea level. the soils are fertile, especially at lower altitudes, because the crumbling substance originating from the calciferous bedrock (diabase-rich or siliceous rocks, i.e., granitegneisses and quartzites; piirainen et al. 1974) alces vol. 44, 2008 härkönen et al.moose and aspen regeneration in finland 33 flow downhill with rainwater and streams of melting snow. the study area belongs to the border area between the southern and middle boreal forest vegetation zone (see kalliola 1973) (fig. 1). the dominant tree species are norway spruce (picea abies) and scots pine (pinus sylvestris). other tree species such as birches (betula pendula and b. pubescens), aspen, rowan (sorbus aucuparia), and alder (alnus incana) occur patchily within the area. the forests outside the park are intensively managed and fragmented by lakes, mires, cultivated land, and small villages. hunting of moose is strictly prohibited in koli national park. the northern karelia game management district estimated the density of moose outside the park, after the hunting season, as 0.3-0.6 moose/km2 in 2002-2006. local hunters estimated by snow tracking that there were about 0.75 moose/km2 within the park in november 2006, and that this density remained about the same in 2002-2006. moose density increases in the park during hunting season (starting the last saturday of september), when they escape hunting pressure from adjacent private forestland, until spring when they disperse back to their traditional summer range. methods moose browsing on aspen was investigated in summer 2006 at 2 spatial scales, the landscape level and the stand level. the park was divided into northern (no) and southern (so) areas at both scales. this was reasonable because the no area was characterized by representative old-growth forests close to the natural stage, whereas, in the so area signs of forestry practices were still clearly visible both in the landscape and forest age structure. at the landscape level, data were collected from systematically located circular sample plots; 132 plots were located in the no area and 141 plots in the so area. the plot size was 113 m2, and plots were placed on a systematic 300 m x 300 m grid. in addition, the stand development class was determined at each plot. aspens were also pooled into three height categories (i: <5 m, ii: 5-15 m, and iii: >15 m). at the stand level, all potential stands where aspen was dominant or was important in the mixture of tree species were identified from aerial photographs. when verified in the field, the stand was included in the sampling procedure, after which 5 young (<5 m in height), 5 middle-aged (5-15 m in height), and 5 old (>15 m in height) aspen-rich stands were randomly selected both in the no and so areas to ensure sufficient variation of stand structures in the data (ericsson et al. 2001). three circular plots of 113 m2 were systematically located within each sampled stand, with the first plot placed in the center of the stand, fig. 1. location of the koli national park in eastern finland. forest vegetation zones after kalela (1970 in kalliola 1973): south finland (1= hemiboreal and 2 = southern boreal), pohjanmaakainuu (3 = middle boreal), and peräpohjola and metsä-lappi (4 = northern boreal). moose and aspen regeneration in finland härkönen et al. alces vol. 44, 2008 34 and the other 2 placed 20 m north, south, west, or east from the center plot depending on the shape of the stand. the condition of each aspen was classified as undamaged or damaged, and alive or dead. each measured aspen was investigated for signs of fresh and older moose browsing. fresh browsing meant that browsing had occurred during winter 2005-2006, and was distinguished from older browsing by its white color at the browsing point. moose browsing was divided into 3 damage categories: twigbrowsing, stem breakage, and bark stripping. the total tree height was measured of all small aspens (i.e., height 0.5-4.0 m). for aspens taller than 4 m, the height was estimated to the nearest 0.5-m class. in order to estimate the moose activity in the area, the number of fecal pellet groups (1 group ≥20 pellets) was counted in each plot. only pellet groups deposited during the winter of 2005-2006 (i.e., those on top of the previous year’s leaf litter) were counted. all statistical analyses were performed with spss package. nonparametric tests were employed because none of the variables had normal distributions. results landscape level the density of aspen was three-fold higher in the so area (152 aspens/ha ± 47 se) than in the no area (48 aspens/ha ± 12 se), but the difference was not significant (mannwhitney u = 8759.5, p = 0.24). the number of fecal pellet groups indicated that moose used the so area (32 pellet group/ha ± 6 se) more than the no area (2 pellet groups/ha ± 1.5 se) (mann-whitney u = 7401.0, p = 0.000). there was no difference in mean aspen height between the two areas (so: 9.7 m ± 1.4 se vs. no: 10.0 m ± 2.4 se, mann-whitney u = 267.5, p = 0.36). moose browsed aspens in the so more often than in the no area. in total, 96% of aspens were damaged in the so area and 62% in the no area; 23% of damaged aspens were dead in the so area, all shorter than 5 m. only 3% of the damaged aspens were dead in the no area. repeated stem breakage was very common. the mean number of stem breakages per aspen did not differ between the two areas (so: 2.7 times/aspen ± 0.3 se vs. no: 2.2 times/aspen ± 0.5 se, mann-whitney u = 62.5, p = 0.12). in order to illustrate the effect of aspen height on moose browsing, a closer examination of moose browsing in three aspen height categories was performed (fig. 2). there was a clear tendency that aspens in height categories i and ii were damaged more often (twig-browsing and stem breakage) than aspens in height category iii. bark stripping was relatively common in height category ii in both areas. in height category i, all aspens in the so area had signs of old damage. the percentage of new damage showed 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % broken stem new broken stem old broken stem 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % tw ig-brow sing tw ig-brow sing, new tw ig-brow sing, old 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % bark stripping new bark stripping old bark stripping 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % dam aged new dam age old dam age fig. 2. proportions of different damage types in different aspen height categories at the landscape level in koli national park, finland (so = southern area; no = northern area). the sum of percentages of new and old damage may exceed the percentage of total damage within a height category due to repeated browsing of the same stems. alces vol. 44, 2008 härkönen et al.moose and aspen regeneration in finland 35 that re-browsing hampered young trees most. the proportion of young aspens (i.e., height category i) decreased from 45% in advanced seedling stands to 17% in mature stands indicating that regeneration of aspen, to some extent, occurs in closed older stands. stand level aspen density declined from young to old stands in both areas (table 1). moose seemingly used stands relatively little as the number of fecal pellet groups was very low. the mean height was similar in same aged stands in both areas (table 1). based on damage type, browsing pressure by moose was very similar in both areas (fig. 3). on average, 79% of the aspens in the so area were damaged, as compared to 73% in the no area. the proportions of new damage showed that re-browsing occurred predominantly in young stands. the proportion of dead aspens was relatively high in middle-aged stands in the so and no areas (fig. 4). the mean number of stem breakages per aspen did not differ between the two areas (so: 2.3 times/ aspen ± 0.3 se vs. no: 2.0 times/aspen ± 0.2 se, mann-whitney u = 48.0, p = 0.43). so <5 m so 5-15 m so >15 m no <5 m no 5-15 m no >15 m aspen density 4,458 ± 725 1,987 ± 253 383 ± 13 2,371 ± 343 2,306 ± 710 896 ± 151 aspen height 1.4 ± 0.1 6 ± 0.6 20.1 ± 1.5 1.4 ± 0.2 4.7 ± 0.4 17.6 ± 1.8 pellet groups 12 ± 7 12 ± 7 table 1. aspen density (trees/ha) and height (m), and number of fecal pellet groups (/ha) in young (<5 m in height), middle-aged (5-15 m in height), and old (>15 m in height) aspen-rich stands at the stand level in aspen-rich forest stands in koli national park, finland. means are given with their standard errors. note: so = southern area of koli national park; no = northern area of koli national park. 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % broken stem new broken stem old broken stem 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % tw ig-brow sing tw ig-brow sing, new tw ig-brow sing, old 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % bark stripping new bark stripping old bark stripping 0 20 40 60 80 100 so < 5 m so 5-15 m so > 15 m no < 5 m no 5-15 m no > 15 m % dam aged new dam age old dam age fig. 3. proportions of different damage types in young (<5 m in height), middle-aged (5-15 m in height), and old (>15 m in height) aspen-rich stands at the stand level in koli national park, finland (so = southern area; no = northern area). the sum of percentages of new and old damage may exceed the percentage of total damage within a stand type due to repeated browsing of the same stems. 0 10 20 30 40 s o < 5 m s o 5 -1 5 m s o > 1 5 m n o < 5 m n o 5 -1 5 m n o > 1 5 m % fig. 4. proportions of dead aspens in young (<5 m in height), middle-aged (5-15 m in height), and old (>15 m in height) aspen-rich stands at the stand level in koli national park, finland (so = southern area; no = northern area). moose and aspen regeneration in finland härkönen et al. alces vol. 44, 2008 36 discussion results clearly showed that the browsing pressure by moose on aspen has been very intense in koli national park. at the landscape level, moose had damaged (twigbrowsing, stem breakage, or bark stripping) 96% of the aspens in the so area and 62% in the no area, and the moose damage at the stand level was 79 and 73% in the no and so areas, respectively. these damage levels are more than four-fold higher than frequencies observed in old-growth forests in eastern finland. latva-karjanmaa et al. (2007) reported that the damage proportions at the stand level varied from 28-44% in managed forests in finland; similarly, härkönen (1998) found them to vary from 35-50%. the proportion of dead aspen further indicated that browsing pressure caused by moose was high. also, at the landscape level all aspens in height category i in the so area had signs of old browsing. bark stripping in height category ii was relatively common in the so and no areas indicating that relatively large aspens are also susceptible to moose damage. at the landscape level, 23% of the damaged aspens (all in height category i) were dead in the so area. at the stand level, the proportion (35%) of the dead aspens was highest in the middle-aged stands. these results could be explained by cumulative stem breaking and browsing which is concentrated on leader shoots and side twigs of young aspens year after year, weakening the ability of aspen to withstand sustained moose browsing. in the case of managed forests, edenius et al. (2002) reported that aspens growing alone are more susceptible to moose browsing than aspens aggregated within stands. hence, a single aspen has the best chance of escaping browsing in a stand with a high aspen density (ericsson et al. 2001). this observation corresponds to our findings because overall proportion of damaged aspens was higher at the landscape level (low aspen densities) than in the stand level (high aspen densities). thus, moose may reinforce the spatially aggregated distribution of aspen both in managed and protected boreal forest landscapes. that aspen density decreased from young to old stands, that low aspen density occurred in old stands at the stand level, and that low overall aspen density occurred at the landscape level were all expected results. however, the observed densities were relatively high in comparison to those found at other conservation areas in eastern finland where aspen densities have been as low as 1 tree/hectare (kouki et al. 2004). nfi results also indicate the relative rareness of mature aspen trees in finnish forests (finnish forest research institute 2006). in young stands, aspen densities (2,300-4,500 trees/ha) were comparable to those in conservation areas measured by heikkilä et al. (2003). stand level volumes of aspen were not of primary interest in this study and they were not calculated, but it has been estimated that an average volume of 5-20 m3/ha is typical for aspen in norway spruce-dominated, old-growth forests in southern and middle boreal fennoscandia (latva-karjanmaa et al. 2007). pellet group counts were used to estimate moose activity in the area. according to neff (1968), the method can provide reliable data under most field conditions. in this study, the observed numbers of pellet groups were surprisingly low when compared to moose browsing pressure on aspens and the estimated moose density in the area, and the numbers reported from managed scots pine-dominated forests in finland (heikkilä and härkönen 1998). this may be explained by overestimation of moose density in the area or by difficulties in separating old pellet groups from new pellet groups during fieldwork (neff 1968, härkönen and heikkilä 1999). however, the pellet group counts were in line with browsing intensity at the landscape level as there were more pellet groups in the so than no area. also, the proportion of damaged aspens was higher in the so area. at the stand level, the number alces vol. 44, 2008 härkönen et al.moose and aspen regeneration in finland 37 of pellet groups was very low, whereas the browsing intensity was relatively high. this weak relationship between browsing incidence on aspen and pellet counts was also observed in sweden (edenius and ericsson 2007). it is evident that norway spruce-dominated forest sites, which are common in koli national park, are less favorable for moose because overall food availability is low in these closed middle-aged or older stands. in this sense, small aggregated patches of young aspens may provide so little food that moose only visit them briefly during a foraging bout. the impact of ungulate populations on regeneration of aspen has been questioned globally, for example, in the rocky mountain region in western north america (kay 1997, suzuki et al. 1999) and in finland (kouki et al. 2004). however, moose or deer browsing is not the only factor behind the poor recruitment of young aspen cohorts in conservation areas. absence of large scale disturbances (e.g., fires and windstorms) that create forest openings in conservation areas strongly hinder the possibility of successful regeneration (latva-karjanmaa et al. 2007) because some disturbance is needed before aspen regeneration starts from root suckers. in addition, young aspen seedlings and saplings are browsed by voles, mountain hare (lepus timidus), and beavers (castor spp.). the impact of these browsers on aspen mortality might have been underestimated in comparison to moose browsing. disease, including “black shoot blight” which is caused by a fungus (venturia tremulae) and kills the leader shoot and leaves, may stunt the height development of young aspens in conservation areas (heikkilä et al. 2003). this disease causes infected aspens to be available longer as potential forage for moose or other browsers, thereby increasing the risk that damaged aspens die due to repetitive browsing. in addition, young aspen seedlings may also die because they cannot tolerate local climate and soil conditions (romme et al. 2005). in managed forests of finland, aspen has been an unwanted species, especially during the 1970s and 1980s, because of its low economic value, fast early growth and high competitive ability with more valuable tree species, and its inclination to host rust disease (melampsora pinitorqua) that is harmful to young scots pine stands. for these reasons foresters have tried to control aspen with silvicultural cleanings in young stands, and by notching and girdling larger individuals in mature forests before clear-cutting. considering these other factors in protected and managed forests, it is evident as edenius and ericsson (2007) have presented, that it is not possible to increase the abundance of aspen only by adjusting ungulate browsing levels. similarly, romme et al. (1995) concluded that low aspen regeneration cannot be explained by any single factor, but involves a complex interaction of many factors. aspen regeneration can be influenced by certain disturbance activities. it has been suggested that prescribed burning and manmade gaps (latva-karjanmaa et al. 2007) will facilitate aspen regeneration. creating large canopy gaps in old-growth forests by killing trees around mature aspens may also promote aspen regeneration. successful aspen regeneration can be facilitated by other human manipulations, for instance, protecting aspen recruitment physically from browsing (angelstam et al. 2000, mclaren et al. 2004, kaye et al. 2005, edenius and ericsson 2007). in managed forests, young aspens could be spared during silvicultural cleanings and thinnings, and mature aspens could be left aside as retention trees during clear-cutting. the proportion of young aspens (<5 m in height) decreased from 45% in advanced seedling stands to 17% in mature stands. however, this clearly indicates that the regeneration of aspen is, at least to some extent, occurring in closed older stands in the study area. it can be concluded that the present browsing pressure by overabundant moose in moose and aspen regeneration in finland härkönen et al. alces vol. 44, 2008 38 koli national park not only kills some proportion of young aspens, but also retards the height development of aspens because of the repeated break-off of main stems and leader shoots (angelstam et al. 2000, heikkilä et al. 2003). this is supported with the observation that where population density is >5 moose/ 1,000 ha, moose browsing depresses the growth of aspen (abaturov and smirnov 2002). at present, it seems that the major proportion of young aspens in koli national park can tolerate both heavy and repeated browsing without dying. however, the persistence of aspen may lead to collapse if moose density increases and stays higher than the current carrying capacity, which is an unforeseeable scenario. the long-term existence, dynamics, and possibilities for natural regeneration of aspen in old-growth forests in koli national park have been determined by the compartmentwise inventory data available for the period 1910–2004 (m. vehmas et al., university of joensuu, unpubl. data). their results indicated that aspen can regenerate naturally in old-growth forests and that its existence is not completely dependent on young successional stages of forests, which corresponds to the findings of this study. however, they assumed that the long-term ecological continuity may be threatened by increased population densities of large mammalian browsers such as overabundant moose. even if their study was not focused on identifying factors related to the success of aspen regeneration, it concluded that the issue between the seemingly increasing population densities of mammalian browsers and success of aspen regeneration needs to be considered in future restoration and management measures planned for koli national park. suzuki et al. (1999) concluded that elk (cervus elaphus) browsing in rocky mountain national park was not causing failure of aspen regeneration at landscape scales, whereas in local areas with the highest browsing pressure from elk, little regeneration has occurred in the past 30 years. a similar observation was made by heikkilä et al. (2003) when evaluating the effects of moose browsing on aspen in finland, and in national parks in newfoundland and labrador (mclaren et al. 2004). our study supports the conclusion that moose browsing will most likely contribute to an altered size distribution of aspen in the boreal landscape (kay 1997, suzuki et al. 1999, ericsson et al. 2001). hence, the number of aspens in koli national park may decrease due to the high browsing pressure by moose, but many aspens may reach maturity once having passed a risky regeneration stage. this should guarantee the spatio-temporal continuum of aspen at a level where biodiversity is maintained in the conservation area. however, aspen may not necessarily continue to exist where foresters and conservationists desire (ericsson et al. 2001). acknowledgements we would like to thank mr. lasse lovén (finnish forest research institute, koli national park) for providing the facilities during fieldwork. we are also grateful to mr. mika venho for his help in extracting the aspen stand information from the forest inventory data of koli national park, and to 2 anonymous referees for their valuable comments on the manuscript. references abaturov, b. d., and k. a. smirnov. 2002. effects of moose population density on development of forest stands in central european russia. alces supplement 2:1-5. angelstam, p., p.-e. wikberg, p. danilov, w. e. faber, and k. nygrén. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland, and russian karelia. alces 36:133-145. baker, w. l., j. a. munroe, and a. e. hessl. alces vol. 44, 2008 härkönen et al.moose and aspen regeneration in finland 39 1997. the effects of elk on aspen in the winter range in rocky mountain national park. ecography 20:155-165. bergström, r., and o. hjeljord. 1987. moose and vegetation interactions in northwestern europe and poland. swedish wildlife research supplement 1:213-228. cumming, s. g., f. k. a. schmiegelow, and p. j. burton. 2000. gap dynamics in boreal aspen stands: is the forest older than we think? ecological applications 10:744-759. edenius, l., and g. ericsson. 2007. aspen demographics in relation to spatial context and ungulate browsing: implications for conservation and forest management. biological conservation 135:293-301. _____, _____, and p. näslund. 2002. selectivity by moose vs spatial distribution of aspen: a natural experiment. ecography 25:289-294. ericsson, g., l. edenius, and d. sundström. 2001. factors affecting browsing by moose (alces alces l.) on european aspen (populus tremula l.) in a managed boreal landscape. écoscience 8:344-349. finnish forest research institute. 2006. finnish statistical yearbook of forestry. 2006. helsinki, finland. 438 pp. gill, r. m. a. 1992. a review of damage by mammals in north temperate forests. 3. impact on trees and forests. forestry 65:363-388. härkönen. s. 1998. effects of silvicultural cleaning in mixed pine-deciduous stands on moose damage to scots pine (pinus sylvestris). scandinavian journal of forest research 13:429-436. _____, and r. heikkilä. 1999. use of pellet group counts in determining density and habitat use of moose alces alces in finland. wildlife biology 5:233-239. heikkilä, r., and s. härkönen. 1998. the effects of salt stones on moose browsing in managed forests in finland. alces 34:435-444. _____, p. hokkanen, m. kooiman, n. ayguney, and c. bassoulet. 2003. the impact of moose browsing on tree species composition in finland. alces 39:203-213. kalliola, r. 1973. suomen kasvimaantiede. werner söderström osakeyhtiö, porvoo, finland. 308 pp. (in finnish). kay, c. e. 1997. is aspen doomed? journal of forestry 95(5):4-11. kaye, m. w., d. binkley, and t. j. stohlgren. 2005. effects of conifers and elk browsing on quaking aspen forests in the central rocky mountains, usa. canadian jour-canadian journal of forest research 15:1284-1295. kouki, j., k. arnold, and p. martikainen. 2004. long-term persistence of aspen – a key host for many threatened species – is endangered in old-growth conservation areas in finland. journal for nature conservation 12:41-52. latva-karjanmaa, t., r. penttilä, and j. siitonen. 2007. the demographic structure of european aspen (populus tremula) populations in managed and oldgrowth boreal forests in eastern finland. canadian journal of forest research 37:1070-1081. mclaren, b. e., b. a. roberts, n. djanchékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40:45-59. neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. journal of wildlife management 32:597-614. piirainen, t., m. honkamo, and s. rossi. 1974. a preliminary report of the geology of the koli area. bulletin of the geological society of finland 46:161-166. romme, w. h., m. g. turner, g. a. tuskan, and r. a. reed. 2005. establishment, persistence, and growth of aspen (populus tremuloides) seedlings in yellowstone national park. ecology 86:404-418. _____, _____, l. l. wallace, and j. s. walker. 1995. aspen, elk, and fire in moose and aspen regeneration in finland härkönen et al. alces vol. 44, 2008 40 northern yellowstone park. ecology 76:2097-2106. rooney, t. p., and d. m. waller. 2003. direct and indirect effects of white-tailed deer in forest ecosystems. forest ecology and management 181:165-176. siitonen, j. 1999. haavan merkitys metsäluonnon monimuotoisuudelle. pages 71-82 in j. hynynen and a. viherä-aarnio (eds.) haapa – monimuotoisuutta metsään ja metsätalouteen. vantaan tutkimuskeskuksen tutkimuspäivä tammisaaressa. tammisaari, finland, november 12, 1998. (in finnish). suzuki, k., h. suzuki, d. binkley, and t. j. stohlgren. 1999. aspen regeneration in the colorado front range: differences at local and landscape scales. landscape ecology 14:231-237. syrjänen, k., r. kalliola, a. puolasmaa, and j. mattsson. 1994. landscape structure and forest dynamics in subcontinental russian european taiga. annales �oo-ian european taiga. annales �oologici fennici 31:19-34. tikka, p. s. 1955. structure and quality of aspen stands. i. structure. finnish forest research institute 44(4):1-33. (in finnish with english summary). 4302.pdf alces vol. 43, 2007 samuel winter ticks and die-offs of moose 39 factors affecting epizootics of winter ticks and mortality of moose w.m. samuel department of biological sciences, university of alberta, edmonton, ab, canada t6g 2e9 abstract: die-offs of moose (alces alces) associated with, or attributed to, winter ticks (dermacentor albipictus) are widespread and have been reported since the early part of the last century. extrinsic factors such as weather and vegetative structure, and host factors such as moose density and, indirectly, tick-induced damage to the hair coat, were examined in an attempt to predict related problems for moose. the proposal that warmer and shorter winters result in increased survival of adult female ticks dropping off moose in march and april, and increased tick populations on moose the following winter, from the air, coincided with annual changes in numbers of ticks on moose, providing managers with a survey tool to monitor and estimate changing numbers of ticks. tick numbers lagged 1 year behind moose numbers in elk island national park over a 12-year period, and many moose died when numbers possibly independent of moose density. the lack of objective and continuous data sets should guide future research efforts. alces vol. 43: 39-48 (2007) key words: alces, density, dermacentor albipictus, epizootics, hair, moose, mortality, ticks, transmission, weather die-offs of moose associated with, or attributed to, winter ticks are numerous and widespread across north america, having occurred since the early part of the last century (summarized by samuel 2004). almost all published and unpublished reports of such sometimes few, often many, dead or dying moose covered with ticks. tick-associated die-offs often are widespread and concurrent, involving many populations of moose in several-to-many jurisdictions. for example, the most recent widespread outbreaks of winter ticks, accompanied by losses of many moose, occurred in late winter and spring, 2002. there were reports from isle royale national park, minnesota, maine, new hampshire, and vermont in the united states, and alberta, manitoba, ontario, and saskatchewan in canada (samuel et al. 2002, peterson and vucetich 2003, samuel and crichton 2003). earlier concurrent widespread die-offs occurred in late winter-spring, 1999 and 1992, and one across alberta in 1982 (samuel 2004). die-offs are often attributed to winter ticks, perhaps because ticks are obvious and numerous on dead and dying moose. unfortunately, direct evidence of the lethal effect of winter ticks on moose populations is lacking. there is, however, good information from samuel et al. 2000) that, at the least, sugin rapid declines of moose numbers. other factors, such as moose numbers or density, habitat, weather, and predation, likely also play a role. the objective here is to review factors potentially contributing to tick-related die-offs of moose using published literature, reports, and observations from central alberta, particularly elk island national park and vicinity; winter ticks and die-offs of moose – samuel alces vol. 43, 2007 40 that might be used to predict epizootics of ticks and die-offs of moose. in an attempt to determine if various weather parameters might be predictive of moose die-offs, i examined parameters of snow cover and temperature at the time of 5 die-offs in alberta, 2 of which (late winter-springs of 1982 and 1999) killed many moose. i also examined the feasibility of using tick-caused damage to the winter hair coat of moose as an indication of tick numbers as done by wilton and garner (1993) and others. weather several aspects of weather are potentially harmful to moose and ticks. for moose, depth of snow and length of severe cold, often in concert with condition of habitat, can be important (mech et al. 1987, see logic in vucetich and peterson 2004). snowfall was much above normal during 3 of 5 years of tick-related moose die-offs in central alberta, including 50 cm above normal during the major die-off of moose in late winter-spring 1982 (table 1). the winter-spring in 4 of 5 die-off years was colder than the 30-year average (table 1). in particular, march and april 1982 and 2002, were much colder than the 30-year average. for ticks, weather in autumn, when young ticks are on vegetation ambushing moose, and early spring, when female ticks are dropping from moose to lay eggs for the next generation of ticks, is critical for survival and to-mid october could bury tick larvae, thus potentially curtailing transmission to moose. that happened the week of october 14, 1991 at elk island national park, near edmonton alberta, when approximately 45 cm of snow fell, burying most low vegetation (aalangdong 1994). in autumn, particularly october, tick larvae tend to quest on low vegetation and ambush passing ungulate hosts (drew and samuel 1985). in 1991, minimum daily temperatures at the park were below –10oc the last 10 days of october. the combination of snow that likely buried tick larvae and cold that inactivated or killed larvae (drew and samuel 1985, aalangdong 1994), decreased potential transmission of ticks to moose by almost half in 1991 (fig. 1). unfortunately, it is impossible to estimate the effect of this snowfall on subsequent tick numbers on moose, because no moose were examined for ticks or tick-caused damage to the winter coat of hair (see below). there was no die-off of moose in the park or other parts of central alberta that winter. there was also no moose die-off in winter 1991-1992 in the peace river country of northwest central alberta, where ticks often cause problems for moose (pybus 1999). there was also no major snowfall or severe cold there in october 1991. weather in late winter-early spring, particularly related to snowfall and temperature the year before a moose die-off, is mentioned future numbers of ticks (e.g., timmermann and whitlaw 1992, wilton and garner 1993, 0 1000 2000 3000 4000 5000 6000 7000 9/ 1 9/ 15 9/ 29 10 /1 3 10 /2 7 11 /1 0 11 /2 4 12 /8 date m e a n n u m b e r la r v a e 1991 1992 fig. 1. mean daily number of tick larvae recovered from vegetation in the main park area of elk island national park, central alberta. larvae were collected by dragging 1-m2 cloths (599 samples in 1991, 1,556 in 1992) along transects on vegetation 1 – 1.5 m in height 4 days each week at 4 sites. transects were 2 km long and divided into eighty 25-m units of which 10 units were selected randomly for sampling (details in aalangdong 1994; data from o. aalangdong). alces vol. 43, 2007 samuel winter ticks and die-offs of moose 41 weather parameter 30-year average 1971-20001 year of late winter-spring die-off 19782 19823 19882 19994 20024 affecting moose: total snowfall (cm) october – april1 115 79.9 164.8 49 132.2 129.8 mean monthly temperature (oc) december – april1 -7.1 -8.5 -11.5 -4.1 -5.8 -8.5 mean minimum monthly temperature (oc) december – april1 -12.8 -14.6 -17.9 -9.8 -13.4 -15.7 mean temperature (oc) march1 -4.5 -3.6 -8.6 1.4 -5.9 -15 mean minimum temperature (oc) march1 -9.9 -8.4 -14.7 -4.7 -10.8 -22 mean temperature (oc) april1 4.3 4.7 -0.6 6 5.1 -2.1 mean minimum temperature (oc) april1 -2.2 -0.5 -7 -2 -0.9 -7.5 affecting ticks: mean temperature (oc) march and march of year previous1 -4.5 -2.8 0.1 -5.1 -3.8 -2 mean temperature (oc) april and april of year previous1 4.3 6.9 4.3 7 6.9 3.9 mean minimum temperature (oc) march and march of year previous1 -9.9 -8.1 -5.7 -9.7 -8.8 -8.8 mean minimum temperature (oc) april and april of year previous1 -2.2 -2.2 -2.9 -0.3 -0.1 -4.9 snowfall (cm) october1,5 8 0 2.4 0.8 17.6 10 snow depth (cm) end october1,5 1 0 0 0 0 4 snow depth (cm) end march and end of march, year previous1 8 0 0 1 0 0 table 1. summary of weather characteristics that potentially impact moose and winter ticks during 5 moose die-offs in the edmonton region (central alberta). 1environment canada, edmonton international airport. 2only in elk island national park. 3throughout alberta, including elk island national park. 4throughout alberta, but not in elk island national park. 5preceding the winter–spring of die-off. winter ticks and die-offs of moose – samuel alces vol. 43, 2007 42 delgiudice et al. 1997). in general, the logic is that a warm late winter-early spring with low precipitation is conducive to survival and egg-laying of female ticks that drop from moose at that time. this results in above normal numbers of young ticks on moose the following autumn, possibly leading to die-offs of moose the subsequent winter. data from drew and samuel (1986) appear to be the genesis of this logic. they put blood-fed, adult, female winter ticks in outdoor, screen-wire cages at elk island national park at biweekly intervals from early march to late april-mid may, 1982 and 1983, and monitored for survival. only 11% of ticks (same percentage, both years) put in cages on snow (i.e., before snowmelt) survived to lay eggs, while 73 and 55% of ticks put in cages on leaf litter (i.e., after snowmelt) in 1982 (~ 10 april) and 1983 (~ 24 april), respectively, survived. drew and samuel (1986) suggested that low survival prior to snowmelt in spring might be due to prolonged exposure to ambient temperatures at the snow surface that are below the tick’s threshold for survival (~ minus 17oc). an extension of this idea is that snow depth or crusting of snow, which forms during the april, prevents ticks from getting to the duff layer, and they die before laying eggs. drew and samuel (1989) found that bloodfed adult female ticks dropped off moose from late february to early may, but the peak drop period was late march. so, what do the data indicate for late march preceding each of the 5 tick-related moose die-offs in the last 30 years; i.e., did female ticks drop on snow or leaf litter? the answer is that most ticks dropped on leaf litter (see bottom, table 1). however, the context of these results is that, even though the 30-year average amount of snow on the ground at end march was 8.0 cm, there has been no snow on the ground at the end of march in 17 of 29 (65%) years, 1977-2005, in the edmonton region. no snow tends to be “normal.” warmer temperatures in march and april should result in higher survival of adult female ticks that have dropped from moose and, in general, it was warmer in march and april in years preceding die-offs (table 1). mean march and april temperatures were at or above the 30-year average 4 of 5 die-off years. mean minimum temperatures in march were warmer than the 30-year average all 5 years, and minimum temperatures in april were at or above average 4 years; conditions good for tick survival and reproduction. in ontario, wilton and garner (1993) saw more severe tick-induced hair damage and loss to moose in years following mean april temperatures above 3oc, suggesting that survival of ticks was good above and poor below this temperature. in the edmonton region the 30-year, daily mean april temperature was 4.3oc. thus, these data indicate that colderthan-average winters in 4 of 5 die-off years, and extensive snow cover in 3 of 5 die-off years might have adversely affected moose in central alberta. further, ticks might have and above normal temperatures during the drop-off period in march-april in all years preceding die-offs. tick numbers and hair damage and loss in studies of moose and ticks at elk island national park, hair loss correlated with rate of tick-induced self-grooming; i.e., moose that groomed more lost more hair (mooring and samuel 1999). in general, annual mean percent hair damage and loss coincided with annual mean number of winter ticks on moose (samuel 2004) indicating that grooming was related to the number of ticks on moose. in a 10-year study in algonquin provincial park, ontario, damage to the winter hair coat of moose was used as an index of tick numbers on moose. years with high numbers of moose carcasses generally coincided with years when hair damage was highest and vice versa alces vol. 43, 2007 samuel winter ticks and die-offs of moose 43 (garner and wilton 1993, wilton and garner 1993). however, a manager has few options to estimate trends in tick numbers. estimating numbers of ticks on a sedated live moose is not reliable because there can be many thousands of ticks on this large mammal with a thick winter coat of hair, making counting impossible (personal observation). counting ticks on dead moose is not an option because the digestion technique used to recover ticks, though accurate (welch and samuel 1989) is laborious. an important management issue is whether data on annual hair damage/loss can provide an early warning of impending dieoffs of moose. the answer might hinge on the accuracy of techniques used during aerial and ground surveys of moose to record tickinduced hair damage and loss. hair damage or aerial survey in 2 ways: (1) use a digitizer to determine the percentage of the lateral silhouette of the torso with hair damage or loss from photographs or diagrams of lateral views of moose (see description of technique in welch et al. 1990, and samuel and welch 1991); or (2) group moose subjectively into several categories of hair-loss severity (see photographs of categories in samuel 1989, cause it demands either taking photographs, or making diagrams of hair damage of the lateral torso of moose. with some experience it is relatively easy to use the second method and assign moose to 1 of 5 categories of hair damage or loss: no damage to hair coat, slight (which is approximately 5 20% of winter hair broken or lost), moderate (~ 20 40%), severe (~ 40 80%), and ghost moose (> 80%). because grooming against ticks, and the resulting hair damage and loss, continues to late april (welch et al. 1990), it is best to do surveys for hair damage as late in the season as possible. unfortunately, moose managers must do their surveys well before mid-april taking advantage of snow and moose behavior. however, annual surveys done earlier, preferably late february, can provide comparative trends in hair damage/loss, as long as they are done at approximately the same time each year (see welch et al. 1990 and samuel and welch 1991 for temporal patterns of the progression of hair loss). hair damage and loss should coincide with tick numbers given that grooming against ticks by moose is proportional to tick bite; i.e., the more ticks present, the more blood-feeding by ticks, which results in more grooming and subsequent hair damaged or lost (mooring and samuel 1998). wilton and garner (1993) calculated a hair-loss severity index (hsi) to summarize category-type data (fig. 2). the two methods, mean % hair loss and hsi, were compared using hair-loss data for moose from elk island national park (fig. 2). yearly changes were parallel, in general, particularly so after the major die-off of moose in late winter-spring of 1982. the same was true for yearly change in hsi and tick numbers (fig. 3) and yearly change in mean % damage to the winter hair coat and mean number of ticks (fig. 4). in summary, annual surveys of tick-caused damage to winter hair of moose, done as late as possible in the winter-spring season, provide good indication of numbers of ticks on moose. unfortunately, more repetitive annual surveys are needed to determine if hair loss surveys are useful to predict moose population response in subsequent years (but see next paragraph on data currently being collected at isle royale national park). numbers of moose samuel (2004) summarized a cascade of events, starting with increasing moose numbers that led to lethargic, ill, or dead moose. the cascade goes as follows: moose numbers increase in a local area; tick numbers increase; tick bite increases, which causes more itching and grooming; hair damage and loss increases; energy expended to grooming increases; blood winter ticks and die-offs of moose – samuel alces vol. 43, 2007 44 loss increases; appetite and feeding by moose possibly suppressed; smaller energy reserves available for blood replacement and to support high rates of grooming; and increased cost of replacing heat energy lost through a damaged hair coat. all of this results in lethargic, ill, or dead moose. this cascade assumes that numbers of ticks track changes in moose numbers. unfortunately, long-term data sets to test this assumption are scant because of the for both moose numbers and tick numbers on an annual basis for many years; 12 years data were collected at elk island national park (fig. 5). currently, similar data are collected only at isle royale national park where hair loss is monitored as an index of tick numbers (delguidice et al. 1997, peterson and vucetich 2006). samuel and colleagues (see samuel 2004) studied winter ticks in a closed system at elk island national park in central alberta, 1978-1996. the park is 195 km2 in size, and surrounded by fences that prevent the large ungulates from leaving. it is dotted with numerous wetlands with aspen (populus spp.) the dominant tree. coyotes (canis latrans) are the only resident large carnivore. the dynamics of the local ungulates were described by blyth and hudson (1987) and blyth (1995). some moose die-offs in the park were concurrent with die-offs of moose populations across central alberta (table 1), most recently in late winter-spring 1982, when tick-related dieoffs occurred throughout alberta. however, there was no major tick-mediated die-off in the park in 1999 (n. l. cool, parks canada, elk island national park, personal communication) when the province was experiencing major losses everywhere (pybus 1999). similarly, though die-offs occurred in local populations throughout much of alberta in 2002, including central alberta, there was no 5 20 35 50 19 78 19 80 19 82 19 84 19 86 19 88 19 90 m e a n % h a ir d a m a g e 1.5 2.5 3.5 s e v e r it y in d e x hair damage hsi fig. 2. comparison of mean % hair damage/loss and hair-loss severity index, two methods used to determine yearly changes in tick-caused damage to the winter hair of moose in elk island national park. categories of hair damage on 327 moose (variable numbers per year), observed from the air and ground, were used to calculate a hair-loss severity index (hsi). moose in categories were grouped subjectively as follows: no damage to hair coat of lateral torso (hsi class value = 1); slight damage (~ 5 – 20% of winter hair broken or lost) (hsi = 2); moderate damage (~ 20 – 40% broken/lost) (hsi = 3); severe damage (~ 40 – 80% broken/lost) (hsi = 4); and ghost moose (> 80% broken/lost) (hsi = 5). the frequency of moose in each category was multiplied by the appropriate class value and the sum divided by the total number of moose sampled for that time period. the second method involved diagrams of the hair damage on the lateral torso of 302 moose (variable numbers per year), surveyed from the air and ground, that were digitized for hair damage using methods in samuel and welch (1991). 10000 25000 40000 55000 70000 19 78 19 80 19 82 19 84 19 86 19 88 19 90 m e a n n u m b e r t ic k s 1.5 2.5 3.5 h a ir lo s s s e v e r it y in d e x ticks hsi fig. 3. annual changes in mean number of winter ticks collected from digested hides and hair-loss severity index (hsi) for moose from elk island national park. hides of 118 moose (variable numbers per year) were digested and ticks collected using techniques in welch and samuel (1989). see figure 2 for information on hsi. alces vol. 43, 2007 samuel winter ticks and die-offs of moose 45 die-off in elk island national park, though a die-off of about 150 moose occurred the next year in the main park area (n. l. cool, parks canada, elk island national park, personal communication). moose and tick numbers (and hair damage and loss, fig. 4) were monitored in the northern part of the park (main park area, 136 km2) in 1978-1990. in general, there was a lag of 1 year in tick numbers tracking moose numbers (fig. 5). die-offs occurred when moose densities approached 3 moose per km2 and mean numbers of ticks on moose approached 50,000 – 60,000. data for estimates of numbers of moose dying in some years are somewhat misleading because various forms of moose management were implemented when moose densities were high. for example, the park culled moose in some years (blyth and hudson 1987, blyth 1995), and culls are included in the annual declines in moose density (fig. 5). the number of culled moose in main park, where the tick research was done, cannot be separated from the estimates of moose dying naturally. culls were done in december 1977 and 1980 in an attempt to balance the moose population with estimated food resources. nonetheless, main park suffered a number of natural, tick-mediated losses, including an estimated 100 moose in winter 1977 (samuel and barker 1979), “extreme levels of mortality” (blyth and hudson 1987) in winter 1982, and smaller natural losses of moose in 1978, 1984, 1988, and 1989 (blyth 1995). chemical reproductive inhibition was administered to 59 adult females in main park in 1987, almost 25% of the mature cows (blyth 1995), and 108 moose were shot from 1978-1990 for research on winter ticks. pybus (1999) summarized 1,130 occurrences involving moose in trouble or found dead (n = 311) during a winter (1999) of reports (n = 1,035) involved winter tick-caused hair loss and indicated increasing occurrence with increasing density of moose. in summary, in spite of a mixture of somewhat subjective data sets, there appears to be a host-density component to tick numbers on moose. however, more reliable long-term data and objective studies are needed to ascertain this relationship. vegetation structure where a blood-fed, adult female winter tick drops from a moose in late winter-early spring, is where she must survive and produce offspring. drew and samuel (1986), aalangdong (1994), and aalangdong et al. (2001) found that different habitats in elk island park, with different microclimatic conditions, 10000 25000 40000 55000 70000 19 78 19 80 19 82 19 84 19 86 19 88 19 90 m e a n n u m b e r t ic k s 0 10 20 30 40 m e a n % h a ir d a m a g e ticks hair damage fig. 4. annual changes in mean number of ticks collected from digested hides and extent of hair damage and loss in spring on moose, from elk island national park. see figures 2 and 3 for samuel (2004). 0 20000 40000 60000 80000 19 77 19 79 19 81 19 83 19 85 19 87 19 89 m e a n n u m b e r ti c k s 0.5 1.5 2.5 3.5 m o o s e d e n s it y ticks moose fig. 5. estimated density of moose (late fall, prior to implementation of management options; see text) and mean number of ticks collected from digested hides in main park area of elk island (2004). winter ticks and die-offs of moose – samuel alces vol. 43, 2007 46 ticks. aalangdong et al. (2001) put ticks in gauze bags ~ 2 cm beneath the litter in various habitat types in elk island national park. all 3 habitat types with open canopies were more suitable for winter tick survival and production of larvae than the 4 habitat types with closed canopies. thus, more ticks survived to produce eggs, and more eggs hatched to larvae, and the larvae survived longer, in habitat types with open canopies. the authors attributed this to mean monthly temperatures at ground level, which in summer were several degrees lower in all 4 ‘closed’ habitats, than in ‘open’ habitats. these results might have broader implications and help explain why moose in parts of alberta with open canopies, such as the aspen-rich peace river country of northwest central alberta, and the aspen parklands and southern edges of the boreal mixed wood forests of central alberta, appear to suffer more from winter ticks than moose in more northerly spruce-dominated forests. summary die-offs of moose are complex events that are probably mediated by several factors including winter weather, habitat conditions, density of moose, and winter ticks. the fact that some tick-associated declines in moose are concurrent in many populations in a number involved (delguidice et al. 1997). the hypothesis that weather in the form of shorter, warmer winters with less precipitation than usual results in more ticks the following year was makes sense, but it is probably more complex than presented here. as holmes (1995) complex biological phenomena” is wishful thinking. more rigorous analysis of climatic variables important to winter ticks in regions with moose die-offs (delguidice et al. 1997, vucetich and peterson 2004) is needed, along with more long-term study of moose and tick numbers via monitoring of damage and loss to winter hair coat of moose; such monitoring is now done only at isle royale national park. assuming future climate change, minimum winter temperatures at high northern latitudes are expected to rise more than temperatures in other seasons (intergovernmental governmental panel on climate change 1996). this will most certainly impact late winter-spring survival and development of off-host stages of ticks (lindgren et al. 2000), in this case, winter ticks. acknowledgements the long-term contribution of many colleagues at elk island national park is appreciated. the natural sciences and engineering research council of canada supported much of the work done in elk island national park. references aalangdong, o. i. 1994. winter tick (dermacentor albipictus) ecology and transmission in elk island national park, alberta. m.sc. thesis, department of biological sciences, university of alberta, edmonton, alberta, canada. _____, w. m. samuel, and a. w. shostak. 2001. off-host survival and reproductive success of adult female winter ticks, dermacentor albipictus in seven habitat types of central alberta. journal of the ghana science association 3:109-116. blyth, c. b. 1995. dynamics of ungulate populations in elk island national park. m.sc. thesis, department of agricultural, food and nutritional science, university of alberta, edmonton, alberta, canada. _____ and r. j. hudson. 1987. a plan for the management of vegetation and ungulates, elk island national park. elk island national park and department of animal science (now, agricultural, food and nutritional science), university of alberta, edmonton, alberta, canada. alces vol. 43, 2007 samuel winter ticks and die-offs of moose 47 delguidice, g. d., r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restriction, ticks, and numbers of moose on isle royale. journal of wildlife management 61:895-903. drew, m. l., and w. m. samuel. 1985. factors affecting transmission of larval winter ticks, dermacentor albipictus (packard), to moose, alces alces (l.), in alberta, canada. journal of wildlife diseases 21:274-282. _____ and _____. 1986. reproduction of the winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64:714-721. _____ and _____. 1989. instar development and disengagement rate of engorged female winter ticks, dermacentor albipictus (acari: ixodidae), following singleand trickle-exposure of moose (alces alces). experimental and applied acarology 6:189-196. garner, d. l., and m. l. wilton. 1993. the potential role of winter tick (dermacentor albipictus) in the dynamics of a south central ontario moose population. alces 29:169-173. holmes, j. c. 1995. population regulation: a dynamic complex of interactions. wildlife research 22:11-19. lindgren, e., l. tälleklint, and t. polfeldt. 2000. impact of climatic change on the nothern latitude limit and population density of the disease-transmitting european tick ixodes ricinus. environmental health perspectives 108:119-123. mech, l. d., r. e. mcroberts, r. o. peterson, and r. e. page. 1987. relationship of deer and moose populations to previous winters' snow. journal of animal ecology 56:615-627. mooring, m. s., and w. m. samuel. 1998. the biological basis of grooming in moose (alces alces): programmed versus stimulus-driven grooming. animal behaviour 56:1561-1570. _____ and _____. 1999. premature winter hair loss in free-ranging moose (alces alces) infested with winter ticks (dermacentor albipictus) is correlated with grooming rate. canadian journal of zoology 77:148-156. peterson, r. o., and j. a. vucetich. 2003. ecological studies of wolves on isle royale. annual report 2002-2003. michigan technological university, houghton, michigan, usa. _____ and _____. 2006. ecological studies of wolves on isle royale. annual report 2005-2006. michigan technological university, houghton, michigan, usa. pybus, m. j. 1999. moose and ticks in alberta: a dieoff in 1998/99. occasional paper no. 20. fisheries and wildlife management division, edmonton, alberta, canada. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1, federation of alberta naturalists, edmonton, alberta, canada. _____ and v. crichton. 2003. winter ticks and winter-spring losses of moose in western canada. the moose call 16:15-16. samuel, w. m. 1989. locations of moose in northwestern canada with hair loss probably caused by the winter tick, dermacentor albipictus (acari: ixodidae). journal of wildlife diseases 25:436-439. _____ and m. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869), on moose, alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15:303-348. _____, _____, and t. leighton. 2002. high moose mortality from winter tick – spring 2002. canadian cooperative wildlife health centre newsletter 9:9. _____, m. s. mooring, and o. i. aalangdong. 2000. adaptations of winter ticks (dermacentor albipictus) to invade moose and moose to evade ticks. alces 36:183195. winter ticks and die-offs of moose – samuel alces vol. 43, 2007 48 _____ and d. a. welch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27:169-182. timmermann, h. r., and h. a. whitlaw. 1992. selective moose harvest in north central ontario-a progress report. alces 28:1-7. vucetich, j. a., and r. o. peterson. 2004. the influence of top-down, bottom-up, and abiotic factors on the moose (alces alces) population of isle royale. proceedings of the royal society of london, series b, 271:183-189. welch, d. a., and w. m. samuel. 1989. evaluation of random sampling for estimating density of winter ticks (dermacentor albipictus) on moose (alces alces) hides. international journal of parasitology 19:691-693. _____, _____, and r. j. hudson. 1990. bioenergetic consequences of tick-induced alopecia on moose. journal of medical entomology 27:656-660. wilton, m. l., and d. l. garner. 1993. preliminary observations regarding mean april temperature as a possible predictor of tick-induced hair-loss on moose in south central ontario, canada. alces 29:197-200. 4211(75-87).pdf alces vol. 42, 2006 rodgers and robins moose detection at night 75 moose detection distances on highways at night arthur r. rodgers1 and patrick j. robins2 1centre for northern forest ecosystem research, ontario ministry of natural resources, 955 oliver road, thunder bay, on, canada p7b 5e1; 2forensic engineering inc., 1439 legion road, burlington, on, canada l7s 1t6 abstract: moose-vehicle collisions are a serious concern in many areas of north america and fennoscandia. in northwestern ontario, more than 400 moose-vehicle collisions occur annually, and 26 fatal collisions have occurred over the last 10 years. to avoid colliding with a moose, a motorist must: (1) successfully see or detect the presence of the animal; (2) determine whether or not the moose poses a threat requiring evasive action; (3) determine what action, if necessary, is required; and (4) implement the action. whereas perception-reaction times of motorists have been studied in detail, allowing calculations of post-detection distances travelled by a vehicle at different speeds, distances to determine the distances at which an animal could be detected at night when it was positioned on each shoulder and in the middle of a highway using high and low beam headlamp settings of different vehicles. overall, we found the mean detection distance across all vehicle types, headlamp settings, tor; on the low beam setting, mean detection distance was 74 m and on the high beam setting it was 137 m. moose decoy location was also important; combining the data for both headlamp settings, mean detection distances were 89 m, 93 m, and 133 m for the left, right, and centre positions, respectively. there was no relationship between headlamp height of different vehicles and moose detection distance. nation capabilities of their headlamps for moose encounters. for drivers using a low beam headlamp alces vol. 42: 75-87 (2006) key words: alces alces, detection distance, moose-vehicle collisions, mvcs, ontario, visibility distance collisions between moose (alces alces) and motor vehicles are a serious concern in many areas of north america and fennoscandia (grenier 1973, child and stuart 1987, lavsund and sandegren 1991, mcdonald bartley 1991, child 1998, joyce and mahoney 2001, lavsund et al. 2003, seiler 2003, timmermann and rodgers 2005). at least 3,000 moose-vehicle collisions occur annually across north america (child 1998); a highly conservative estimate since many accidents are not reported and most jurisdictions do not maintain accurate records (child and stuart 1987, romin and bissonette 1996, sullivan and messmer 2003, transport canada 2003). in northwestern ontario alone, more than 400 moose-vehicle collisions were reported in 2002 (staff sergeant r. beatty, ontario provincial police, unpublished data 2004). siderable numbers of moose and can result in substantial property damage, human injury, moose detection at night – rodgers and robins alces vol. 42, 2006 76 and death; 20% of moose-vehicle collisions result in injuries with a 0.5% human fatality rate (garrett and conway 1999, transport canada 2003) and 26 human fatalities have resulted from collisions between vehicles and wildlife in northwestern ontario over the last 10 years (transport canada 2003). the economic costs associated with moosevehicle collisions include the material loss of vehicles, human injuries (ambulances, medical expenses, disability payments), human fatalities (life insurance, funeral expenses), call-out costs for police, veterinarians, and moose, loss of meat and hunting opportunities, delays (seiler 2003, timmermann and rodgers 2005); at an average cost of cdn$4,500 per accident, including only vehicle damage and loss of meat value (transport canada 2003), the economic cost of reported moose-vehicle collisions is at least cdn$13,500,000 annually in north america. notwithstanding the potentially severe social and economic consequences of moosevehicle collisions, these accidents can directly reduce moose population numbers locally or affect their productivity through alteration of sex and age ratios (leopold 1933, peterson 1955, child 1998). in north america, moose mortalities resulting from collisions with vehicles correspond to about 4% of the annual allowable moose harvest, ranging from 0.3% in manitoba to 196% (i.e., almost double the annual allowable harvest) in new hampshire (child 1998). of 1,673 non-hunting moose mortalities recorded in northeastern ontario over a 10-year period (1983-1991), 48% were attributed to motor-vehicle collisions; total incidental fatalities were almost double the combined losses to predation, subsistence (child 1998). clearly, there is good reason to consider the importance of moose-vehicle collisions in the development of sustainable moose population management programs and the setting of harvest objectives. moreover, in areas where collisions with motor vehicles populations, additional management actions accidents. a wide range of measures to reduce moose-vehicle collisions have been applied in various jurisdictions, with greater or lesser degrees of success, including; public education programs (e.g., pamphlets, posters, bumper create high quality habitat in areas away from highway corridors, vegetation management to widen transportation routes and improve roadside visibility, adjustments of travel speed, improved lighting and signage, construction of physical structures (i.e., fencing, one-way mirrors and ultrasonic warning devices, ultraviolet (uv) headlamps, and, more recently, development of intelligent transportation systems (e.g., microwave radar, infrared images, imaging) (child 1998, forman et al. 2003, jhwf 2003, transport canada 2003, timmermann and rodgers 2005). of these, properly maintained fencing appears to be the most effective, but is impractical for extensive use because of high installation and maintenance and bartley 1991, forman et al. 2003, jhwf 2003, transport canada 2003). alternatively, a combination of vegetation management and 1991, child 1998). however, management of vegetation may only provide temporary reductions in moose-vehicle collisions and if not maintained to limit the growth of early seral vegetation that may attract moose to highway corridors (child 1998). regulating vehicle speed, on the other hand, is inexpenalces vol. 42, 2006 rodgers and robins moose detection at night 77 sive to implement and maintain relative to other measures. most moose-vehicle collisions occur between 1800-0200 hrs on straight and relatively where visibility is limited by encroaching vegetation (stuart 1984, child et al. 1991, del jhwf 2003). to avoid an accident, drivers must successfully: (1) detect the presence of a moose; (2) determine whether or not threat that will require an evasive response; (3) determine what action (e.g., steering or and (4) if necessary, implement the chosen action (olson 1996, olson and farber 2003). some amount of time will pass from when action is completed, during which the vehicle will cover some or all of the distance between the vehicle and the moose. how much of that distance will be traversed depends on: (1) the (2) how fast the vehicle is travelling; (3) how an evasive manoeuvre. whereas perceptionreaction time (olson 1996, olson and farber 2003) and the time and distance required to given travel speed (russell 1999) have been documented, no data are available pertaining to actual driver detection distances for moose at night. the intent of this study was to determine the distance at which a driver operatthe presence of a moose. we also attempted to ascertain whether or not detection distance was related to variation in headlamp heights of different vehicle types. this information was then used in comparisons with previously ing data to estimate travel speeds that may be implemented along highway corridors to motor vehicles and moose. study area the study was conducted on an 800 m straight and level section of highway 527 canada. the 2-lane segment of highway used in the tests was asphalt covered with opposand roadway edges demarcated by 3m-wide cuts through natural forest (primarily balsam poplar, populus balsamifera, trembling aspen, p. tremuloides, and white spruce, picea glauca on the west side of the highway was cleared distance of about 7 m and on the east side to almost 20 m. the section of highway used was intersected by several game trails showof use by moose, thereby providing a realistic setting for the study. a s m a l l c l e a r i n g ( n 4 8 º 3 3 ’ 4 6 ” , w89º08’08”) on the west side of highway 800 m section of highway 527, approximately canada, used in determinations of moose detection distances on a highway at night. moose detection at night – rodgers and robins alces vol. 42, 2006 78 test site, was used as a staging area for drivers from view of the test segment by an almost and rolling topography. methods moose surrogate as it would have been impractical to control the behaviour of a live moose for the purpose of this study, a decoy was employed bull moose decoy constructed from foam, real antlers. the moose-hide covering was critical in simulating the luminance properties of a moose at night. the moose surrogate was located about 600 m from the start of the test section of the highway, and just north of an existing natural game trail. during the trials, the moose surrogate was set up on one of the shoulders or in the centre of the highway and always faced west. drivers the test subjects in this study consisted of 14 drivers from the local geographic area who ranged in age from 20 to 55 yrs. the mean and median ages of the tested drivers were 38 and 40 yrs, respectively. four subjects were female, 10 were male. three subjects required no corrective eyewear while driving, 3 wore contact lenses, and 8 wore eyeglasses while driving. vehicles and headlamp heights seven vehicle types were used in this study (table 1). the vehicles were chosen to can highway motor vehicles and represented a variety of standard headlamp types and heights above ground (measured to the middle of the headlamp on each vehicle). the only common vehicle type not included in the test surface similar to that of a highway bus. the for proper alignment prior to the trials. test conditions at the time of the tests (2300-0430 hrs) with a quarter moon that set at about 0100 hrs. and the road was still damp in sections. as a result, when the air cooled through the night from +5ºc at the beginning of the trials to -4ºc at their conclusion, a sporadic low rolling fog condition was observed throughout the test area. the section of road near the moose surrogate target, however, was clear during all trials and the pavement was dry for most of the tests. during the tests, the highway was closed vehicle type year model headlamp type headlamp height (cm) motorcycle 2002 yamaha v-star 1100 classic sealed beam 87 highway tractor 1998 international 90s halogen 103 minivan 2004 dodge caravan halogen 74 automobile (halogen) 2003 ford focus halogen 65 automobile (hid) 2004 kia amanti xenon hid 71 2004 ford f-150 halogen 98 sport utility vehicle 1995 jeep halogen 85 table 1. headlamp characteristics of test vehicles used in determinations of moose detection distances on a highway at night. alces vol. 42, 2006 rodgers and robins moose detection at night 79 north of the test area. this allowed test vehicles to move safely at slow speeds and prevented any effects on visibility that might be caused test procedure in total, there were 42 test trials. each driver was randomly assigned to a single vehicle (2 drivers per vehicle type) and to a single headlamp setting condition (high beam or low beam) for that vehicle, with the exception of 4 subjects, who by virtue of requiring cial tractor or motorcycle, were assigned to the appropriate vehicle type. each of these 4 drivers, however, was assigned either the high beam or low beam condition on a random basis. thus, each subject drove one of the test vehicles on either the high beam setting or the low beam setting, but not both. each of the 14 subjects drove their assigned test vehicle 3 times, one for each moose location (left shoulder, centre of driving lane, and right shoulder). the order of the trials with respect to driver, vehicle type, and headlamp setting was randomly determined. to drive slowly through the test area using the assigned high beam or low beam headlamp setting until the moose surrogate was visually detected, then bring the vehicle to a full and immediate stop. one of the investigators acto ensure that the subjects were able to judge and maintain an approach speed of about 10-15 once the vehicle was fully stopped, luminance readings for the moose surrogate target the vehicle with a hagner universal photometer model s2 (b. hagner ab, solna, sweden) capable of detecting light levels as low as 0-1 lux with an accuracy of ± 3%. however, in spite of the sensitivity of the photometer moose surrogate to the photometer placed at the front of the vehicle was so low (< 1 lux) that these measurements were abandoned after moose to the front surface of the vehicle was measured with a laser technology impulse laser model 200xl (laser technology, inc., centennial, colorado, usa) that can measure up to 2,200 m with a typical accuracy of ± 1 m and an accuracy of ± 2 m at the maximum distance. following these measurements, the test subject turned the vehicle around and returned to the staging area. statistical analysis the dependent variable in this study, as the linear distance, to the nearest meter, between the moose target and the front surface of the vehicle at the point where the driver stopped the vehicle after visually detecting the presence of the moose surrogate on the highway. in addition to presenting the means (± sd) and medians (range) of these data from the trials, results are expressed in terms of the 15th percentiles to denote the visibility distances at which most drivers would be able to detect the presence of a moose on a highway under these test conditions. in this type of drivers within the 85th percentile, thereby excluding the 15% of tested subjects with the shortest detection distances for a particular set of conditions (olson and farber 2003). the study employed a 2 x 3 factorial design with repeated measures of detection distance independent variable was headlamp setting, for which 2 levels were established: high beam and low beam. the second independent variable was moose location, for which there were 3 levels: left side of the highway, centre of the driving lane, and right side of the highway. each subject experienced all 3 moose location conditions, producing repeated measures on the moose location variable. subsequently, a 2 x 3 factorial anova (spss 13.0, spss inc., moose detection at night – rodgers and robins alces vol. 42, 2006 80 the detection distance data. vehicle type was not directly analysed as a variable of interest because of limited confounded with other variables such as driver, type of headlamp system, etc. we were also unable to determine any relationships between detection distances and the types of headlamp systems (i.e., sealed beam vs halogen vs high intensity discharge) of different vehicles simple linear regression (spss 13.0, spss inc., chicago, illinois, usa) was used to explore the relationship between detection distances and variation in headlamp heights among different vehicle types. distances required to bring a vehicle to a safe stop from selected speeds were calculated data measured in previous studies (russell 1999, olson and farber 2003). these required stopping distances were compared to fully adjusted detection distances of test drivers to determine whether or not there would be sufa collision when travelling on a straight and adjustment of detection distances for test drivers before data obtained in this study could be used in comparisons with drivers in the real world, measured detection distances needed to be adjusted for the perception-reaction and stopping distances of the test drivers, as well as their expectancy of encountering the moose surrogate. from the time a driver detects the presence of an unexpected object-of-interest, to the time the driver is able to initiate some evasive response, perception-reaction will require about 0.50 – 1.25 secs for most (i.e., 85%) drivers (olson and farber 2003). react. thus, the minimum perception-reaction time of 0.5 secs is appropriate for test subjects. since the speed of the vehicle during or less, it would have travelled as much as to react to the moose surrogate and apply the (russell 1999), an additional distance of vehicle to a comfortable but decisive stop from accordingly, the measured detection distances were adjusted by adding 5 m to account for activities of test drivers. the test subjects in this study. real drivers night-time visibility studies suggest an additional 0.5 secs is a reasonable adjustment to the expected detection time for real drivers at night compared to experimental test drivby drivers in the real world is subsequently calculated as a function of vehicle speed; e.g., would be 6.95 m closer to the moose surrogate when it was detected than one of the test drivers. these distances were calculated for real drivers travelling at a range of different detection distances previously adjusted to account for the perception-reaction processes calculation of stopping distances for real drivers whereas a minimum perception-reacalces vol. 42, 2006 rodgers and robins moose detection at night 81 tion time of 0.5 secs may be suitable for test the highway at night, as above, a maximum perception-reaction time of 1.25 secs, as measured in previous studies (olson and farber 2003), is more appropriate for most (i.e., 85%) real-world drivers. thus, from the time that a moose is detected on a highway at night, it is expected that all but the 15% of drivers with the slowest perception-reaction times will be able to initiate an evasive manoeuvre within 1.25 secs. the distance travelled by a vehicle during the driver’s perception and reaction is speed dependent and is simply the arithmetic tion-reaction time of 1.25 secs (table 2). to a stop as the evasive action. some other action such as steering, a speed reduction, or sounding a warning with the horn would complete stop, so our calculations account achieve an evasive response. the following directly incorporate the gravitational rate of complete stop from a particular speed: f s d 9.25 2 (1) where, d = distance required to stop (m); s f = deceleration rate and sliding on a well-travelled dry asphalt surface; russell 1999), the distances required to bring the vehicle to a complete stop from selected speeds are given in table 2. results main effects on detection distance were found for both headlamp setting (f = 35.77; df = 1, 12; p = 0.00006) and moose location (f = 6.56; df = 2, 16; p = 0.008), but there was these two variables. headlamp setting the mean (± sd) and median (range) moose detection distances for the low beam headlamp setting were 74 m (± 29 m) and 75 m (23 – 124 m), respectively. the 15th percentile value was 47 m, indicating that most (85%) of the tested subjects were able to detect the presence of the moose from 47 m away or greater. the mean and median distances for the high beam headlamp condition were 137 m (± 51 m) and 147 m (28 – 210 m), respectively, with a 15th percentile value of 74 m. moose location the moose surrogate was set up at 3 locations on the highway. when data were combined for high and low beam headlamp conditions, the mean (± sd) and median (range) detection distances, respectively, for these moose locations were 89 m (± 55 m) and 64 m (23 – 189 m) for the left shoulder; 93 m (± 37 m) and 84 m (28 – 172 m) for the right shoulder; and 133 m (± 54 m) and 124 m (40 – 210 m) for the centre of the 50 60 70 80 90 100 110 120 distance travelled (m) perception-reaction 17 21 24 28 31 35 38 42 14 20 28 36 46 56 68 81 table 2. distances travelled by a vehicle during a driver’s perception-reaction time of 1.25 seconds moose detection at night – rodgers and robins alces vol. 42, 2006 82 driving lane. the 15th percentile values for the 3 moose location conditions were 46 m, 73 m, and 79 m for the left, right, and centre positions, respectively. vehicle type and headlamp height there was no linear relationship between headlamp height of different vehicles and moose detection distance on either the low (r = 0.001; f = 0.000, df = 1, 19, p = 0.997) or high beam setting (r = 0.167; f = 0.543, df = 1, 19, p = 0.470). nor was there any relationship between headlamp height and moose detection distance when the surrogate was located on the left (r = 0.018; f = 0.004, df = 1, 12, p = 0.951), right (r = 0.360; f = 1.783, df = 1, 12, p = 0.207), or centre (r = 0.014; f = 0.002, df = 1, 12, p = 0.962) of the driving lane. total data set comparisons with required stopping distances, we combined the detection data across all vehicle types, headlamp settings, and moose location conditions, which produced mean and median detection distances of 105 m (± 52 m) and 99 m (23 – 210 m), respectively, with a 15th percentile value of 54 m. adjusted detection distances for test drivers although moose location was found distance, drivers in the real world obviously cannot predict or control the position of a live moose on a highway. on the other hand, real drivers can control the headlamp setting of their vehicle. thus, adjustments were made to the 15th percentile detection distances of test drivers for the low and high beam headlamp setting conditions, as well as the total data set, for vehicles travelling at different speeds. as previously outlined, detection disaccount for the perception-reaction processes example, the 15th percentile value for moose detection distance in the total data set is increased to 59 m; for low beam and high beam settings, values are adjusted to 52 m and 79 m, respectively. next, adjusted test values were reduced by the additional distance required for real world travelling at a particular speed; i.e., the distance travelled in an additional 0.5 secs at a given speed. thus, for a vehicle travelling and the fully adjusted 15th percentile values were 45m for the low beam setting, 72 m for the high beam setting, and 52 m for the combined data set (table 3). fully adjusted moose detection distances estimated for real-world drivers travelling at other selected speeds are given in table 3. for comparisons with fully adjusted detection distances of test drivers, distances travelled by a vehicle during a real-world driver’s estimated perception-reaction time (table 2) to estimate the total distances required, following detection of a moose, to bring a vehicle to a safe stop from selected speeds (table 3). discussion when the distance required to perceive a moose on the road, react, and stop a vehicle exceeds the available detection distance at a given speed (table 3), then a collision will above that speed, the greater the impact and potential consequences of a collision. conversely, if the moose detection distance exceeds that required by a driver to perceive, react, and bring a vehicle to a safe stop from a given speed, then it is expected that a moosevehicle collision can be avoided. based on the setting, moose location, vehicle type, driver, etc.), the required stopping distance exceeds alces vol. 42, 2006 rodgers and robins moose detection at night 83 or more (table 3). thus, drivers can avoid a moose-vehicle collision 85% of the time by at night. in most jurisdictions, however, it is recommended that drivers use the high beam headlamp setting on their vehicle when travelling on highways at night; e.g., in ontario, drivers are expected to use the high beam headlamp setting at night whenever possible and switch to the low beam setting within 150 m of an oncoming vehicle or when following a vehicle within 60 m. in the high beam condition, the required detection distance to perceive, react, and stop a vehicle exceeds the available distance at speeds of 80-90 the required detection distance exceeds the available detection distance at speeds of 60-70 section of highway where the visibility trials is in agreement with the required detection distance on the high beam setting described in this study but too high for the low beam condition or combined data. although moose location is unpredictable in real-world situations, we found that visibility distance was affected by the location of the moose surrogate on the highway. based on the 15th percentile values, the surrogate was detected further away when placed in the centre of the driving lane (79 m) or on the right shoulder (73 m), than on the left shoulder of the highway (46 m). this is consistent with transport canada (2001) regulations that ensure headlamps are aligned so the light does not project up or towards low beam setting; high beam headlamps are aimed so the brightest spot is centred at the same height as the headlamp. thus, reduced detection distances when the moose surrogate was located on the left shoulder of the highway were largely the result of measurements made on the low beam headlamp setting. lamp heights of different vehicles and moose detection distance, regardless of headlamp setting or the location of the moose surrogate headlamp alignment according to transport canada (2001) regulations. although we expected detection distance might increase with 50 60 70 80 90 100 110 120 required distance (m) 31 41 52 64 77 91 106 123 available distance (m) low beam 45 44 42 41 39 38 37 35 (n = 21) high beam 72 71 69 68 66 65 64 62 (n = 21) total 52 51 49 48 46 45 44 42 (n = 42) to a complete stop from selected speeds (sum of distances travelled during a perception-reaction 2) with distances available to complete the evasive manoeuvre based on moose detection distances expectancy, while travelling on a highway at night using low or high beam headlamp settings (n = number of trials). when the distance required exceeds the available detection distance at a given moose detection at night – rodgers and robins alces vol. 42, 2006 84 height of the headlamps above the roadway surface, any potential improvement afforded to vehicle types with higher headlamps was negated by angling them downward to prevent low beam setting; this downward projection would also affect the aim of headlamps on the high beam setting. this study to real life depends on the degree to which the subjects, conditions, and procedures employed, correspond to those that would be expected in the real world. the vehicles used on canadian roadways. the subjects were real drivers from the same geographic locale as the study and a real section of highway, which is normally frequented by moose, was used for the tests. the moose surrogate was conditions reproduced in the investigation procedures were as close an approximation to what a real driver would face, as would be possible in a study such as this. nonetheless, there are a multitude of factors that might diminish the ability of drivers in the real world to detect, perceive, react, and avoid colliding matter, on a highway at night. the drivers in this study lived in a region in which moose encounters are frequent and many of them had experienced moose encounters on the roadway in the past. as such, this group of drivers, as a whole, could be considered ideal. many of these subjects were able to identify the presence of the surrogate moose target when only the lighter coloured lower legs were illuminated by the headlamps. an inexperienced motorist with regard to moose encounters might require additional detection time and therefore be closer to a moose upon completion of detection, leaving less room to additionally, data collected in this study sober, alert, and unusually attentive drivers. a driver who is fatigued, momentarily distracted by a passenger or in-vehicle device, or otherwise momentarily inattentive, would be expected to be closer to the moose at the point of detection than has been determined herein. these test trials involved a static target. the moose surrogate was placed on the roadway and “stood still” for each trial. live moose are highly unpredictable and often in motion during encounters with vehicles; they can enter the roadway quite suddenly. through their movement, the light leg colours, vulva moose may be easier for motorists to detect than the stationary moose surrogate used in these trials. while a moose in motion might provide additional visual cues that assist in would nonetheless face more complex and challenging avoidance situations in moosevehicle encounters than did the test subjects in this investigation. the results of this study must also be considered preliminary because practical constraints restricted the variables examined to headlamp height, setting, and moose location. vehicle types varied, but not systematically. numbers of subjects to permit independent, rather than repeated measures across the moose location variable. ideally, sample unique detection distance summary data for headlamp illumination, moose location, vehicle type, and headlamp type conditions, since these characteristics are normally fairly moose-vehicle collision. of course, these trials should be conducted under a variety of weather conditions at different times of the year and on various road surfaces. while luminance values were not a critical measure of interest in this study, it might be useful in the future to alces vol. 42, 2006 rodgers and robins moose detection at night 85 ascertain the luminance differences between the moose surrogate employed in this study and that expected from a live animal. it is possible that the sheen from a clean coat or the drab appearance of a wet and dirty coat of a live animal could present visual cues for real drivers that are different from those inherent in the decoy used in the present investigation. follow-up investigations should also consider ways to test differences in detection distances for moose at night between stationary and moving targets. in the present endeavour the target remained stationary during each trial and only its location across the highway was varied. a moving moose might present a motorist with a different set of visual and cognitive challenges. management implications our results (table 3) suggest that most drivers travelling at speeds in excess of about to be overdriving the illumination capabilities of their headlamps for moose encounters. even when the high beam headlamp setting is previous studies that have found an increase in moose-vehicle collisions when travel speed al. 2003). thus, it would be prudent to suggest that along highway corridors where collisions with motor vehicles present a serious threat pacts on local moose populations, speed limits unfortunately, lowering speed limits is not generally favoured or supported by motorists or road authorities (lavsund and sandegren 1991). speed limit signs do little to change driver behaviour and motorists travel at speeds determined by their perception of roadway ed speed limits (romin and bissonette 1996, putman 1997). thus, it may be more acceptable to recommend diurnal or seasonal speed limit reductions (jhwf 2003). for example, approximately 70% of wildlife-vehicle collisions in northwestern ontario occur between june and october (staff sergeant r. beatty, ontario provincial police, unpublished data 2004), suggesting a reduction of speed limits during that period. both texas and montana appear to be too high, according to the present study, to effectively reduce wildlife-vehicle collisions, and only montana has moose. based on previous studies (stuart 1984, forman et al. 2003) and our results, it would be more reasonable to recommend speed limits of these recommended speeds will not prevent moose-vehicle collisions from occurring but incidents, particularly if combined with other for speeding through areas where collisions with motor vehicles present a serious threat may get a driver’s attention and remind them to slow down, automated radar speed detectors and public service announcements of where, when, and why speed reductions are being implemented (jhwf 2003). acknowledgements of the ontario provincial police, especially staff sergeant bob beatty and rod brown, and members of the ministry of transportation of ontario, especially tom marinis, without ing of this study would not have been possible. we are grateful to manitoba conservation, vince crichton for arranging the loan of their motors in thunder bay, ontario, for the loan of the xenon lamp equipped test vehicle and moose detection at night – rodgers and robins alces vol. 42, 2006 86 highway tractor. a special debt of gratitude is expressed to the many volunteer drivers who suffered a long cold night to complete the brad allison, and 2 anonymous reviewers for their comments and suggestions on an earlier version of this manuscript. references child, k. n. 1998. incidental mortality. pages 275-301 in management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, s. p. barry, and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27: 41-49. _____, and k. m. stuart. 1987. vehicle and train collision fatalities of moose: some management and socio-economic considerations. swedish wildlife research supplement 1: 699-703. del frate, g. g., and t. h. spraker. 1991. moose vehicle interactions and an associated public awareness program on the forman, r. t. t., d. sperling, j. a. bissonette, a. p. clevenger, c. d. cutshall, v. h. dale, l. fahrig, r. france, c. r. goldman, k. heanue, j. a. jones, f. j. swanson, t. turrentine, and t. c. winter. 2003. road ecology: science and solutions. island press, washington, d.c., usa. garrett, l. c., and g. a. conway. 1999. characteristics of moose-vehicle collijournal of safety research 30: 219-223. grenier 1972. proceedings of the north ameri155-194. (jhwf) jackson hole wildlife foundation. crossing study, teton county, wyoming. wildlife foundation by biota research wyoming, usa. joyce, t. l., and s. p. mahoney. 2001. spatial and temporal distributions of moose-vehicle collisions in newfoundland. wildlife society bulletin 29: 281-291. lavsund, s., t. nygrén, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. _____, and f. sandegren. 1991. moosevehicle relations in sweden: a review. alces 27: 118-126. leopold, a. 1933. game management. mcdonald, m. g. 1991. moose movement and mortality associated with the glenn alces 27: 208-219. olson, p. l. 1996. forensic aspects of driver perception and response. lawyers and judges publishing company, limited, _____, and e. farber. 2003. forensic aspects of driver perception and response. second edition. lawyers and judges publishing company, limited, tucson, _____, and m. sivak. 1986. perceptionresponse time to unexpected roadway factors 28: 91-96. oosenbrug, s. m., e. w.e. w. mercer, and s. h. ferguson. 1991. moose-vehicle collisions in newfoundland management considerations for the 1990’s. alces 27: 220-225. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. putman, r. j. 1997. deer and road trafalces vol. 42, 2006 rodgers and robins moose detection at night 87 journal of environmental management 51: 43-57. romin, l. a., and j. a. bissonette. 1996. deer-vehicle collisions: status of state monitoring activities and mitigation efforts. wildlife society bulletin 24: 276-283. russell, c. g. 1999. equations and formulas reconstructionist. lawyers and judges publishing company, limited, tucson, schwartz, c. c., and b. bartley. 1991. reducing incidental moose mortality: considerations for management. alces 27: 227-231. seiler, a. 2003. the toll of the automobile: wildlife and roads in sweden. ph.d. thesis, swedish university of agricultural sciences, uppsala, sweden. stuart, k. m. 1984. wildlife-vehicle collisions in british columbia. report prepared for the british columbia ministry of environment and ministry of transportation and highways, victoria, british columbia, canada. sullivan, t. l., and t. a. messmer. 2003. perceptions of deer-vehicle collision management by state wildlife agency and department of transportation administrators. wildlife society bulletin 31: 163-173. timmermann, h. r., and a. r. rodgers. 2005. moose: competing and complementary values. alces 41: 85-120. transport canada. tive devices and associated equipment. technical standards document no. 108, revision 3. transport canada, standards and regulations division, motor vehicle standards and research branch, road safety and motor vehicle regulation directorate, transport canada, ottawa, ontario, canada. _____. 2003. collisions involving motor vehicles and large animals in canada. final report. prepared for transport canada road safety directorate by l-p tardif and associates, incorporated, ottawa,ottawa, ontario, canada. 101 demography and sustainable harvest rates of low-density moose in northern british columbia ian w. hatter nature wise consulting, #49-640 upper lakeview road, invermere, british columbia v0a 1k3, canada abstract: numerous moose (alces alces) populations throughout alaska, yukon, and the northwest territories occur at low-density, a condition that often persists for decades and is referred to as a low-density equilibrium (lde). the demographic conditions for these populations include low-density (≤ 0.4 moose/km2), low annual recruitment of calves (~ 0.25 calves/cow), and static population growth (λ ~ 1.00). i used data from aerial surveys and hunter harvest surveys to assess if these conditions applied to 4 moose populations in northern british columbia, canada over a 20-year period from 1996/97–2015/16. all populations exhibited low-density, low recruitment, and static growth suggesting that moose in this part of the province exist within a lde state. harvest and moose densities were positively related. harvest rates from survey data ranged from 2.4–3.2% of the total population and 6.1–10.5% of the bull population. a stochastic model was used to estimate sustainable harvest rates defined as rates where the harvest risk was ≤ 10% probability that the post-hunt bull:cow ratio dropped below a given adult sex ratio threshold after 50 years of harvest. sustainable harvest rates averaged ≤ 2.4% of the total population or ≤ 8.4% of the bull population with 0.50 bulls/cow as the threshold, ≤ 3.2% of the population or ≤ 13.0% of bulls with 0.40 bulls/cow as the threshold, and ≤ 4.1% of the population or ≤ 20.4% of bulls with 0.30 bulls/cow as the threshold. modelling indicated that even small changes in harvest rates could greatly affect the probability of bull:cow ratios dropping below adult sex ratio thresholds. research focussed on specific factors contributing to low moose density and increased population survey effort should improve estimates of sustainable harvest rates and management of moose in northern british columbia. alces vol. 58: 101–112 (2022) key words: adult sex ratio threshold, alces alces, bull:cow ratio, calf recruitment, demography, low-density equilibrium, moose, predation, sustainable harvest rate numerous moose (alces alces) populations throughout alaska, yukon, and the northwest territories occur at low-density (≤ 0.4 moose/ km2) in a state often referred to as a low-density equilibrium (lde) or low-density dynamic equilibrium (ldde), that may persist for decades (bergerud 1992, gasaway et al. 1992, stenhouse et al. 1995, lake et al. 2013, joly 2017). in this state, body condition, productivity, and survival of adult moose are relatively high and sufficient to allow population growth. however, the population does not increase because wolf (canis lupus) and bear (ursus spp.) predation limit calf recruitment. harvest is typically restricted to bull-only to prevent further decline in these predator-prey systems (gasaway et al. 1992). bergerud (1992) hypothesized that for each low-density population there is also a stabilizing recruitment of calves (rs) where the growth rate is static. in unhunted or lightly hunted moose populations rs ~ 0.25 calves/cow (bergerud 1992, bergerud and elliott 1998). messier (1994) modelled moose-wolf interactions over a broad spectrum of moose densities in north america and predicted that moose would demography and harvest of low-density moose – hatter alces vol. 58, 2022 102 stabilize at ~ 2.0 moose/km2 in the absence of predation and at ~ 1.3 moose/km2 in the presence of wolves. however, if moose productivity was diminished through deteriorating habitat quality or bear predation on calves, then a lde (0.2–0.4 moose/km2) was predicted. these demographic conditions for lde populations have received general support from additional reviews on the effects of predation on moose numbers (van ballenberghe and ballard 1994, ballard and van ballenberghe 1997, van ballenberghe and ballard 1997). hatter (1999) reviewed moose harvest management in central and northern british columbia and suggested that the demography of 4 northern moose populations adjacent to yukon and the northwest territories may exist in a lde state, and that sustainable harvest rates may be substantially lower than in higher density populations in central british columbia. the objectives of this analysis were to: 1) assess if the demographic conditions for lde populations (i.e., low-density [≤ 0.4 moose/km2], low calf recruitment [r ~ 0.25 calves/cow], and static population growth [λ ~ 1.0)]) applied to these 4 populations from 1996/97–2015/16; 2) estimate average harvest rates for each population based on survey and harvest densities during this 20-year period; and 3) determine sustainable harvest rates for moose existing in a lde state based on a stochastic demographic model. study area the study area included game management zones (gmz) 6f, 6e, 7pd, and 7pe in northern british columbia, canada (fig. 1). gmzs are amalgamations of adjacent wildlife management units which share similar ecological characteristics and hunter harvest patterns, and thus provide a suitable framework for assessing moose demography and harvest rates over large geographical areas (hatter 1999, kuzyk et al. 2018). gmzs 6f, 6e, 7pd, and 7pe comprised 29,256 km2, 44,587 km2, 38,330 km2, and 61,204 km2 of habitable moose range, respectively. six moose ecotypes have been identified in british columbia based on ecological, climatic, and physiographic differences in their habitats (eastman and ritcey 1987). the boreal upland ecotype occupies gmzs 6f, 6e, and 7pd while the fig. 1. location of game management zones 6f, 6e, 7pd, and 7pe in northern british columbia. alces vol. 58, 2022 demography and harvest of low-density moose – hatter. 103 boreal lowland ecotype occupies gmz 7pe. the boreal upland ecotype primarily occurs within the boreal black and white spruce, the subalpine spruce-willow-birch, and the alpine tundra biogeoclimatic zones. gmz 7pe is comparatively homogeneous and the boreal lowland ecotype almost exclusively occupies the boreal white and black spruce biogeoclimatic zone. in general, moose habitat quality is higher in the spruce-willowbirch zone of gmzs 6f, 6e, and 7pd and lowest in the boreal black and white spruce zone in gmz 7pe. the climate for all 4 gmzs is characterized by long, cold winters and cool, short summers. moose co-exist with wolves, grizzly bears (u. arctos), and black bears (u. americanus) in all 4 gmzs. wolf density ranged between 5–15 wolves/1000 km2 with lower densities in gmz 7pe (bc flnro 2014, serrouya et al. 2015). grizzly bear density ranged from 10–30 bears/1000 km2 in gmzs 6f, 6e, and 7pd with ≤ 10 bears/1000 km2 in gmz 7pe (bc flnro 2020). black bears occur at unknown density in all 4 gmzs. caribou (rangifer tarandus) occur throughout much of the study area, with mountain goats (oreamnos americanus) and thinhorn sheep (ovis dalli) in mountainous regions. elk (cervus canadensis), mule deer (odocoileus hemionus), and white-tailed deer (o. virginianus) are moderately abundant in the extreme southern portion of gmz 7pe, and either absent or occur at very low to low density elsewhere. small numbers of reintroduced bison (bison bison) occupy portions of gmzs 7d and 7e. licensed hunting seasons were restricted to bull-only (bag limit of 1 moose/hunter) in all gmzs with season dates generally between 15 august and 30 november. bull hunting was primarily regulated by general open seasons although limited-entry seasons with no antler restrictions, or a combination of general open and limited entry seasons occurred in some areas (kuzyk et al. 2018). only bulls with spike or fork antlers and bulls with tri-palm antlers could be harvested during september – october from 1996– 2002 in gmzs 7pd and 7pe. after 2002, bulls with at least 10 points on one antler could also be harvested. indigenous people also harvest moose at unknown levels throughout the 4 gmzs. harvest is generally tied to access from roads and waterways, and although considerable road access exists within gmz 7pe, it is limited elsewhere. methods moose and hunter harvest surveys moose were counted in helicopter surveys in each gmz, generally in mid-december through late february in suitable weather and snow conditions. stratified random block or distance sampling survey methods were used to enumerate bulls (1+ year-old males), cows (1+ year-old females), and calves (kuzyk et al. 2018); unclassified animals were typically ≤5% of total count. the survey areas were assumed representative of moose density and composition within a gmz. although herd composition counts were also conducted periodically, they were deemed less reliable and not used in this analysis. a total of 20 surveys were used to estimate density, and bull:cow and calf:cow ratios from 1996/97–2015/16 in the 4 gmzs. annual hunter surveys were used to estimate the licensed resident harvest, number of hunters, and hunter days from 1996–2015. annual mail questionnaires were sent to 15,477 (average) provincial hunters chosen randomly, with an average response rate of 68% (kuzyk et al. 2018). reporting of non-resident licenced harvest was mandatory and obtained from guide declarations. moose demography moose within each gmz were assumed to comprise a discrete population (hatter 1999, demography and harvest of low-density moose – hatter alces vol. 58, 2022 104 kuzyk et al. 2018). i used the resident licensed kills/100 hunter days of effort (kpue) as an annual population index for each gmz, and estimated the growth rate (λ) by regressing ln(kpue) against year from 1996–2015 where λ = er and r is the exponential rate of growth. the 95% cis were estimated by bootstrapping with 2000 samples (haddon 2001). i considered the population to be stable if the 95% cis for λ included 1.00. i determined the average winter moose density, calf:cow ratio and bull:cow ratio for each gmz from 1996/97– 2015/16 by weighting each estimate by the size of the survey area. this assumed larger survey areas provided a more accurate estimate of the population parameters at the gmz level, that the population was stable, and that each survey area was randomly placed within the gmz. i used 2000 bootstrap samples to estimate 95% cis on the weighted (by area) estimates of the population ratios and moose density. harvest density was calculated as the bull harvest/1000 km2 of moose range within each gmz and the 95% cis with 2000 bootstrap samples. harvest rates were calculated for each gmz as (bull harvest/1000 km2)/(moose/km2 ×1000) for the population harvest rate and (bull harvest/1000 km2)/(bulls/km2 ×1000) for the bull harvest rate. sustainable harvest rates sustainable harvests were characterized as those where the post-hunt bull:cow ratio remained above a specified adult sex ratio threshold. i considered 3 ratio thresholds: 0.3, 0.4, and 0.5 bull:cow. i used these thresholds as both 0.3 and 0.5 bull:cow ratios are used as management targets in british columbia (bc moe 2010), and 0.4 is used in yukon (czetwertynski 2015). i assessed the licensed harvest only (bulls), and did (could) not consider the indigenous harvest as it was unknown. i estimated harvest sustainability with a demographic model parameterized on the average density and population composition from moose winter surveys of the 4 gmzs; density dependence was not modelled. nutritional density dependence affecting population performance was unlikely due to the low moose density, and inadequate information existed to model density dependence of mortality due to predation. the modelled posthunt population (n) was partitioned into bulls (b), cows (c), and calves (ca) as follows: nt+1 = bt+1 + ct+1 cɑt+1 (1) where: bt+1 = btsmt + 0.5(catsjt) – (mhtnt) (2) or bt+1 = btsmt + 0.5(catsjt) – (mhtbt) (3) ct+1 = ctsft + 0.5(cɑtsjt) (4) cɑt+1 = ct+1 × rpst,t+1 (5) where t was year, sm was the annual bull survival rate, sf was the annual cow survival rate, sj was the annual calf survival rate (i.e., from post-hunt or 0.5 years-ofage to the next post-hunt period or 1.5 years-of-age), rpst was the post-hunt recruitment rate (i.e., the average calf:cow ratio from moose density surveys of the 4 gmzs), and mh was the harvest rate. the post-hunt calf sex ratio was assumed to be 1:1 (ballard et al. 1991, boer 1992). sf was set to 0.89 which was the average survival rate from 5 cow mortality studies within low-density moose populations (sf = 0.91, larsen et al. 1989; sf = 0.91, gasaway et al. 1992; sf = 0.88, stenhouse et al. 1995; sf = 0.88, bertram and vivian 2002; sf = 0.90, joly et al. 2017). previous studies of naturally fluctuating moose populations reported adult sex ratios near parity, alces vol. 58, 2022 demography and harvest of low-density moose – hatter. 105 suggesting that sm and sf are equal or 1:1 (timmerman 1992). conversely, the annual survival rate of bulls was lower than cows in the unhunted moose population on isle royale (peterson 1977), with the average adult sex ratio of 0.8 bull:cow on the island from 1950–1981 (page 1992). i therefore modelled 3 bull survival rates corresponding to unhunted adult sex ratios: 1.0, 0.9, and 0.8 bull:cow. the annual calf survival rate was determined by iteratively adjusting sj until the modelled population achieved a stable age distribution with λ = 1.0. i used this value of sj in the model projections as i was interested in sustainable harvest rates for stable, low-density moose populations. bergerud (1992) considered moose calves at 6–9 months of age to be fully recruited, or that calves and cows have similar winter survival rates. however, several studies found that winter survival rates of calves are lower than cows (ballard et al. 1991, joly et al. 2017, kuzyk et al. 2019), suggesting that recruitment should be measured when calves are 1 year-ofage (hatter 2020). i determined the annual recruitment rate (r) from rpst and the ratio of the winter calf (sjw) and cow (sfw) survival rates where: r = rpst × sjw ∕ sfw (6) (hatter 2020). mortality of cows was assumed negligible from calving to the posthunt period (i.e., sfw = sf ). similarly, mortality from 1 year-of-age to the post-hunt period was assumed negligible (i.e., sjw = sj). rs was equal to r when λ = 1.0. harvest rates were applied to either the total population (bull harvest/post-hunt population, eq. 2) or the bull population (bull harvest/posthunt bulls, eq. 3) because both metrics are commonly used by moose managers in british columbia (kuzyk et al. 2018). i considered harvest rates up to 6% of the total population and up to 25% of the bull population. i made the demographic model stochastic to assess harvest risk (i.e., the probability that the harvest rate was not sustainable) by including the se for each parameter based on estimates of the coefficient of variation (cv). i set cv(sj) = 0.15 (ballard et al. 1991) and cv(sf) = 0.017 from the 5 cow mortality studies. the cv for the winter calf:cow ratio was calculated from survey estimates in the 4 gmzs. the cv for implementation uncertainty (i.e., annual variation in harvest rates) was set to 19% to match annual changes in reported harvests for each of the 4 gmzs during 1996–2015. the cv for survey uncertainty of the population size and bull:cow ratios was set to the survey standard by gasaway et al. (1986) where a cv of 15.2% equals + 25% of the true value 90% of the time. for the survival rates, i portioned the total variance for sj, sf, and sm into seenvironmental and sesampling by assuming 50% of the variance was due to environmental uncertainty and 50% was sampling error. finally, i generated correlated random winter survival rates (burgman et al. 1993) with r = 0.75 for bulls and cows, r = 0.5 for cows and calves, and r = 0.5 for bulls and calves. all random variables were drawn from a normal distribution. i conducted 2000 monte carlo simulations of the model for each bull survival rate and harvest rate during a 50-year period. i chose 50 years because mean annual bull:cow ratios changed very little after that. survey uncertainty was incorporated into the initial simulation year. the harvest rate was considered sustainable if the percentage of the simulations that resulted in final (year 50) bull:cow ratios below the adult sex ratio threshold was ≤ 10%. simulations were performed with the microsoft excel add-in program poptools 3.2 (hood 2011). results moose demography six moose surveys ranging from 1943–7766 km2, 7 surveys ranging from 1533–7319 demography and harvest of low-density moose – hatter alces vol. 58, 2022 106 km2, 3 surveys ranging from 5675–8917 km2, and 4 surveys ranging from 11,022– 34,260 km2 were conducted in gmz 6f, gmz 6e, gmz 7pd, and gmz 7pe, respectively, in 1996/97–2015/16 (fig. 2). moose population growth rate based on kpue was approximately stable in the 4 gmzs over the 20-year period (table 1). although gmz 7pe had a slightly declining growth rate (λ = 0.98), the 95% cis included λ =1.00. further, no decline occurred in the density estimates in gmz 7pe from the 4 moose surveys conducted in 2004–2016, and calf:cow ratios were among the highest in the 4 gmzs (table 1). the average moose density was 0.27/ km2, 0.36/km2, 0.16/ km2, and 0.09/km2 in gmz 6f, gmz 6e, gmz 7pd, and gmz 7pe, respectively. the winter calf:cow ratio ranged from 0.24 (gmz 7pd) to 0.32 (gmz 7pe), averaging 0.28 (table 1); the cv was 0.27 from the 20 moose surveys in the 4 gmzs. the bull:cow ratio ranged from 0.58 (gmz 7pe) to 0.97 (gmz 7pd), averaging 0.75 across the 4 gmzs (table 1). assuming a stable population (λ = 1.0), 0.28 calf:cow ratio in early winter, a density of 0.22 moose/km2, and a cow winter survival rate of 89%, the modelled winter calf survival rate was 79% and stabilizing recruitment (rs) was 0.25 calves/cow. r for each moose population, after adjustment for winter survival rates of cows and calves, was 0.25 (gmz 6f), 0.26 (gmz 6e), 0.21 (gmz 7pd), and 0.28 (gmz 7pe). harvest density based on hunter survey data was lowest in gmz 7pe (3.0/1000 km2) and highest in gmz 6e (8.5/1000 km2) (table 1). population harvest rates from moose and hunter survey data were 2.5% (gmz 6f), 2.4% (gmz 6e), 2.9% (gmz 7pd), fig. 2. survey and harvest parameters for game management zones 6f, 6e, 7pd, and 7pe from 1996/97– 2015/16. n/km2 = moose/km2, b:c = bull:cow ratio, ca:c = calf:cow ratio, and kpue = resident kill per 100 resident hunter days. alces vol. 58, 2022 demography and harvest of low-density moose – hatter. 107 and 3.2% (7pe); bull harvest rates were 7.3% (gmz 6f), 6.1% (gmz 6e), 6.7% (gmz 7pd), and 10.5% (gmz 7pe). the average harvest and moose density were highly correlated among gmzs (r = 1.00, p = 0.001, fig. 3), but harvest rate and bull:cow ratio were not (population harvest rate: r = −0.18, p = 0.82; bull harvest rate: r = −0.75, p = 0.25). sustainable harvest rates sustainable harvest rates were affected by different bull survival rates, with lower survival rates associated with reduced harvests (table 2). population harvest rate up to 2.0% of the modelled post-hunt moose population posed little risk of reducing the bull:cow ratio below threshold values (table 2a). sustainable population harvest rates, averaged across the 3 estimates of sm, were ≤2.4% with the 0.5 bull:cow ratio threshold, ≤3.2% with the 0.40 threshold, and ≤4.1% with the 0.30 threshold. the average population harvest rate was 2.8% based on survey data which was the upper limit of sustainability for the 0.5 bulls/cow threshold. however, the bull:cow ratios from winter moose surveys were ≥0.5, suggesting that the modelling results are conservative. simulated bull harvest rates up to 7% of the modelled post-hunt bull population presented little risk of lowering adult sex ratios below bull:cow thresholds (table 2b). sustainable bull harvest rates averaged ≤8.4% with the 0.5 bull:cow threshold, ≤13.0% with the 0.40 threshold, and ≤20.4% with the 0.30 threshold. the table 1. demographic status of low-density moose populations within 4 game management zones (gmzs) located in northern british columbia, 1996/97–2015/16. moose/km2 and sex/age composition are from moose density surveys during winter. growth rate and harvest are from hunter harvest surveys. numbers in brackets are 95% cis. gmz moose/km2 bulls/cow calves/cow growth rate (λ) harvest/1000 km2 6f 0.27 0.65 0.28 1.01 6.5 (0.20–0.35) (0.57–0.74) (0.24–0.34) (0.99–1.02) (6.1–7.0) 6e 0.36 0.81 0.29 1.00 8.5 (0.29–0.43) (0.69–0.90) (0.24–0.35) (0.98–1.01) (7.9–9.1) 7pd 0.16 0.96 0.24 0.99 4.5 (0.07–0.34) (0.73–1.16) (0.21–0.29) (0.97–1.02) (4.2–4.7) 7pe 0.09 0.58 0.32 0.98 3.0 (0.06–0.12) (0.51–0.69) (0.25–0.36) (0.97–1.00) (2.6–3.4) x̅ 0.22 0.75 0.28 1.00 5.6 sd 0.12 0.17 0.03 0.01 2.4 fig. 3. relationship between average surveyed hunter kill density (harvest/1000 km2) and average surveyed moose density (moose/ km2) in four game management zones in northern british columbia, 1996/97– 2015/16. error bars display minimum and maximum values. demography and harvest of low-density moose – hatter alces vol. 58, 2022 108 average bull harvest rate based on survey data was 7.7% which was within the range of modelled sustainable rates for the 0.5 bull:cow threshold. harvest risk rose sharply as the modelled harvest rates were increased. for example, a population harvest rate of 2.5% had a 0% probability of reducing the bull:cow ratio below 0.5 bulls/cow, whereas a harvest rate of 3.5% had a 95% probability (fig. 4). similarly, a bull harvest rate of 8% had a 0% chance of reducing the adult sex ratio below 0.5 bull:cow, but a 12% harvest had an 81% chance (fig. 5). discussion moose from all 4 gmzs in northern british columbia met the demographic conditions for lde populations. all populations were at low density (≤0.4 moose/km2) and their long-term (20 year) population growth rate was static (λ ~ 1.0). across the gmzs, surveys indicated that the winter calf:cow ratio averaged 0.28, the modelled winter calf table 2. sustainable harvest rates (mh) where bull survival rates (sm) are 0.890 (unhunted sex ratio = 1.0 bull/cow), 0.878 (unhunted sex ratio = 0.9 bulls/cow), and 0.862 (unhunted sex ratio = 0.8 bulls/cow). a. harvest rate applied to total population (mh = bull harvest/post-hunt population) threshold value sm = 0.890 sm = 0.878 sm = 0.862 0.3 bulls/cow mh ≤ 4.4% mh ≤ 4.1% mh ≤ 3.8% 0.4 bulls/cow mh ≤ 3.6% mh ≤ 3.1% mh ≤ 2.8% 0.5 bulls/cow mh ≤ 2.8% mh ≤ 2.4% mh ≤ 2.0% b. harvest rate applied to bull population (mh = bull harvest/post-hunt bulls) threshold value sm = 0.890 sm = 0.878 sm = 0.862 0.3 bulls/cow mh ≤ 21.6% mh ≤ 20.6% mh ≤ 19.1% 0.4 bulls/cow mh ≤ 14.2% mh ≤ 13.2% mh ≤ 11.5% 0.5 bulls/cow mh ≤ 9.6% mh ≤ 8.4% mh ≤ 7.0% fig. 5. relationship between harvest risk and bull population harvest rate for post-hunt adult sex ratio thresholds of 0.30, 0.40, and 0.50 bulls/ cow. the harvest risk is the estimated probability that a specified harvest rate will result in a bull:cow ratio (b/c) below the adult sex ratio threshold. the adult bull survival rate was 0.890. fig. 4. relationship between harvest risk and total population harvest rate for post-hunt adult sex ratio thresholds of 0.30, 0.40, and 0.50 bulls/ cow. the harvest risk is the estimated probability that a specified harvest rate will result in a bull:cow ratio (b/c) below the adult sex ratio threshold. the adult bull survival rate was 0.890. alces vol. 58, 2022 demography and harvest of low-density moose – hatter. 109 survival was 78%, and annual calf recruitment averaged 0.25 calves/cow. bergerud (1992) hypothesized that the mechanism for a lde in a moose population is density-dependent wolf predation on calves, augmented by density-independent bear predation on neonates; food was not considered to be a limiting factor. joly et al. (2017), however, studied a low-density (0.06–0.12 moose/km2) population in north-central alaska and found that productivity was significantly lower in areas with less high-quality habitat, indicating that forage resources/nutrition could be a contributing factor in a lde population. they suggested that the relative roles of predation on young calves, winter weather, and nutritional constraints likely interact to maintain low moose density. much of the moose range in gmz 7pe consists of low to moderate habitat capability (thiessen 2010), suggesting that habitat quality may be a local limiting factor in this population. alternatively, the lower moose density in gmz 7pe may reflect a lack of suitable habitat for moose to spaceout and avoid predation (bergerud 1992). two important considerations in managing low-density moose populations are to maintain an appropriate bull:cow ratio that provides hunting opportunities for large-antlered bulls and to ensure adequate numbers of bulls are available to locate and breed all receptive cows (timmermann 1992). given these considerations, adult sex ratios for low-density populations (≤0.2 moose/km2) in british columbia are managed with an objective to maintain a bull:cow ratio ≥0.5, with higher density populations managed with a threshold bull:cow ratio ≥0.3 (bc moe 2010). in yukon, low-density moose are managed to maintain an adult bull:cow ratio of at least 0.4 (czetwertynski 2015), while in parts of alaska the management objective is a ratio of 0.3 (young and boertje 2008). i used a stochastic model to assess harvest risk, or the probability that a specified harvest rate will reduce the bull:cow ratio below the adult sex ratio threshold. i considered the harvest rate as sustainable if the harvest risk was ≤10%. sustainable harvest rates averaged ≤2.4% of the total population or ≤8.4% of the bull population at the 0.50 bull:cow ratio threshold, ≤3.2% of the population or ≤13.0% of bulls at the 0.40 threshold, and ≤4.1% of the population or ≤20.4% of bulls at the 0.30 threshold. similarly, a harvest rate of 10% of adult bulls or 2.2–3.3% of the total population was sustainable in low-density yukon moose populations with a bull:cow ratio objective (threshold) of 0.4 (czetwertynski 2015). while the surveyed harvest and moose densities were highly correlated, harvest rates and bull:cow ratios were not. this may indicate that the surveyed bull:cow ratios were not representative of the gmz, or some compensatory mortality existed in the licensed harvest. in addition, unmeasured variation in the mortality rates of cows could have affected the bull:cow ratio. i used kpue to ascertain long-term (20 year) stability of the moose population within each of the 4 gmzs. while several studies have found that kpue is correlated with moose abundance (crête et al. 1981, fryxell et al. 1988), there is potential for error or bias in growth rates measured with kpue (fryxell et al. 1988, bowyer et al. 1999, hatter 2001, decesare et al. 2016). i assumed that the survey areas provided a representative sample of the moose density and bull:cow and calf:cow ratios within each gmz. another data concern was that the indigenous harvest was unknown and not an annual harvest metric (kuzyk et al. 2018). i considered 10% as an acceptable, although subjective level of harvest risk. while higher risk levels such as 15% or 20% could have been chosen with greater risk tolerance, sustainable harvest rates would have demography and harvest of low-density moose – hatter alces vol. 58, 2022 110 only increased slightly. for example, selecting a 20% harvest risk versus 10% risk only increased the sustainable population harvest rate from 2.8% to 2.9% for the 0.5 bull:cow ratio threshold with sm = 0.870. computation of stabilizing recruitment required an estimate of the winter cow survival rate. i assumed all cow mortality occurred during winter such that the winter survival rate was equal to the annual survival rate. however, studies of low-density moose populations have documented that some cow mortality occurs during summer (larsen et al. 1989). calculation of rs would have been improved with a more reliable estimate of the winter cow survival rate. i did not include density-dependent nutritional effects in the demographic model as moose density in the 4 gmzs ranged from 0.1–0.4 km2, and thus were well below the habitat carrying capacity of 1.5–2.0 moose/ km2 for north american populations (bergerud 1992, messier 1994). despite this, nutritional influences may have been present within gmz 7pe where habitat quality was lower. inclusion of density dependence in the model would have increased sustainable harvest rates (caughley 1977). the stochastic modelling indicated that minor changes in harvest rates could greatly affect the probability of bull:cow ratios dropping below adult sex ratio thresholds, revealing that small changes in harvest management could affect the resulting bull:cow ratios. sustainable harvest rates were also affected by different bull survival rates with lower survival rates sustaining reduced harvests. i did not model non-stationarity in environmental variation although global warming is expected to be highly influential to ungulate population dynamics in northern latitudes where variation in demographic rates, and thus sustainable harvest rates, are closely aligned to annual cycles in climate and primary productivity (brown 2011). for these reasons, wildlife managers should be cautious about applying estimates of sustainable harvest rates from this study to other low-density moose populations. despite these limitations, i was able to offer support for the lde hypothesis for all 4 gmzs in northern british columbia and provide additional insight on sustainable harvest rates for moose within these northern ecosystems. more study is recommended to identify the relative roles of hunting, predators (wolves and bears), habitat quality, and winter weather in maintaining low densities of moose across northern british columbia. further understanding of sustainable harvest rates would benefit from additional moose surveys, indigenous harvest surveys, and studies of moose survival rates. acknowledgements m. bridger, d. heard, and h. schindler kindly reviewed an early draft of the manuscript and made numerous helpful and insightful suggestions. the manuscript was also greatly improved by constructive reviews from n. decesare and 2 reviewers. references ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114: 3–49. _____, and v. van ballenberghe. 1997. predator/prey relationships. pages 247– 273 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north america moose. wildlife management institute, washington, dc, usa. bergerud, a. t. 1992. rareness as an antipredator strategy to reduce predation risk for moose and caribou. pages 1008–1021 in d. r. mccullough and r. h. barrett, editors. wildlife 2001: populations. elsevier applied science publishers, london, united kingdom. alces vol. 58, 2022 demography and harvest of low-density moose – hatter. 111 _____, and j. p. elliott. 1998. wolf predation in a multiple-ungulate system in northern british columbia. canadian journal of zoology 76: 1551–1569. doi: 10.1139/z98-083 bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior alaska. journal of wildlife management 66: 747–756. doi: 10.2307/3803140 boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1 (1992): 1–10. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–89. british columbia ministry of environment (bc moe). 2010. moose harvest management procedure. fish and wildlife branch, victoria, british columbia, canada. british columbia ministry of forests, lands and natural resource operations (bc flnro). 2014. management plan for the grey wolf (canis lupus) in british columbia. fish and wildlife branch, victoria, british columbia, canada. _____. 2020. british columbia grizzly bear population estimate for 2018. fish and wildlife branch, victoria, british columbia, canada. brown, g. s. 2011. patterns and causes of demographic variation in a harvested moose population: evidence for the effects of climate and density-dependent drivers. journal of animal ecology 80: 1288–1298. doi: 10.1111/j.1365-2656.2011.01875.x burgman, m. a., s. ferson, and r. akçakaya. 1993. risk assessment in conservation biology. chapman and hall, new york, new york, usa. caughley, g. 1977. analysis of vertebrate populations. john wiley and sons, new york, new york, usa. crête, m., r. j. taylor, and p. a. jordon. 1981. optimization of moose harvest in southwestern quebec. journal of wildlife management 45: 598–611. doi: 10.2307/3808693 czetwertynski, s. 2015. sustainable harvest rates for moose in yukon, draft report. fish and wildlife branch, department of environment, government of yukon, whitehorse, yukon, canada. decesare, n. j., j. r. newby, v. j. boccadori, t. chilton-radandt, t. thier, d. waltee, k. podruzny, and j. a. gude. 2016. calibrating minimum counts and catch-per-unit effort as indices of moose population trend. wildlife society bulletin 40: 537–547. doi: 10.1002/wsb.678 eastman, d. s., and r. ritcey. 1987. moose habitat relationships and management in british columbia. swedish wildlife research supplement 1: 101–117. fryxell, j. m., w. e. mercer, and r. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52: 14–21. doi: 10.2307/ 3801050 gasaway, w. c., r. d. boertje, d. grangaard, d. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120: 3–59. _____, s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. institute of arctic biology, university of alaska, fairbanks, alaska, usa. haddon, m. 2001. modelling and quantitative methods in fisheries. chapman and hall/crc, london, united kingdom. hatter, i. w. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. _____. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces 37: 71–77. demography and harvest of low-density moose – hatter alces vol. 58, 2022 112 _____. 2020. revisiting the recruitment-mortality equation to assess moose growth rates. alces 56: 39–47. hood, g. m. 2011. poptools version 3.2.h. (accessed september 2022). joly, k., t. craig, m. d. cameron, a. e. gall, and m. s. sorum. 2017. lying in wait: limiting factors on a low-density ungulate population and the latent traits that can facilitate escape from them. acta oecologica 85: 174–183. doi: 10.1016/j.actao.2017.11.004 kuzyk, g., i. hatter, s. marshall, c. procter, b. cadsand, d. lirette, h. schindler, m. bridger, p. stent, a. walker, and m. klaczek. 2018. moose population dynamics during 20 years of declining harvest in british columbia. alces 54: 101–119. _____, c. procter, s. marshall, h. schindler, h. schwantje, m. scheideman, and d. hodder. 2019. determining factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife working report. no. wr-127. progress report: february 2012–april 2019. british columbia ministry of forests, lands and natural research operations and rural development, victoria, british columbia, canada. lake, b. c., m. r. bertram, n. guldager, j. r. caikoski, and r. o. stephenson. 2013. wolf kill rates across winter in a low-density moose system in alaska. journal of wildlife management 77: 1512–1522. doi: 10.1002/jwmg.603 larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rate of moose mortality in southwest yukon. journal of wildlife management 53: 548–557. doi: 10.2307/3809175 messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478–488. doi: 10.2307/1939551 page, r. e. 1992. can we expect to achieve natural moose sex ratios? pages 26–34 in d. f. hatler, editor. proceedings of the 1991 moose harvest management workshop. wildlife branch, british columbia ministry of environment, victoria, british columbia, canada. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. united states national park service scientific monograph no. 11. superintendent of documents, washington, dc, usa. serrouya, r., h. van oort, and c. demars. 2015. wolf census in three boreal caribou ranges in british columbia; results from 2015. report prepared by the alberta biodiversity monitoring institute, edmonton, alberta, canada. stenhouse, g. b., p. b. latour, l. kutny, n. maclean, and g. glover. 1995. productivity, survival, and movements of female moose in a low-density population, northwest territories, canada. arctic 48: 57–62. thiessen, c. 2010. horn river basin moose inventory, january/february 2010. peace region technical report. british columbia ministry of environment, fort st. john, british columbia, canada. timmermann, h. r. 1992. moose sociobiology and implications for harvest. alces 28: 59–77. young, d. d., and r. d. boertje. 2008. recovery of low bull:cow ratios of moose in interior alaska. alces 44: 65–71. van ballenberghe, v., and w. b. ballard. 1994. limitation and regulation of moose populations: the role of predation. canadian journal of zoology 72: 2071–2077. doi: 10.1139/z94-277 _____, and _____. 1997. population dynamics. pages 223–245 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, dc, usa. https://poptools.software.informer.com/3.2/ https://poptools.software.informer.com/3.2/ alces vol. 44, 2008 ritchie challenges of pine beetle infestation 127 management and challenges of the mountain pine beetle infestation in british columbia chris ritchie environmental stewardship, british columbia ministry of environment, 4051 18th avenue, prince george, british columbia, canada v2n 1b3 abstract: central british columbia is currently subject to the largest outbreak of mountain pine beetle (dendroctonus ponderosa) ever recorded in british columbia. the massive expansion of this natural disturbance agent is a result of both natural and human-associated influences including milder winter weather and fire suppression policy. resource managers are grappling with a response to the infestation that considers economic, social, and ecological factors. in british columbia the response has moved from a control or sanitation phase, to an economic recovery or salvage phase. the condition of the landscape resulting from the insect and the management associated with each phase will impact wildlife populations. distribution and abundance of certain species will either increase or decline in response to changes in the forest vegetation and hydrologic regime. caribou (rangifer tarandus caribou), fisher (martes pennanti), marten (martes americana), woodpeckers, and pygmy nuthatches (sitta pygmaea) are considered species with high sensitivity to mortality of pine trees that will adversely affect their forage, cover, and nesting/denning habitat. moose (alces alces) will probably benefit in the short-term from increased forage resources, but may decline long-term from intensive forest management to recover mature forest stands. the impact of larger and more dispersed moose and wolf (canis lupus) populations could harm the recovery and stability of threatened caribou populations in british columbia. alces vol. 44: 127-135 (2008) key words: dendroctonus ponderosa, epidemic, forest management, habitat, lodgepole pine, moose, pinus contorta. central british columbia is currently subject to the largest recorded outbreak of mountain pine beetle (mpb; dendroctonus ponderosa). the massive expansion of this natural disturbance agent is a result of a combination of natural and human influences. land managers are grappling with a balanced response that considers economic, social, and ecological factors. the initial response in british columbia was to control the infestation. however, most of the province is now under salvage logging to recover economic timber value. the condition and changes in the landscape resulting from the insect and associated forest management has and will continue to influence wildlife populations, including moose (alces alces). distribution and abundance of certain wildlife species will increase in response to change in the forest cover and hydrologic regime, while others will decline. effective and responsive forest and wildlife management to the mpb infestation will depend on understanding and predicting such changes. life history of the mountain pine beetle pine forests in interior british columbia are currently suffering the largest mpb epidemic in recorded history (british columbia ministry of forests 2007). the mpb is a bark beetle the size of a rice-grain and native to pine forests of north america (safranyik and carroll 2006). its primary host in british columbia is mature (>60 years old) lodgepole pine (pinus contorta) trees. in the 1-year challenges of pine beetle infestation ritchie alces vol. 44, 2008 128 cycle typical of the beetle, adults leave trees in summer and fly to adjacent pine hosts. adults bore through the outer bark and create vertical galleries in the inner bark (phloem), where they lay their eggs. pheromones produced during gallery boring attract other beetles to the tree. the eggs hatch and the larvae feed on the phloem, excavating lateral tunnels through the inner bark that girdles the tree. when mpb populations are at epidemic levels, pheromone-mediated mass attacks can result in sufficient larval tunnelling to kill the tree by disrupting the flow of water and nutrients. the beetle benefits from a symbiotic blue stain fungus (ophiostoma spp.) (carroll and safranyik 2004, rice et al. 2007) that it introduces to the tree, which further disrupts sap movement and compromises the tree’s ability to defend itself against or “pitch-out” beetles. successful attacks by mpb can be identified by the numerous pitch pockets on the stem of a tree or by the sawdust (frass) from gallery excavation at the base of a tree. the infestation stage when larvae feed is often called “green attack” because the foliage has not lost its green color. over the fall and winter, the foliage starts to fade to a pale green or yellow, reaching a brilliant red color, or the “red attack” phase, by the time of flight the next summer. in the subsequent 2-3 years, needles fall from the tree resulting in the “gray-attack” phase. at endemic levels the beetle attacks a few stressed trees in a stand producing an irregular “salt and pepper” appearance to the stand. if appropriate conditions exist, population levels may grow rapidly to epidemic proportions as occurs currently in british columbia. the mass attack of trees produces a more contiguous carpet of red forest cover. these mass attacks affect nearly all pine trees in the watershed, not just older or stressed trees (shore et al. 2006). mountain pine beetle epidemic in british columbia the mpb is a natural element of british columbia pine forests and epidemics have occurred numerous times, notably in 1976-81 in the flathead in southeastern british columbia (young 1988) and 1986-88 in the chilcotin in central british columbia (alfaro et al. 2004). the current epidemic, however, is projected to be the largest in recorded history. while no single epicentre has been identified, conditions on the nechako plateau in north central british columbia were ideal for mpb in the late 1990s. in 1999, less than 10 million m3 of new “red attack” pine forest was recorded throughout british columbia (walton et al. 2007). this outbreak was followed by a near exponential increase that peaked at 140 million m3 of new “red attack” in 2004 (fig. 1). computer models project that 78% (>1 billion m3) of the mature pine in british columbia will be killed by 2018 (walton et al. 2007). two factors have historically limited mpb population growth in british columbia. sudden cold snaps of -30º to -40º c in early winter cause high larval mortality (carroll and safranyik 2004). however, such cold conditions must occur before the beetle produces its natural antifreeze, or before deep snow insulates the base of pine trees. these conditions have not occurred in central british columbia for 30 years (carroll et al. 2004). mild winters in recent years have been coupled with warm, dry summers that produce more stress on trees, reducing their ability to repel mpb. the frequent fire interval (i.e., 80-125 yr) typical of lodgepole pine forests in british columbia limits the amount and contiguity of the beetle’s primary host (i.e., >60 year old pine trees). aggressive fire fighting in the last 50 years has reduced the number and size of fires that otherwise would have naturally reduced and spatially disrupted the host supply (mackillop and holt 2004). extremely suitable conditions favouring population growth of mpb have resulted in the unprecedented epidemic currently sweeping through interior pine forests. the rocky mountains have historically alces vol. 44, 2008 ritchie challenges of pine beetle infestation 129 been viewed as a possible barrier to the eastward movement of mpb. however, since 2005 the mpb epidemic has become firmly established east of the rocky mountains in the peace portion of british columbia and adjacent portions of alberta. the distribution of the mpb’s primary host ends in western alberta where the predominant lodgepole pine forest changes to a lodgepole-jack pine (pinus banksiana) mix, with transition to pure jack pine in eastern alberta and eastward across canada (ono 2004). jack pine was believed to be a natural barrier to eastern movement of mpb. however, laboratory trials have demonstrated that mpb can successfully reproduce in jack pine, and its symbiotic blue stain fungus is equally as virulent in jack pine as in lodgepole pine (rice et al. 2007). it is unclear whether winter conditions in the prairie are severe enough to curtail the mpb epidemic. mountain pine beetle impacts to wildlife the major change arising from mpb infestation is death of all mature pine in a stand, and thus the loss of the dominant tree canopy cover. because mpb is a natural element of lodgepole pine forests, wildlife in those forfig. 1. infestation area (dark areas) of the mountain pine beetle in western canada, 2006 (natural resources canada 2008). the infected area in british columbia increased >10 fold from 1999-2004 (walton et al. 2007). challenges of pine beetle infestation ritchie alces vol. 44, 2008 130 ests have adapted to periodic outbreaks and epidemics. outbreaks have severe impacts on the mature pine trees in a stand, but do not usually kill all the pine. usually pines <20 cm diameter breast height (dbh) and non-pine vegetation are unaffected in a stand, and trees <60 years old were thought immune to mpb attack (safranyik 2004). however, the current epidemic is attacking pine trees as young as 15 years and only 7 cm dbh (robert hodgkinson, british columbia ministry of forest and range, pers. comm.) the reallocation of resources (e.g., water, nutrients, sunlight) that results from the death of pine trees promotes growth of other vegetation in the stand (williston et al. 2006). the resultant stand is typical of an early stage of forest succession over an area proportional to the size of the epidemic. shrubs and forbs in the understory may flourish providing a benefit to resident wildlife that use such vegetation. an abundance of standing dead trees and snags may benefit cavity nesters and species that forage on insects. substantial negative impact should be limited to those species heavily dependent on pine trees or stands of monoculture mature pine. examples of such species in british columbia include woodland caribou (rangifer tarandus caribou) that feed on terrestrial lichen, furbearers dependent on old or mature forest, and birds that depend on pine seeds. the northern ecotype of woodland caribou forages on terrestrial lichens found in stands of low elevation pine during winter. lichen dominates the forest understory in these forest stands with nutritionally poor soils (williston et al. 2006). loss of the predominant pine canopy will change growing conditions in the understory to the detriment of lichen as more sunlight and soil moisture is available to shrubs (williston et al. 2006). also, a barrier to caribou movement and reduced access to lichen may occur if the dead pine canopy blows down or falls over in extensive portions of their winter range. although these forest types are maintained by periodic fire, an abnormally high fuel load due to high blow down could create conditions for severe fires that could further reduce the low soil productivity. although caribou are adapted to mpb epidemics, the magnitude of the current epidemic and associated habitat changes in the surrounding managed landscape could have an extremely negative impact on caribou (cichowski 2007). fisher (martes pennanti) and marten (martes americana) exhibit a strong dependence on mature or old forest habitat. forests with large old trees provide security cover, abundant small mammals as prey, subnivian access, and denning and resting sites (ruggiero et al. 1994). in the short term between “red attack” and recovery of understory vegetation (i.e., 1-5 years), furbearers will likely experience reduced security cover from avian predators, and a change of prey type and abundance (weir 2003). in the medium term (i.e., 20-50 years), the abundance of snags should decline and convert to horizontal coarse woody debris (cwd). this transition will reduce the number of elevated cavities but increase the number of ground dens. in the longer term (i.e., 70100 years), cwd will eventually decay and disappear from the stand, thus reducing den sites and access to prey below snow cover. because the mpb attacks all pine species in british columbia, even bird species with high dependence on ponderosa pine (pinus ponderosa) for food or cover will be impacted. for example, pygmy nuthatches (sitta pygmaea) depend on the large seeds and nesting cavities in these trees, as well as the insects that inhabit them (kingery and ghalambor 2001). loss of all large ponderosa pine to mpb will drastically reduce the seed supply for many years, important winter food for nuthatches. nuthatch numbers may be reduced until the younger pine that survive the infestation can produce abundant seed and provide nesting cavities. williamson’s sapsucker (sphyrapicus thyroideus) and lewis’ alces vol. 44, 2008 ritchie challenges of pine beetle infestation 131 woodpecker (melanerpes lewis) rely on large ponderosa pine trees for nesting sites (cooper 1995, cooper et al. 1998). the largest trees in a stand are used repeatedly by these species over the period they occupy the site. the number of suitable nest trees may increase in the short term, however, there may be a long period of poor nesting habitat after large trees drop from the canopy. moose are generally anticipated to be “winners” as a result of the mpb infestation (janz 2006), but could suffer some consequences. removal of the pine canopy will increase forage for moose in the short-term (williston et al. 2006). however, as affected stands recover their overstory, the shrub layer could be less abundant than at pre-infestation. the extensive and sometimes uniform nature of this pattern of succession may reduce habitat heterogeneity that benefits moose (peek 1998). further, canopy loss will affect thermal conditions in a stand by increasing sunlight on the forest floor and within stand temperatures. moose are sensitive to heat stress (schwartz and renecker 1998) and dead standing pine in “gray attack” stands will provide reduced shade value. although thermal conditions in winter may also worsen from less cover, moose are better adapted for extremes in cold and snow than heat. forest management to address mpb epidemic while mpb epidemics ultimately stop due to natural factors, a variety of measures have been undertaken in british columbia to control the spread of the infestation. aerial surveys are conducted during fall to detect the rate of expansion when "red attack" or red trees (previous year’s attack) become visible. stands of red trees are used to locate suitable areas for ground surveys to identify “green attack” trees. these are felled and burned at the stump to destroy insects in the tree prior to flight. during the early phase of the infestation, and at the leading or expanding edge of the epidemic, “fall and burn” programs have been employed in an attempt to control the spread of the infestation. helicopters are sometimes used to haul trees to a central burning location. where the expanding edge is close to a road network, small scale “snip and skid” or small patch logging is conducted to remove infested trees and recover some economic value. however, control of mpb is no longer feasible in most of british columbia. dead pine trees can be salvage-logged using conventional harvest methods to recover some economic value. however, there is some urgency to recover such trees before they dry and crack, and lose most of their economic value as sawlogs. this period known as “shelf life” varies depending on environmental and site conditions. significant loss in economic value of sawlogs is forecast within 1-3 years after death (byrne et al. 2006). as a result, the british columbia ministry of forests and range has increased the allowable annual harvest of dead pine (mof 2007). in addition, the government and forest industry are actively exploring other uses and products of dead pine. these products include oriented strand board, wood stove pellets, and bio-fuel. shelf life for such products is believed to be considerably longer than for sawlogs. impacts of mpb management on wildlife the most severe impacts associated with mpb management result from salvage harvest, roads, and post-harvest site treatment. many of these impacts are similar to those from the epidemic alone, but are often more pronounced. for example, the forest canopy is removed by logging, but so are most standing dead trees. logging also damages understory vegetation, including advanced tree regeneration and cwd. in an effort to address shelf life and find economic efficiencies, salvage logging of affected pine stands tends to be more intensive and extensive resulting in large cutblock openings. these openings and the challenges of pine beetle infestation ritchie alces vol. 44, 2008 132 associated road networks may fragment habitat of certain wildlife species. road networks, either re-activated or new roads that access the massive landscapes of dead pine may be open longer in order to access fiber for secondary (non-sawlog) industries. this increased human presence in the forested landscape may result in the displacement of species sensitive to human activity (e.g., wolverine (gulo gulo) and grizzly bear (ursus arctos); ruggiero et al. 1994, ciarniello et al. 2007). finally, preparing logged sites for reforestation and tending the next tree crop may damage some important habitat elements. cwd may be trampled or piled and burned, and planting of commercial species can reduce the diversity of vegetation in a stand. use of herbicides and mechanical thinning can reduce the period of herb and shrub-dominated early succession, and potentially reduce the deciduous component in the stand. both activities may reduce the amount or duration of forage available to wildlife in a stand. these impacts from site preparation are common to most clear-cut logging operations, but salvage programs are anticipated to be at a much larger scale than typical commercial logging in the region. moose management moose are predicted to receive a net benefit as a result of the mpb infestation and associated forest management (janz 2006), but both negative and positive impacts will occur. removal of the pine canopy by mpb or logging will certainly increase forage resources for moose (williston et al. 2006). however, salvage logging will increase the rate of canopy loss and amplify the effect of higher thermal conditions in a stand. the massive scale of salvage harvests could augment heat stress of moose over very large areas. the stand tending activities on these large salvage openings will truncate the period of early seral shrub growth, and the forage benefit from removal of the pine canopy could be negated by subsequent stand management. finally, extensive road networks for salvage logging may indirectly reduce moose numbers through over-harvest, and increased disturbance, displacement, vehicle collisions, and predator mobility (stotyn et al. 2008). of concern is the possibility that an increased moose population may have negative implications for woodland caribou in the region. both ecotypes of caribou are listed as threatened under the canadian species at risk act and are in recovery planning. clearly, northern caribou may be adversely affected by reduction of lichen, however, both northern and mountain (arboreal lichen feeders) caribou may be adversely affected by changes in predator-prey relationships and dynamics (wittmer 2004). the gray wolf (canis lupus) is the principal predator of moose in central british columbia, and a higher moose population as a result of improved forage conditions will presumably allow a higher wolf population. wolves also prey upon caribou, but generally at a lower rate because of their spatial separation (i.e., caribou frequent elevations above valley bottom where wolves are active) and low density (seip and cichowski 1996). caribou habitat is not typically used by moose because of relatively poor forage conditions in nutrientpoor, pine-lichen stands in northern caribou habitat and deep snow in mountain caribou habitat. canopy reduction in pine stands may release shrub growth (williston et al. 2006) attracting more moose and wolves, effectively reducing the spatial separation between wolves and caribou and exposing caribou to increased predation risk. population control of moose has been identified as one possible measure to promote caribou recovery (mountain caribou technical advisory committee 2002). the mpb infestation has peaked when measured by the rate of annual expansion (walton et al. 2007), yet its impact will affect the ecosystem and forest management for decades. moose will initially benefit alces vol. 44, 2008 ritchie challenges of pine beetle infestation 133 from increased forage in salvage cut-blocks and provide increased benefit to hunters and the non-hunting public, but negative impacts on caribou populations are likely. however, salvage cutblocks will be intensively managed in an effort to reduce the shortfall in sawlogs as a result of the mpb infestation. stand tending and eventual canopy closure in the large plantations will gradually reduce moose forage and moose populations should reflect this reduction. a critical challenge will be the development of wildlife management objectives that address habitat response to the mpb infestation and related forest management. flexible management strategies will be necessary to maintain stable moose and wolf populations while promoting threatened caribou populations. acknowledgements this paper benefited from helpful comments of doug heard and doug wilson. references alfaro, r. i., r. campbell, p. vera, b. hawkes, and t. l. shore. 2004. dendroecological reconstruction of mountain pine beetle outbreaks in the chilcotin plateau of british columbia. pages 245256 in t. l. shore, j. e. brooks, and j. e. stone, editors. mountain pine beetle symposium: challenges and solutions. kelowna, british columbia. october 30-31, 2003. british columbia ministry of forests. 2007. mountain pine beetle. (accessed may 2007). byrne, t., c. stonestreet, and b. peter. 2006. characteristics and utilization of postmountain pine beetle wood in solid wood products. pages 236-254 in l. safranyik and b. wilson, editors. the mountain pine beetle: a synthesis of biology, management and impacts on lodgepole pine. natural resources canada, canadian forest service, pacific forestry centre, victoria, british columbia. carrol, a. l., s. w. taylor, j. regniere, and l. safranyik. 2004. effects of climate change on range expansion by the mountain pine beetle in british columbia. pages 233-232 in t. l. shore, j. e. brooks, and j. e. stone, editors. mountain pine beetle symposium: challenges and solutions. kelowna, british columbia. october 30-31, 2003. _____, and l. safranyik. 2004. the bionomics of the mountain pine beetle in lodgepole pine forests: establishing a context. pages 21-32 in t. l. shore, j. e. brooks, and j. e. stone, editors. mountain pine beetle symposium: challenges and solutions. kelowna, british columbia. october 30-31, 2003. ciarniello, l., m. boyce, d. heard, and d. seip. 2007. components of grizzly bear habitat selection: density, habitats, roads, and mortality risk. journal of wildlife management 71:1446-1457. cichowski, d. 2007. literature review effects of mountain pine beetles on caribou prepared for: sunil ranasinghe, alberta sustainable resource development public lands and forests division, forest health section. edmonton, alberta. cooper, j. m. 1995. status of the williamson’s sapsucker in british columbia. british columbia. ministry of environment, lands, and parks, wildlife branch. working report wr-69. _____, c. siddle, and g. davidson. 1998. status of the lewis’s woodpecker (melanerpes lewis) in british columbia. british columbia ministry of environment, lands, and parks, wildlife branch. working report wr-91. janz, d. w. 2006. mountain pine beetle epidemic – hunted and trapped species sensitivity analysis. british columbia ministry of environment, environmenchallenges of pine beetle infestation ritchie alces vol. 44, 2008 134 tal stewardship, prince george, british columbia. kingery, h., and d. c. ghalambor. 2001. pygmy nuthatch (sitta pygmaea). in a. poole and f. gill (editors.). the birds of north america, no. 567. the birds of north america, inc., philadelphia, pennsylvania. mackillop, d., and d. r. f. holt. 2004. mountain pine beetles (dendroctonus ponderosae) and old growth forest characteristics in the moist interior plateau, vanderhoof forest district. prepared for: west fraser sawmills, fraser lake, british columbia. (mof) british columbia ministry of forest. 2007. timber supply and the mountain pine beetle infestation in british columbia 2007 update. forest analysis and inventory branch, victoria. mountain caribou technical advisory committee. 2002. a strategy for the recovery of mountain caribou in british columbia. british columbia ministry of water, land, and parks. victoria. natural resources canada. 2008. the mountain pine beetle in british columbia. < h t t p : / / m p b . c f s . n r c a n . g c . c a / b i o l o g y / introduction_e.html> (accessed october 2008). ono, h. 2004. the mountain pine beetle: scope of the problem and key issues in alberta. pages 62-66 in t.l. shore, j.e. brooks, and j.e. stone, editors. mountain pine beetle symposium: challenges and solutions. kelowna, british columbia. october 30-31, 2003. peek, j. m. 1998. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d. c. rice, a. v., m. n. thormann, and d. w. langor. 2007. virulence of, and interactions among, mountain pine beetle associated blue-stain fungi on two pine species and their hybrids in alberta. canadian journal of botany 85: 3160-323. ruggiero, l. f., k. aubry, s. buskirk, l. j. lyon, and w. zielinski. 1994. the scientific basis for conserving forest carnivores: american marten, fisher, lynx and wolverine. u.s. department of agriculture, forest service general technical report rm-254. safranyik, a., and a. l. carroll. 2006. the biology and epidemiology of the mountain pine beetle in lodgepole pine forests. pages 3-66 in l. safranyik and b. wilson, editors. the mountain pine beetle: a synthesis of biology, management and impacts on lodgepole pine. natural resources canada, canadian forest service, pacific forestry centre, victoria, british columbia. safranyik, l. 2004. mountain pine beetle epidemiology in lodgepole pine. pages 33-40 in t. l. shore, j. e. brooks, and j. e. stone, editors. mountain pine beetle symposium: challenges and solutions. kelowna, british columbia. october 30-31, 2003. schwartz, c., and l. renecker. 1998. food habits and feeding behaviour. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d. c. seip, d. r., and d. b. cichowski. 1996. population ecology of caribou in british columbia. rangifer, special issue number 9: 73-80. shore, t., l. safranyik, and r. whitehead. 2006. principles and concepts of management. pages 117-122 in l. safranyik and b. wilson, editors. the mountain pine beetle: a synthesis of biology, management and impacts on lodgepole pine. natural resources canada, canadian alces vol. 44, 2008 ritchie challenges of pine beetle infestation 135 forest service, pacific forestry centre, victoria, british columbia. stotyn, s., m. setterington, and d. p. tobler. 2008. the impacts of roads on wildlife and metrics for assessment. british columbia ministry of environment, environmental stewardship. williams lake, british columbia. walton, a., j. hughes, m. eng, a. fall, t. shore, b. riel, and p. hall. 2007. provincial-level projection of the current mountain pine beetle outbreak: update of the infestation based on the 2006 provincial aerial overview of forest health and revisions to the model (bcmpb. v4). british columbia ministry of forest, victoria. weir, r. d. 2003. status of the fisher in british columbia. wildlife bulletin number b-105. british columbia ministry of sustainable resource management, conservation data centre, and british columbia ministry of water, land, and air protection, biodiversity branch, victoria, british columbia. williston, p., d. cichowski, and s. haeussler. 2006. the response of caribou terrestrial forage lichens to mountain pine beetles and forest harvesting in the east ootsa and entiako areas: final report – 2005 – years 1 to 5. a report to morice-lakes innovative forest practices agreement, prince george, british columbia, the bulkley valley centre for natural resources research and management, smithers, british columbia, and british columbia parks, smithers, british columbia. wittmer, h. u. 2004. mechanisms underlying the decline of mountain caribou (rangifer tarandus caribou) in british columbia. ph. d. thesis. university of british columbia. vancouver, canada. young, c. 1988. coming of age in the flathead. how the british columbia forest service contended with the mountain pine beetle infestation in southeastern british columbia. pest management report number 10. british columbia ministry of forests, victoria. p147-160_4111.pdf alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 147 proposal to combine cree and scientific knowledge for improved moose habitat management on waswanipi eeyou astchee, northern québec hugo jacqmain1, louis bélanger1,2, réhaume courtois2, thomas beckley3, solange nadeau4, christian dussault2, and luc bouthillier1 1faculté de foresterie et de géomatique, pavillon abitibi-price, université laval, québec, pq, canada g1k 7p4; 2ministère des ressources naturelles et de la faune du québec, direction de la recherche sur la faune, 930 chemin sainte-foy, 4e étage, québec, pq, canada g1s 2l4; 3faculty of forestry and environmental management, university of new brunswick, fredericton, nb, canada e3b 5a3; 4natural resources canada, canadian forest service, atlantic forestry centre, p.o. box 4000, fredericton, nb, canada e3b 5p7 abstract: first nations involvement in forest management is necessary to achieve sustainability, even more in northern québec where the cree have constitutional rights on the land. an innovative research approach has been undertaken to improve forest management on eeyou astchee, the cree territory. this project targets moose (alces alces) because of its importance to the cree people and because it is a representative species of the northern black spruce ecosystem. the research aims at combining common vision of moose habitat needs in this northern area. in this poorly known ecosystem, combinforest management. based on new knowledge of this common vision, socioecologicaly-adapted habitat management strategies will be proposed for the study area. the involvement of key stakeholders, and recognition of their knowledge, should promote better support for the research project and better social acceptability of the proposed management recommendations. alces vol. 41: 147-160 (2005) key words: alces alces, cree, forest management, local ecological knowledge, moose, québec, waswanipi résumé : l’implication des peuples autochtones dans la gestion forestière est une prémisse au développement durable, tout particulièrement dans le nord du québec, où les cris détiennent des droits constitutionnels qui protègent leur utilisation distincte du territoire. dans la présente recherche, nous proposons une approche novatrice permettant d’améliorer l’aménagement forestier sur l’eeyou astchee, le territoire des cris. le projet cible l’orignal (alces alces) par l’importance qu’il revêt pour les cris et par le fait qu’il est considéré comme une espèce représentative de la pessière noire nordique. la en habitat de l’orignal. dans cet écosystème encore peu étudié, la combinaison des connaissances cries la base de cette vision commune, des stratégies d’aménagement adaptées au contexte socio-écologique seront proposées pour le territoire à l’étude. l’implication adéquate des parties prenantes, et la prise en considération de leurs connaissances respectives devraient promouvoir un support accru au projet de recherche en cours et une meilleure acceptabilité sociale des stratégies d’aménagement proposées. alces vol. 41: 147-160 (2005) mots clés: aménagement forestier, alces alces, connaissance écologique locale, cri, orignal, québec, waswanipi knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 148 large-scale forest harvesting, as done in northern québec, is viewed by some as a conflict with the fundamental needs of aboriginal peoples and has sparked significant conflict among native people, government authorities and the forest industry (tobias 1991, beaulieu 2000). one of the primary causes of these conflicts is the negative impact that forest harvesting has on certain wildlife habitats and populations that are important for the cree (mce 1998). under certain circumstances, large-scale forest harvesting induces immediate and substantial loss of productive wildlife hunting territory (johnston and elliot 1996, morel and bélanger 1998), and decreases local wildlife populations (messier 1993, potvin et al. 1999) for cree hunters and trappers. this is particularly true for moose (alces alces), which seem to avoid recent clearcut areas (thompson and vukelich 1981, courtois 2002). the waswanipi cree first nation, in this conflicting context, created a model forest in order to understand and improve the coexistence of their intensive use of the land with large-scale forest exploitation. this model forest is a partnership led by the cree, with forest industries, provincial and federal governments, and universities, contributing to sustainable forest management. management of moose habitat on eeyou astchee (cree territory, fig. 1) has always been one of the major subjects of disagreement between cree and government representatives (mce 1998) and thus turned out to be a research priority for the model forest. despite their constitutional rights on the land (gagnon 1973), the cree have had little influence on management policies for moose habitat even though they have requested input. undertaken in 2003 with the support of the waswanipi cree model forest, this project aims at developing a framework to include cree concerns and knowledge in the science-based management structure. to do so, we are trying to link native and non-native natural resources management systems by combining cree and scientific knowledge about moose. moose are certainly one of the most interesting wildlife species on which to focus collaboratively since it is one of the most important in the cree culture (feit 1999, jacqmain and bélanger 2002) and represents the main ‘’bush food’’ intake for cree hunters (gagnon 1973, feit 1999). moose can also be considered a representative species of the boreal forest (courtois et al. 1998) and can be used to quantify and qualify this ecosystem (jackson et al. 1991). specific seasonal needs of moose, in terms of habitat composition and structure, make this species interesting for tallymen (official cree land managers of family hunting grounds) and forest managers as an important component in the implementation of integrated resource management on a family hunting ground (trapline), at both the stand and the landscape level (hénault et al. 1999, potvin et al. 1999, jacqmain and bélanger 2002). furthermore, habitat at the northern limit of the species’ range, where moose populations have been at low densities since the 1980s (messier 1993), could be improved or at least protected to reduce moose vulnerability to predation (joyal 1987). this paper presents the existing theory on aboriginal participation in forestry and suggests a mechanism to implement it on waswanipi eeyou astchee. a literature review of moose habitat needs, according to both scientific and cree perceptions, reveals some knowledge gaps which restrain sustainable management of moose habitat in the northern black spruce ecosystem. more important, this review allows a better understanding of similarities and differences between the cree and the scientific visions of moose needs and behaviour. at the end, a brief description of the research protocol ilalces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 149 lustrates how this project could contribute to building a common understanding between both visions of the biological requirements of moose and fill knowledge gaps to better serve cree and non-cree forest managers. since the project will end in 2006, no results are presented in this paper. shifting paradigm in natural resources management the existence of aboriginal and treaty rights in the canadian constitution, and the formal recognition of their attachment to and dependence on nature (gagnon 1973, croteau 1999), sustainable forest management on crown lands should involve distinct participative processes for aboriginals (western and wright 1994). many now suggest that complementary native and scientific management systems could generate improved sustainable natural resource management strategies (taiepa et al. 1997, duerden and kuhn 1998). integration of local indigenous collective knowledge (brassard 2001), needs, and values in a science-based management framework that recognizes the legitimacy of aboriginal peoples in the decision-making process, could facilitate resolution of management conflicts (daniels and walker 2001). aboriginal knowledge about the natural environment could complement scientific knowledge gaps about northern ecosystems (fast and berkes 1994), and enhance the resilience of aboriginal socioecological systems (begossi 1997). considering humans as a component of the ecosystem (gerlach and bengston 1994), this approach is in line fig. 1. eeyou astchee (waswanipi cree family hunting grounds). knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 150 with the practical definition of ecosystem management for which social acceptability and economic profitability must be attained within the limits of historical ecosystem variability and integrity (gilmore 1997, leduc et al. 2000). goal and objectives the goal of this project is the development and evaluation of the potential outcomes of a management process which uses two distinct sources of information (the cree and scientific knowledge) and evaluates their convergence to define moose habitat needs on eeyou astchee. based on this learning, innovative management strategies for moose habitat will be proposed to research partners. the first objective is to fill an inherent lack of scientific information about moose habitat needs in northern ecosystems. the study is mainly oriented toward establishing habitat preferences in terms of composition and structure of habitat types, to evaluate annual fidelity to specific areas (e.g., winter yards, calving sites), and to assess the importance of riparian habitats for moose. the second objective is to evaluate the impact of forest harvesting on moose habitat. this objective focuses on the duration of the negative impact of forest harvesting on moose (delay in habitat restoration) and the influence of habitat spatial patterns (e.g., buffer strips, mosaic cutting) on moose habitat use. the third objective is to elaborate on socioecologically-adapted moose habitat management strategies that will encourage participation of the cree in resource management and improve moose habitat. study site the research was conducted on land occupied by the waswanipi cree (eeyou astchee), and covers 35,000 km2 of boreal forest 600 km north of québec city, canada (fig. 1). the waswanipi cree have occupied this land since time immemorial (gagnon 1973) and trapped and hunted on it for subsistence needs: food, clothing, and tools (krech 1999). they took part in fur trade with a post in waswanipi island operated by the hudson bay company from 1819 to 1960 (marshall 1987). now, the small northern community numbers 1,200 people and the land is divided into 53 distinct family hunting grounds (fig. 1). in this area, the québec government granted the first industrial timber license in the early 1970s, which represented the first contact of the cree with modern forest exploitation (feit 1978). after a 30-year cutting period, the rift between the needs of the cree hunters and trappers and those of the forest industry has forced the provincial government to implement an adapted and exclusive forest regime in eeyou astchee. the objective of this new forest regime is to have better participation from the cree in the forest management planning process and better habitat management for featured species, such as moose. the court case (cree vs. québec and forest industry) which led to the adoption of this new forest regime, began on a family hunting ground where a yarding area for moose, protected for a long time by the tallyman, had been logged. eeyou astchee, which is part of the spruce–moss bioclimatic domain, is located in the most northern forest harvesting zone in québec (mrnq 2000), where the average annual climate is cold and humid with a mean annual temperature around -0.1 °c (beauchesne et al. 2000). more than 17% of forested land is unproductive (from a wood fibre aspect) and is comprised of muskeg, swamp, exposed rock, and open areas. the productive forest is composed of coniferous (89.2%), mixed (9.2%), and deciduous stands (1.5%). coniferous stands are dominated by black spruce (picea mariana) in association with balsam fir (abies balsamea) and jack pine (pinus banksiana). alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 151 in mixed stands, balsam fir is the dominant species, and is found in association with deciduous species such as paper birch (betula papyrifera) and trembling aspen (populus tremuloïdes) (bergeron et al. 1998). the climax forest is composed of large homogeneous stands of black spruce interspersed with small stands of intolerant deciduous trees (joyal 1987). state of the knowledge on moose habitat in northern ecosystems (from a scientific point of view) food-cover trade-off within their annual home range, moose travel between preferred habitat types according to specific seasonal needs (peek et al. 1992) and the degree of interspersion of different habitat components that can meet these needs (leresche 1974). generally, moose respond more to food availability than to cover accessibility compared with other ungulates (westworth et al. 1989). usually, prime food habitats and prime cover habitats are found in different areas within the animal’s range which may explain density changes between large areas (lima and dill 1990, dussault 2001). at a small scale, corresponding to a forest stand, open areas which are visited by moose for their high food availability usually do not provide adequate cover from adverse climatic elements; conversely, good shelter usually offers low food availability (lieffers et al. 1999). trade-offs between stands offering sufficient food and cover are more pronounced for females with calves than for males (thompson and vukelich 1981, courtois et al. 2002, dussault et al. 2005) or females without calf (dussault 2001). food availability has a significant influence on habitat selection by moose, at both coarse and fine scales (joyal 1987, crête 1989, courtois et al. 2002). in the boreal forest, moose are mostly associated with mixed and deciduous stands (grenier and audet 1974, joyal 1987, crête and courtois 1997) and young stands of post fire origin, windfall, insect outbreaks, and forest harvesting (loranger et al. 1991). these stand types supply an abundant shrub layer of deciduous twigs and balsam fir, which are the main food intake for moose in winter (girard and joyal 1984, crête 1989). the importance of mixed and deciduous stands is a significant variable that explains the variation in moose densities in northern environments (gingras et al. 1989). in northern coniferous forests, moose must extend their home range to travel between suitable patches of available food (crête and courtois 1997). scanty distribution of forage can lead to suboptimal nutrition and high in utero or perinatal mortality (verme 1977) and, consequently, can influence moose productivity in northern regions following unfavourable weather conditions (crête and courtois 1997). the quality of the cover has an impact on moose mortality due to hunting and natural predation (girard and joyal 1984, eason 1989), which are the two most important limiting factors for moose populations (courtois 1993). good cover is also important for thermoregulation (dussault 2001), as moose are easily stressed by warm temperatures throughout the year. coniferous canopies which retain snow are generally preferred over deciduous cover for deer (ozoga 1968) during periods of high snow accumulation. while coniferous cover does not seem to be a strong limiting factor for moose, this species does show a higher preference for such habitats in late winter presumably due to the higher interception properties of conifers (dussault et al. 2005). forestry and habitat management although forestry can be a useful wildlife management tool in rejuvenatknowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 152 ing old forest stands (peek et al. 1976, hundertmark et al. 1990) creating young productive feeding sites in terms of browse availability (crête 1977, girard and joyal 1984), large cutovers can have negative impacts on moose habitat, productivity, and population densities (joyal 1987, eason 1989). moose generally avoid large clear cuts and this negative impact can last more than 10–15 years post cutting (potvin et al. 1999). reduction of forest cover can increase moose vulnerability to hunting and natural predation (girard and joyal 1984, joyal 1987, rempel et al. 1997). moose usually react to clear cutting by increasing daily movements and avoiding clear cut patches having a sparse shrub layer (courtois et al. 2002). concentration of moose in patches of residual forest in a clear cut area can theoretically increase predation risk (girard and joyal 1984, joyal 1987). corridors and riparian habitats conventional forest harvesting in québec produces a patchwork of large cut blocks divided by 60-100m residual forest strips between cutovers and 20 m riparian reserves along lakes and streams. for some species, corridors composed of residual mature forest in a recently harvested area fulfill an inherent need for movement (hobbs and hopkins 1991) and can be considered as habitats (e.g., riparian communities; johnson 1989). however, there is little information on the importance of movement corridors for moose in the literature. it seems that moose do not seek buffer strips in conventionally cut landscapes (potvin and courtois 1998). on the other hand, some authors have documented the use of shoreline timber reserves by moose as resting cover and travel corridors (brusnyk and gilbert 1983) and the use of frozen, narrow river channels as bedding and feeding sites (hundertmark et al. 1990). in migrating moose populations in northern habitats, sandegren et al. (1983) determined that routes between seasonal ranges are located in valleys near rivers, which contain accessible shrubby growth (audet and grenier 1976). thus, riparian habitats can be considered as an important part of moose range even in the supposedly non-migrating populations of northern québec (joyal 1987). particular habitats, annual fidelity, and site reutilization unlike white-tailed deer (odocoileus virginianus), moose generally do not use exactly the same areas year after year (crête and jordan 1981). however, in northern regions, where the core matrix of the landscape is composed primarily of poor habitats interspersed with small scattered patches of good habitat, moose may seek these good habitats and thus use them repeatedly (nault and martineau 1983, potvin et al. 2001). the exclusivity and the specificity of these habitats can make them essential for moose (crête and courtois 1997), and thus may be reused year after year (jacqmain et al. 2003). state of the knowledge on moose habitat in northern ecosystems (from a cree point of view) food-cover trade-off according to the cree point of view, moose seem to associate more with undisturbed forests (jacqmain and bélanger 2002). in this view, forest fires are perceived mainly as a habitat destroyer and as a cause of starvation rather than as a nutrient recycling natural phenomenon that can rejuvenate over-mature forests and benefit moose (marshall 1987, dupont et al. 2005). however, regenerating stands are good feeding grounds for moose if they are close to coniferous cover (jacqmain and bélanger 2002, saganash 2004). in mature forest landscapes, moose are seen alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 153 as preferring mixed and deciduous mature stands, which offer simultaneously good cover and available food (feit 1987, lajoie et al. 1993, jacqmain et al. 2003, dupont et al. 2005). forestry and habitat management for the cree, forest harvesting is not perceived a priori to benefit moose in the short term (mce 1998, dupont et al. 2005), although the cree now visit some old clearcuts because of their good hunting potential (saganash 2004). compared with the foresters’ view, in which moose are perceived as benefiting in the mid-term from logging, this observation does not appear to be well accepted or confirmed by the cree (mce 1998). however, some moose hunters have noticed that moose may use regenerated areas about 20 years after harvesting when deciduous regeneration and available browse near good cover are present (jacqmain and bélanger 2002, saganash 2004). another cree concern relative to logging is that the riparian strips (residual mature tree corridors along water bodies) do not seem wide enough to be used by moose (hébert and bélanger 2004, saganash 2004, dupont et al. 2005). width and effectiveness of buffer strips between cut blocks are also criticized for the same reason, since habitat contiguity is a concern for native hunters (jacqmain and bélanger 2002, saganash 2004, dupont et al. 2005). for the cree, some special habitat types must be protected from logging, and need to be surrounded by a large residual buffer with traveling corridors to be continually used in a clearcut landscape (notion of ecozone) (lajoie et al. 1993, mce 1998, jacqmain and bélanger 2002, saganash 2004). for some tallymen, the management vision of a healthy trapline that integrates with forest harvesting would be represented by several permanently protected seasonal habitats (ecozone), surrounded by large residual mature stands, and interconnected by large mature forest corridors, including riparian habitats (lajoie et al. 1993, jacqmain and bélanger 2002, saganash 2004). furthermore, access to the land must be controlled to reduce the impact of over harvesting and poaching of wildlife (mce 1998, saganash 2004). corridors and riparian habitats the cree people suggest that moose annual cycle can be divided into 4 main seasons (spring–summer, mating time, winter, and late winter) for which needs and habitat preferences differ (feit 1987, jacqmain and bélanger 2002). moose need connectivity between these specific habitats to travel safely and to find enough food and water while traveling. in uncut landscapes, riparian habitats are thus very important because they offer good availability of deciduous twigs and accessible water and are bordered by dense mature cover for protection (jacqmain and bélanger 2002). particular habitats, annual fidelity, and site reutilization to fulfill some particular needs, the cree recognize that there are a few restricted, specific habitats (such as wintering and calving areas) that are critical. calving sites are normally situated in isolated areas, such as swamps or peninsulas, where moose can find protection and quiet. females may sometimes travel a long distance to find such places (feit 1987). wintering areas, known as moose yards, are typically described as elevated terrain, intersected by valleys, with mature mixed or deciduous stands used for food and mature coniferous stands for cover (lajoie et al. 1993, jacqmain and bélanger 2002, dupont et al. 2005). topography is also an important characteristic of moose yards where moose, in late winter, are primarily located in hilly terrain for several knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 154 reasons such as reduced risk of predation and lower snow depth (feit 1987). the cree have observed that, over time, moose consistently use these preferred and critical seasonal habitats (mce 1998, jacqmain et al. 2003). research approach since the project’s objective is to gather new information about moose habitat which will give a better portrait of such habitats in northern ecosystems, both cree and scientific visions must be investigated. this will allow filling gaps in the scientific and cree knowledge previously enunciated. to do so, two distinct but complementary research processes are used. these strategies tackle the same topics but will generate different information to help identify the similarities and differences in knowledge and build a common vision about moose habitat requirements. research hypotheses have been oriented based on several issues presented by the cree which seem to contradict available scientific literature and/or for which little scientific information is available for the study area. scientific investigation of moose habitat needs moose habitat use is being assessed using global positioning system (gps) radio collars. gps technology is the most appropriate tool for studying large mammals’ habitat use that otherwise has a high cost-benefit ratio (rodgers et al. 1997). furthermore, as proposed by weber (2000), a participatory management process will greatly benefit from using the “best” available science to bring new information to the decision-making table. this study targets only adult females as sample units (aebischer et al. 1993), because they are known to have more specific habitat requirements than adult males, particularly when they are accompanied by calves (thompson and vukelich 1981, courtois et al. 2002, dussault et al. 2005). moose were selected on the hunting grounds of cree families from whom we received permission by the tallymen to radio collar. this selection was made to optimally cover the study area (ecological variability and different forest management practices). net-gunning will be used for capturing moose, as tranquillizers require a 45-day consumption ban, which is incompatible with the winter subsistence hunting of the cree. the same 15 females will be monitored for 3 years with annual replacement of the collar battery packs. cree hunters’ needs and knowledge about moose habitat cree knowledge about moose, moose habitat needs, and the impact of forest activities on moose will be collected through interviews with recognized knowledgeable tallymen and cree hunters. hunters and trappers have based this knowledge on that of their ancestors and have amended it with their personal experiences, following an ecological empirical understanding process (gauthier 2002). although this knowledge is characterized as “less technological” (berkes 1993), its value and usefulness in natural resource management for data acquisition and for elaboration of management guidelines is factual, demonstrated, and recognized (mcdonald 1988, johnston and ruttan 1992, healy 1993, craig and smith 1996). gathering local ecological knowledge (davis and wagner 2003) involves following fundamental requirements when working with first nations. the research ethic includes developing a complete and understandable presentation of the project, gaining the support of the local community, involving a cree assistant in the project, presenting results, validating the results, and making sure that useful results will be available to the community. as the desired information concerns alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 155 moose habitat and needs of moose hunters, questions are mainly addressed to full-time moose hunters and tallymen who are considered to be dependent on local natural resources (gagnon 1973, mongeon 1993, croteau 1999). the majority of the interviews are in cree, with the cree research assistant translating from english to cree, as our targeted public (cree who live in the bush) are more at ease in the cree language. development of a common understanding and proposal of socio-adapted management strategies all interview questions are developed and formulated to obtain qualitative and quantitative information on the same topics tackled with scientific habitat analysis. although some comparisons can obviously be made, one must keep in mind that the goal of the project is not to rank cree knowledge versus scientific information, but rather to evaluate the degree of convergence between these two, and to understand those parts that do not appear to be congruent. based on new knowledge about moose habitat and the impact of forest harvesting, and with respect to the needs of tallymen and cree hunters, moose habitat management strategies will be proposed and developed. by considering scientific and cree knowledge, we expect that these management strategies will better suit stakeholders’ needs, while ensuring a suitable environment for moose. through a working committee within the model forest (cree, governments, industries, and university), results will be presented to the authority in charge of natural resources management in the study area for analysis and hopefully field implementation. potential outcomes of the project to our knowledge, there are few ongoing studies involving a true combination of aboriginal and scientific knowledge related to strategies for natural resource management. the use of gps collars and the involvement of the principal and knowledgeable stakeholders, allow managers to predict that results will be suitable and accurate within the socioecological context. better management of northern moose habitat, at the periphery of the animal’s range where the habitat may be restrictive (crête and courtois 1997), in combination with an adjusted population harvesting strategy will certainly benefit the species and allow densities to increase. the high degree of local, regional, and provincial support and the participation of principal stakeholders are good signs that the proposed approach has a high potential to resolve management conflicts. all the processes will be monitored and reported so they can be used or adapted in other areas involving resource management and first nations. acknowledgements we express thanks to the waswanipi cree model forest and its partnership; waswanipi cree first nation, tembec industry, ministère des ressources naturelles et de la faune du québec (especially the technical assistance of laurier breton), canadian forest service, laval university, québec wildlife foundation, fonds sur la nature et les technologies du québec, and wildlife habitat canada. we are also indebted to jacques robert and denis audette for their trust and wisdom. we finally thank the participation, confidence, and great patience of waswanipi tallymen and moose hunters. references aebischer, n. j., p. a. robertson, and r. e. kenward. 1993. compositional analysis of habitat use from animal radio-tracking data. ecology 74:1313-1325.ecology 74:1313-1325. audet, r., and p. grenier. 1976. habitat knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 156 hivernal de l’orignal dans la région de la baie james, étude préliminaire. ministère du tourisme de la chasse et de la pêche, québec, québec, canada. beauchesne, p., v. girardin, and j. p. ducruc. 2000. cadre écologique de référence etcadre écologique de référence et stations forestières. ministère de l’environnement, québec, québec, canada. québec, canada.. beaulieu, a. 2000. les autochtones du québec, des premières alliances aux revendications contemporaines. bibliothèque nationale du québec, québec, canada. québec, canada.. begossi, a. 1997. resilience and neo-traditional populations: the caicaras (atlantic forest) and coboclos (amazon, brazil). pages 120-157 in f. berkes and c. folke, editors. linking social and ecological systems. cambridge university press, cambridge, england, u.k. bergeron, j. f., p. grondin, and j. blouin. que du sous-domaine bioclimatique de la pessière à mousses de l’ouest. ministère des ressources naturelles du québec, direction des inventaires forestiers, québec, québec, canada.. berkes, f. 1993. traditional ecological knowledge in perspective. pages 35-51 in j. t. inglis, editor. traditional ecological knowledge: concepts and cases. international development research centre, ottawa, ontario, canada. brassard, m. j. 2001. la construction des savoirs collectifs locaux: un outil de transformation sociale pour les petites communautés? doctoral thesis. univer-autés? doctoral thesis. université du québec à chicoutimi, chicoutimi, québec, canada. brusnyk, l. m., and f. f. gilbert. 1983. use of shoreline timber reserves by moose. journal of wildlife management 47:673-685. courtois, r. 1993. description d’un indice de qualité d’habitat pour l’orignal (alces alces) au québec. ministère du loisir, de la chasse et de la pêche, direction générale de la ressource faunique, québec, québec, québec, canada.. _____. 2002. a preliminary assessment on structure on moose density in clear-cuts of north-western québec. alces 38:167176. _____, c. dussault, f. potvin, and g. daigle. 2002. habitat selection by moose (alces alces) in clear-cut landscapes. alces 38:177-192. _____, j.-p. ouellet, and b. gagné. 1998. characteristics of cutovers used by moose (alces alces) in early winter. alces 34:201211. craig, j., and r. smith. 1996. “a rich forest”: traditional knowledge, inventory and restoration of culturally important plants and habitats in the alteo river watershed. ahousaht ethnobotany project, university of victoria, victoria, british columbia, canada. crête, m. 1977. importance de la coupe forestière sur l’habitat de l’orignal dans le sud-ouest du québec. canadian journal of forest research 7:241-257. _____. 1989. approximation of k carrying capacity for moose in eastern québec. canadian journal of zoology 67:373-380. _____, and r. courtois. 1997. limitinglimiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal ofjournal of zoology 245:765-781. _____, and p. a. jordan. 1981. régime alimentaire des orignaux du sud-ouest québécois pour les mois d’avril à octobre. canadian field-naturalist 95:50-56. croteau, juge j. j. 1999. jugement de la cour supérieure du québec. gouvernement du québec, québec, canada. http://www. azimut.soquij.qc.ca/. daniels, s. e., and g. b. walker. 2001. worthe collaborative learning approach. praeger, westport, connecticut, usa. alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 157 davis, a., and j. r. wagner. 2003. who knows? on the importance of identifying ‘experts’ when researching local ecological knowledge. human ecology 31:463-489. duerden, f., and r. g. kuhn. 1998. scale, context, and application of traditional knowledge of canadian north. polarpolar record 34:31-38. dupont, p. p., r. roy, and l. imbeau. 2005.2005. modalités d’aménagement pour les aires forestières d’intérêt pour la faune dans la communauté de waswanipi. centre technologique des résidus industriels pour la forêt modèle crie de waswanipi, amos, québec, canada.. dussault environnementales sur la sélection de l’habitat de l’orignal. doctoral thesis. université laval, québec, québec, canada. _____., j.-p. ouellet, r. courtois, j. huot, l. breton, and h. jolicoeur. 2005. linking moose habitat selection to limiting factors. ecography 28:619-628. eason, g. 1989. moose response to hunting and 1 km² block cutting. alces 25:6374. fast, h., and f. berkes. 1994. native land use, traditional knowledge, and sustainable economy in the hudson bay bioregion. technical paper for the hudson bay programme, university of manitoba, natural resources institute, winnipeg, manitoba, canada. feit, h. 1978. waswanipi realities and adaptations: resource management and cognitive structure. doctoral thesis. department of anthropology, mcgill university, montreal, québec, canada. _____. 1987. north american native hunting and management of moose populations. swedish wildlife research supplement 1:25-42. _____. 1999. james bay cree. pages 34-78 in r.b. lee and r. daly, editors. the cambridge encyclopaedia of hunters and gatherers. cambridge university press, cambridge, england, u.k. gagnon, a. 1973. la baie james indienne. texte integral du jugement du juge albert malouf. éditions du jour, montréal, québec, canada. gauthier, b. 2002. recherche sociale: recherche sociale, de la problématique à la collecte des données. bibliothèque du québec, presse de l’université du québec, québec, canada. gerlach, l. p., and d. n. bengston. 1994. if ecosystem management is the solution, what’s the problem? journal of forestry 92:18-21. gilmore, d. w. 1997. what is ecosystem management? conservation biology 8:27-38. gingras, a., r. audy, and r. courtois. 1989. inventaire aérien de l’orignal dans la zone de chasse 19 à l’hiver 1987-88. ministère du loisir, de la chasse et de la pêche du québec, direction régionale de la côte-nord et direction de la gestion des espèces et des habitats, québec, québec, canada. girard, f., and s. joyal. 1984. l’effet des coupes à blanc sur les populations d’ori-nc sur les populations d’orignaux du nord-ouest du québec. alces 20:40-53. grenier, p., and r. audet. 1974. inventaire aérien de l’orignal et étude à petite échelle de son habitat dans le secteur nord du territoire de la société d’énergie de la baie james. ministère du tourisme de la chasse et de la pêche, québec, québec,, québec, canada.. healy cation of tek. pages 34-87 in n. m. and g. b. williams, editor. traditional ecological knowledge: wisdom for sustainable development. centre for resource and environment studies, australian national university, canberra, act, australia. hébert, j., and l. bélanger. 2004. riparian knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 158 zone protection, preliminary results. laval university for the waswanipi cree model forest, québec, québec, canada. hénault, m., l. bélanger, a. r. rodgers, g. redmond, k. morris, f. potvin, r. courtois, s. morel, and m. mongeon. 1999. moose and forest ecosystem management: the biggest beast but not the best. alces 35:213-225. hobbs, r. j., and a. j. m. hopkins. 1991. the role of conservation corridors in a changing climate. pages 281-290 in d.a. saunders and r.j. hobbs, editors. the role of corridors. chipping norton, surrey beatty, new south wales, australia. hundertmark, k. j., w.l. eberhardt, and r. e. ball. 1990. winter habitat use by moose in southeastern alaska: implications for forest management. alces 26:108-114. jackson, g. l., g. d. racey, j. g. mcnicol, and l. a. godwin. 1991. moose habitat interpretation in ontario. ontario ministry of natural resources, toronto, ontario, canada. jacqmain, h., and l. bélanger. 2002. ndoho istchee project, understanding, documenting and structuring the notion of ecozone laval university for the waswanipi cree model forest, québec, québec, canada _____, r. dion, and l. belanger. 2003. projet de caractérisation de l’habitat de l’orignal dans la pessière noire à mousses de l’ouest compris dans la zone de chasse 17. université laval pour l’administration régionale crie et la forêt modèle crie de waswanipi, québec, québec, canada. johnson, a. s. 1989. the thin green line: riparian corridors and endangered species in arizona and new mexico. pages 3546 in g. mackintosh, editor. preserving communities and corridors. defenders of wildlife, washington, d.c., usa. johnson, m., and r. a. ruttan. 1992. traditional environmental knowledge of the dene: a pilot project. pages 15-39 in m. johnson, editor. capturing traditional environmental knowledge. dene cultural institute and the international development research centre, ottawa, ontario, canada. johnston, m. h., and j. a. elliott. 1996. imblack spruce community in northwestern ontario. environmental monitoring and assessment 39:283-297. joyal s. 1987. moose habitat investigations in québec and management implications. swedish wildlife research supplement 1:139-152. krech, s. 1999. the ecological indian, myth and history. w.w norton & company, inc., new york, new york, usa. lajoie, g., r. beaulieu, and r. dion. 1993. caractérisation des ravages d’orignaux toire de la baie james (secteur waswanipi) à l’aide d’un système d’information géographique (sig). administration régionale crie, montréal, québec, canada. leduc, a., y. bergeron, p. drapeau, b. harvey, and s. gauthier. 2000. le régime2000. le régime naturel des incendies forestiers: un guide pour l’aménagement durable de la forêt boréale. l’aubelle, novembre-décembre:13-22. leresche, r. e. 1974. moose migration in north america. naturaliste canadien 101:393-415. lieffers, v. j., c. messier, k. j. stadt, f. gendron, and p. g. comeau. 1999. predicting and mapping light in the understory of boreal forests. canadian journal of forest research 29:796-811. lima, s. l., and l. m. dill. 1990. behavioural decisions made under the risk of predation: a review and prospectus. canadian journal of zoology 68:619-640. loranger, a. j., t. n. m. bailey, and w. w. larned. 1991. effects of forest succeson the kenai peninsula, alaska. alces alces vol. 41, 2005 jacqmain et al. knowledge on moose habitat 159 27:100-109. marshall, s. 1987. light on the water, a pictorial history of the people of waswanipi. waswanipi band, waswanipi, québec, canada. mcdonald, m. 1988. an overview of adaptive management of renewable resources. pages 12-43 in m. r. milton and l. n. carbyn, editors. traditional knowledge and renewable resource management in northern regions. iucn commission on ecology and the boreal institute for northern studies, university of alberta, edmonton, alberta, canada. messier, f. 1993. a review of moose management plan for hunting zones 17 and 22 in northern québec. university of saskatchewan for the grand council of the crees, saskatoon, saskatchewan, canada. (mrnq) ministère des ressources naturelles du québec. 2000. la limite nordique des forêts attribuables. gouvernement du québec, québec, québec, canada. (mce) ministère du conseil exécutif.du conseil exécutif. 1998. mario lord et grand conseil de cris du québec contre le procureur général du québec et al. cour du québec. gouvernement du québec, québec, québec, québec, québec, canada.. mongeon, m. 1993. l’appel du territoire. forêt conservation:20-25. morel, s., and l. bélanger. 1998. an integrated wildlife/forest management model: accommodating traditional innu activities and forest management practices. forestry chronicle 74:363-366. nault, r., and r. martineau. 1983. etude1983. etude de l’orignal (alces alces) de la région du futur réservoir d’eastmain. direction de l’environnement, société d’énergie de la baie-james, québec, québec, canada., québec, canada. ozoga, j. j. 1968. variations in microclimate in a conifer swamp deeryard in northern michigan. journal of wildlife management 32:574-585. peek, j. m., r. j. mackie, and g. l. dusek. 1992. over-winter survival strategies of north american cervidae. alces supplement 1:156-161. _____, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. potvin, f., and r. courtois. 1998. effets à1998. effets à court terme de l’exploitation forestière sur la faune terrestre: synthèse d’une étude de cinq ans en abitibi-témiscamingue et implications pour l’aménagement forestier. ministère de l’environnement et de la faune, québec, québec, canada., québec, canada. _____, _____, and l. bélanger. 1999. short-term response of wildlife to clearcutting in québec boreal forest: multiscale effects and management implications. canadian journal of forest research 29:1120-1127. _____, _____, and c. dussault. 2001. fréquentation hivernale de grandes aires de coupe récente par l’orignal en forêt boréale. société de la faune et des parcs, québec, québec, canada., québec, canada.. rempel, r. s., p. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. rodgers, a. r., r. s. rempel, r. moen, j. paczkowski, c. schwartz, e. j. lawson, and m. j. gluck. 1997. gps collars for moose telemetry studies: a workshop. alces 33:203-209. saganash, a., jr. 2004. draft directives on the protection and management of wildlife habitats. waswanipi forest authority, waswanipi band, waswanipi, québec, canada. sandegren, f., r. bergstrom, g. cederlund, and e. dansie. 1983. spring migration of female moose in central sweden. alces 19:210-234. taiepa, t., p. lyver, p. horsley, j. davis, m. knowledge on moose habitat – jacqmain et al. alces vol. 41, 2005 160 bragg, and h. moller. 1997. co-management of new zealand’s conservation estate by maori and pakeha: a review. environmental conservation 24:236-250. thompson, i. d., and m. f. vukelich. 1981. use of logged habitats in winter by moose cows with calves in northeastern ontario. canadian journal of zoology 59:2103-2114. tobias, j. l. 1991. protection, civilization, assimilation: an outline history of canada’s indian policy. pages: 23-56 in j. r. miller, editor. sweet promises: a reader on indian-white relations in canada. university of toronto press, toronto, ontario, canada. verme, l. j. 1977. assessment of natal mortality in upper michigan deer. journal of wildlife management 41:700-708. weber, e. p. 2000. a new vanguard for the environment: grass-roots ecosystem management as a new environmental movement. society and natural resources 13:237-259. western, d., and m. wright. 1994. natural connections: perspectives in community-based conservation. island press, washington, d.c., usa. westworth, d. a., l. brusnyk, j. roberts, and h. veldhuzien. 1989. winter habitat use by moose in the vicinity of an open pit copper mine in north-central british columbia. alces 25:156-166. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice alces vol. 45, 2009 larter – moose monitoring program 89 a program to monitor moose populations in the dehcho region, northwest territories, canada nicholas c. larter department of environment & natural resources, government of the northwest territories, po box 240, fort simpson, nt x0e 0n0, canada. abstract: moose (alces alces) are an important traditional and spiritual resource for residents of the dehcho region of the northwest territories. maintaining healthy and sustainable populations of moose for future generations is a goal of the department of environment and natural resources (enr). following a regional wildlife workshop with dehcho first nations, the need for a program to determine baseline information on moose populations and to foster community-based monitoring of moose in the dehcho was identified. such a program needed to be established prior to future proposed developments including the mackenzie gas project. after extensive community consultation between local first nations and enr, a baseline aerial survey over a large area of the dehcho was designed, and was to be followed by an annual monitoring program. two key components identified for the annual monitoring program were an aerial survey and harvest sampling. the aerial survey would provide information on moose density and calf production, and harvest sampling would provide information on the relative health and physical condition of animals consumed by local residents. in light of increasing developmental pressures in the region, such information collected over time is important to harvesters, first nations, wildlife managers, and land use planners alike because it should document change in the quantity and quality of a key traditional wildlife resource. population estimates from the aerial surveys indicated that the estimated population density and calf:cow ratios were reasonable. harvest data indicated low incidence of diseases and parasites, low levels of cadmium in organ tissue, and that moose were mostly in good or excellent body condition based on observation and fat indices. this study is an example of successfully combining the knowledge and cooperation of first nation moose harvesters with the technical support of government biologists to secure valuable biological information for baseline data to monitor change associated with development in a region. alces vol. 45: 89-99 (2009) key words: aerial surveys, dehcho first nations, land development, monitoring, moose, northwest territories, physical parameters. moose (alces alces) are an important traditional and spiritual resource for residents of the dehcho region (dehcho) in the southwestern northwest territories of canada. prior to april 2002, the dehcho had no regional biologist and wildlife research programs were very limited. there had been a few surveys for moose and dall’s sheep (ovis dalli) conducted in the 1980s shortly after the establishment of the all-season liard highway (decker and mackenzie 1980, treseder and graf 1985, case 1989). an additional moose survey was conducted in the liard valley near nahanni butte in the late 1990s. canada’s northern boreal forest is often viewed as one of the last great wilderness in the world. increasingly, however, northern people are realizing that this wilderness is and will be altered by large-scale factors generated by outside processes beyond their direct control. mining and oil and gas development are major concerns in communities that rely on native species or “country foods” for sustenance. with increased resource development and access to the boreal forest comes the potential for increased harvest and changes in the relative moose monitoring program – larter alces vol. 45, 2009 90 health and condition of moose. in september 2002 at a regional wildlife workshop in fort simpson hosted by dehcho first nations and the then department of resources, wildlife & economic development (rwed), first nations delegates expressed similar concerns and made it clear that there was an immediate need for study of moose in the region. the paucity of information about moose in dehcho was recognized as an important omission/gap in the biophysical assessment for the proposed mackenzie valley pipeline. through many community meetings with local first nations and government biological staff, a survey area was determined and an extensive baseline population survey for moose was conducted during winter 2003-2004. at the 2nd biannual dehcho wildlife workshop in october 2004 there was high community interest in the results of the baseline survey, the need to continue annual monitoring, and to expand the program to record measures of animal health and condition. after several community meetings with local first nations, a more extensive moose monitoring program was designed. in this paper i describe the program and provide some preliminary results of the surveys. study area the study area included the dehcho administrative region of the southwestern northwest territories as defined by the department of environment and natural resources (enr), government of the northwest territories (ca. 150,000km2). it included the lower portions of the mackenzie river valley to the north and east, the liard river valley to the south, and the mackenzie mountains to the west. there were 6 communities in the region: fort liard, nahanni butte, trout lake, jean marie river, fort simpson, and wrigley. two of these communities (nahanni butte and trout lake) were accessible by road only during winter; the others were on an all-weather road system. the study area fig. 1. the dehcho study area in the northwest territories; depicted are community locations and the road system, including winter roads. alces vol. 45, 2009 larter – moose monitoring program 91 also included a portion of northeastern british columbia because the traditional harvesting area of residents of the acho dene koe band of fort liard extends south into northeastern british columbia (fig. 1). the vast majority of the study area is part of the northern boreal forest and is dominated by the taiga plains ecoregion. the taiga shield ecoregion is found to the east, and the mackenzie mountains that make up the extreme western portion of the study area are part of the tundra cordillera ecoregion. moose and woodland caribou (rangifer tarandus caribou), of both the boreal and northern mountain ecotypes, are the dominant ungulates and wood bison (bison bison athabascae) occur in the liard valley from nahanni butte south into northeastern british columbia. timber wolves (canis lupus), black bear (ursus americanus), grizzly bear (ursus arctos), wolverine (gulo gulo), and lynx (lynx canadensis) are the key predators. methods baseline population surveys logistics and cost precluded the use of the typical gasaway survey technique (gasaway et al. 1986) given the huge, relatively uninhabited area to be surveyed. a workshop on moose survey techniques in may 2003 brought together biologists from yukon, alaska, and northwest territories to discuss new refinements in moose survey methodology. a geospatial technique (ver hoef 2001, 2002) was in use as the primary survey method in alaska and the yukon. after discussions with first nations, a consensus was reached that this method would be used for baseline surveys of moose in the dehcho. the geospatial survey method consists of 5 basic elements: 1) defining the survey area, 2) stratifying the area, 3) determining sample sizes, 4) surveying a random sample of sample units within the area, and 5) analyzing the data. the major differences between this approach and the gasaway technique are that the size of the survey area can be much larger, sample unit boundaries are delineated by latitude (2°) and longitude (5°), correction for sightability is unnecessary, and the analysis utilizes a spatial statistics model. importantly, data analysis could still follow the basic approach of gasaway et al. (1986). we defined 2 survey areas: 1) the mackenzie river valley (mrv) extending from blackwater river in the north to jean marie river in the south including the communities of wrigley, fort simpson, and jean marie river, and 2) the liard river valley (lrv) extending from poplar river in the north and the british columbia-northwest territories boundary (60o n latitude) in the south including the communities of nahanni butte and fort liard (fig. 2). to determine survey areas, we sent maps to each of the communities and requested that they indicate areas they would like included in the survey. the maps were then combined digitally on a gis system, and a composite map was drafted and circulated. once the composite map was finalized, a 2o latitude x 5o longitude grid of all the sample units was created (fig. 2); sample units averaged ca.16 km2. we evaluated the size of the survey areas, aircraft availability, and length of daylight and decided to survey the mrv (23,281 km2) in november 2003 and the lrv (9,585 km2) in february 2004. the sample units in each survey area were stratified by expected low and high moose density almost exclusively from traditional knowledge of first nations harvesters after many meetings between communities and environment and natural resources (enr) biological staff. there was no conflict in stratifying areas that overlapped traditional hunting areas used by neighboring communities. however, there were some isolated parts of the survey areas where harvesters had insufficient knowledge to comfortably stratify the survey blocks. for those areas we relied on previous survey and/or landcover classification information to stratify the survey blocks. a draft of the proposed stratification moose monitoring program – larter alces vol. 45, 2009 92 was then circulated to first nations for ratification. survey blocks were randomly chosen; 100 blocks totaling 1,595 km2 for the mrv survey and 78 blocks totaling 1,313 km2 for the lrv survey. the analysis followed ver hoef (2001, 2002). we used 2 cessna 185 fixed-wing aircraft with pilot and 3 passengers. a survey crew consisted of 1 pilot, 1-2 local harvesters as observers, and 1-2 enr staff as recorder/observers. we programmed the locations of all sample units into handheld portable and fixed gps units before each flight, and aircraft flew to the designated sample units. we employed a total count for all sample units, which usually consisted of flying transects of varying distances apart across the sample unit. we attempted to keep airspeed at approximately 160 km/h and elevation at 125-175 m above ground level; altitude was influenced by ground cover. the amount of time required to survey each sample unit varied by topography and ground cover. a waypoint was taken for all observed moose to determine whether or not the animal was located inside the sample unit. any tracks were followed to determine if the moose was within the unit boundary. a track log of the flight path was recorded on a handheld gps. we classified moose into 5 different classes: calf, female, and small, medium, and large male based upon antler and body size. monitoring surveys after completing the initial baseline surveys it was apparent that costs and logistics would be prohibitive to continue annual large-scale surveys. however, there was need to monitor the moose population density, the number of calves:100 adult females, and male:female ratios between periodic largescale surveys. following discussions with first nations and biostatician jay ver hoef, a smaller area was designated for annual surveys fig. 2. the two baseline survey areas separated by the hatched line. the mackenzie river valley (mrv) contained 1459 sample units covering 23,281 km2 in the north and the liard river valley (lrv) contained 569 sample units covering 9,585 km2 in the south. alces vol. 45, 2009 larter – moose monitoring program 93 to provide an adequate subset of the original sample units used in the baseline surveys. the annual surveys were conducted in november to minimize classification errors of sex/age classes as compared to february when calves are larger and males are antlerless. accurate sex/age classification to estimate calf:100 adult females and male:female ratios is an important component of this population monitoring program. five subsets containing 17-66 sample units from the baseline mrv survey area, and 4 subsets containing 18-69 sample units from the baseline lrv survey area were delineated (fig. 3). we used the same sample unit identification numbers and stratification in these subsets as in the baseline surveys. in november 2004 we flew 34 of a possible 268 sample units in the mrv, and 20 of a possible 169 sample units in the lrv. following the 2004 survey, we realized we could fly more sample units cost-effectively and increase coverage to ca. 16% of both mrv and lrv. we flew 43, 40, and 43 sample units in the mrv and 27, 28, and 25 sample units in the lrv in 2005, 2006 and 2007, respectively. all sample units surveyed were selected randomly, and survey flights were conducted as in the original baseline surveys except that we used only one cessna 185. health and condition sampling of moose tissue and organs first nation harvesters in the dehcho were requested to provide biological samples and general information from harvested moose starting in the winter of 2004-2005. posters indicating the required samples were circulated and labeled sampling kits were provided by enr to local dene band and métis offices for distribution to local harvesters. because part of the program was designed to measure the level of various elements in consumed moose, fig. 3. location of sample units used in the monitoring surveys; the mackenzie river valley (mrv) and liard river valley (lrv) survey areas are separated by the hatched line. monitoring surveys used a subset of the original baseline sample units; 268 sample units covered 4,281 km2 in the mrv, and 169 sample units covered 2,848 km2 in the lrv. moose monitoring program – larter alces vol. 45, 2009 94 we required submission of a complete kidney and a sample of liver. kidneys are considered a delicacy and harvesters were initially quite unwilling to provide these samples unless there was some form of compensation. following discussions with first nations delegates at the dehcho regional wildlife workshop in 2004, it was agreed that harvesters would be paid $50 (cad) for submitting a complete suite of samples. the following information and samples were requested by enr: name of hunter, date of harvest, location of harvest, sex, estimated age (calf, yearling, adult), general body condition (excellent, good, fair, poor), pregnant (yes, no), lactating (yes, no), an entire kidney including the surrounding attached fat, the lower jaw or incisor bar, a minimum 5 cm x 5 cm piece of liver, a minimum 5 cm x 5 cm piece of muscle, an intact ankle bone with marrow, and a handful of fecal pellets. each sampling kit contained a pencil and sampling checklist that provided room for additional comments (e.g., presence of abnormalities and parasites). it was important to compile information about the body condition of harvested moose from the perspective of the harvesters who have a wealth of past experience with moose in the area. samples were collected until march 2007 to attain a sample size of at least 40 to assess contaminant levels. sample kits were kept frozen and transferred to the enr office in fort simpson for processing. since march 2007, enr has requested a reduced number of biological samples including only the incisor bar or front teeth, an intact ankle bone with marrow, and fecal pellets; sample kits are provided but samples are submitted on a voluntary basis. harvesters are happy not to have to submit the highly valued kidneys. to supplement the samples from moose harvested by first nations residents that were almost exclusively from the main river drainages in the taiga plains ecoregion, enr requested 2 outfitting operations in the mackenzie mountains (tundra cordillera ecoregion) to provide an entire kidney including surrounding attached fat, the front incisor bar, a minimum 5 cm x 5 cm piece of liver, and a minimum 5 cm x 5 cm piece of muscle from moose harvested by their clients during fall (september-october) hunts, on an opportunistic basis. samples were collected from fall harvests in 2004-2006 and were processed at the enr office in fort simpson. age and bone marrow a first incisor was extracted from each moose and sent to matson’s laboratory (milltown, montana, usa) where they were aged by counting cementum annuli from the root of the premolar (matson 1981). ankle bones were collected only from moose harvested by first nation hunters. the bone marrow fat was extracted from a 5-10 cm length of bone and placed on a petri dish. the petri dish and the wet marrow fat were weighed on an ohaus electronic balance (±0.005 g) and then placed in a drying oven at 100 oc for a minimum of 48 h until the weight of the dried marrow fat and petri dish was constant. i calculated the % fat content in ankle bone marrow with the formula: % fat content = (dry weight of fat + petri dish/wet weight of fat + petri dish) x 100 kidney and fat analysis kidneys and accompanying fat were thawed and weighed on an ohaus electronic balance (±0.005 g). the fat was trimmed following riney (1955) and the kidney with remaining fat was weighed. we then peeled the fat and capsule off the kidney and reweighed the organ. the kidney fat index (kfi) was calculated for each kidney with the formula: kfi = (weight of fat remaining after trimming/weight of kidney) x 100 we also calculated the ratio of the total fat weight (before trimming) to kidney weight for each kidney. alces vol. 45, 2009 larter – moose monitoring program 95 for the elemental analyses the kidney was rinsed in distilled water, and cut in half bilaterally. one half of the kidney was placed in a ziploc bag, labeled, and frozen. each liver sample was rinsed with distilled water and trimmed to not exceed 300 g wet weight. the rinsed sample was placed in a whirl-pak bag, labeled, and frozen. all frozen samples were shipped for analysis to the environment canada laboratory at the aquatic ecosystem protection research division, burlington, ontario, canada. organ tissue analysis all organ samples were thawed, thoroughly homogenized, and subsamples of each tissue (wet) were digested in a closed teflon vessel using ultrapure nitric acid. the resulting digestate was analyzed for the concentrations of 31 elements including arsenic, cadmium, lead, selenium, and zinc with inductively coupled plasma-mass spectrometry. the digestate was analyzed for mercury by cold vapour atomic absorption spectroscopy. there is little consistency in the reporting of heavy metal levels in moose tissue from canadian, american, and scandinavian studies. some report in wet weight (e.g., gamberg et al. 2005, venäläinen et al. 2005, arnold et al. 2006) and others dry weight (e.g., paré et al. 1999, crichton and paquet 2000); therefore, results are in both wet and dry weights. the moisture content of each tissue sample was determined prior to acid digestion. the results were expressed in mg/ kg or ppm wet weight and converted to mg/kg or ppm dry weight with the formula: ppm dry weight = ppm wet weight/((100 % moisture content)/100) fecal samples frozen fecal samples (525 g wet weight) were forwarded to the bow valley research lab in calgary. subsamples were screened for the presence of giardia and cryptosporidium by the sucrose flotation method. briefly, fecal material is suspended with phosphate buffered saline solution (pbss) and filtered, the filtrate is added to a sucrose solution (specific gravity 1.13) with methylene blue added, and then centrifuged. the resulting supernatant and pellet are decanted in pbss. immunofluorescent staining is applied to the suspended pellet to assist with the microscopic examination. presence was reported as the number of oocytes/g of feces. subsamples were also screened for parasites using the modified wisconsin fecal floatation technique, based upon a concentrated sugar solution. briefly, fecal material is centrifuged in water, the resulting homogeneous solution is filtered, and centrifuged again. the supernatant is resuspended in sugar flotation solution (specific gravity1.270) and centrifuged once more before the contents are examined under a microscope. parasite presence was reported as the number of eggs/g of feces. results and discussion baseline population surveys the mrv and lrv surveys were done after the traditional fall moose harvest in september-october. the mrv survey was flown 11-16 november 2003 (fig. 2). we surveyed 63 high density and 37 low density sample units, providing approximately 7% coverage of the survey area. an average time of 19 min 45sec (range 9 min 12 sec to 48 min 19 sec) was spent surveying and 140 moose were observed, 74 in sample units. the density estimates were 4.4 moose/100 km2 and 32.1 calves:100 adult females. the lrv survey was flown 16-19 february 2004 (fig. 2) and 52 high density and 26 low density sample units were surveyed providing approximately 14% coverage of the survey area. an average time of 17 min 23 sec (range 9 min 3 sec to 30 min 25 sec) was spent surveying and 90 moose were observed, 65 in sample units. the density estimates were 4.9 moose/100 km2 and 44.6 calves:100 adult females. in november moose were more active and visible, in larger groups, and easier to classify by sex and age classes than in februmoose monitoring program – larter alces vol. 45, 2009 96 ary. in contrast, ground cover was completely frozen, the weather was generally better for flying, and length of daylight was increasing in february, but moose were in more closed habitats. based upon the range in number of moose observed in high density areas (010) and low density areas (0-2), i feel that the stratification process and decisions were appropriate, and that the geospatial survey technique used for the baseline surveys provided, at worst, a conservative estimate of the minimum population density of moose in the study area following the traditional fall (september-october) moose harvest. monitoring surveys coverage of the annual monitoring surveys ranged from 12-16% of the baseline survey areas; 4,281 km2 for mrv and 2,848 km2 for lrv (fig. 3). annual surveys ranged from 34-43 blocks in the mrv area and 20-28 blocks in the lrv area. the average time spent surveying sample units ranged from about 14 min 14 sec to 20 min 59 sec in the mrv and 15 min 16 sec to 20 min 45 sec in the lrv area; 60-82 moose were observed during these annual surveys. a minimum ratio of 35 calves:100 adult female moose was estimated from the survey data, a value midrange of the those in the baseline surveys. health and condition sampling of moose tissue and organs i was prepared to pay a monetary reimbursement to first nation harvesters for providing a suite of biological samples that included organs that were highly valued traditional food. i also wanted to ensure that an adequate sample size was collected within a reasonable time frame because voluntary sampling programs often suffer from inadequate sample sizes collected over extended time (years) raising issues of confidence (thomas et al. 2005). a total of 43 complete suites of tissue samples were fig. 4. the locations of 61 harvested moose from which a full suite of biological samples was collected; the origin was designated as ● from the mackenzie mountains, x from the mackenzie river valley, and + from the liard river valley. alces vol. 45, 2009 larter – moose monitoring program 97 collected from moose harvested throughout the study area in january 2005-march 2007. in addition, 18 suites of tissue samples were collected by outfitters in the mackenzie mountains where moose were harvested in september and october 2004-2006 (fig. 4). harvesters rated most moose as excellent or good condition (table 1); there were no reports of winter ticks or papillomas on any harvested moose. age as anticipated, the average age of moose harvested by first nation hunters (x = 4.3 yr, range = 0-12, n = 43) was lower than that of moose harvested by guided hunters in the mackenzie mountains (x = 7.4 yr, range = 4-14, n = 17), because the latter harvested exclusively trophy males (table 1). the average age of females (5.6 years) harvested by first nation hunters was higher than that of males (3.6 years) even though the age range was the same for both sexes (0-12 yr; table 1). bone marrow ideally i would have preferred to collect femur bones for measuring marrow fat content because it is a commonly reported index for assessing starvation and the capability of habitat to support moose (franzmann 1998). however, only ankle bones were made available. few harvested moose had low marrow fat regardless of sex or when they were harvested; just 4 of 39 had marrow fat content <50.0%. the average marrow fat content of males (69.6%) was lower than that of females (78.4%; table 1). the lowest marrow fat was from a 12 yr old male harvested in january 2007; although the hunter described the moose as skinny, it was ranked as fair not poor condition. kidney and fat analysis although there is some question about the relationship between total body fat and kfi (mcgillis 1972), for first nation harvesters the amount of fat associated with the kidneys is a visual indicator of overall animal condition. based upon their reports of body condition, moose rated as excellent body condition had an average kfi of 75.0, those rated as good body condition had an average kfi of 47.6, and those rated as fair body condition had an average kfi of 31.6. no harvested moose was rated as poor body condition. not surprisingly, males taken by guided hunters in september-october had lower kfi, averaging 29.0; 1 male had no kidney fat. females harvested by first nation hunters had a somewhat higher average kfi than males (table 1). organ tissue analysis the levels of cadmium found in the kidneys of moose harvested by first nation hunters was similar to those reported in moose harvested elsewhere in north america and scandinavia (e.g., paré et al. 1999, macdonald 2002, venäläinen et al. 2005). the majority of kidneys from these animals had renal cadmium levels <60μg/g table 1. a summary of results from various analyses conducted on biological samples collected from harvested moose in september 2005-march 2007, dehcho, northwest territories, canada. data are categorized male and female and are from first nation (fn) or guided harvesters in the mackenzie mountains (mt). calves were denoted as 0 years of age. the % marrow fat was from ankle bones submitted by first nation harvesters. body condition was reported by the harvester and ranked as excellent, good, fair, or poor. n average range age 60.0 5.2 0-14 fn male 28.0 3.6 0-12 fn female 15.0 5.6 0-12 mt male 17.0 7.4 34.0 kidney fat index 48.0 47.1 0-155.8 fn male 25.0 50.0 5.2-155.8 fn female 12.0 57.5 15.3-142.2 mt male 11.0 29.0 0-57.8 % marrow fat 39.0 72.7 10.8-96.6 fn male 25.0 69.6 10.8-92 fn female 14.0 78.4 53.3-96.6 body condition 61.0 excellent (21) good (34) fn male 28.0 7.0 17.0 fn female 15.0 6.0 7.0 mt male 18.0 8.0 10.0 moose monitoring program – larter alces vol. 45, 2009 98 dry weight (<20μg/g wet weight). however, the levels of cadmium found in the kidneys of moose harvested by guided hunters in the mackenzie mountains was higher than found in moose harvested from the major river drainages; 50% of moose sampled from the mackenzie mountains had renal cadmium levels >1000μg/g dry weight (>225μg/g wet weight). gamberg et al. (2005) found similarly high levels of cadmium in kidneys of moose harvested in yukon. as expected, cadmium levels were higher in older than younger moose. interestingly, there was a relatively strong linear relationship between the level of cadmium in the liver and kidneys of moose. this finding could be important because it could indicate that collecting samples of liver only would be sufficient to measure and monitor the relative cadmium level. local harvesters are much more willing to provide samples of liver than an entire kidney from their moose. fecal samples presence of disease and parasites was low; none of the 41 fecal samples tested positive for giardia or cryptosporidium. although nematodirus eggs were present in 31 of 41 fecal samples (76%), most positive samples had <10 eggs/g. the greatest infestation was 21 eggs/g found in a 2-year old male. based upon samples from harvested moose, the incidence and prevalence of diseases and parasites is low in dehcho moose. summary concern for the impact of increased resource use, access, and development in dehcho resulted in a cooperative research program to provide baseline information about resident moose. enr conducted aerial surveys and first nations harvesters and guided hunters and outfitters provided information from harvested moose. population density estimates and calf:cow ratios, low incidence and prevalence of diseases and parasites, low levels of cadmium in organ tissue, and fat indices and physical assessments indicated that moose were productive and in goodexcellent body condition and provide a highly valued and healthy traditional food resource. this study serves as an example of combining the knowledge and cooperation of first nation moose harvesters, with the technical support of government biologists to secure baseline data for monitoring change in a moose population. acknowledgements this paper is dedicated to the memory of cam lancaster of nahanni butte outfitters. cam provided samples and insight to the moose programs in the dehcho. cam passed away in a plane crash in the mackenzie mountains shortly after this paper was presented at the 6th international moose symposium in yakutsk, russia. i am indebted to all of the many dehcho traditional harvesters whose input was critical to designing and implementing this program; without their support and input this program would never have materialized – mahsi cho. i thank those harvesters and outfitters that provided biological samples for the program and the many harvesters who have assisted in the aerial surveys conducted as part of the program. jay ver hoef is acknowledged for his assistance with the design and statistical analyses of aerial survey data. deborah johnson, dean cluff, and lynda yonge provided discussion and critique of the program and earlier drafts of this manuscript. comments from ken child and another anonymous reviewer improved this manuscript. erin bayne completed the statistical analyses of the trace element data. danny allaire produced the figures. funding for this program has come from the government of the northwest territories, indian and northern affairs canada, parks canada, and the northern contaminants program. references arnold, s. m., r. l. zarnke, t. v. lynn, m. alces vol. 45, 2009 larter – moose monitoring program 99 -a. r. chimonas, and a. frank. 2006. public health evaluation of cadmium concentrations in liver and kidney of moose (alces alces) from four areas of alaska. science and the total environment 357: 103-111. case, r. 1989. distribution and abundance of dall’s sheep in the southern mackenzie mountains, northwest territories. department of renewable resources file report 81, yellowknife, northwest territories, canada. crichton, v., and p. c. paquet. 2000. cadmium in manitoba’s wildlife. alces 36: 205-216. decker, r., and j. mackenzie. 1980. the population, distribution and density of moose in the liard valley 1978. northwest territories wildlife service file report 7, yellowknife, northwest territories, canada. franzmann, a. w. 1998. restraint, translocation, and husbandry. pages 519-557 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. gamberg, m., m. palmer, and p. roach. 2005. temporal and geographic trends in trace element concentrations in moose from yukon, canada. science and the total environment 351-352: 530-538. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska no. 22, fairbanks, alaska, usa. macdonald, c. r. 2002. summary of field and contaminant data for the 2002 collection of bluenose-east caribou near deline, nt. report submitted to the sahtu renewable resources board, tulita, northwest territories, canada. matson, g. m. 1981. matson’s workbook for cementum analysis. milltown, montana, usa. mcgillis, j. r. 1972. the kidney fat index as an indicator of condition in various age and sex classes of moose. north american moose conference workshop 8: 105-114. paré, m., r. prairie, and m. speyer. 1999. variations in cadmium levels in moose tissues from the abitibi-témiscamingue region. alces 35: 177-190. riney, t. 1955. evaluating condition of freeranging red deer (cervus elaphus) with special reference to new zealand. new zealand journal of science and technology section b. 36: 429-463. thomas, p., j. irvine, j. lyster, and r. beaulieu. 2005. radionuclides and trace metals in canadian moose near uranium mines: comparison of radiation doses and food chain transfer with cattle and caribou. health physics 88: 423-438. treseder, l., and r. graf. 1985. moose in the northwest territories – a discussion paper. department of renewable resources manuscript report 13, yellowknife, northwest territories, canada. venäläinen, e-r., m. anttla, and k. peltonen. 2005. heavy metals in tissue samples of finnish moose, alces alces. bulletin of environmental contamination and toxicology 74: 526-536. ver hoef, j. m. 2001. predicting finite populations from spatially correlated data. pages 93-98 in proceedings of the 2000 joint statistical meetings in statistics and the environment section. american statistical association, indianapolis, indiana, usa. _____. 2002. sampling and geostatistics for spatial data. ecoscience 9: 152-161. 4002.p65 alces vol. 40, 2004 toweill and vecellio shiras moose in idaho 33 shiras moose in idaho: status and management dale e. toweill1 and gary vecellio2 1 idaho department of fish and game, p.o. box 25, 600 s. walnut street, boise, id 83720, usa; 2 idaho department of fish and game, 4279 commerce circle, idaho falls, id 83204, usa abstract: limited data indicate that shiras moose (alces alces shirasi) occurred in low numbers in idaho throughout the 19th century. harvest was allowed in idaho during 1893-1898, after which seasons were closed. shiras moose were fully protected in idaho from 1899-1945. moose populations increased during the 20 th century and harvest seasons resumed in 1946. harvest has focused on mature males, allowing continued population growth through the end of the 20 th century. rapid population growth during 1980-2000 resulted in moose dispersing westward from the rocky mountains and southward from the panhandle region of idaho. the management goal for moose in idaho is to provide opportunities for recreational hunting and harvest of mature male moose. although some managers assess moose populations directly by aerial survey, most managers rely on indirect measurements (e.g., hunter success rate and antler spread of bulls harvested) to assess the impact of harvest on moose populations. other population indicators (e.g., dispersal into previously unoccupied areas, damage to private property) have been used as indicators of social tolerance for expanding moose populations. where moose have approached the limit of social tolerance, attempts to stabilize or reduce populations by harvest of females and translocation of ‘problem’ moose have been utilized. both a historic perspective of moose abundance and a revised statewide population estimate are provided. alces vol. 40: 33-43 (2004) key words: antler measurements, controlled harvest, idaho, management, moose, shiras typical moose habitat in idaho encompasses timbered western slopes of the rocky mountains. in idaho, moose occupy all western slopes of the rocky mountains westward to hells canyon and isolated mountain ranges south of salmon, idaho along the border with montana and wyoming southward to utah. moose are managed as a game animal in idaho. the idaho department of fish and game (idfg) holds management authority and has identified moose as a trophy species; a big game animal whose population is sufficient to support only strictly regulated annual harvest. in addition to regulating harvest, idfg has responsibility to respond to depredation complaints caused by moose (toweill 1988). moose occupied slightly more than half (51%) of idaho, an area of 109,668 km2 (42,343 mi2) in 2002. moose are hunted in all administrative regions of idaho, and in about two-thirds of idaho game management units (gmu) (fig. 1). the recent expansion of moose in idaho has allowed the idfg to increase moose hunting opportunity from < 20% of gmus during 1946-1982 to > 60% of gmus by 2000 (fig. 2). we describe recent range expansion of moose, summarize idfg harvest data, and provide a revised population estimate for shiras moose in idaho. historic distribution the distribution of moose in 2002 was much greater than at any previous time in shiras moose in idaho – toweill and vecellio alces vol. 40, 2004 34 0% 20% 40% 60% 80% 100% 19 46 19 50 19 54 19 58 19 62 19 66 19 70 19 74 19 78 19 82 19 86 19 90 19 94 19 98 20 02 fig. 2. percent of game management units with moose permits offered, idaho 1946-2002. moose in northern idaho exist prior to 1900. moose apparently became established in the area of yellowstone national park soon after 1850, and were reported in the salmon river mountains in 1891 (merriam 1891). the first hunting season for moose was established in idaho in 1893, but was closed in 1898 due to concern about dwindling herds. writing in 1905, brooks reported that moose occurred in southeastern idaho in a range bounded by “the eleventh auxillary meridian on the west and the fall or cascade creek on the east” and by “the southern branch of the warm river on the north and the big robinson on the south” (brooks 1905:201), an area known as big black mountain or moose mountain that “barely measures ten miles in diameter” (brooks 1905:202). he reported that moose had formerly ranged as far south as jackson hole and east of the north fork of the snake river in idaho, wyoming, and montana, but that the range had become progressively restricted within the previous decade (1895-1905). elimination of moose hunting seasons in idaho beginning in 1899 may have allowed moose populations to grow. bailey (1935) reported that there were “numbers” of moose in the chamberlain basin and salmon river watershed in 1902. davis (1939) reported that idaho moose numbered about 500 in 1910. citing reports of increasing moose in the upper snake river valley in 1935 and an estimate of 528 moose in national forests of northern idaho in 1925, davis (1939) estimated that idaho had 1,000 moose in 1939. thirty permits authorizing the harvest of bull moose in fremont county only were authorized by the idaho department of fish and game (idfg) in 1946, and again in 1947. during that period, fremont county was believed to include the range of more than half the moose in idaho (biladeau 1949). an aerial survey of moose in fremont county in 1949 yielded observations of 536 2002 moose hunting bull bull and cow southwest clearwater panhandle salmon upper snake southeast magic valley fig. 1. state of idaho, department of fish and game administrative regions and game management units showing availability of bull and cow moose permits, 2002. r e c o r d e d h i s t o r y . e x p l o r e r s w i t h merriwether lewis and william clark’s corps of discovery failed to observe moose, although they were informed by native americans in 1806 that there were “… plenty of moos (sic) to the s.e. of them on the east branch [salmon river] of lewis’s [snake] river …” (thwaites 1959, vol. 5:99). journals of the fur trappers and explorers that traveled throughout the western rocky mountains between 1806 and 1850 failed to mention the occurrence of moose (compton and oldenburg 1994). houston (1968) concluded that few if any moose occupied the area of jackson hole and yellowstone national park prior to 1850. few records of alces vol. 40, 2004 toweill and vecellio shiras moose in idaho 35 fig. 3. percent (%) of first-choice applications for bull and cow moose permits being drawn, idaho 1990-2002. 0% 20% 40% 60% 80% 100% 19 90 92 94 96 98 00 20 02 19 90 92 94 96 98 00 20 02 bulls cows moose (biladeau 1949). records from states adjacent to idaho provide additional indication of moose population expansion. moose from eastern idaho apparently expanded southward into utah by 1906 or 1907, although a population was not considered established until 1947 (durrant 1952, utah division of wildlife resources 2000). in similar fashion, moose populations expanded westward from the priest lake basin by 1954, establishing a population in northeastern washington (poelker 1972). moose likely crossed hells canyon and the snake river from idaho into the blue mountains of washington (ingles 1965) and oregon (verts and carraway 1998), although there is no evidence that these movements resulted in establishment of new populations to date. moose incursions into oregon have continued with increasing frequency, with 25 records since 1960, 18 of those since 1990 (vic coggins, oregon department of fish and wildlife, file data, november 2002). moose management moose are managed by idfg to provide high quality hunting opportunities and associated recreation, while encouraging expansion of moose populations into suitable habitat in idaho (leege et al. 1990). harvest of moose is strictly controlled. permits are issued randomly to applicants. successful applicants become ineligible for life following harvest of one moose of either sex. allocation of hunt permits harvest of moose is regulated by controlled hunt permits allocated by random draw. each permit is restricted to either antlered or antlerless moose (hereafter bull or cow) within a particular hunt area. every hunter is required to have each harvested moose checked by a representative of idfg. hunter demand for moose permits is high. in 1980, idfg received 25,524 applications for 140 moose permits (leege et al. 1990), with the result that only 1 person among 182 applicants obtained a moose hunting permit (at that time, all permitees were limited to harvest of antlered moose). to reduce competition after 1980, applicants were required to submit funds for the purchase of their permit and tag with their application. this requirement reduced the number of applicants by over half (from 25,524 to 11,649) in 1981. increases in the number of permits offered annually has resulted in a higher likelihood of being drawn for a permit since that time. likelihood of drawing a permit for antlered moose was about 10% from 19901999, and has been near 20% since 2000 (fig. 3). the number of applications for antlerless moose permits has expanded rapidly since 1990, when drawing success was similar to that for antlered moose (about 15%). however, the number of antlerless moose permits offered annually has increased even more rapidly, so that by 1999 the number of applicants was less than the number of antlerless permits available. permits not fully subscribed in the annual drawing have been sold on a ‘first-come’ basis following the drawing. permittees unsuccessful in harvesting a moose must wait 2 years before becoming eligible for another moose tag. regulations are reviewed and permit levels established on alternate, odd-numbered years. shiras moose in idaho – toweill and vecellio alces vol. 40, 2004 36 successful moose hunters must have their animal checked by an idfg representative within 10 days of harvest. unsuccessful hunters are required to submit their unused moose tag as proof of non-use (failure to do so is presumptive evidence of harvest and exclusion from future draw opportunity). most moose hunting in idaho occurs on public land. a summary of land ownership in areas open to moose hunting (fig. 4) indicates that 94% of the land area is managed by federal or state government. the vast majority of federal and state land in idaho is open to hunting. controlled harvest idfg moose management philosophy is to allow harvest of antlered moose at levels which allow populations to continue to expand. therefore, harvest quotas for antlered moose (i.e., moose having at least one antler longer than 15.2 cm or 6 inches) are limited, and adjusted as necessary to achieve a mean maximum antler spread of harvested bull moose > 89 cm (35 inches). at this harvest level, the mean age of harvested moose is believed to be approximately 4 years of age (gasaway et al. 1987). harvest of antlerless moose is designed primarily to reduce moose population growth, promote human health and safety where moose occur in suburban settings, and limit moose depredations. moose hunting seasons are long. hunting seasons for bull moose extend 86 days, from august 30 to november 23 annually. hunting seasons for cow moose typically extend 40 days (october 15-november 23). long seasons allow successful applicants maximum opportunity for hunting recreation and opportunity to harvest. opening dates for cow seasons were delayed until october 15 in an effort to reduce losses of orphaned calves by allowing them an addifig. 4. area (km2) open to moose hunting in idaho by administrative region, and land ownership, 2002. tional 10 weeks to mature. since 1990, moose hunters have averaged 5.4-8.2 days of hunting before harvesting an antlered (bull) moose, and 2.65.2 days before harvesting an antlerless moose (fig. 5). more days hunting for each bull harvested reflects reduced availability due to lower numbers of bulls versus cows and great selectivity in choosing a bull to harvest for this once-in-a-lifetime trophy. mean number of days prior to harvest has stayed relatively constant in the last 12 years for both bulls and cows (fig. 5). moose harvest success has ranged from > 60% to > 80% annually (fig. 6). the most common cause identified by unsuccessful hunters for failure to harvest a moose is lack of participation during the hunting season. harvest data are used to monitor the effect of hunting on moose populations. the statewide objective for mean antler fig. 5. mean number of days hunted prior to harvest for bull and cow moose, by year in idaho, 1990-2002. 0 5,000 10,000 15,000 20,000 25,000 30,000 35,000 pa nh an dle cl ea rw ate r so ut hw es t ma gic v all ey so ut he as t up pe r s na ke sa lm on s q ua re k m private state federal 0 1 2 3 4 5 6 7 8 9 10 19 90 92 94 96 98 00 20 02 19 90 92 94 96 98 00 20 02 # da ys h un te d bulls cows alces vol. 40, 2004 toweill and vecellio shiras moose in idaho 37 fig. 6. moose permits and harvest including all zones and tags statewide, idaho 1990-2002. percent harvest success labeled above permits. spread is > 89 cm (35 inches) among all harvested bulls, and has been in place since 1990. antler spread of harvested moose has been maintained at that level since 1990 (fig. 7). maximum antler spread recorded in idaho has been 152 cm (60 inches), and each year a few moose are harvested that approach this size (fig. 7). annual harvest of antlered moose is generally believed to account for 15% of known bulls, although data are limited. based on file data from the northeastern portion of gmu 1 (jim hayden, personal communication, idfg), the population of moose was 0.31 moose/km2 (0.80 moose/ mi2) during february 1993. bull moose density was 0.093 bull moose/km2 (0.24 bull moose/mi2) in this area, and bull moose harvest density was 0.015 bull moose/km2 (0.04 bull moose/mi2). this equated to an estimated annual hunting mortality rate of 14% [0.015/(0.015 + 0.093)]. some areas are more heavily exploited. in gmu 2 near the washington border, annual harvest was estimated to account for 38% of the bull moose present in 1996, and 33% of the bull moose in 2000. surveys of gmu 2 conducted in february 1996 resulted in an estimate of 0.104 moose/km2 (0.27 moose/mi2) and 0.031 bull moose/km2 (0.08 bull moose/mi2). harvest accounted for 0.019 bull moose/km2 (0.05 bull moose/ mi2) in 1996, for a harvest rate of 38% [0.019/(0.019 + 0.031)]. moose populations had increased to 0.193 moose/km2 (0.50 moose/mi2) in 2000, with an estimated 0.039 bull moose/km2 (0.10 bull moose/mi2). annual harvest accounted for 0.019 bull moose/ km2 (0.05 bull moose/mi2), yielding an annual harvest rate of 33% [0.019/(0.019 + 0.039)]. estimates of comparatively higher annual harvest in gmu 2 were reflected in smaller average antler spread from this gmu, although sample sizes are small (jim fig. 7. mean antler spread and 95% confidence interval for moose in idaho, 1990-2002. sample sizes shown above range, height of wide box is 95% ci. 74 % 80 % 77 % 80 % 76 % 78 % 75 % 74 % 61 % 74 % 82 % 80 % 85 % 0 300 600 900 1200 1500 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 permits harvest 38 9 36 2 36 1 42 5 35 0 51 3 51 6 57 2 56 5 66 1 67 8 73 8 63 1 0 10 20 30 40 50 60 70 80 90 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 an tl er s pr ea d (i n. ) 0 50 100 150 200 an tl er s pr ea d (c m ) shiras moose in idaho – toweill and vecellio alces vol. 40, 2004 38 12 6 23 46 13 00 419 9 17 9610 52 0 10 20 30 40 50 60 70 80 90 panhandle clearwater southwest magic valley southeast upper snake salmon idfg administrativ e region an tle r sp re ad (i n. ) 0 50 100 150 200 an tle r sp re ad (c m ) fig. 8. mean antler spread and 95% confidence interval for moose in idaho by administrative region, 1990-2002. sample sizes are shown above range, height of wide bar is 95% ci. hayden, personal communication, idfg). moose populations and harvests are greatest in northern idaho (panhandle and clearwater regions) and extreme eastern idaho (upper snake and southeast idaho) (fig. 1). among all regions, mean antler spread ranges from 89.9 cm (35.4 inches) in the salmon region to 94.0 cm (37.0 inches) in the panhandle region (fig. 8). mean antler measurements do differ (p < 0.001) among regions, with the panhandle and upper snake regions being similar and slightly greater than clearwater and southeast regions (fig. 8). among the moose harvested during seasons designated for antlerless harvest, a portion (3-22%) are males (primarily calves). since 1990, the portion of antlerless harvest consisting of males has averaged 7.6% (table 1). unregulated harvest and mortality this category includes all recorded annual losses of moose to human activity. major elements of these types of losses include vehicle accidents and illegal hunter harvest. the extent of these losses is difficult to measure, because there is no central repository for this information and reporting is sporadic. in addition to these causes of mortality, other factors may also impact local moose populations. one of these factors is translocation of moose by idfg. idfg has legal responsibility to respond to wildlife depredation concerns (toweill 1988), and one means of addressing these concerns is translocation of moose within idaho. methodology for translocating moose was described by naderman (1994). although the number of translocations of moose varies annually depending on severity of winter weather, during the winter of 2001-2002 approximately 104 moose were physically relocated away from idaho falls and nearby areas in eastern idaho. among 527 moose deaths recorded in fremont county between 1969 and 1975 (ritchie 1978), legal harvest accounted for 217 (41%). the balance of losses was comprised of 165 moose illegally harvested (31%), 32 moose allocated to indian harvest (6%), and 113 moose deaths attributed to natural causes, accidents, and unknown alces vol. 40, 2004 toweill and vecellio shiras moose in idaho 39 table 1. antlerless moose permits, harvests, and % male calves in the antlerless harvest in idaho, 1993-2001. table 2. documented human-caused and natural/unknown moose mortalities not considered legal harvest for idaho, 1990-2002. year antlerless permits total harvest % males in harvest (n) 1993 65 54 22.2% (12) 1994 65 40 10.0% (4) 1995 81 63 7.9% (5) 1996 81 63 3.2% (2) 1997 98 73 11.0% (8) 1998 98 66 6.1% (4) 1999 123 109 3.9% (4) 2000 123 87 6.9% (6) 2001 142 93 4.3% (4) total 876 648 7.6% (49) category mortality factor number human-caused vehicle & train 452 illegal kill 416 indian harvest 97 other human-caused 48 natural/unknown unknown 177 natural mortality 71 winter kill 46 predation 5 causes (21%). research conducted on moose between june 1979 and december 1980 in central idaho near elk city (pierce et al. 1985) documented cause of death for 40 moose. of these, 10 (25%) were legally harvested. of the balance, 21 (50%) were illegally killed, 6 (15%) were harvested by tribal members, and 3 (8%) moose deaths were due to accidents and natural causes. pierce et al. (1985) reported that 7 of 20 moose radio-collared by one of the authors (kuck, unpublished) near soda springs in southeastern idaho died during 1978-1981. six of those animals (86%) were illegally harvested. pierce et al. (1985) concluded that unregulated harvest from all causes was largely unreported and often underestimated. a review of all recorded mortality other than legal hunting during the period 19902002 revealed that mortality due to vehicle (including train) collisions and illegal harvest were the dominant causes of nonhunting related mortality (table 2). mortality due to vehicle collisions is significantly underestimated, since there is no comprehensive effort to collect moose-vehicle collision data and mortally injured moose capable of moving away from the scene of an accident under their own power are rarely recorded as mortalities. if located, postmortem cause of death for these animals is usually categorized as either natural or unknown. given the relatively high likelihood of vehicle accidents going unreported to idfg and post-collision mortality of moose struck but able to leave the scene of a collision, it is suspected that reported moose mortality due to vehicle collisions may represent half of actual mortality. while losses of approximately 50 moose/year due to collisions have been reported since 1990, annual losses were estimated to be more than twice that number by local conservation officers, and increasing as both moose and roads proliferate. illegal harvest is also believed to be significantly under-reported. illegal harvest and wounding of moose by hunters seeking elk and deer are rarely reported by individuals responsible, most of whom are fearful of receiving a citation. many of the people who illegally harvest moose do so in locales where the potential for discovery is low (private lands, remote sites, etc.), and such individuals may hide evidence of their activity (pierce et al. 1985). although 3040 illegal kills have been recorded annually statewide since 1990 (table 2), pierce et al. (1985) estimated that 5-10% of moose populations in 2 study areas died annually as a result of recorded illegal kills. annual losses due to illegal harvest are likely increasing as expanding moose populations provide additional opportunities. shiras moose in idaho – toweill and vecellio alces vol. 40, 2004 40 we believe (based on reports from conservation officers statewide and investigations of illegal harvest) that annual illegal kill of moose averages 50 moose/region, or 350-400 moose statewide. in addition to illegal kills, moose in idaho may also be legally harvested by members of several indian tribes holding subsistence or harvest treaty rights. such harvest is rarely reported to idfg. since 1990, 97 incidents of moose harvest by indians have been reliably reported, which accounts for only 7% of all moose mortalities recorded due to causes other than idfg-regulated harvest (table 2). natural losses losses of moose due to natural causes (predation, disease, accidents, malnutrition, etc.) are rarely reported. most occur away from human habitations or roads, and many occur during seasons (i.e., winter) when few humans are active in remote portions of moose habitat. natural mortality of moose older than calves is believed similar to that reported for adult cow moose in alaska by ballard et al. (1991), where an annual mortality of 5.2% was recorded. bangs et al. (1989) recorded a slightly higher rate of mortality (8%), with mortality of animals aged 1-5 years only 3%. since 1990, natural and unknown-caused moose mortalities account for 299 cases (23%) of all nonharvest mortalities (table 2). in idaho, potential predators on moose include black bears (ursus americanus), mountain lions (felis concolor), and wolves (canis lupus). data relative to predation on moose in idaho is scarce; only 5 of 1,312 known non-harvest mortalities since 1990 have been attributed to predators (table 2). mountain lions are suspected as the cause of 3 of the 5 recorded predator kills in idaho (big game mortality reports, idfg, boise, idaho, usa). population estimation population estimates for moose are difficult to obtain, even in relatively small areas, and total counts are impossible over large areas. helicopter surveys have been used to provide a means of estimating moose numbers over large areas in idaho, but large areas occupied by moose occur in steep, heavily-vegetated terrain where aerial surveys are impossible. the first statewide estimates of idaho’s moose population were 500 moose in 1910, and 1,000 moose in 1939 (davis 1939). hatter (1949) reported a population of 1,000 moose in idaho, based on an aerial survey of moose in fremont county conducted in 1949. it is unclear whether hatter considered herds in northern idaho (where very few moose may have been present at that time) in his estimate, which was reported as a statewide total population estimate. wildlife managers of idfg, using a variety of data and input from local conservation officers, estimated the moose population in each gmu in idaho during 1981, 1985, and 1990 (idfg 1981, hayden et al. 1985, leege et al. 1990). other estimates of idaho’s moose population (table 3) appear in karns (1998) and timmermann and buss (1995, 1998). with population surveys unavailable, biologists typically employ indices (relative measures of some object such as pellet groups or tracks) to detect trends in populations. only rarely can such indices be correlated to population number except in a very general sense. in idaho, statewide population trends are monitored using a combination of aerial survey estimates over small areas, and indices based on mandatory check of hunter harvested moose and antler measurements of bull moose. since current harvests are inconsistent with published estimates of moose populations in idaho, we reviewed available data in an effort to derive an updated statewide estimate of idaho’s moose population. alces vol. 40, 2004 toweill and vecellio shiras moose in idaho 41 table 3. historic estimates of moose in idaho. 1wildlife species management plans; idaho department of fish and game 1981, hayden et al. 1985, leege et al. 1990. 2karns 1998. 3timmermann and buss 1995, 1998. year idfg1 karns2 timmermann & buss3 1960 4,100 1965 4,400 1970 4,600 1975 4,700 1980 4,900 1981 3,530 1982 3,600 1985 4,385 5,100 1990 4,565 5,100 5,500 population estimate based on occupied range and population density one way to estimate idaho’s moose population is to derive a population density then expand that to population area. moose densities in wyoming, immediately east of idaho, were estimated using fixed-wing and helicopter surveys designed to produce confidence intervals within 10% (hnilicka 1994). estimates averaged 0.042 moose/km2 (0.11 moose/mi2) of occupied habitat, and ranged from 0.04 – 0.52 moose/ km2 (0.10 – 1.34 moose/mi2) (hnilicka 1994). in areas where comparable surveys have been flown in idaho, comparable moose densities have been recorded. aerial survey data from the caribou national forest of eastern idaho (idfg 2002) yielded estimates of moose densities of 0.24-0.40 moose/km2 (0.63-1.04 moose/mi2). similar data obtained from aerial surveys in northern idaho’s priest river drainage (jim hayden, personal communication, idfg) indicated that moose densities may reach 0.42 moose/km2 (1.1 moose/mi2). if we assume that idaho moose densities are bracketed by the minimum density for moose dispersal of 0.2 moose/km2 reported by gasaway et al. (1980) and the average density of 0.29 moose/km2 reported for wyoming, then idaho would have a statewide moose population between 20,000 and 30,000 moose (0.2 * 109,038 = 21,808 moose, and 0.29 * 109,038 = 31,621 moose). this is based upon an estimated occupied range equal to the area of gmus now having a moose harvest season (fig. 1). population estimate based on harvest and estimated mortality moose populations remain stable if annual recruitment equals annual losses. since we know or can estimate annual losses of the male portion of the population, and since we have samples from the population that reflect the relative proportions of males, females, and calves within the population, we can derive a crude but conservative estimate of population size—crude because harvest (the best monitored mortality factor) is dependent on the number of permits issued annually, and conservative since we assume population stability despite evidence that the statewide moose population is expanding. to derive this estimate, we need to know the proportion of the population comprised of males (34%, based on aerial survey data collected in 2000 and 2002), the number of bull moose removed annually by hunters (733 plus 4 male calves in 2001), and the proportion of the males removed by harvest (estimated to be 15%). then, the number of males in the population can be estimated (737/0.15 = 4,913). since males comprised 34% of the total population, the population can be estimated (4,913/0.34 = 14,450). a population of 14,450 moose in idaho would equate to 0.13 moose/km2 (0.34 moose/mi2). while both of these estimates are crude approximations, we believe they provide bounds on idaho’s moose population, and that idaho moose conservatively numbered between 15,000 and 25,000 animals in 2002; shiras moose in idaho – toweill and vecellio alces vol. 40, 2004 42 approximately 3 times population estimates published in 1990 (table 3). acknowledgements the authors appreciate the thoughtful comments provided by jim hayden and pete zager, idfg, and janet rachlow of the university of idaho. references bailey, r. g. 1935. river of no return. b a i l e y b l a k e p r i n t i n g c o m p a n y , lewiston, idaho, usa. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. bangs, e. e., t. n. bailey, and m. f. porter. 1989. survival rates of adult female moose on the kenai peninsula, alaska. alces 21:17-35. biladeau, t. d. 1949. idaho big game kill report, 1947-1948. mimeo. idaho department of fish and game, boise, idaho, usa. brooks, h. 1905. the idaho moose. pages 201-216 in new york zoological society tenth annual report, brooklyn, new york, usa. compton, b. b., and l. e. oldenburg. 1994. the status and management of moose in idaho. alces 30:57-62. davis, w. b. 1939. the recent mammals of idaho. the caxton printers limited, caldwell, idaho, usa. durrant, s. d. 1952. mammals of utah: taxonomy and distribution. university of kansas, lawrence, kansas, usa. gasaway, w. c., s. d. dubois, and k. l. brink. 1980. dispersal of subadult moose from a low density population in interior alaska. proceedings of the north american moose conference and workshop 16:314-337. _____, d. j. preston, d. j. reed, and d. d. roby. 1987. comparative antler morphology and size of north american moose. swedish wildlife research supplement 1:311-325. hatter, j. 1949. the status of moose in north america. transactions of the north american wildlife conference 14:492-501. hayden, j. a., b. ritchie, and b. davidson. 1985. moose species management plan, 1986-1990. idaho department of fish and game, boise, idaho, usa. hnilicka, p. 1994. the status and management of moose in wyoming. alces 30:101-107. houston, d. b. 1968. the shiras moose in jackson hole, wyoming. grand teton natural history association, moose, wyoming, usa. technical bulletin 1. (idfg) idaho department of fish and game. 1981. trophy species: moose, bighorn sheep, mountain goat, pronghorn antelope species management plan, 1981-1985. idaho department of fish and game, boise, idaho, usa. _ _ _ _ _ . 2 0 0 2 . p r o j e c t w 1 7 0 r 2 5 ; statewide surveys and inventories, moose. federal aid in wildlife restoration report. idaho department of fish and game, boise, idaho, usa. ingles, l. g. 1965. mammals of the pacific states: california, oregon, and washington. stanford university press, stanford, california, usa. karns, p. d. 1998. population distribution, density and trends. pages 125-139 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. leege, t. a., c. anderson, d. aslett, and t. lucia. 1990. moose species management plan, 1991-1995. idaho department of fish and game, boise, idaho, usa. merriam, c. h. 1891. results of a biologialces vol. 40, 2004 toweill and vecellio shiras moose in idaho 43 cal reconnaisance of south-central idaho. u.s. department of agriculture, government printing office, washington, d.c., usa. north american fauna 5. naderman, j. 1994. methodology for relocating moose. alces 30:109-115. pierce, d. j., b. w. ritchie, and l. kuck. 1985. an examination of unregulated harvest of shiras moose in idaho. alces 21:231-252. poelker, r. j. 1972. the shiras moose in washington. washington department of game, olympia, washington, usa. ritchie, b. w. 1978. ecology of moose in fremont county, idaho. wildlife bulletin 7. idaho department of fish and game, boise, idaho, usa. thwaites, r. g., editor. 1959. original journals of the lewis and clark expedition, 1804-1806. antiquarian press limited, new york, new york, usa. timmermann, h. r., and m. e. buss. 1995. the status and management of moose in north america—early 1900’s. alces 31:1-14. _____, and _____. 1998. population and harvest management. pages 559-615 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. toweill, d. e. 1988. wildlife depredation plan. idaho department of fish and game, boise, idaho, usa. utah division of wildlife resources. 2000. statewide management plan for moose. utah division of wildlife resources, salt lake city, utah, usa. verts, b. j., and l. n. carraway. 1998. land mammals of oregon. university of california press, berkeley, california, usa. p37-48_3925.pdf alces vol. 41, 2005 stéen et al. diseases in moose 37 diseases in a moose population subjected to low predation margareta stéen1, ing-marie olsson1, and emil broman2 1veterinary service and food control, county administrative board of gaevleborg se 801 70 gaevle, sweden; 2department of applied environmental science, göteborg university, box 464, se 405 30 gothenburg, sweden abstract: well-publicized studies on two moose diseases, elaphostrongylosis and moose wasting syndrome, conducted across sweden from 1985 to 1994 resulted in an increased number of reports of sick and dead moose. a sample of 724 moose were investigated, including 426 females, 208 males, and 90 of unknown sex with an average age of 3.7 years (sd = 4.9, range 0-20). prominent diagnoses were elaphostrongylosis (18%), moose wasting syndrome (11%), and accidental death (11%). other important diagnoses were neoplasm (5%), parasitic (6%), nervous system (5%), infectious (4%), eye and ear diseases (4%), and predation (3%). from the beginning to the middle of the 20th century approximately 10 wolves, 130 bears, 175 lynx, and 100 wolverines were present in sweden (449,000 km2). currently, the scene is quite different with wolf, bear, lynx, and wolverine populations all increasing. the total number of large predators and scavengers is estimated at 2,500-3,000. we believe that the diversity of moose diseases seen in the future will differ from that observed during the 1980s and 1990s by being less visible due to increasing predation. alces vol. 41: 37-48 (2005) key words: alces alces, bear, canis lupus, diseases, elaphostrongylosis, moose wasting syndrome, predation, ursus ursus, wolf statistics on regulated hunting (culled) of moose (alces alces) have been recorded in sweden since 1881, mirroring the population density. the population began to increase in the late 1920s, rose rapidly in the 1970s, and peaked in the 1980s (cederlund and markgren 1987, cederlund and bergström 1996). today, the summer population is estimated to be 300,000-400,000 animals (stéen et al. 1998b) resulting in an approximate density of 1.0 – 1.5 moose/km2. the moose harvest in sweden is substantial with approximately 100,000 animals being culled each year and the meat annually comprising 4-5% of the country’s total meat production (stéen et al. 1998b). because moose are an important natural resource of considerable economic value to tourism, hunting, and meat production, a long-term overview of diseases affecting their health is of great societal interest. routine investigations of wildlife diseases have a long history in sweden (stéen et al. 1997, 1998b). diseases in moose have mostly been described as single or clusters of cases (borg 1975, 1987; stéen et al. 1998b). climate, especially snow depth, and nutritional stress due to limited food resources have been regarded as the main causes of natural (i.e., ing the 1980s another picture emerged, when large numbers of sick or dead moose were found throughout sweden. two diseases, elaphostrongylosis (ela), caused by the parasite elaphostrongylus alces, and moose wasting syndrome (mws), the etiology of which is still unknown, gave rise to public concern and interest. in 1985, two projects were initiated at the swedish university of agricultural sciences (slu) to investigate both diseases and a number of reports and papers have subsequently been published (stéen and diseases in moose – stéen et al. alces vol. 41, 2005 38 rehbinder 1986; feinstein et al. 1987; stéen and diaz 1988; stéen et al. 1989; stéen and johansson 1990; stéen and roepstorff 1990; rehbinder et al. 1991; stéen 1991; olsson et al. 1993; stéen et al. 1993; frank et al. 1994; merza et al. 1994; stéen et al. 1994; olsson et al. 1995; stéen et al. 1997; frank 1998; lankester et al.1998; olsson et al. 1998; stéen et al. 1998a; frank et al. 2000a, 2000b, 2000c, 2000d; gajadhar et al. 2000; olsson 2001; of disease in wild animals is in accordance with that of wobeser (1981) who described diseases as any impairment that interferes with tions, including responses to environmental factors such as nutrition, toxicants, climate, infectious agents, inherent or congenital defects, or combinations of these factors. an overview of the diseases seen in moose examined from 1985 to 1989, when non-human predators, including wolf (canis lupus), brown bear (ursus arctos), wolverine (gulo gulo), and lynx (lynx lynx), were few is presented in this paper. diseases and other causes of morbidity and mortality were grouped into 19 diagnostic categories. the results are compared to previous reports of disease and mortality in moose. methods study period and material whole carcasses or organs from approximately 1,000 moose were examined from 1985 to 1989. in this paper we describe moose examined by stéen (table 1). data collection and necropsy data describing where and how each animal was found or killed and the circumstances surrounding the case accompanied each sample. post-mortem and follow-up investigations were performed as described in stéen et al. (1997, 1998a). evaluation of physical condition was done (n = 642) by visual inspection of the body fat, its location, and appearance. condition categories included normal condition, below normal (poor), serous atrophy, or absence of adipose tissue (emaciated). moose were aged by tooth wear and eruption (skunke 1949, reimers and nordby 1968). diagnosed causes of disease were categorized (table 2). statistics inferential statistics were performed using sas®. analyses were considered statistically p < 0.05. results samples originated from across sweden, with the majority of cases being from north of stockholm (59°n). samples of sick or dead moose were submitted year-round, although lowed by fall, winter, and summer (table 1). samples from hunter harvested animals were taken in the fall. age and sex the average age of moose examined was table 1. summary of moose samples examined in the study. parameters number of animals total 724 type of sample carcass 315 organ 405 unknown 4 sex females 426 males 208 sex unknown 90 season fall 294 winter 154 spring 200 summer 76 manner of death euthanized 230 found dead 304 culling 171 unkown 19 alces vol. 41, 2005 stéen et al. diseases in moose 39 3.7 years (sd = 4.9, range 0-20, n = 617). grouped into 4 age classes, the distribution was: calf (41%), yearling (11%), adults 2-11 years (49%), and seniors 12-20 years (9%). the overall sex ratio was 1 bull/2cows, a female-biased adult ratio which makes cows more numerous than bulls (fig. 1). the sex was unknown in 12% of the cases. diagnoses diagnoses were grouped into: blood, lymphatic and cardiovascular systems (blc), digestive system (dig), endocrine system (end), eye and ear (ee), infectious diseases (inf), metabolic disturbances (met), muscleskeletal system (mus), neoplasm (neo), nervous system (ner), parasitic diseases tive and urinary systems (rep), respiratory system (res), and skin and connective tissue (skn) (merck veterinary manual 1979). additional diagnoses were malformation (mal), elaphostrongylosis (ela), moose wasting syndrome (mws), predation (pred), and miscellaneous causes (mis). no pathological of the 609 diagnosed cases, the most frequently diagnosed condition was ela (22%), followed by mws (14%), and accidental diagnoses were par, neo, ner, inf, and ee. predation was seen in 3% of cases and was comprised of 11 calves, 3 yearlings, 7 adults, 3 seniors, and 1 of unknown age. the age distribution of the predated cases did not differ from that of cases with other diagnoses 2 = 2.5280, 3 df, p = 0.4703). the relative risk (proportion of diagnosis table 2. number of animals diagnosed per disease category based on age classes. disease category calves adults (2-11 years) seniors (12-20 years) unknown total blood, lymphatic and cardiovascular systems (blc) 1 1 3 2 1 8 digestive system (dig) 7 1 6 1 1 16 endocrine system (end) 3 1 4 eye and ear (ee) 5 2 14 5 26 infectious diseases (inf) 7 20 3 2 32 metabolic disturbances (met) 1 1 muscle-skeletal system (mus) 3 1 13 1 2 20 neoplasm (neo) 1 3 24 6 5 39 nervous system (ner) 11 24 3 38 parasitic diseases (par) 11 12 16 1 3 43 25 12 36 3 2 78 reproductive and urinary systems (rep) 2 4 1 7 respiratory system (res) 5 11 2 2 20 skin and connective tissue (skn) 1 1 2 malformation (mal) 4 8 1 13 elaphostrongylosis (ela) 94 15 16 2 5 132 moose wasting syndrome (mws) 16 2 50 13 2 83 predation (pred) 11 3 7 3 1 25 miscellaneous causes (mis) 7 4 8 2 1 22 total 211 57 265 48 28 609 diseases in moose – stéen et al. alces vol. 41, 2005 40 among necropsied moose) from ela was greater in calf/yearlings than in adults/seniors 2 = 73.4419, 1 df, p < 0.0001, 2 = 18.8946, 1 df, p < 0.0001, fig. 2b). for mws the opposite pattern was seen in respect to age classes (fig. 2a,b), however differences were not statisti2 = 14.1872, 1 df, p 2 = 0.6691, 1 df, p = 0.4134, females and males, respectively). animals with neo and inf were over-represented in 2 = 17.0127, 1 df, p < 0.0001, 2 = 7.6095, 1 df, p = 0.0058, neo and inf, respectively). among calves, sex was not related to 2 = 0.1374, 1 df, p 2 = 0.2476, 1 df, p = 0.6188) (fig. 2a, b). two yearlings, 1 male and 1 female, were diagnosed with mws and the relative risk did not differ between the sexes 2 = 0.0037, 1 df, p = 0.9516, fig. 2a, b). among older animals, adults and seniors together, bulls appeared to be more prone to 2 = 4.253, 1 df, p = 0.0392, fig. 2a, b) but the opposite pattern was seen for mws; i.e., the relative risk was higher for cows 2 = 6.3486, 1 df, p = 0.0117, fig. 2a, b). the frequency of both ela and mws 2 = 33.1174, 3 df, p 2 = 13.699, 3 df, p = 0.0033, ela and mws, respectively). moose with ela were over-represented in springtime while cases with mws were most prevalent in winter (fig. 3). predated carcasses were found most frequently in spring and in areas along the norwegian border. the occurrence of animals with ela and mws differed geographically. the relative risk of ela was greatest in northern sweden (fig. 4a) while the relative risk of mws was 0.00 0.20 0.40 0.60 0 4 8 12 16 20 age (years) p ro p o rt io n o f m o o se female male fig. 1. age distribution of examined moose (female n = 426; males n = 208). female 0.00 0.20 0.40 0.60 0.80 1.00 c a lv e s y e a rl in g s a d u lts s e n io rs p ro p o rt io n s o f d ia g n o se s others mws ela male 0.00 0.20 0.40 0.60 0.80 1.00 c a lv e s y e a rl in g s a d u lts s e n io rs p ro p o rt io n s o f d ia g n o se s others mws ela fig. 2. proportion of moose diagnosed as ela, mws, and ‘others’ based on age class and sex; (a) female (calves n = 125, yearlings n = 27, adults n = 220, and seniors n = 47) and (b) males (calves n = 96, yearlings n = 34, adults n = 69, and seniors n = 3). (a) (b) alces vol. 41, 2005 stéen et al. diseases in moose 41 highest in the south (fig. 4b). condition a disease was not always manifested by diminished body condition. emaciation, poor, and normal condition were almost equally represented among the cases (33%, 31%, and 36%, respectively). condition was related to season; poor/emaciated being most prevalent in spring, followed by winter, summer, and autumn. excluding animals culled in the fall, this pattern did not change (fig. 5). the poor/emaciated categories were over-represented in moose showing ela and mws 2 = 28.0245, 1 df, p 2 = 10.237, 1 df, p < 0.05, respectively), but there was a tendency for inf to be under-represented 2 = 3.7057, 1 df, p = 0.0542). about 50% of animals with inf were in normal body condition, compared to other disease diagnoses where the corresponding value was < 40% (fig. 6). moose with tumors (neo), as well as predated animals, were in all categories of body condition. discussion diagnoses the proportion of different diagnoses (i.e., relative risks) varied between age-class and sex. mws is more common among older animals and more common among cows than bulls. for ela, the opposite pattern was observed. it appears that both age and sex can explain some of the variance in susceptibility such conclusions are equivocal, as we do not know the size and structure of the population from which the dead moose came, which is essential to estimate the absolute risk of death. therefore, one can only estimate the relative risk of death. with reservations about discrepancies between relative and absolute risks, this study indicates that adult bulls were more prone to ela than adult cows, but no differences were found between male and female calves and yearlings. infections with elaphostrongylus spp. normally occur in summer and fall, with (1986) studied ela in reindeer and concluded that male calves belonging to dominant mothers are more heavily infected with e. rangiferi than females. the largest calves eat more and therefore experience a higher risk of ingesting gastropods with e. rangiferi. stuve (1986) found a higher prevalence of ela in male than female moose calves, also suggesting that males were more likely to be infected. moose calf weights were dependent on summer browse in the cows’ home range, the quality of which is related to the cows’ status. for older animals, stuve (1986) attributed the difference in infection between the sexes to physiological changes during the rut in accordance with age is related to ela, with yearling and calves being most frequently infected. earlier studies have shown that moose shed most e. alces lowed by a sharp drop in larval shedding and reduced level of adult worms in older animals (stuve 1986, stéen 1991, olsson et al.1995, stéen et al. 1997). our results show that emaciation is associated with ela. this could of course be a spurious correlation or a cause/effect of 0.00 0.20 0.40 0.60 0.80 1.00 s p ri n g s u m m e r a u tu m n w in te r p ro p o rt io n s o f d ia g n o s e s others mws ela fig. 3. proportion of moose diagnosed as ela, mws, and ‘others’ based on season (spring n = 200, summer n = 76, autumn n = 294, and winter n = 154). diseases in moose – stéen et al. alces vol. 41, 2005 42 ela. that emaciation might be an effect is supported by the fact that e. alces can cause a nervous disorder, with lack of co-ordinarehbinder 1986, stéen et al. 1989, stéen and roepstorff 1990). stuve (1986) found that general condition and he found that the difference in carcass weights between infected and non-infected animals increased with age. conversely, moose experimentally infected with e. alces retained their normal weight when fed ad libitum (stéen et al. 1998a). inf differed from other diseases in being positively related to condition. animals with inf probably die acutely before they loose condition or become emaciated, which is in contrast to animals with ela and mws. health status the disease pattern, including deaths caused by wild predators seen in our sample differs from that seen elsewhere (lankester 1987, guilazov 1998, van ballenberghe and ballard 1998), suggesting that moose populations in sweden might be different from other populations. for example, densities and human harvest rates are higher in sweden than in russia and north america, but predation is lower. we believe that the disease patterns observed are masked worldwide by predation. the scenario of the swedish moose loss, with the exception of hunting, will probably change over time and become more similar to the causes of deaths (e.g., predation) observed in other countries. another difference between the swedish fig. 4. proportion of moose examined between 1985 and 1989 with signs of (a) ela and (b) mws for each swedish county. (a) (b) alces vol. 41, 2005 stéen et al. diseases in moose 43 and north american moose populations is the origin of the diseases (borg 1956, nilsson 1971, borg and nilsson 1985, borg 1987, lankester 1987, lankester and samuel 1998, stéen et al. 1998b). an interesting and notable observation is that moose in north america have few eurasian parasites but have acquired new parasites and diseases from indigenous wild ungulates and livestock (lankester 1987, lankester and samuel 1998). the diseases observed in sweden are, as far as we can ted from livestock or deer, with the exception of malignant catarrhal fever (warsame and stéen 1989). most of the diseases in american moose have not been diagnosed in swedish moose. the viroses epizootic haemorrhagic disease, bluetongue, western equine and st. louis encephalitis, norway virus, california encephalitis virus, and contagious ecthyma have not been diagnosed in swedish moose or bovine rhinotracheitis are, however, known in swedish livestock (moreno-lopéz 1979, sjv 1994) and in reindeer (rockborn et al. 1990) but not moose. further, the bacterial and parasitic diseases in american moose (leptospirosis, brucellosis, necrobacillosis, toxoplasma gondii, entamoeba bovis, paramphistomum spp., fascioloides magna, taenia ovis, t. krabbei, echinoccocus granulosus, thysanosoma actinioides, orthostrongylus macrotis, parelaphostrongylus tenuis, elaeophora schneideri, onchocerca cervipedis, setaria yehi, , dermacentor albipictus, cephenemyia jellisoni, c. phobifera, and haematobosca alcis), have not been found nor reported for swedish moose (nilsson 1971, stéen et al.1998b). despite all the diseases and parasites enumerated, lankester (1987) and lankester and samuel (1998) proclaim american moose to be generally healthy. they explain this status of health partly by the fact that american moose occur at a low densities (0.1-0.6/km2), which reduces the transmission rate of parasites and diseases. also van ballenberghe and ballard (1998) state that moose host a variety of diseases and parasites that are seldom a major limiting factor for population growth. on the other hand, wobeser (1994; 3) declared: “although most infectious agents do not result in obvious disease, the host must pay a price for harboring parasites that live, grow, and reproduce at expense of the host. interactions between parasites and other stress factors can in wild animals are often considered only in terms of death or obvious physical disability, parameters. in other words, the effect of diseases on wild populations may be much greater than is evident by simply counting the dead or 0.00 0.20 0.40 0.60 0.80 1.00 s p ri n g s u m m e r a u tu m n w in te r p ro p o rt io n s o f m o o se emaciated poor normal fig. 5. proportion of moose in normal condition, poor condition, or emaciated split by season (spring n=183, summer n=73, autumn n=166, and winter n=131). moose culled during regular sports hunting excluded. 0.00 0.20 0.40 0.60 0.80 1.00 e l a m w s in f t o ta l p ro o ri tio n s o f m o o se emaciated poor normal fig. 6. proportions of moose in normal condition, poor condition, or emaciated split by diagnosis (ela n=126, mws n=81, inf n=28, and ‘total’ n=642). diseases in moose – stéen et al. alces vol. 41, 2005 44 maimed. the swedish moose population is dense, with up to 1.0-1.5 moose/km2, which may increase the risk of disease and parasite transmission as discussed by lankester and samuel (1998). it is unknown how our data relate to density, but for mws there are indications that density dependence might be the ultimate reason for its appearance (broman et al. 2002a). does the large number of diseased moose seen in sweden during the latest 2 decades indicate anything about the health status of swedish moose? the opportunity to see and report abnormal moose is greater in the intensely managed forests of sweden than in the wilderness of north america and russia. from 1985 to 1989, approximately 200 moose per year were necropsied and diagnosed in sweden, compared to a total of 420 between 1947 and 1982 (borg 1987, 1991). currently, up to 100 moose per year are examined at the national veterinary institute, uppsala, sweden (mörner 2001). if the necropsies ever, the moose population increased sharply in the 1980s (cederlund and markgren 1987, cederlund and bergström 1996), making it thus, increased observations do not necessarily indicate a higher absolute mortality risk. broman et al. (2002b) estimated natural death < 4% for adults in the area where the highest incidences of natural moose mortalities were recorded (community of mark) between 1991 and 1998. there were no wild predators in mark during this period implying that natural mortalities were synonymous with mortality caused by disease. without predators it appears that the mortality risk due to disease has been quite low in the 1980s and 1990s, but the relative risk of ela and mws was high. our description of diseases and natural mortalities differs from that of guilazov (1998) who described predation as the primary mortality factor in moose of northern russia. based on our results, the risk of being killed by predators was low in the late 1980s. only 25 of the moose in the entire sample were killed by wolves suggesting that predation was not common. future scenario currently, winter populations of moose and roe deer in sweden are approximately 250,000 and 1.5 million, respectively (stéen et al. 1998b). predators have increased substantially in recent decades. wolves were protected in sweden when estimated numapproximately 67-81 (aronsson et al. 2001). in the 1930s, 130 brown bears were known, and in 60 years (1996) they had increased to 800-1,300 (sou 1999). it is realistic to believe that, despite future increase in numbers and range expansion, harvesting of predators will remain banned or be highly regulated by the from predation will no doubt result. interactions between risk of disease and predation may result in compensatory rather than additive death. for moose, the risk of being killed by wolves or bears depends on age (e.g., van ballenberghe 1987, ballard and van ballenberghe 1998, sou 1999, wikenros 2001). ballard and van ballenberghe (1998) showed that calves and old cows are the primary target of wolves. predation by bears was the most frequent cause of early calf mortality (franzmann and schwartz 1986; boertje et al. 1987, 1988). franzmann and schwartz (1986) estimated bear density in alaska to be 19.0/100 km2. this compares to a desirable density of at least 0.7/100 km2 anticipated outside the swedish reindeer husbandry area (61°n to 69°n) (sou 1999). swedish studies of bear predation on moose calves indicate a 20-25% loss, while bears accounted for 0.5-1.5% of adult mortalities (sou 1999). studies in both sweden and north america indicate that bear predation on calves is additive, at approxialces vol. 41, 2005 stéen et al. diseases in moose 45 mately 3 calves per bear/year (sou 1999). sou (1999) reports that natural mortality of adults was approximately 5% in an area with no predation and that the additive loss of adults by bear predation was 0.5-1.5% per bear/year. while predation on calves is mostly likely additive, van ballenberghe (1987) stated that predation on adults was mostly compensatory, with the various mortality factors tending to substitute more for each other. ballard and van ballenberghe (1998) cite mech et al. (1995) whose results show that wolf-predated calves and adults during winter have low marrow fat values, indicating poor condition. also, peterson (1977) reported that wolves from isle royale prey on heavily parasitized, diseased, or otherwise inferior moose. this information suggests that in the future, the weak, vulnerable, sick, and old moose will be preyed upon before dying from a disease. we suggest that the panorama of moose diseases seen in the future will differ from that seen during the 1980s and 1990s by being less visible due to increasing predation. references aronsson, å., p. wabakken, h. sand, o. k. steinset, and i. kojola. 2001. varg i skandinavien. statusrapport, för vintern 2000/2001. (the wolf in scandinavia: status report of the 2000/2001 winter). (in swedish with english summary). ballard, w. b., and v. van ballenberghe. 1998. predator/prey relationships. pages 247-273 in a. w. franzmannn and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. boertje, r. d., w. c. gasaway, d. v. granagaard, and d. g. kelleyhouse. 1988. predation on moose and caribou radiocollared grizzly bears in east-central alaska. canadian journal of zoology 66:492-499. _____, _____, _____, _____, and r. o. stephenson. 1987. factors limiting moose population growth in subunit 20 e. federal aid in wildlife restoration report. alaska department of fish and game, juneau, alaska, usa. borg, k. 1956. vilt och viltsjukdomar. pages 314-342 in i. lauritzon, editor. lantbrukets djurbok iii. strömberg, stockholm, sweden. (in swedish). _____. 1975. viltsjukdomar. lts förlag, stockholm, sweden. (in swedish). _____. 1987. a review of wildlife diseases from scandinavia. journal of wildlife diseases 23:527-533. _____. 1991. rådjur. dödsorsaker, miljöpåverkan och rättsmedicin. swedish environmental protection agency, naturvårdsverkets rapport 3921. (in swedish). ____, and p. o. nilsson. 1985. silbenstumörer hos älg och rådjur (ethmoid tumors in elk and roe deer). nordique veterinary medecine 37:145-160. (in swedish with english summary). broman, e., k. wallin, m. stéen, and g. cederlund. 2002a. a wasting syndrome in swedish moose (alces alces): background and current hypotheses. ambio 31:409-416. _____, _____, _____, and_____. 2002b. “mass” deaths of moose alces alces in southern sweden: population level characterization. wildlife biology 8:209218. cederlund, g., and r. bergström. 1996. trends in the moose-forest system in fennoscandia, with special reference to sweden. pages 265-281 in r.m. de graaf and r. i. miller, editors. conservation of faunal diversity in forested landscapes. _____, and g. markgren. 1987. the development of the swedish moose population, 1970-1983. viltrevy, swedish wildlife research supplement 1:55-61. diseases in moose – stéen et al. alces vol. 41, 2005 46 feinstein, r., c. rehbinder, e. rivera, t. nikkilä, and m. stéen. 1987. intracytoplasmic inclusion bodies associated with vesicular, ulcerative and necrotizing lesions of the digestive mucosa of a roe deer (capreolus capreolus l.) and a moose (alces alces l.). acta veterinaria scandinavica 28:197-200. frank, a. 1998. 'mysterious' moose disease in sweden: similarities to copper deficiency and/or molybdenosis in cattle and sheep. biochemical background of clinical signs and organ lesions. science of the total environment 209:17-26. _____, m. anke, and r. danielsson. 2000a. experimental copper and chromium deficiency and additional molybdenum supplementation in goats. i. feed consumption and weight development. science of the total environment 249:133-142. _____, r. danielsson, and b. jones. 2000b. the 'mysterious' disease in swedish moose. concentrations of trace elements in liver and kidneys and clinical chemistry. comparisons with experimental molybdenosis and copper deficiency in the goat. science of the total environment 249:107-122. _____, _____, and _____. 2000c. experimental copper and chromium deficiency and additional molybdenum supplementation in goats. ii. concentrations of trace and minor elements in liver, kidneys and try. science of the total environment 249:143-170. _____, v. galgan, and l. r. petersson. 1994. secondary copper deficiency, chromium deficiency and trace element imbalance in moose (alces alces l.): effects of an anthropogenic activity. ambio 23:315317. _____, d. s. sell, r. danielsson, j. f. fogarty, and v. m. monnier. 2000d. a syndrome of molybdenosis, copper deficiency, and type 2 diabetes in the moose population of south-west sweden. science of the total environment 249:123-131. franzmann, a. w., and c. c. schwartz. 1986. black bear predation on moose calves in highly productive versus marginal moose habitat on the kenai peninsula, alaska. alces 22:139-153. gajadhar, a., t. steeves-gurnsey, j. kendall, m. lankester, and m. stéen. 2000. differentiation of dorsal-spined elaphostrongyline larvae by polymerase chain reaction amplification of its-2 of rdna. journal of wildlife diseases 36:713-722. guilazov, a. s. 1998. causes of reindeer (rangifer tarandus) and moose (alces alces) mortality in the lapland reserve and its surroundings. alces 34:319-327. halvorsen, o. 1986. epidemiology of reindeer parasites. parasitology today 2:334-339. lankester, m. 1987. pests, parasites and diseases of moose (alces alces) in north america. swedish wildlife research supplement 1:461-490. _____, i-m. olsson, m. stéen, and a. a. gajadhar. 1998. extra-mammalian larval stages of elaphostrongylus alces (nematoda: protostrongylidae), a parasite of moose (alces alces) in fennoscandia. canadian journal of zoology 76:33-38. _____, and w. samuel. 1998. pests, parasites and diseases. pages 479-517 in a. w. franzmannn and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. mech, l. d., t. j. meier, j. w. burch, and l. g. adams. 1995. patterns of prey selection by wolves in denali national park, alaska. in d. seip, editors. ecology and conservation of wolves in a changing world. proceedings of the 2nd international wolf symposium. canadian circumpolar institute, university of alberta, edmonton, alces vol. 41, 2005 stéen et al. diseases in moose 47 alberta, canada. merck veterinary manual book of diagnosis and therapy for the veterinarian. fifth edition. merck & co., inc., rahway, new jersey, usa. merza, m., e. larsson, m. stéen, and b. morein. 1994. association of a retrovirus with a wasting condition in the swedish moose. virology 202:956-961. moreno-lopéz, j. 1979. a serosurvey of viruses during outbreaks of acute respiratory and/or enteric disease in swedish b 26:634-640. mörner, t. 2001. var tog älvsborgssjukan vägen? (where did “the älvsborg disease” disappear?) svensk jakt 11:64-65. (in swedish). nilsson, o. 1971. the interrelationship of endoparasites in wild cervids (capreolus capreolus l. and alces alces l.) and domestic ruminants in sweden. acta veterinaria scandinavica 12:36-68. olsson, i.-m. 2001. elaphostrongylus alces – transmission, larval morphology and tissue migration. m.sc. thesis, swedish university of agricultural sciences, uppsala, sweden. _____, r. bergström, m. stéen, and f. sandgren. 1995. a study of elaphostrongylus alces in an island moose population with low calf body weights. alces 31:61-75. _____, m. w. lankester, a. a. gajadhar, and m. stéen. 1998. tissue migration of elaphostrongylus spp. in guinea pigs (cavia porcellus). journal of parasitology 84:968-975. _____, m. stéen, and h. mann. 1993. gastropod hosts of elaphostrongylus spp. (protostrongylidae, nematoda). rangifer 13:53-55. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. u.s. national park service, science monograph series 11. rehbinder, c., k. gimeno, k. belak, s. belak, m. steen, m. rivera, and t. nikkilä. 1991. a bovine viral diarrhoea/mucosal disease-like syndrome in moose (alces alces): investigations on the central nervous system. veterinary record 129:552-554. reimers, e., and o. nordby. 1968. relationship between age and tooth cementum layers in norwegian reindeer. journal of wildlife management 32:957-961. rockborn, g., c. rehbinder, b. klingeborn, m. leffler, k. klintevall, t. nikkilä, a. landén, and m. nordkvist. 1990. the demonstration of a herpesvirus related to bovine herpesvirus 1, in reindeer with ulcerative and necrotizing lesions of the upper alimentary tract and nose. rangifer 10:373-384. saether, b.-e., and m. heim. 1993. ecological correlates of individual variation in age at maturity in female moose (alces alces); the effects of environmental variability. journal of animal ecology 62:482-489. (sjv) swedish board of agriculture. 1994. viktiga smittsamma sjukdomar. swedish board of agriculture, jönköping, sweden. (in swedish). skunke, f. 1949. älgen. studier, jakt och vård. stockholm, sweden. (in swedish). (sou) statens offentliga utredningar. 1999. sammanhållen rovdjurspolitik. slutbetänkande av rovdjursutredningen. statens offentliga utredningar (sou 1999:146) and appendix, the swedish government, stockholm, sweden. (in swedish). stéen, m. 1991. elaphostrongylosis. a clinical, pathological, and taxonomical study with special emphasis on the infection in moose. ph.d. thesis, swedish university of agricultural sciences, uppsala, sweden. _____, c. g. m. blackmore, and a. skorping. 1997. cross infection of moose (alces alces) and reindeer (rangifer tarandus) with elaphostrongylus alces and diseases in moose – stéen et al. alces vol. 41, 2005 48 elaphostrongylus rangiferi (nematoda, protostrongylidae): effects on parasite morphology and prepatent period. veterinary parasitology 71:27-38. _____, a. g. chaubaud, and c. rehbinder. 1989. species of the genus elaphostrongylus parasite of swedish cervidae. a description of e. alces n.sp. annales de 64:134-142. _____, and r. diaz. 1988. studies of a bovine virus diarrhea/mucosal disease like syndrome in swedish moose (alces alces l.). m.sc thesis, swedish university of agricultural sciences, uppsala, sweden. _____, _____, and w. e. faber. 1993. an erosive/ulcerative alimentary disease of undetermined etiology in swedish moose (alces alces l.). rangifer13:149-156. _____, w. e. faber, and a. oksanen. 1998b. disease and genetical investigations of fennoscandian cervids – a review. alces 34:287-310. _____, and c. johansson. 1990. elaphostrongylus spp. from scandinavian cervidae – a scanning electron microscope study (sem). rangifer 1:39-46. _____, s. persson, and l. hajdu. 1994. protostrongylidae in cervidae and ovibos muscatus: a clustering based on isoelectric focusing on nematode body proteins. applied parasitology 35:193-206. _____, and c. rehbinder. 1986. nervous tissue lesions caused by elaphostrongylus in wild swedish moose. acta veterinaria scandinavica 27:336-342. _____, and l. roepstorff. 1990. neurological disorder in two moose calves (alces alces l.) naturally infected with elaphostrongylus alces. rangifer 3:399-406. _____, i. warsame, and a. skorping. 1998a. experimental infection of reindeer, sheep and goats with elaphostrongylus spp. (nematoda, protostrongylidae) from moose and reindeer. rangifer 18:73-80. stuve, g. 1986. the prevalence of elaphostrongylus cervi infection in moose (alces alces) in southern norway. acta veterinaria scandinavica 27:397-409. van ballenberghe, v. 1987. effects of predation on moose numbers: a review of recent north american studies. swedish wildlife research supplement 1:431-460. _____, and w. b. ballard. 1998. population dynamics. pages 223-246 in a. w. franzmannn and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. warsame, i. y., and m. stéen. 1989. malignant catarrhal fever in swedish moose (alces alces l). rangifer 9:51-57. wikenros, c. 2001. wolf winter predation on moose and roe deer in relation to pack size. department of conservation biology, grimsö wildlife research station, riddarhyttan, sweden, no. 75. wobeser, g. 1981. diseases of wild waterfowl. plenum publishing corporation, _____. 1994. investigation and management of disease in wild animals. plenum << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 4009(133-149).pdf alces vol. 42, 2006 tyers – yellowstone population history 133 moose population history on the northern yellowstone winter range daniel b. tyers u.s. forest service, gardiner ranger district, p.o. box 5, gardiner, mt 59030, usa abstract: moose probably colonized the northern yellowstone winter range (nywr) in the latter half of the 19th century. euro-american settlement of the nywr occurred at roughly the same time. the northern boundary of yellowstone national park, authorized in 1945 in response to perceived damage by moose to willow stands, evidently reduced the moose population quickly and maintained it at ecosystem and impacted old growth forest important for moose survival during winter. the moose population associated with the nywr declined by 75% or more and has shown no sign of recovery by 2002. several techniques for assessing population trend for moose on the nywr were tested. given the problems associated with monitoring a species at low densities with a dispersed social organization and occupying habitats where visibility is limited, aerial population censuses were not useful. a horseback trail survey, a road survey, and counts of moose in early winter or late spring in larger willow stands had greater potential as indices to moose population changes. alces vol. 42: 133-149 (2006) key words: monitoring, moose, northern yellowstone winter range, population history, yellowstone the northern yellowstone winter range (nywr) (fig. 1) supports over half the wintering ungulates that utilize yellowstone national park (ynp) during summer (national academy of sciences 2002). while elk (cervus elaphus) and bison (bison bison) constitute more than 80% of the ungulate biomass on the nywr during winter (national academy of sciences 2002), this winter range is essential to several less common ungulates, including moose (alces alces), bighorn sheep (ovis canadensis), mule deer (odocoileus hemionus), and pronghorn antelope (antilocapra americana) (yellowstone national park 1997). in 1985, a study (tyers 2003) was initiated to identify moose habitat needs and population status. this paper summarizes information collected on the history of moose on the nywr and gives recommendations for monitoring the nywr moose population. accurate assessment of ungulate population dynamics and factors that regulate populations is essential to sound population management (gasaway et al. 1986, van ballenberghe and ballard 1998). obtaining reliable demographic information on any (mccullough 1984, saether 1987), but moose monitor because they are the least social north american deer and frequently occupy habitats with poor observability (schladweiler 1973, houston 1974). moose population size is typically assessed in 3 waystotal area counts, sample estimates, and indices (timmermann and buss 1998). timmermann and buss (1998) advocated multiple information sources to assess population status. i used historic documents to trace the history of moose populayellowstone population history – tyers alces vol. 42, 2006 134 tions on the nywr and multiple population monitoring methods, including aerial surveys, horseback surveys, road surveys, and spatially restricted counts, to determine if vegetation the yellowstone ecosystem in 1988 precipitated changes in moose population size. the results of my population monitoring efforts during 1985 – 2001 allowed me to evaluate ing moose population indices and to identify reasonable techniques for monitoring future population trends. study area the nywr includes parts of ynp, the southern third of the gardiner ranger district, and state lands (fig. 1). the boundary of the nywr is based on winter distribution of elk (houston 1982). during this study, elk were the dominant ungulate species (10,000 – 25,000), but mule deer (2,000 – 3,000), bighorn sheep (100 – 200), bison (500 – 1,000), and pronghorn antelope (100 – 300) also occupied the nywr. moose numbers were unknown, but they wintered throughout the study area in scattered areas of suitable habitat, usually at higher elevations than elk. vegetation on the nywr varies from low elevation (< 2,000 m) sage (artemisia spp.) steppe to high elevation (3,000 m) coniferous forests. willow (salix spp.) stands occur along streams and in wet areas within forests. fig. 1. map of the northern yellowstone winter range study area showing prominent features and sampling areas. bcsu=bear creek study unit; ypsu=yellowstone park study unit; scsu=slough creek study unit; sbsu=soda butte study unit. alces vol. 42, 2006 tyers – yellowstone population history 135 lodgepole pine (pinus contorta), engelmann spruce (picea engelmannii (abies lasiocarpa pseudostuga menzieisii), and whitebark pine (p. albicaulus) are the most common coniferous species in the mature conifer forest present in the nywr in 1988, thus converting about 30% of mature forest to early seral stages (tyers 2003). methods historical documents to the study area. documents not considered by other authors that provided an historical special interest. population monitoring techniques horseback transect index — in 1947, 1948, and 1949, montana fish and game biologist joe gaab looked for moose each september in the absaroka primitive area (now the absaroka beartooth wilderness) on about 177 km of trail. from 1985 to 2001, other observers repeated his route through the hellroaring, buffalo fork, and slough creek drainages 34 times between july and late ocgaab rode primarily to look for moose. during 1985 through 2001, observers conducted other tasks along the routes (trail maintenance, hunter compliance checks, and moose observed. gaab and more recent observers recorded age (calf or > 1 year of age) and gender (for moose > 1 year of age) of all moose sighted. from 1985 to 2001, the days spent covering the route ranged from 5 to 32, and trails were not traveled in any particular sequence. in both periods (1947-1949 and 1985-2001) observations were restricted to daylight hours and sightings were reported as number of moose seen per day per observer group. observer group size varied from 1 to 6. road transect index — moose sightings along the 89-km stretch of road from gardiner to cooke city, the only road in ynp maintained for wheeled vehicles year-round, and abundance. each trip was considered one sample regardless of the direction of travel. no attempt was made to standardize time of day, but at least 4 trips per month were completed in all months. data collected between january 1987 – december 1992 and january 1995 – december 1997 were used to determine if there were differences in the number of moose seen seasonally and if moose numbers seen along the road differed before and after the to determine if changes between pre across the nywr, the road was divided into 5 sections. each section consisted of a road segment that traversed similar vegetation to mammoth (8.0 km), included the gardner river canyon. topography was broken and the surrounding vegetation was arid grasslands and dry sagebrush unaffected by the 1988 the road (1,585 m). the second section was from mammoth to tower junction (29.1 km). topography and vegetation were diverse. vegetation included stretches of stunted willow and aspen, and 1 in 1988 raced across this area leaving a mosaic burn pattern in which many old tree stands were converted to young seral stages. junction to round prairie (30.9 km). this stretch was mostly a broad open valley with the lamar river. because most of this section yellowstone population history – tyers alces vol. 42, 2006 136 much change in vegetative structure. the 1988 (13.2 km) through mature lodgepole pine. city (8.0 km), followed soda butte creek through the largest willow stands along the transect. the rest of the vegetation was the area north of the road burned in 1988. cooke city was the highest point along the road (2,134 m). — barwhere moose were frequently observed during aerial elk counts on the nywr during 19681970. two of the largest, frenchy’s meadow in the slough creek drainage and the willow stands along soda butte creek outside the eastern boundary of ynp west of cooke city (fig. 1), were selected for systematic sampling month year-round from 1987 through 1990. three radio-collared moose in the frenchy’s meadow area and 4 in the cooke city area were available for use as a check on survey to the willow stands were counted on each animals were located to determine what proportion of radio-marked animals available in the drainage were in the willow stand. two indices of abundance were calculated (uncollared and radio-collared); and (2) the percent of available radio-collared moose seen. there were too few radio-collared animals to make valid estimates of total moose numbers in willow stands using mark-recapture methodology (lancia et al. 1994), but the proportion of radio-collared animals seen did provide an estimate of the proportion of animals in the vicinity of the willow stands that were visible. data from moose counts in willow stands were used to determine if moose numbers in favored willow stands varied among months or among years (including years before and daily willow stand observations — beited in number and were restricted to morning hours, ground observations were used to better delineate the time of year and time of day that moose were most easily observed in willow stands. from april 1996 through june 1997, moose were counted and numbers recorded in the willow stand between silver gate and cooke city. observations were limited to a standardized segment of the stand. these data wing aircraft were optimally timed (diurnally and seasonally) and provided another potential in number of daylight hours through the year and occasional gaps in data collection, data were standardized as number of moose seen per number of observation attempts. data were used to determine if moose were more — data collected from ed that moose were most observable around pilots were told to follow transects (0.4-km al. (1986). this approach was abandoned on lowing transects due to wind and topography, limited visibility due to dense forest canopy, and frustration by observers along unproduclimited to areas where moose were most likely northern boundary of ynp (inside and outside ynp). stands were covered carefully on all alces vol. 42, 2006 tyers – yellowstone population history 137 about 97 – 113 kph at 61 – 152 m above the ground, depending on obstacles. statistical tests when assumptions on sample distribution were met (zar 1999), the anova module in the statistica software package (statsoft 1995) was used to test for temporal (hourly, monthly, seasonal, and/or annual) differences (p < 0.05) willow stands. to test assumptions, data were dressed homogeneity and shapiro-wilks’ w test assessed normality. questionable data terplots of means versus standard deviations. data that did not meet the assumptions of anova were tested using a kruskal-wallis test by ranks, for multiple treatments involving nonparametric data. because the f-statistic is considered robust, the central limit theorem was invoked and anova was used if sample size was large (n > 100). a post hoc test, student-newman-keuls, was used to look for differences among treatments when p, f, z, and h statistics are reported in this paper. in analysis of road transect data, a 0/1 (present/absent) measurement scale was used moose as an observer drove the road between gardiner and cooke city. each trip (gardiner – cooke city or cooke city – gardiner) was considered an independent trial resulting in a set of 1,020 observations. seasonal likelihoods were determined from analysis of 2-month periods (november/december, january/february, march/april, may/june, july/august, and september/october). an estimate of likelihood of sighting a moose each year during 1987 – 1992 and 1995 – 1997 was calculated by dividing trips with moose sightings by the total trips in a calendar year. to determine if differences between preand consistent over the entire transect route, individual road sections were compared over time using a z-test (p < 0.05). results historical documents the earliest reports on moose located in ments of population status in the northern the early 1900s (tyers 1981). mcdowell and moy (1942) reported that “old timers” regarded moose as a rarity in drainages along the north boundary of ynp between 1907 and 1915 while rush (1942) reported that moose were considered “fairly common” by 1913 in the same area. in 1920, stevenson (1920) noted that there were 13 moose wintering in 2 drainages currently designated as prime moose winter habitat in the nywr (12 in hellroaring and 1 in buffalo fork) and that the habitat would support more wintering moose than were present. in 1921, the u.s. forest service began snowshoe surveys conducted in december – april) to deter poaching and monitor wildlife near the northern boundary of ynp. crane (1922) counted 16 moose during the winter of 1921 – 1922. uhlhorn (1923) estimated 25 moose for the winter of 1922 – 1923. johnson's (1925) report for 1924 – 1925 stated he could account for 65 moose. he noted that calf survival was high and believed the population was increasing. by 1936, u.s. forest service reports (usda 1936, mcdowell and moy 1942) of willow stands associated with the nywr and with the moose population that used them. these reports noted that willow condition was positively related to elevation and negatively related to access by elk and moose. the moose population wintering along the northern ynp yellowstone population history – tyers alces vol. 42, 2006 138 boundary in 1935 – 1936 was estimated at 193 (54 in the hellroaring, 80 in the buffalo fork, and 60 in the slough creek drainage). over-winter utilization of willow in stands used by moose was estimated at 90%, and 75% of the willows in moose winter range were described as recently dead. montana fish and game department personnel surveyed drainages north of ynp from summer through autumn 1942 (mcdowell and moy 1942). from june through october, they covered 341 miles (549 km) on foot and 1,341 miles (2,158 km) on horseback. they reported 194 unduplicated moose and suginto the area from ynp and that the population was increasing. they noted that >50% of willow plants were severely damaged in some areas where ungulates wintered while little or no degradation in willow stands was observed at elevations above ungulate winter range. they called for a controlled harvest of moose to prevent further damage to willow. cooney et al. (1943) re-surveyed part of the area covered by mcdowell and moy (1942) in 1943 and reported an increase in moose numbers over that reported in the same area during the 1942 survey. in 1942 and 1944, montana fish and game department employees conducted december or january moose surveys in the nywr north of the ynp boundary (parsell and mcdowell 1942, mcdowell and page 1944). they found in and around frenchy's meadow (slough creek drainage) but were surprised at the large numbers of moose occupying forested slopes dowell (1942) estimated that elk and moose had utilized 90% of current willow growth by december 1942 and reported moose foraging on alder (alnus incana), engelmann spruce, the 1945 montana state legislature passed substitute bill no. 41, which authorized the montana fish and game commission to “remove and dispose of moose increasing in numbers and damaging property by the limited license method” (montana fish and autumn 1945, mcdowell (1946) reported that 40 permits were issued and hunters killed 35 moose across a hunting area that included most of the nywr north of ynp (including the hellroaring, buffalo fork, and slough creek drainages, and the cooke city area). reports of the impacts of hunting on moose varied. mcdowell (1946) noted that a forest service employee reported 18 moose on a winter survey following the 1945 season, where cooney et al. (1943) had counted 31 in winter 1943. mcdowell (1946) believed this decreased count was most likely due to moose moving from one drainage (hellroaring) to another (slough creek) due to declining willow production in hellroaring. in a report submitted by mcdowell and smart (1945) describing a 1945 winter survey, the authors noted that 90% of the current year’s willow production in some stands was utilized despite the harvest. in 1946, only 20 of 30 permits and guides concerned about declining moose numbers, permit numbers were further reduced in 1947 (couey 1947). in 1947, 1948, and 1949 montana fish and game biologist joe gaab conducted horseback moose abundance (gaab 1948, 1949, 1950). he traveled about 110 miles (177 km) of trail during september to count moose, using the same trails each year. he recorded 106, 71, and 30 independent moose sightings, respectively. in his opinion, the moose population was in a decline, which he attributed, in part, to a continued deterioration of willow stands (gaab 1948, 1949, 1950). when interviewed of quota hunting, hunters shot “many more” moose than permits allowed, although 50 years later he could recall anecdotes but not actual alces vol. 42, 2006 tyers – yellowstone population history 139 gaab, montana fish and game department, personal communication). agency reports on moose population surveys and hunting seasons during most of the 1950s and 1960s were scarce. in 1963, montana fish and game regulations listed a moose harvest quota in districts along the northern boundary of ynp of 45 with no restrictions on age or gender. a 1964 wildlife management plan for the gardiner ranger district, gallatin national forest (kehrberg 1964), noted that addressing the “moose problem” (declining moose populations and deteriorating willow stands) in the hellroaring-slough creek area was a management priority. a different perspective on moose population/habitat trends from the 1920s to the 1960s was provided by tony bliss, co-owner of a small private parcel in slough creek near the large willow stand in frenchy's meadow. he summarized his observations of moose population trends (kehrberg 1964: 9-10) as follows: “1926 to 1935 lots of tall willow and few moose, elk and moose fed hay by yellowstone park in lower slough creek; 1935 to 1945 more moose, still lots of willow, feeding ended about 1936; 1941 to 1945 away at war; 1955 to 1962 fewer and fewer moose stable values for indices of hunter effort (such as hunting days per moose harvested) suggested that the moose population remained relatively stable through the 1970s and early 1980s (t. lemke, montana fish, wildlife and parks, personal communication). when this hunting districts along the northern boundary of ynp was 55 with no restriction on age or gender. quotas were reduced and restricin the yellowstone area in 1988. in 1990, the montana department of fish, wildlife and parks issued 42 harvest permits (23 antlered and 19 antlerless) (t. lemke, montana fish, wildlife and parks, personal communication). the quota was reduced to 21 (13 antlered, 8 antlerless) in 1991, in response to population declines observed during this study. permits were reduced to 13 in 1996 (all antlered). population indices horseback transect index — gaab's (1948, 1949, 1950) 177 km transect in the absaroka beartooth wilderness was re-run 34 during the years 1985 – 2001 (fig. 2). the number of moose observed per day declined between 1947 and 2001. only values in 1988 and 1989 approached sighting rates reported by gaab. the total number of moose seen on surveys also declined. gaab’s counts averaged 69.0 (sd = 38.0, n = 3). total counts in the 0 1 2 3 4 1947 1949 1986 1988 1990 1992 1994 1996 1998 2000 year m oo se /d ay fig. 2. average number of moose seen per party per day during horseback surveys in the yellowstone ecosystem 1947 – 1949, 1985 – 1992, and 1995 – 2001. in years with >1 survey (1992, 1995 – 2001), values are the mean of multiple surveys. yellowstone population history – tyers alces vol. 42, 2006 140 15.0 (sd = 4.4, n late 1980s averaged 44.5 (sd = 6.4, n = 2). counts in the 1990s averaged 6.0 (sd = 5.8, n = 20), and counts in 2000 – 2001 averaged 2.0 (sd = 2.8, n = 9). road transect index — the overall likelihood of seeing at least 1 moose while traveling the gardiner to cooke city road (n = 1,020) was 0.26 during the 9 years data were collected. the likelihood of seeing at least 1 moose per trip varied seasonally, with the during may/june when moose were observed on 50.4% of trips and lowest during september/october when moose were observed on only 7% of trips (fig. 3). because numbers of trips were relatively consistent across the tion effects were based on pooled data for individual years. the likelihood of sighting a moose during a drive between gardiner and cooke city was highest in 1989 (n = 84 trips, likelihood = 0.49), the year immediately subsequent years (fig. 4). the lowest likelihood of sighting a moose (0.02) occurred in decline in moose sightings (z-test, p < 0.05) was included in the test. moose distribution along the road between gardiner and cooke city was not uniform before or after the section 1 (gardiner to mammoth) before or to roosevelt junction), moose were observed z-test, p = 0.04). in section 3 (roosevelt junction to round prairie), the sighting incidence was p < 0.01). in section 4 (round prairie to warm creek), incidences of sighting were similar before and after the p = 0.95). the percentage of trips in which moose were observed in section 5 (warm creek – cooke p = 0.15). — seventy-eight aerial searches of willow stands in frenchy’s meadow and 73 in the cooke tween june 1987 and december 1990. the 1988, 1989, and 1990 (h = 1.95, p = 0.58). was in 1988 (4.9), followed by 1989 (3.1). results were the same for 1987 and 1990 0 10 20 30 40 50 60 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 year li ke lih oo d (% ) fig. 4. likelihood (%) of seeing at least 1 moose while traveling the road from gardiner to cooke city, montana (89 km) for each of the years 1987 – 1992 and 1995 – 1997. 0 10 20 30 40 50 60 ja n/ f eb m ar /a pr m ay /j un e ju ly /a ug s ep /o ct n ov /d ec month li ke lih oo d (% ) fig. 3. likelihood (%) of seeing at least 1 moose while traveling the road from gardiner to cooke city, montana (89 km) by 2-month period for the years 1987 – 1992 and 1995 – 1997. alces vol. 42, 2006 tyers – yellowstone population history 141 age number of moose seen per month over 4 h = 18.89, p = 0.026) (fig. 6). the highest average number seen by december (8.6), and may (7.6). the percent of radio-collared moose available for observation (i.e., alive, in the drainage, and with operational radio-collars) seen per years (h = 5.26, p = 0.15). means for years varied from 0 (1987) to 12% (1988). although lared moose observed by month were detected (h = 13.41, p = 0.15), the highest percent seen was in may (18.0%), followed by december (13.8%), and november (13.1%) (fig. 6). this implies that in the late spring and early winter periods, when moose were most visible, < 20% of moose in a drainage were likely to daily willow stand observations — daily counts of the number of moose in a willow stand near cooke city were made at half-hour intervals for 15 months. the mean number of moose seen per half-hour of (f = 10.76, p < 0.001). the highest average number seen per half-hour was in june 1997 (0.9), followed by december 1996 (0.6), and may 1996 (0.6) (fig. 7). counts of moose were highest on average at two times of day, between 0600 – 0930 hours and 2030 – 2130 hours (fig. 8). when daylight, moose were most visible in the hours near sunrise and sunset. for months in late spring (may and june) and early winter (november and december) when highest numbers 0 4 8 12 16 20 1 2 3 4 5 road section number li ke lih oo d (% ) before fires after fires fig. 5. likelihood (%) of seeing at least 1 moose while traveling the 5 sections of road between gardiner and cooke city, montana, prior to 1 = gardiner – mammoth (8.0 km); section 2 = mammoth – tower junction (29.1 km); section 3 = tower junction – round prairie (30.9 km); section 4 = round prairie – warm creek (13.2 km); section 5 = warm creek – cooke city (8.0 km). 0 2 4 6 8 10 jan mar may july sep nov month a ve ra ge # m oo se pe r fl ig ht 0 5 10 15 20 % of c ol la re d m oo se s ee n collared and uncollared moose % of collared moose seen yellowstone population history – tyers alces vol. 42, 2006 142 of moose per half-hour were recorded, the optimum times for moose observation were: may, 0600 – 0700; june, 0600 and 2130; november, 0730; and december, 0830. – even though the north and south halves of the study area could not a sharp decrease in moose sightings between november 1989 and may 1990 (fig. 9). the highest number seen on a single survey was 59 in november 1989. the lowest count (13) occurred in may 1992. discussion population history long-term studies in north america support the idea that moose populations consistently erupt, crash, and then stabilize at various densities depending on prevailing ecological conditions (mech 1966, peek et al. 1976, schwartz and franzmann 1989, loranger et al. 1991, messier 1991, van ballenberghe and ballard 1998). geist (1974) attributed this pattern to a basic response by moose populations to changes in habitat quality. in his opinion, over the species’ evolutionary history, moose have typically occupied limited areas of perhas created transient habitat, they have rapidly colonized these areas and reached comparatively high densities. population eruptions can also be triggered by plant succession following logging and reduction of hunting or predation pressure, if a population is being held at low densities due to predation or hunting (mech 1966, peek et al. 1976, messier 1991). moose evidently colonized the nywr in the 1800s and initially increased in numbers in a manner similar to colonizing moose in other areas in north america, but the population did not respond positively to the massive 0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 a pr -9 6 m ay -9 6 ju n96 ju l-9 6 a ug -9 6 s ep -9 6 o ct -9 6 n ov -9 6 d ec -9 6 ja n97 f eb -9 7 m ar -9 7 a pr -9 7 m ay -9 7 ju n97 month h ou r (m ili ta ry ) fig. 8. hours in which highest counts of moose in a willow stand near cooke city, montana were recorded by month, april 1996 – june 1997. values are based on counts at half-hour intervals during all daylight hours. months with 2 values indicate ties. hour of the day follows standard conventions for mountain standard and daylight savings time. fig. 7. average number of moose seen per month (based on daily counts at half-hour intervals during daylight hours) in a willow stand near cooke city, montana, april 1996 – june 1997. 0 0.5 1 1.5 2 a pr 96 m ay ju n ju l a ug s ep t o ct n ov d ec ja n 97 f eb m ar a pr m ay ju n month n um be r pe r 1/ 2 ho ur alces vol. 42, 2006 tyers – yellowstone population history 143 ecosystem during 1988, as might have been (schwartz and franzmann 1989). when moose invaded the nywr, they encountered an environment in transition due to european settlement. human predation was initially important and then curtailed. forest succession was altered with attempts habitats on the nywr by the middle of the 20th century. reports of negative impacts on willow stands (usda 1936, mcdowell and moy 1942) indicate that, at least in some drainages, moose numbers may have stabilized or over-populated the area by the late 1930s. regulated permit-based hunting, introduced in the 1940s to alleviate damage to willow stands on the nywr, may have ended a population eruption triggered by a ban on hunting, dating from the early 1900s, and reduced predation resulting from concerted efforts to eliminate predators from the yellowstone ecosystem during the 1910s – 1930s. because no organized monitoring of moose populations was conducted by agencies from 1950 to 1985, the population trends during this 35-year period will never be known, but the horseback surveys conducted for this study from 1985 to 1987 produced similar moose sighting rates as gaab’s (1949) survey in 1949, perhaps indicating that the population remained relatively stable from 1949 to 1987. affected moose habitat and population levels at a landscape scale. in the period immediand summer 1990, some indices produced by winter 1990 – 1991, however, all indices indicated substantial declines in moose numthe reduction in numbers was greater than no sign of population recovery was evident through 2001; the last year data for 1 or more indices was collected. population monitoring the horseback surveys, road transects, and 0 10 20 30 40 50 60 dec88 nov-89 dec90 may92(#1)flight date n um be r of m oo se north south total fig. 9. number of moose seen during aerial surveys of the complete northern yellowstone winter range (nywr) and in 2 segments of the nywr (north and south of the yellowstone river) from december 1988 to may 1992. yellowstone population history – tyers alces vol. 42, 2006 144 decline from 1987 to 1990, but revealed a similar pattern of change (relatively low in 1987, high in 1988 and 1989, low in 1990) to that provided by the horseback survey and the road transect. sighting numbers per day in 1988 and 1989 and consistent, very low sighting rates from 1995 to 2001. the high numbers of moose seen in 1988 and 1989 were probably due to into unburned willow stands along the route. data on moose movement and survival (tyers 2003) indicate that data collected from ably under-represented actual moose numbers before 1988. rored results from the horseback survey; an increase in sighting likelihood in 1988 – 1989 and a decline thereafter. the decrease was most pronounced on the section where forround prairie) and least pronounced where areas bisected by the road were not burned. apparent as early as 1990 while values from the were similar to values for 1985 – 1987 (prewas more sensitive to population changes than the horseback survey or it may only be an artifact of sampling greater areas of burned terrain or more marginal habitat on the road transect than on the horseback survey. systematic aerial surveys were not initicounts decreased substantially by 1990 and tinued in 1992. by 1992, numbers of moose low and limited to a few large willow stands, including those monitored in willow stand was detected between 1987 and 1990. study could potentially be improved by timing sampling to optimize moose sightability. time of year can affect sightability of moose (lynch 1975, crête et al. 1986, gasaway et al. 1986, bisset and rempel 1991). february and march are considered the most difmore likely to be in dense cover (leresche and rausch 1974, novak and gardner 1975, novak 1981). sightability in november and december may be higher because moose form larger groups and have stronger preferences for vegetation with low, open canopies. this has been found in alaska (peek et al. 1974, gasaway et al. 1986), minnesota (peek et al. 1974, mytton and keith 1981), michigan (peterson and page 1993), alberta (lynch 1975), and ontario (bisset and rempel 1991). however, 34 consecutive years of aerial surveys in saskatchewan were successfully conducted in january and february (stewart and gauthier 1988). in ynp, barmore (1980) found seasonal variation in moose sightability during attempts to count moose incidental to elk distribution environment. most moose barmore (1980) saw were associated with willow, and he was communities in may, early june, and december. in my study, radio-collared and uncollared moose were more likely to be observed from ber and december) and may than in other seasonal periods. a similar seasonal pattern was observed during intense ground sampling in willow stands near cooke city. alces vol. 42, 2006 tyers – yellowstone population history 145 of moose (leresche and rausch 1974). timmermann (1974) suggested from 1000 to 1400 hours as the optimal time for moose aerial surveys in ontario. peterson and page (1993) after sunrise. data from half-hour counts in a willow stand near cooke city for this study indicated that early morning (0600 – 0930 hours) and late evening (2030 – 2130 hours) were the best times of the day to see moose in the yellowstone area. would aerial surveys in early winter or late spring, concentrated in early morning hours, moose associated with the nywr at current population levels? aerial surveys of moose rausch 1974, stevens 1974, novak 1981), but despite problems, counting moose on winter ranges from aircraft is still considered the most practical method for estimating moose numbers over large areas in north america (mantle 1972, gasaway et al. 1986, gasaway and dubois 1987, timmermann and buss 1998). in some areas, aerial surveys are 78% of moose located during intense ground surveys were seen from the air. evans et al. aircraft saw 94% of moose observed by crews in helicopters. gasaway et al. (1978) noted that 91% of radio-collared moose available to be seen were found during intensive searches from the air. used in a systematic survey of the nywr would locate a high proportion of the moose population. even in the months with highest sightability (november, december, and may), < 20% of radio-collared moose known to be in drainages containing preferred willow high variability in both percent of radio-collared animals observed and in total animals observed indicate that using a large number of radio-collared moose to develop a sightability ity in estimating elk numbers (samuel et al. 1987), is not likely to yield good results given the low density and low visibility of moose associated with the nywr. low density and low sightability would also limit the utility of helicopter surveys. stands does have potential for tracking changes in the moose population associated with the nywr, if counts are made in early winter or late spring and limited to early morning hours. boundaries of key willow stands are cover relatively small areas (most are < 40 ha). counts of moose along the highway between gardiner and cooke city during early winter and late spring may also provide a relatively cheap means of monitoring population trends. summer – autumn horseback surveys, especially when costs can be mitigated by combining counts with required tasks, such as trail maintenance and hunter management, may also be useful in tracking trends in moose populations. although indices are less intellectually satisfying as a base for management of moose than statistically valid population estimates, indices may provide a reasonably reliable mechanism for determining population trends in situations where logistical constraints preclude accurate estimates of moose numbers. acknowledgements of the northern yellowstone cooperative wildlife working group (nycwwg). the nycwwg is comprised of agencies with management or research responsibilities for the nywr and includes montana fish, wildlife, and parks, yellowstone national park, gallatin national forest, and the u.s. geological survey northern rocky mountain research center. funding was provided by yellowstone population history – tyers alces vol. 42, 2006 146 gerry bennett, dick ohman, jim wood, john and kathryn harris, safari club international montana chapter, gallatin national forest, montana fish, wildlife and parks, and yellowstone national park. i also wish to thank a number of individuals that, by providing helped make this study possible, namely: phyllis wolfe, tami blackford, john varley, frank singer, tom lemke, sam reid, patrick hoppe, steve yonge, jeremy zimmer, par hermansson, sarah coburn, reyer rens, kim olson, maria newcomb, tom nelson, rob st. john, tris hoffman, ellen snoeyenbos, anna carin andersson, bill chapman, tom drummer, bill gasaway, emma cayer, vanna boccadori, mark petroni, tom puchlerz, lynn irby, jack taylor, tom mcmahon, harold picton, peter gogan, kurt alt, harry whitney, mike ross, kelsey gabrian, and randy wuertz. references barmore, w. j. 1980. population characteristics, distribution, and habitat relationships national park. final report. national park service, yellowstone national park, mammoth hot springs, wyoming, usa. bisset, a. r., and r. s. rempel. 1991. linear analysis of factors affecting the accuracy of moose aerial inventories. alces 27:127-139. cooney, r. f., w. k. thompson, and l. e. mcdowell. 1943. montana moose survey slough creek-hellroaring unit. supplement no. 2. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. couey, f. m. 1947. summary of montana’s 1946 moose kill. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. crane, h. m. 1922. report on game protection and patrol of winter elk range, winter of 1921-1922. unpublished report. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. crête, m., l. rivest, h. jolicouer, j. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose aircraft in southern quebec. journal of applied ecology 23:751-761. edwards, r. y. 1954. comparison of aerial and ground census of moose. journal of wildlife management 18:403-404. evans, c. d., w. a. troyer, and c. j. lensink. 1966. aerial census of moose by quadrant sampling units. journal of wildlife management 30:767-776. gaab, j. e. 1948. absaroka unit moose survey-1947. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. _____. 1949. absaroka unit moose survey1948. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. _____. 1950. absaroka unit moose survey1949. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. gasaway, w. c., and s. d. dubois. 1987. estimating moose population parameters. swedish wildlife research supplement 1:603-617. _____, _____, s. j. harbo, and d. g. kelleyhouse. 1978. preliminary report on accuracy of aerial moose surveys. proceedings of the north american moose alces vol. 42, 2006 tyers – yellowstone population history 147 conference and workshop 14:32-55. _____, _____, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska. number 22. university of alaska, fairbanks, alaska, usa. geist, v. 1974. on the evolution of reproductive potential in moose. naturaliste canadien 101:527-537. houston, d. b. 1974. aspects of the social organization of moose. pages 690-696 in v. geist and f. walther, editors. the behavior of ungulates and its relation to management. volume ii. publication 24. union of conservation, nature, and natural resources, morges, switzerland. _____. 1982. the northern yellowstone elk: ecology and management. mcmillian, new york, new york, usa. johnson, c.h. 1925. report of game studies and game protection on the park district of the absaroka national forest, winter of 1924-25. unpublished report. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. kehrberg, e. v. 1964. wildlife management plan, gardiner ranger district. unpublished report. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. lancia, r. a., j. d. nichols, and k. h. polluck. 1994. estimating the number of animals in wildlife populations. pages 215-253 in t. a. bookhout, editor. research and management techniques for wildlife and habitats. fifth edition. the wildlife society, bethesda, maryland, usa. le resche, r. e., and r. a. rausch. 1974. accuracy and precision of aerial moose censusing. journal of wildlife management 38:175-182. loranger, a. j., t. n. bailey, and w. w. larned. 1991. effects of forest succeson the kenai peninsula, alaska. alces 27:100-109. lynch, g. m. 1975. best timing of moose surveys in alberta. proceedings of the north american moose conference and workshop 11:141-152. mantle, e. f. 1972. a special moose inventory, 1971aubinadong moose study area, sault ste. marie forts district, ontario. proceedings of the north american moose conference and workshop 8:124-137. mccullough, d. r. 1984. lessons from the george reserve, michigan. pages 211-242 in l.k. halls, editor. whitetailed deer: ecology and management. stackpole books, harrisburg, pennsylvania, usa. mcdowell, l. e. 1946. report on 1945 moose season to date. montana fish and game department, wildlife restoration division. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. _____, and m. moy. 1942. montana moose survey, hellroaring-buffalo fork-slough creek unit. montana fish and game department, wildlife restoration division. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. _____, and p. page. 1944. montana moose survey slough creek-hellroaring unit. supplement no. 3. montana fish and game department, wildlife restoration division. montana department of fish, wildlife, and parks, region 3, bozeman, montana, usa. _____, and r. smart. 1945. montana moose survey slough creek-hellroaring unit. supplement no. 4. montana fish and game department, wildlife restoration division. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. mech, l. d. 1966. the wolves of isle royale. yellowstone population history – tyers alces vol. 42, 2006 148 fauna service no. 7. national park serwashington, d.c., usa. messier, f and regulating factors on the demography of moose and white-tailed deer. journal of animal ecology 60:377-393. montana fish and game department. 1945. forword to moose hunter harvest surveys. unpublished report. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. mytton, w. r., and l. b. keith. 1981. dynamics of moose populations near rochester, alberta, 1975-1978. canadian field-naturalist 95:39-49. national academy of sciences. 2002. ecological dynamics of yellowstone’s northern range. yellowstone science 10(2):3-11. novak, m. n. 1981. the value of aerial inventories in managing moose populations. alces 17:282-315. _____, and j. gardner. 1975. accuracy of aerial moose surveys. proceedings of the north american moose conference and workshop 11:154-180. parsell, j., and l. mcdowell. 1942. montana moose survey slough creek-hellroaring unit. supplement no. 1. montana fish and game department, wildlife restoration division. montana department of fish, wildlife and parks, region 3, bozeman, montana, usa. peek, j. m., r. e. le resche, and d. r. stevens. 1974. dynamics of moose aggregation in alaska, minnesota, and montana. journal of mammalogy 55:126-137. _____, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. l., and r. e. page. 1993. detection of moose in mid-winter from wildlife society bulletin 21:80-86. rush, w. 1942. agency correspondence: w. m. rush to l. mcdowell and m. moy. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. saether, b. e. 1987. patterns and processes in the population dynamics of scandinavian moose (alces alces): some suggestions. swedish wildlife research supplement 1:525-537. samuel, m. d., e. o. garton, m. w. schlegel, and r. g. carson. 1987. visibility bias during aerial surveys of elk in northcentral idaho. journal of wildlife management 51:622-630. schladweiler, p. 1973. ecology of shiras moose in montana. big game research report. montana department of fish and game, helena, montana, usa. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose, and forest succession, some management considerations on the kenai peninsula, alaska. alces 25:1-10. statsoft. 1995. statistica for windows, volume 1. general conventions and statistics 1. second edition. tulsa, oklahoma, usa. stevens, d. r. 1974. rocky mountain elk shiras moose range relationships. naturaliste canadien 101:505-516. stevenson, d. h. 1920. hellroaring slough creek game patrols, 1919-20. unpublished report. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. stewart, b. r., and d. a. gauthier. 1988. temporal patterns in saskatchewan moose population, 1955-1988. alces 24:150-158. timmermann, h. r. 1974. moose inventory methods: a review. naturaliste canadien 101:615-629. _____, and m. e. buss. 1998. population harvest and management. pages 559-616 in a. alces vol. 42, 2006 tyers – yellowstone population history 149 w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington d.c., usa. tyers, d. b. 1981. the condition of the northern yellowstone winter range in yellowstone national park a discussion of the controversy. m.sc. thesis, montana state university, bozeman, montana, usa. _______. 2003. winter ecology of moose on the northern yellowstone winter range. ph. d. thesis, montana state university, bozeman, montana, usa. ulhorn, c. f. 1923. winter game conditions of the hellroaring, buffalo fork, and the slough creek drainages absaroka national forest winter of 1922-1923. unpublished report. u.s. forest service, gallatin national forest, gardiner ranger district, gardiner, montana, usa. (usda) u. s. department of agriculture. 1936. absaroka winter game studies 1935-1936. unpublished report. u.s. forest service, gallatin national forest, livingston ranger district, livingston, montana, usa. van ballenberghe, v., and w. b. ballard. 1998. population dynamics. pages 223-246 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington d.c., usa. yellowstone national park. 1997. yellowchange in a wildland ecosystem. national park service, yellowstone national park, mammoth hot springs, wyoming, usa. zar, j. h. 1999. biostatistical analysis. fourth edition. prentice hall, upper saddle river, new jersey, usa. alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 109 the effects of human activity on summer habitat use by moose odd n. lykkja1,2, erling johan solberg3, ivar herfindal1,4, jonathan wright1, christer moe rolandsen1 and martin g. hanssen1 1centre for conservation biology, department of biology, norwegian university of science and technology, n-7491 trondheim, norway; 2nina naturdata as, n-7898 limingen, norway; 3norwegian institute for nature research, n-7485 trondheim, norway; 4museum of natural history and archaeology, norwegian university of science and technology, n-7491 trondheim, norway abstract: non-fatal disturbance by humans can be analogous to predation risk because animal response to both directly reduces time available for other fitness-increasing activities such as foraging, maternal care, and reproductive behaviour. we studied the effects of human disturbance on moose (alces alces) by examining hourly locations and movement patterns of 41 gps-marked moose relative to human activity in central norway during summer 2006. our results indicated that moose moved further from inhabited houses and to areas of lower housing density in periods of high human activity as compared to periods of low human activity, and that this behavioural response was closely related to the level of human activity in the area used by moose. we also detected significant differences between responses of males and females with calves; males were more willing to use areas near houses and with higher housing density during periods of low human activity. this differential response was likely due to the higher perceived risks of foraging associated with maternal protection of non-independent offspring. our study supports the idea that indirect cost associated with human disturbance is analogous to the influence of perceived predation risk on animals. we suggest that such indirect effects on moose should be accounted for when planning human construction and activity in prime moose habitat. alces vol. 45: 109-124 (2009) key words: alces alces, gps, habitat use, human disturbance, moose, norway, perceived predation risk. in the past 20 years a number of studies have investigated the effects of human activity and infrastructure on behaviour, habitat selection, life history, and ultimately population dynamics of large ungulates (singer and beattie 1986, andersen et al. 1996, dussault et al. 2005). these studies have increased our understanding of the evolutionary background and underlying mechanisms responsible for why and how ungulate species respond to different stimuli. an important principle receiving increased support is that non-fatal disturbance by humans can be analogous to increases in perceived predation risk (frid and dill 2002). the reason is that responses to both human disturbance stimuli and predation risk involve a decrease in the time available for other fitness-enhancing activities like foraging, maternal care, and reproductive behaviour (frid and dill 2002). humans have been an important predator of large ungulates for thousands of years; it follows that most ungulates have good reason to avoid humans. accordingly, responses to human disturbances can be regarded as a trade-off against preferred resource use, in the same way that predator risk influences individual choice of home range and habitat use (white and berger 2001, creel and christianson 2008). this approach can be used to explore whether human disturbance has limiting effects on the number of individuals that utilize an area, and to predict change in local density following human disturbance (gill and sutherland 2000). moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 110 previous studies have shown that habitat use of several ungulate species is affected by human infrastructure and activity (wolfe et al. 2000, nellemann et al. 2001, setsaas et al. 2007), particularly for intensively harvested populations and for open-dwelling species like reindeer (rangifer tarandus) (nellemann et al. 2001, strand et al. 2006). less is known about the response of forest-dwelling species because they are more secretive, but it is assumed that they are more tolerant of human activity because forests provide better security cover. this is reflected in their typical response to potential danger of an approaching predator (including humans) that often involves immobility and hiding rather than rapid flight (andersen et al. 1996). however, that does not mean that they are unaffected. areas close to human settlements and roads can be used less by forest-dwelling ungulates, despite such areas often containing high density, quality forage (histøl and hjeljord 1995). despite the fact that human infrastructure and activity seem to affect the temporal and spatial distributions of various ungulate species, avoidance of humans is not equal among all species or populations and may differ seasonally. in particular, ungulates seem to be able to tolerate human activity to a higher extent during periods of food shortage (strand et al. 2006), conforming to the prediction of “the predation risk allocation hypothesis” which states that allocation of risk should be responsive to changes in an animal’s energetic state over time (lima and bednekoff 1999). males and females appear to be affected differently by human activity (childress and lung 2003, ciuti et al. 2004), presumably because of different life history and reproductive traits that cause contrasting time and energy budget trade-offs. for example, we might expect males to be more active and risk-taking than females because males are not impeded by protecting offspring. further, moose (alces alces) have a polygynous mating system (andersen and sæther 1996) and their reproductive success often depends on large body size (weckerly 1998). males are expected to maximize energy intake at all times to achieve highest possible growth rate, social status, and mating success, whereas females trade-off growth for reproduction and maternal care (clutton-brock et al. 1988). male fallow deer (dama dama) had high use of areas with high anthropogenic disturbance during day and night, whilst females frequented disturbed areas only at night (ciuti et al. 2004). thus, disturbance reactions by animals can be perceived as part of a dynamic process that reflect individual trade-offs between sex-specific consequences of foraging under human disturbance, versus the pay-off associated with foraging in undisturbed areas with possibly more resource competition (gill et al. 2001a, b, frid and dill 2002). moose are not evenly distributed in their environment, but rather show habitat preferences (andersen and sæther 1996). large individual differences in dispersal distance and home range size occur, but for long periods often lasting several years, some moose rarely move outside an area of only 4-5 km2 (cederlund et al. 1987). moose with large and small home-ranges co-exist, but little is known about the mechanisms behind this difference or any consequence upon reproductive success (andersen and sæther 1996). one possible reason for this large variation is that food resources are heterogeneously distributed across a landscape and some individuals need larger home ranges to meet their nutritional requirements (tufto et al. 1996, anderson et al. 2005). individuals may adjust their dis-). individuals may adjust their distribution relative to habitat quality to balance acquisition of resources; this may reflect their competitive ability in the population. this theoretical pattern of individual distribution according to resource abundance has been termed the “ideal free distribution” (fretwell 1972, parker and sutherland 1986, koops and abrahams 2003) because it assumes that animals are “free” to go wherever they will alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 111 do best, and that all animals are “ideal” in having complete information about the availability of resources. an alternative explanation may be that human disturbance influences the utilization of the various resources inside established home ranges, without any largerscale competitive adjustments in adjacent and/ or overlapping home ranges. after near extermination, wolf (canis lupus) and brown bear (ursus arctos) populations are slowly increasing in norway but their effects on moose populations are regarded as negligible (solberg et al. 2003). given the high legal harvest and accidental kill on roads and rails, humans can be assumed to be the most important predator of moose in norway. generally, predation risk varies seasonally (lucas et al. 1996), within a day (fenn and macdonald 1995), or even minute to minute during an encounter with a predator (dill and gillett 1991). thus, the fact that animals are able to detect and respond to temporal variation in the risk of predation (kats and dill 1998) is not surprising. responses to risk can be morphological (tollrian and harvell 1999) or behavioural (lima 1998), including changes in habitat use (creel et al. 2005), vigilance (childress and lung 2003), foraging (winnie and creel 2007), aggregation (barta et al. 2004), movement patterns, (fortin et al. 2005) and sensitivity to environmental conditions (winnie et al. 2006). the costs of these responses can be manifested by reduced survival, growth, or reproduction (pangle et al. 2007, creel and christianson 2008). we examined the extent to which moose respond to human activity by analyzing how they use their home range at (1) various distances from human activity centres, and (2) with variable housing density. moose may be expected to choose home ranges and habitats within home ranges away from human activity centres to avoid contact with humans. however, because habitat use is assumed to have large influence on body growth and lifetime reproductive success, and because habitats with high forage quality are often associated with human infrastructure, moose may have to compromise between rich habitats and associated high levels of human disturbance. they could optimise this compromise by using their preferred habitats during periods of reduced (perceived) risk of predation, i.e., during periods of low human activity. we predicted that moose would regulate their proximity to humans based on human activity, and that their use pattern would vary based on the mean distance to humans or the density of humans in the local vicinity. moreover, according to sexual selection theory, males should have less to lose and more to gain from risky behaviour than females (clutton-brock et al. 1988). hence, we predicted that males would be more willing to approach houses and use areas of higher housing density at times of human activity. study area the study was conducted in the central part of norway in the county of nord-trøndelag, and the southern part of nordland, between 63◦ and 65◦ n and 10◦ and 14◦ e (fig. 1). the study area ranged in altitude from sea level to mountainous areas with peaks to 1500 m. below the woodland limit at about 500-700 m, the area was covered by boreal forests, mainly norwegian spruce (picea abies) and scots pine (pinus sylvestris) (moen et al. 1999). the forest was managed for timber and pulp with a high frequency of clear cutting; this patchy environment had large areas in early succession providing high quality forage and a high carrying capacity for moose. combined with more restrictive hunting, the moose population had increased substantially for 50 years and was the most important game species in the area (lavsund et al. 2003). the annual moose harvest in nord-trøndelag in recent years was about 5000 (i.e., about 5 moose per 10 km2 forest and bog; solberg et al. 2006). human settlements and agricultural land in the study area were confined to areas along moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 112 the western coast and along trondheim fjord, as well as inland valley floors (fig. 1). areas with higher human density were characterized by higher primary production and a longer growing season compared with more elevated areas (moen et al. 1999). about 6 humans and 0.26 km public road per km2 were found in nord-trøndelag at the time of the study (ssb 2009). materials and methods data collection we used 41 moose (28 females each with 1 or 2 calves, and 13 males) marked with gps/ gsm collars in february-march 2006; observations occurred 1may-20 september the same year. the data were restricted to observations during weeks when moose were within 4000 m of the nearest house, on average, to exclude extreme outliers. observations in areas where moose were likely not to be affected by humans (forman et al. 2003) were included to serve as a control. the excluded data constituted <10% of the observations. the gps/gsm collar (gps plus, vectronic aerospace, berlin) provided almost instantaneous locations by sending positional data by sms via the cell phone network that covers most of norway. one position was logged hourly, and every 5 h the 5 latest fixes were sent to a server computer at the norwegian institute for nature research (nina). if a moose stayed in an area outside the cell phone network for a period of time, all logged positions were sent as soon as connection was regained. the overall fix rate during the study period was 87.7%. topography, vegetation, position fix-rate, and animal posture may influence signal transmission between gps-satellites and receivers, thus influence fix acquisition and location error and cause systematic bias in location data (cain et al. 2005). however, because the proportion of successful fix attempts was high, we expected no reduced accuracy or bias. females were approached on foot during the calving season to document calf production and again at the end of the study period. we could not verify whether 18% of the cows still had their calf/ calves in the autumn, but believed they did based on indirect observations (foot prints, hunter observations, movement pattern; see sæther et al. 2003 for more details about methodology). the data on human activity were obtained from the norwegian mapping authority as digital map layers with information about roads and buildings. the spatial precision of these data corresponded to printed maps at the1:50.000 scale. buildings were represented as points and villages and cities as polygons. buildings associated with infrequent visits by humans (cabins, summer farms, and outhouses) were not included in the analyses. we also considered using densities and distances to roads, but because distance to inhabited houses was highly correlated with distance to public roads (r = 0.945, p < 0.001), whereas the correlation between density of houses and densities of roads was more moderate (r = 0.350, p < 0.001), we used only distance to and density of houses in the analyses. habitat analysis we made a raster map with 100 m2 resolution covering the study area. each cell was assigned the euclidian distance to the nearest inhabited house; housing density was estimated with the quadratic kernel function described fig. 1. maps of norway (left) and the study area with public roads represented by black lines (right). alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 113 by silverman (1986, p. 76, equation 4.5) with a search radius of 564.2 m. this produced a search area of 1 km2 that we considered a reasonable scale to estimate housing density in a given cell. by overlaying the gps positions of moose on the 2 maps, we assigned distance and density for each moose location. statistical analysis the main question of our study was if the spatial distribution pattern of moose during the day varied in accordance with the level of local human activity. we examined this by analyzing the temporal variation of moose locations to distance to houses and housing density during the day relative to the average distance or density measurements the same day. to control for the large individual variation in response variables, we calculated relative distance and density on a daily scale by centring the response variables on daily means per individual. these response variables were approximately, normally distributed around zero and controlled for the variation in mean distance or density among individuals. the predictor variables were simply the individual mean distance or density per day (24 h). in subsequent analyses the relative within-individual variation in distance or density are denoted as centred distance and centred density, whilst the mean values are denoted mean used distance and mean used density. we analyzed the variation in centred distance and centred density based on the values taken each hour during the 24-h day, as well as on average values for 2 periods within the day that reflected 2 distinct levels of human activity, high and low. the timing of all observations was rounded to the nearest hour. the “high human activity” period began at 0700 hr when most people go to work and school, and ended at 2000 hr. the “low human activity period” began at 2100 hr and included observations until 0600 hr the following morning. in each 24-h period there were 14 observations categorized as “high human activity” and 10 as “low human activity.” this categorization corresponded reasonably with summer traffic-counts at main roads in the study area (lykkja 2008). the short time lag between successive observations introduced problems regarding spatial and temporal autocorrelation. gps positions collected within short time intervals are usually not statistically independent. one solution is to average away any possible pseudoreplication by analyzing means calculated over longer time periods/larger areas (crawley 2002). we chose a conservative approach where we averaged the data on a weekly basis; the statistical analyses regarding the daily pattern were based on the mean values per individual/h/week (n = 16643), or individual mean values per human activity period/week (n = 1395). hence, the distance to nearest house at 1200 hr was the average of the daily centred distance to nearest house at 1200 hr for the entire week (monday-sunday). we examined which of the explanatory variables were related significantly to our response variables by applying linear mixed effect models (lme) to the data using the lme4 package (bates 2007) in r version 2.6.0 for windows (r development core team 2007). moose individual was included as a random factor (random intercept). by using mixed effect models, we avoided estimating intercepts for each individual that would require additional degrees of freedom (crawley 2002), and we also accounted for the non-independence in observations per moose due to individual differences in behaviour. we first analyzed variation in centred distance/density during the 24-h period in relation to mean used distance/density at 3 different levels (factor with 3 levels). the 3 categories of distance were close (<1000 m, n = 9117 observations), intermediate (1000-2000 m, n = 4097 observations), and far (>2000 m, n = 3429 observations) from nearest house based on mean used distance. corresponding categories for density were high (>2 houses/km2, moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 114 n = 3781 observations), intermediate (0.2-2 houses/km2, n = 4135), and low (0-0.2 houses/ km2, n = 8727) mean used densities. we also included the fixed effects of hour (continuous, including up to third order polynomials) and moose sex (factor with 2 levels) in the models. in the second step, we analyzed the variation in centred distance/density during the human activity periods in relation to moose sex, human activity period (factor with 2 levels), and mean used distance/density (continuous). to investigate possible non-linear relationships, 2 polynomials (second and third order mean used distance/density) were also included in the models. in addition, the global models included all possible 2-way interactions. however, we did not include 3-way interactions in the models. this was based on our expectation that the influence distance categories or mean distances had on diurnal patterns or human activity patterns was similar among males and females, even if the diurnal patterns or human activity differed between males and females. moreover, the inclusion of 3-way interactions would lead to a substantial number of parameters and candidate models, increasing the possibility of over-fitting the data. we performed model selection by using the akaike information criterion (aic) with second order adjustment of the aic (aicc) to correct for small sample sizes. all combinations of explanatory variables from the global model were allowed as candidate models, with the exception that if an interaction or a polynomial term was included in the model, the main effect or lower-grade polynomial was always included in the same model. the aiccvalue is based on the principle of parsimony to find the best fitted models (i.e., the parameters with the lowest aicc-values; burnham and anderson 2002). models that differed from the best model with ∆aicc ≤ 2 were considered to have similar empirical support by the data (burnham and anderson 2002). we also calculated akaike weights (aiccw) for each model and interpreted the weights as the probability that the model is best for the situation given the data set and the candidate models (burnham and anderson 2002). evidence ratios were calculated as aiccwratio = aiccw(model1) / aiccw(model2) to examine the strength of evidence for one model in favour of another. aicc values were computed based on log-likelihood from models fitted with maximum-likelihood (ml), whereas the parameter estimates and their corresponding uncertainty estimates were based on models fitted with restricted maximum likelihood (reml) (crawley 2002). all tests were run using the statistical program r 2.6.0 for windows (r development core team 2007). results distance to houses the best model describing centred distance to nearest house during the 24-h period included all main effects and 2-way interactions except for the interaction distance category * sex and the interaction between the third order polynomial of hour and sex (model 1, table 1). the second best model also included the interaction between the third order polynomial of hour and sex (model 2, table 1), but this model was >2x less supported than the best model (aiccwratio of model 1 compared with model 2 = 2.43; table 1). according to model 1, moose showed marked differences in centred distance from nearest house during the 24-h period (fig. 2). when close to houses (<1000 m), both males and females showed a pronounced diurnal pattern where shorter distances than average were used at night and longer distances than average were used during day (fig. 2). this pattern was the same, but less pronounced, when moose were situated at intermediate distances (1000-2000 m) from houses. however, when situated far from houses (>2000 m), this model predicted that slightly shorter distances than average were being used in the afternoon. for females especially, there was an apparent change in the diurnal pattern in this category, where longer alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 115 distances than average were used at night and shorter distances than average during day. in the second step, we found the variation in centred distance from housing between periods of high and low human activity was best explained by moose sex, human activity, mean distance, the second order polynomial of mean distance, in addition to a sex-specific effect of these 3 variables (model 1, table 2). only 2 additional models were within ∆aicc ≤ 2 of the best model (table 2). the second best model (∆aicc=1.31) also included an interaction between mean used distance and sex, whereas the third best model included the additional third order polynomial of mean used distance (∆aicc=1.92). the simplest and best model was almost 2x better supported than the second best model (aiccwratio of model 1 compared to model 2 = 1.92; table 2), whereas the third best model was almost 3x less well supported than the best model (aiccwratio of model 1 compared to model 3 = 2.72; table 2). according to model 1, moose were observed farther from houses than the average during periods of high human activity and closer than the average during periods of low human activity, especially when the mean used distance was short (fig. 3). this effect decreased with increasing mean used distance from houses. males approached houses to a greater extent than females when humans were less active, regardless of their mean used distance from m odel sex d istance category sex : d istance category h our h our 2 h our 3 h our : sex h our 2 : sex h our 3 : sex h our : d istance category h our 2 : d istance category h our 3 : d istance category ∆ a ic c a ic c-w eight 1 x x x x x x x x x x 0 0.418 2 x x x x x x x x x x x 1.77 0.172 table 1. the aicc-based ranking of models explaining daily variation in centred distance using linear mixed effect models with individual as a random factor. variables included in the candidate models are marked by an x. the best model (model 1) had an aicc-value of 193207.70. ∆aicc refers to the difference in aicc between the best model and the candidate model. the global model had an aiccvalue of 193213.3 (∆aicc= 5.64); only models with ∆aicc <3 are presented. m odel sex h um an activity m ean distance m ean distance 2 m ean distance 3 sex : h um an activity m ean distance : sex m ean distance 2 : sex m ean distance 3 : sex m ean distance : h um an activity m ean distance 2 : h um an activity m ean distance 3 : h um an activity ∆ a ic c a ic c -w eight 1 x x x x x x x 0 0.102 2 x x x x x x x x 1.31 0.053 3 x x x x x x x x 1.92 0.039 4 x x x x x x x x x 2.54 0.029 table 2. the aicc-based ranking of models explaining variation in centred distance during day periods of high and low human activity using linear mixed effect models with individual as a random factor. variables included in the candidate models are marked by an x. the best model (model 1) had an aicc-value of 14597.46. ∆aicc refers to the difference in aicc between the best model and the candidate model. the global model had an aicc-value of 14605.56 (∆aicc= 8.17); only models with ∆aicc <3 are presented. moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 116 houses. they also seemed to move further away relative to their mean distance during periods of high human activity. housing density the best model explaining variation in centred density for moose observations during the 24-h period included all main and interaction effects, except for the interaction between density category and moose sex. the second best model, which was also the global model, also included this effect but had a ∆aicc-value of 3.95 and was less supported. according to this model, moose showed a daily pattern of centred density (fig. 4) that was consistent with the results of centred distance from housing (fig. 2). when mean used density was high (>2 houses/km2), both males and females used areas of lower than average density during day, and higher than average density at night. this pattern was similar for males but less pronounced both at intermediate (0.2-2 houses/km2) and low density (00.2 houses/ km2). there was little or no daily variation for females in these latter categories. in the second step, the best model included all main and interaction effects, except for the effect of moose sex on the second and third order polynomial terms of mean used density (table 3). an alternative model, that also included the interaction between the second order polynomial of mean used density and sex (model 2; table 3), had approximately the same aicc-value (∆aicc=0.01) and was equally supported based on aiccweight (aiccwratio of model 1 compared to model 2 = 1.00; table 3). this indicates that the difference between males and females may have been more pronounced than indicated by fig. 5. however, the parameter estimate for the effect of moose sex on the second order polynomial term of density was associated with a high uncertainty (beta = -0.006 ± 0.004 se), and to avoid over-parameterization (burnham and anderson 2002), the simpler model (model 1) was regarded as best. a third model with fig. 2. model estimates of daily variation in centred distance from inhabited houses for moose during summer. solid, dashed, and dotted lines represent short (0-1000 m), intermediate (1000-2000 m) and long (2000-4000 m) mean used distance, respectively. panel (a) shows the predicted response for males, and (b) the predicted response for females. fig. 3. centred distance from inhabited houses in relation to the daily mean distance used by moose during summer. black lines represent males, grey lines females. solid and dashed lines represent high and low human activity periods, respectively. the lines indicate the predicted response from the most parsimonious model (model 1 in table 2). alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 117 a ∆aicc ≤ 2 included an interaction between the third order polynomial of mean used density and sex. however, this model was almost 3x less supported than the best model (aiccwratio of model 1 compared to model 3 = 2.83; table 3). fig. 5 depicts the predicted response from the best model for variation in centred density relative to mean used density. the difference between centred densities used during periods of high versus low human activity increased with increasing mean used density up to 10 houses per km2. both males and females used higher housing density than average in the low human activity period, and males m odel sex h um an activity m ean density m ean density 2 m ean density 3 sex : h um an activity m ean density : sex m ean density 2 : sex m ean density 3 : sex m ean density : h um an activity m ean density 2 : h um an activity m ean density 3 : h um an activity ∆ a ic c a ic c-w eight 1 x x x x x x x x x x 0 0.017 2 x x x x x x x x x x x 0.01 0.017 3 x x x x x x x x x x x x 2.05 0.006 table 3. the aicc-based ranking of models explaining variation in centred density between periods of high and low human activity using linear mixed effect models with individual as a random factor. variables included in the candidate models are marked by an x. the highest ranked model (model 1) had an aicc-value of 2309.83. ∆aicc refers to the difference in aicc between the best model and the candidate model. the global model (model 3) had an aicc-value of 2311.88 (∆aicc= 2.05); only models with ∆aicc <3 are presented. fig. 4. model estimates of the daily variation in centred density of inhabited houses in the vicinity of moose. solid, dashed, and dotted lines represent high (>2 houses/km2), intermediate (0.2-2 houses/km2) and low (0-0.2 houses/km2) mean used density, respectively. panel (a) shows the predicted response for males, and (b) the predicted response for females. fig. 5. centred density of inhabited houses in the vicinity of moose in relation to the daily mean density of housing in areas used by moose during summer. black lines represent males, grey lines females. solid and dashed lines represent high and low human activity periods, respectively. the lines indicate the predicted response from the most parsimonious model presented in table 3 (model 1 in table 3). moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 118 used higher housing density than females regardless of mean used density. in the high human activity period, the model predicted that both sexes used lower density than average, except for males at very high used density. there was less difference between sexes in this period compared with the low human activity period. males seemed to use lower density than females when mean used density was <5 houses/km2, and higher density above. moreover, the model showed an upward slope for both sexes at 10-15 houses/km2; however, this should be interpreted with caution because of limited sample size. discussion our results seem to suggest that nonfatal disturbance by humans is analogous to increases in perceived predation risk (frid and dill 2002). habitat use was limited by human activity within certain distances to and densities of human activity centres. moose seemed to respond to increased levels of human activity during day by retreating to relatively safer habitats or locations of longer distance to nearest house or lower housing density. moose were affected when closer than 1500 m to the nearest house (fig. 3), and when housing density was approximately <2 houses / km2 (fig. 5). at shorter distance and higher density, moose responded by relocating to less disturbed areas at the time of day when humans were most active. we interpreted this as anti-predator behaviour that increased travel costs to move away from disturbance (formaniwicz and bobka 1988), and perhaps more importantly, reduced opportunity to forage in optimal habitat when humans are most active (creel et al. 2005). to some extent, these results may provide a conservative picture of behaviour relative to human activity due to limitations of our analytical approach. at very short mean used distances, there was a constraint to express avoidance of humans because of a lack of options to change distance to humans. when moose averaged 200 m from the nearest house, they simply could not approach closer than 200 m during the low human activity period. this effect also explains the funnel-like shape in the right hand side of fig. 5. moreover, because moose did not distribute themselves evenly in terms of mean used density, many observations were made at house density near zero, resulting in an extremely skewed distribution. in fact, as many as 467 observations had a mean used density of exactly zero, indicating there was no variation in either the response or the predictor variable. this methodological limitation concerning extremely low mean used densities (left part in fig. 5) implies that the effect of human activity may have been more pronounced than indicated in fig. 5. however, the observed distribution indicates that moose often avoided areas of high-density housing. many prey species alter their use of habitats in response to predation risk through trade-off of forage quality/quantity for increased security (abramsky et al. 1996, heithaus and dill 2002). such responses are likely to reduce fitness (werner et al. 1983). elk (cervus elaphus) in alberta, canada responded to human hunters by moving from nutritionally profitable grassy meadows to forests (morgantini and hudson 1985), provoking a significant change in diet; intake of rough fescue decreased from 87% to 34% as browsing increased. following the hunting season, elk reverted to grazing in fescue-dominated meadows. when behavioural responses to predation cause pronounced dietary change, effects on fitness are expected (creel and christianson 2008). for example, nelleman et al. (2001) investigated effects of infrastructure and associated human activity on the distribution of wild reindeer during winter in a mountain region in western norway, and found that density of reindeer was 79% lower within 2.5 km of power lines compared to background areas. available forage in terms of lichen cover declined 15-30 fold with distance, and alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 119 was lowest in the undisturbed areas with the highest density of grazing animals. they concluded that intensified grazing in zones away from disturbance lead to increased grazing pressure and lower carrying capacity. forest-dwelling species like moose are assumed to be more tolerant of human activity because forests provide more security. however, our results indicate that habitats close to human activity are used less despite the fact that such areas are often associated with higher primary productivity and potentially higher density/quality forage (solberg et al. 2006). vegetation and browsing surveys at variable distance from humans would best assess the extent to which human disturbance affects carrying capacity. our results most likely reflect a flexible response to temporal variation in predation risk, where moose seem to use areas with human disturbance when humans are less active. it is generally accepted that selection will favour individuals that appropriately balance the benefits and costs of anti-predator behaviour (lima 1998). by allocating the most intensive foraging period when predation risk is lowest, moose close to humans might obtain forage as effectively as they could in an entire day in more distant lower quality habitat. in addition, the high number of observations at relatively short mean used distances from housing (fig. 3) indicates that moose are attracted to areas with high human activity. this could be due to availability of high quality browse, but may also relate to availability of readily eaten agricultural crops. although not measured, it is possible that moose close to human activity centres engage more frequently in anti-predator behaviours such as increased vigilance (winnie and creel 2007). assessment of time spent vigilant versus feeding along gradients of temporal and spatial predation risk would be informative. the “predation risk allocation hypothesis” (lima and bednekoff 1999) provides a number of testable predictions to evaluate the proportion of time spent vigilant under such circumstances. an alternative explanation an alternative explanation for the observed trends in fig. 3 and 5 may be that moose close to humans prefer to feed on agricultural land and are indirectly attracted to human activity centres, whilst more distant moose select forest feeding patches that are disconnected to human presence. moose quite often feed in agricultural fields that are concentrated in valleys close to houses. because moose are primarily crepuscular and nocturnal foragers (andersen and sæther 1996, hanssen 2008), they move into such areas at night. what is assumed to be an anti-predator behaviour may simply reflect feeding behaviour regulated by cues like light conditions that correlate with human activity. however, the question remains why moose move away from these foraging areas during daylight hours, rather than remain nearby. we believe this is in response to human disturbance and moose perceive forest habitat as more secure to ruminate and rest. to investigate this idea we also carried out analyses where the 24-h period was divided into 2 light regime periods, light and dark, based on sunrise and sunset (results not presented here for reasons of brevity). interestingly, this approach provided less convincing results than those presented here, indicating that moose responded primarily to diurnal variation in human activity, rather than light level per se. therefore, we suggest that the daily movement pattern of moose in areas of high housing density is mainly an anti-predator behaviour. differential response of males and females the best models explaining centred distance (fig. 3) and density (fig. 5) indicated that males stay closer to houses and in areas with higher housing density than females during the low human activity period. males moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 120 also tolerated higher housing density than females during the high human activity period, especially in areas of highest house density (>5 houses/km2; fig. 5). similarly, winnie and creel (2007) found that male elk in montana, usa, showed weaker anti-predator responses than females, despite facing greater risk of predation. they concluded that antipredator behaviours carry substantial costs, which males, because of their poorer body condition, were less willing or able to pay than females. ramsrud (2007) found that parturient female moose select habitats characterized by relatively low levels of human activity (roads and houses) and of rather poor forage quality, whereas males were rather indiscriminate with respect to human activity and land cover. thus, maternal females seem to avoid humans more than males at this critical time of year. a similar pattern was observed for male and female lynx (lynx lynx) in south-eastern norway (bunnefeld et al. 2006). ultimately, these differences are likely to be a product of sexual selection. moose have a polygynous mating system (andersen and sæther 1996) where their reproductive success often depends on large body size (weckerly 1998). they are expected to maximize energy intake, growth rate, and social status to ensure mating success, whereas females trade-off growth against reproduction and maternal care (clutton-brock et al. 1988). overall, our results seem to support these expectations and we suggest that the difference in the pattern of habitat use was due to differential sex-specific risks and benefits of foraging in human disturbed areas. conclusions this study supports the hypothesis that habitat use by moose is limited by human activity in areas with high housing density. we suggest that this variation in spatial allocation is a behavioural response to higher perceived predation risk associated with humans. human infrastructure and activity may thus have negative effects on the number of moose that an area can sustain. indirect effects of predation, as demonstrated here, can be large but difficult to detect when the influence is reduced reproduction rather than survival (creel and christianson 2008). we suggest that in cost-benefit analyses, indirect effects of human activity on moose populations should be recognized and included with direct effects such as harvest and vehicular collisions. acknowledgements we are grateful to the directorate for nature management, the county governor in nord-trøndelag county, and the research council of norway for their financial support of this study, and all the financial contributors to the project “elgundersøkelsene i nordtrøndelag, bindal og rissa 2005-2009.” we also thank vidar grøtan (ntnu) for help with statistics, morten heim (nina) for organization of data, and 3 anonymous referees for constructive comments. references abramsky, z. b., e. strauss, and a. subach. 1996. the effect of barn owls (tyto alba) on the activity and microhabitat selection of gerbillus allenbyi and g. pyramidum. oecologia 105: 313-319. andersen, r., j. d. c. linnell, and r. langvatn. 1996. short term behavioural and physiological response of moose alces alces to military disturbance in norway. biological conservation 77: 169-176. _____, and b.-e. sæther. 1996. elg i norge. (moose in norway). n.w. damm & søn a.s teknologisk forlag. (in norwegian). anderson, d. r., m. g. turner, j. d. forester, j. zhu, m. s. boyce, h. beyer, and l. stowell. 2005. scale-dependent summer resource selection by reintroduced elk in wisconsin, usa. journal of wildlife management 69: 298-310. alces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 121 barta, z., a. liker, and f. monus. 2004. the effects of predation risk on the use of social foraging tactics. animal behaviour 67: 301-308. bates, d. 2007. lme4: linear mixed-effects models using s4 classes. r package version 0.99875-9. bunnefeld, n., j. d. c. linnell, j. odden, m. a. j. van duijn, and r. andersen. 2006. risk taking by eurasian lynx (lynx lynx) in a human-dominated landscape: effects of sex and reproductive status. journal of zoology 270: 31-39. burnham, k. p., and d. r. anderson. 2002. model selection and multimodel inference: a practical information-theoretical approach. second ed. springer, new york, new york, usa. cain, j. t., p. r. krausman, and b. d. jansen. 2005. influence of topography and gps fix interval on gps collar performance. wildlife society bulletin 33: 926-934. cederlund, g., f. sandegren, and k. larsson. 1987. summer movements of female moose and dispersal of their offspring. the journal of wildlife management 51: 342-352. childress, m. j., and m. a. lung. 2003. predation risk, gender and the group size effect: does elk vigilance depend upon the behaviour of conspecifics? animal behaviour 66: 389-398. ciuti, s., s. davini, s. luccarini, and m. apollino. 2004. could the predation risk hypothesis explain large-scale sexual segregation in fallow deer (dama dama)? behavioral ecology and sociobiology 56: 552-564. clutton-brock, t. h., s. d. albon, and f. e. guinnes. 1988. reproductive success in male and female deer. pages 325-343 in t. h. clutton-brock, editor. reproductive success: studies of individual variation in contrasting breeding systems. university of chicago press, chicago, illinois, usa. crawley, m. j. 2002. statistical computing: an introduction to data analysis using s-plus. john wiley & sons, chichester, england. creel, s., and d. christianson. 2008. relationships between direct predation and risk effects. trends in ecology and evolution 23: 194-201. _____, j. jr., winnie, b. maxwell, k. hamlin, and m. creel. 2005. elk alter habitat selection as an antipredator response to wolves. ecology 86: 3387-3397. dill, l. m., and j. f. gillett. 1991. the economic logic of barnacle balanus glandula (darwin) hiding behaviour. journal of experimental marine biology and ecology 153: 115-127. dussault, c., j.-p. ouellet, r. courtois, j. huot, l. breton, and h. jolicoeur. 2005. linking moose habitat selection to limiting factors. ecography 28: 619-628. fenn, m. g. p., and d. w. macdonald. 1995. use of middens by red foxes risk reverses rhythms of rats. journal of mammalogy 76: 130-136. forman, r. t. t., d. sperling, j. a. bissonette, a. p. clevenger, c. d. cutshall, v. h. dale, l. fahrig, r. france, c. r. goldman, k. heanue, j. a. jones, f. j. swanson, t. turrentine, and t. c. winter. 2003. road ecology: science and solutions. island press, washington d. c., usa. formaniwicz, d. r., and m. s. bobka. 1988. predation risk and microhabitat preference: an experimental study of the behavioural response of prey and predator. american midland naturalist 121: 379-386. fortin, d., h. l. beyer, and m. s. boyce. 2005. wolves influence elk movements: behavior shapes a trophic cascade in yellowstone national park. ecology 86: 1320-1330. fretwell, s. d. 1972. populations in a seasonal environment. princeton university moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 122 press, princeton, new jersey, usa. frid, a., and l. dill. 2002. human-caused disturbance stimuli as a form of predation risk. conservation ecology 6: 11. gill, j. a., k. norris, and w. j. sutherland. 2001a. the effects of disturbance on habitat use by black-tailed godwits limosa limosa. the journal of applied ecology 38: 846-856. _____, _____, and _____. 2001b. why behavioural responses may not reflect the population consequences of human disturbance. biological conservation 97: 265-268. _____, and w. j. sutherland. 2000. predicting the consequences of human disturbance from behaviour decisions. pages 51-64 in l. m. gosling and w. j. sutherland, editors. behaviour and conservation. cambridge university press, cambridge, uk. hanssen, m. 2008. summer movement patterns of moose (alces alces) in central norway. m. sc. thesis, norwegian university of science and technology, trondheim, norway. heithaus, m. c., and l. m. dill. 2002. food availability and tiger shark predation risk influence bottlenose dolphin habitat use. ecology 83: 480-491. histøl, t., and o. hjeljord. 1995. sørnorske elgbeiter, kvalitet og bæreevne. en vurde-en vurdering av sørnorske elgbeiter ut fra variasjon i slaktevekt, vegetasjon og klima. (moose ranges of southern norway, quality and carrying capacity). agricultural university of norway, ås. (in norwegian). kats, l. b., and l. m. dill. 1998. the scent of death: chemosensory assessment of predation risk by prey animals. ecoscience 5: 361-394. koops, m. a., and m. v. abrahams. 2003. integrating the roles of information and competitive ability on the spatial distribution of social foragers. the american naturalist 161: 586-600. lavsund, s., t. nygren, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. lima, s. 1998. nonlethal effects in the ecology of predator-prey interactions what are the ecological effects of anti-predator decision-making? bioscience 48: 2534. _____, and p. a. bednekoff. 1999. temporal variation in danger drives antipredator behavior: the predation risk allocation hypothesis. the american naturalist 153: 649-659. lucas, j., r. d. howard, and j. g. palmer. 1996. callers and satellites: chorus behaviour in anurans as a stochastic dynamic game. animal behaviour 51: 501-518. lykkja, o. n. 2008. the limiting effects of human infrastructure and activity centres on moose (alces alces) habitat use during the summer season. m. sc. thesis, norwegian university of science and technology, trondheim, norway. moen, a., a. lillethun, and a. odland. 1999. national atlas of norway: vegetation. norwegian mapping authority, hønefoss, norway. morgantini, l. e., and r. j. hudson. 1985. changes in diets of wapiti during a hunting season. journal of range management 38: 77-79. nellemann, c., i. vistnes, p. jordhøy, and o. strand. 2001. winter distribution of wild reindeer in relation to power lines, roads and resorts. biological conservation 101: 351-360. pangle, k. l., s. d. peacor, and o. e. johannsson. 2007. large nonlethal effects of an invasive invertebrate predator on zooplankton population growth rate. ecology 88: 402-412. parker, g. a., and w. j. sutherland. 1986. ideal free distributions when individuals differ in competitive ability: phenotypelimited ideal free models. animal behavalces vol. 45, 2009 lykkja et al. – moose, habitat use, and human activity 123 iour 34: 1222-1242. ramsrud, j. k. 2007. calving site selection by moose: anti-predation versus feeding conditions. m. sc. thesis, norwegian university of science and technology, trondheim, norway. r development core team (2007) r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. isbn 3-900051-07-0, url . (accessed april 2009). setsaas, t., t. holmern, g. mwakalebe, s. stokke, and e. røskaft. 2007. how does human exploitation affect impala populations in protected and partially protected areas? a case study from the serengeti ecosystem, tanzania. biological conservation 136: 563-570. silverman, b. w. 1986. density estimation for statistics and data analysis. crc press, new york, new york, usa. singer, f. j., and j. b. beattie. 1986. the controlled traffic system and associated wildlife responses in denali nationalpark. arctic 39: 195-203. solberg, e. j., c. rolandsen, m. heim, v. grøtan, m. garel, b.-e. sæther, e. b. nilsen, g. austrheim, and i. herfindal. 2006. elgen i norge sett med jegerøyne – en analyse av jaktmaterialet fra overvåkningsprogrammet for elg og det samlede sett elg-materialet for perioden 1966-2004. (moose in norway – an analysis of material collected by moose hunters 1966-2004). nina rapport 125, trondheim, norway. (in norwegian). _____, h. sand, j. d. c. linnell, s. brainerd, r. andersen, j. odden, h. brøseth, j. swenson, o. strand, and p. wabakken. 2003. utredninger i forbindelse med ny rovviltmelding. store rovvilts innvirkning på hjorteviltet i norge: økologiske prosesser og konsekvenser for jaktuttak og jaktutøvelse. (reports for the large preda-(reports for the large predator policy statement. the effects of large carnivores on wild ungulates in norway: implications for ecological processes, harvest and hunting methods). nina fagrapport 63, trondheim, norway. (in norwegian). (ssb) statistics norway. 2009. . (accessed april 2009). strand, o., k. bevanger, and t. falldorf. 2006. reinens bruk av hardangervidda. sluttrapport fra rv7-prosjektet. (wild reindeer habitat use at hardangervidda. final report from the hw7 project). nina rapport 131, trondheim, norway. (in norwegian). sæther, b.-e., e. j. solberg, and m. heim. 2003. effects of altering adult sex ratio and male age structure on the demography of an isolated moose population. journal of wildlife management 67: 455-466. tollrian, r., and d. harvell. 1999. the ecology and evolution of inducible defenses. princeton university press, princeton, new jersey, usa. tufto, j., r. andersen, and j. d. c. linnell. 1996. habitat use and ecological correlated of home range size in a small cervid: the roe deer. journal of animal ecology 65: 715-724. weckerly, f. w. 1998. sexual-size dimor-sexual-size dimorphism: influence of mass and mating systems in the most dimorphic mammals. journal of mammalogy 79: 33-52. werner, e. e., j. f. gilliam, d. j. hall, and g. g. mittelbach. 1983. an experimental test of the effects of predation risk on habitat use in fish. ecology 64: 1540-1548. white, k. s., and j. berger. 2001. antipredator strategies of alaskan moose: are maternal trade-offs influenced by offspring activity? canadian journal of zoology 79: 2055-2062. winnie, j. jr., d. christianson, s. creel, and b. maxwell. 2006. elk decision-making rules are simplified in the presence of moose, habitat use, and human activity – lykkja et al. alces vol. 45, 2009 124 wolves. behavioral ecology and sociobiology 61: 277-289. _____, and s. creel. 2007. sex-specific behavioural responses of elk to spatial and temporal variation in the threat of wolf predation. animal behaviour 73: 215-225. wolfe, s., b. griffith, and c. a. g. wolfe. 2000. response of reindeer and caribou to human activities. polar research 19: 63-73. 3903.p65 alces vol. 39, 2003 ericsson human dimensions in moose research 11 of moose and man: the past, the present, and the future of human dimensions in moose research göran ericsson department of animal ecology, swedish university of agricultural sciences, se 901 83 umeå, sweden abstract: there is a gap between a growing interest to study the moose/human interface (mhi) and the actual effort made to understand this human dimension (hd) component in moose research. a content analysis of alces 1974-2001 showed that the relative contribution of hd-papers increased until 1991 but decreased thereafter. of 66 hd-articles published, 68% of the papers covered how “man affects moose” with hunting and collisions the single most important topics, and 15% were about “values” (economic and attitudes). outside alces, articles appeared that were underrepresented in alces or in the proceedings of the north american moose conference and workshop. i identify four priority hd-areas for future studies: (1) how do people react to changing moose densities and distributions?; (2) which management alternatives are acceptable for managing the urban and suburban mhi and what makes them acceptable?; (3) how important are moose to non-consumptive users?; and (4) what are the population dynamics and attributes of the consumptive moose user and what makes moose important to consumptive users? a scientific challenge is to further merge ecological and social science to integrate this in management strategies. alces vol. 39: 11-26 (2003) key words: consumptive, economy, human, literature review, moose, natural resource use, nonconsumptive, science, sociology, wildlife management a major focus of wildlife management is managing the interactions between people and wildlife; e.g., moose (alces alces). research is a foundation for all wildlife management, and research needs to encompass mhi to properly fulfill the needs of management. all wildlife management is based on human values, with “management” itself being a human construct (decker et al. 2001). we manage moose and other wildlife because the society we live in views them as a resource. studying the human dimension (hd) component of wildlife management is not new; in fact the field arose when humans made the first attempt to manage wildlife. the western world witnessed a rise of public concern over conservation issues in the 1800s (e.g., brusewitz 1992, decker et al. 2001). the usa saw the birth of the “wise-use” movement (e.g., glifford pinchot 1865-1946) and the wildlife preservation movement (e.g., john muir 1838-1914). although these movements’ perspectives differed widely pinchot envisioned a sustainable use and muir envisioned preservation with minimal human involvement they agreed that human activity and behavior had to be regulated in relation to natural resource use. out of pinchot’s ideas grew natural resource policies which have influenced, and continue to influence, natural resource policy worldwide. pinchot’s ideas resulted in policies that centered on the resource itself and promoted development and inclusion of scientific biological knowledge. with respect to moose, active management began during the pinchot/muirera when people started to recognize that human activity (i.e., hunting) had to be human dimensions in moose research ericsson alces vol. 39, 2003 12 regulated to prevent moose from becoming extinct. in eastern north america, moose populations went towards local extinction because of unregulated commercial and subsistence hunting during the 1800s (wolfe 1987). in the swedish-norwegian union, the parliament imposed a 10-year ban on moose hunting in 1825 when moose were at very low numbers (brusewitz 1992). thus, as early as the 1850s, people had started to manage moose both in europe and north america (karns 1998). the hd-core issues how and why people value wildlife and wildlife management actions, and peoples’ motivations behind consumptive and non-consumptive use of wildlife first emerged in the early 1970s (hendee and potter 1971). the 1970s has been termed “an era of discovery and organization” (brown and decker 2001). hendee (1969) and hendee and potter (1971) opened up and vitalized the hd-field by suggesting several topics for research such as hunting satisfaction, population dynamics and characteristics of hunters, nonconsumptive use of wildlife, and wildlife economics. hendee thereafter stimulated the discipline with several important contributions about consumptive as well as nonconsumptive use of game and wildlife (e.g., potter et al. 1973, hendee 1974, hendee and burdge 1974). during 1973-1984 the hd-field of wildlife continued to expand and attracted new thinkers and environments that further developed the field (brown and decker 2001). most of this research occurred in north america. europe was a few steps behind: there were important non-peer reviewed national reports which were significant, but the international, peer-reviewed arena saw few, if any, contributions before the mid1980s (see norling 1987). starting in 1987, s e v e r a l i m p o r t a n t p a p e r s o n t h e bioeconomics of moose came out of sweden and norway (e.g., norling 1987; mattsson 1990 a, b; storaas et al. 2001). in 1984, the hd-field of wildlife management expanded but remained vaguely defined and largely lacked qualitative and quantitative studies (wolfe 1987, brown and decker 2001). still, 1984 was a landmark year for the first time there were enough papers for specific socioeconomic sessions at the north american wildlife and natural resource conference and at the second international moose symposium. now, 18 years later i revisit the concerns and predictions that wolfe (1987) identified at the second international moose symposium in 1984. wolf concluded “that existing information on most of the topics [e.g., how humans affect moose; values] covered is inadequate” (wolfe 1987). however, he was not that worried over the scarce existing knowledge and saw no reason for “embarrassment” as the discipline was young but fast emerging. his closing sentence was full of trust for the future: “hopefully, a decade hence, the amount and quality of information will have improved to the point, where we can provide more definitive answers to the questions raised here” (wolf 1987:673). what did the future actually bring? first, i address the hypothesis that the hd-component of moose management has increased between 1984 and 2001. my prediction, based on the general expansion of hd in wildlife management (brown and decker 2001), is that the amount of socioeconomic information about moose has increased since 1984. i test this by content analyses of published articles in alces 1974-2001 and in the proceedings of the international moose symposia 1984, 1990, 1998, and 2002 (papers presented at the 2002 meeting). also, i predict that the hd-field received relatively more attention in years with international moose symposia (sensu wolfe 1987). second, testing information quality is alces vol. 39, 2003 ericsson human dimensions in moose research 13 difficult to do objectively. i visited the peerreviewed periodic literature dated 1984 through 2001 to see if outlets other than alces had published hd-articles regarding moose and man. i assumed increasing outlets for moose hd-research results may indicate an improved quality as well as expansion of moose hd expertise. my prediction, based on the general expansion of the field in wildlife management (brown and decker 2001), is that the quality of hd and mhi information has increased since 1984 as evidenced by increasing numbers of outlets for publications. third, i identify areas for future research in the hd/mhi field. methods i have organized the paper according to the most widely adopted definition of human dimensions (decker et al. 2001). human dimensions is how people value moose, how people want moose to be managed, and how people are affected by or affect moose or moose management decisions. i reviewed all papers in alces volumes 10-37, the proceedings of the international moose symposia 1984, 1990, 1998, and presented oral contributions at the 2002 symposium. i classified only full-length articles, not summaries from workshops or abstracts. articles about moose where classified as a hd contribution if they focused on: (1) values (economic, attitudinal); (2) human effects on moose (hunting, infrastructure, and traffic/collisions with vehicles); (3) moose hunters; or (4) miscellaneous hd-studies. to category (2) how humans affect moose i classified articles about selective harvest. because of the definition adopted, study systems where humans are a driving force behind the population dynamics of moose are not included in my review (e.g., ericsson 2001). originally, i planned for a fifth category (moose affects man) but no articles specifically addressed the topic (besides vehicle collisions which i placed in category (2)). articles describing general moose management in a particular state/ country, surrogate biology (e.g., hunter surveys to assess moose harvest), predator control, and potential human health impacts due to moose meat consumption were not classified as hd-articles (but see, for example, cretê et al. 1987, danell et al. 1989, kim et al. 1998). using the electronic databases agricola, agris, biosis, cab, econlit, fsta, and treecd, i searched the peer-reviewed periodic literature for papers about moose and human interactions published from 19842001. when testing for trends over time in the publication alces, the number of hd-articles was described as a proportion of all articles. the random variation in a time series can make it difficult to detect an underlying trend, thus smoothing is a standard technique of emphasizing or describing a trend (brown and rothery 1993). in my case, as the publication year of a single article may be a random event, a better measure of development of hd in moose research may be a 5-year moving average. by that i implicitly assume that a single year is representative of what had been done the previous 2 years and at least 2 years thereafter. i performed all statistical analyses with sas (sas institute 1989). results content analyses alces.— in alces 1974-2001 (volumes 10-37) 584 articles were published of which 11% (n = 66) were hd-articles (table 1). in the proceedings of the international symposia (1984, 1990, 1998), and of articles presented at the 5th international symposium 2002, 9% (17 of 148) were hdarticles (table 2). i found no difference in the proportion of hd-articles in alces compared to the proceedings (p = 0.46, t-test). human dimensions in moose research ericsson alces vol. 39, 2003 14 t ab le 1 . c on te nt a na ly si s o f a lc es v ol um es 1 037 (1 97 420 01 ). v al ue s m an a ff ec ts m oo se y ea r/v ol . pa pe rs h d -p ap er s h d -r at io 5yr . m ea n ec on om ic a tt it ud es h un tin g in fr as tr uc tu re tr af fic m oo se h un te rs m is c 20 01 /3 7 23 4 17 % 1 1 1 1 20 00 /3 6 21 0 0% 19 99 /3 5 19 3 6% 10 % 1 2 19 98 /3 4 23 1 4% 8% 1 19 97 /3 3 19 2 11 % 12 % 1 1 19 96 /3 2 17 2 12 % 10 % 2 19 95 /3 1 26 4 15 % 11 % 1 2 1 19 94 /3 0 23 2 9% 13 % 1 1 19 93 /2 9 31 3 10 % 18 % 1 2 19 92 /2 8 22 4 18 % 16 % 4 19 91 /2 7 25 10 1 40 % 20 % 1 9 19 90 /2 6 20 1 5% 20 % 1 19 89 /2 5 19 5 26 % 18 % 1 2 1 1 19 88 /2 4 26 3 12 % 13 % 1 1 1 19 87 /2 3 14 1 7% 13 % 1 19 86 /2 2 21 3 14 % 9% 1 1 1 19 85 /2 1 26 1 4% 12 % 1 19 84 /2 0 16 1 6% 12 % 1 19 83 /1 9 17 5 29 % 11 % 4 1 19 82 /1 8 17 1 6% 11 % 1 19 81 /1 7 18 2 11 % 12 % 1 1 19 80 /1 6 31 1 3% 7% 1 19 79 /1 5 16 2 13 % 6% 2 19 78 /1 4 15 0 0% 5% alces vol. 39, 2003 ericsson human dimensions in moose research 15 t ab le 1 . c on ti nu ed v al ue s m an a ff ec ts m oo se y ea r/v ol . pa pe rs h d -p ap er s h d -r at io 5yr . m ea n ec on om ic a tt it ud es h un tin g in fr as tr uc tu re tr af fic m oo se h un te rs m is c 19 77 /1 3 22 1 5% 6% 1 19 76 /1 2 14 1 7% 5% 1 19 75 /1 1 26 2 8% 2 19 74 /1 0 17 1 6% 1 t ot al 58 4 66 11 % 4 6 31 1 13 4 7 % o f h d -a rt ic le s 6% 9% 47 % 2% 20 % 6% 11 % 1 i nd ic at es s pe ci al th em e is su e. t ab le 2 . c on te nt a na ly si s of p ro ce ed in gs o f t he in te rn at io na l m oo se s ym po si a 19 84 , 1 99 8, 1 99 0, a nd 2 00 2. v al ue s m an a ff ec ts m oo se y ea r/v ol . pa pe rs h d -p ap er s h d -r at io ec on om ic a tt it ud es h un tin g in fr as tr uc tu re tr af fic m oo se h un te rs m is c 20 00 /5 th 55 9 16 % 1 2 2 2 2 19 98 /4 th 21 0 0% 19 90 /3 rd 22 1 5% 1 19 84 /2 nd 50 7 14 % 4 2 1 t ot al 14 8 17 9% 5 0 4 0 3 2 3 % o f h d -a rt ic le s 29 % 0% 24 % 0% 18 % 12 % 18 % human dimensions in moose research ericsson alces vol. 39, 2003 16 in alces, the number of hd-articles were not simply linearly related to year (model: #hdarticles = #articles year; p model = 0.23; fig. 1). using the moving average publication rate in the time-series analysis, the proportion of hd articles has increased over time (p model = 0.02, r 2 adj = 0.188, n = 26; fig. 2). inspection of fig. 2 suggests two trends, with 1990 as the cutoff year; an increasing trend 1976-1990 and a decreasing trend 1991-1999. of the 66 hd-articles published in alces 1974-2001, the majority (68%, n = 45) were published in category (2) “man affects moose” (hunting, n = 31; traffic, n = 13; and, infrastructure, n = 1). the second largest contributing category was category (1) “values” with 15% (n = 10) of the hd-articles. “miscellaneous” hd-articles (category 4) with 11% (n = 7), and hunting hd-articles (3) 6% (n = 4). outside the alces world. —thirtytwo hd-articles were found in alternate journals published 1984-2001. a majority (66%, n = 21) of the hd-articles were from category (1), 25% (n = 8) were in category (2), 6% (n = 2) from category (3), and 1 in category (4). articles dealing with “values” were more heavily represented in the alternate literature than alces (66% vs. 15%) journals that published more than one peer-reviewed article in the hd-area were arctic (n = 2), canadian journal of forestry (n = 3), journal of wildlife management (n = 2), scandinavian journal o f f o r e s t e c o n o m i c s ( n = 2 ) , scandinavian journal of forestry research (n = 3), society and natural resources (n = 3), and the wildlife society bulletin (n = 5). literature review values. — to make proper and socially justifiable decisions about consumptive and non-consumptive use, values have to be assigned and moose management is no exception. when we talk about “values” we normally mean either peoples’ thoughts and actions (i.e., attitudes and behavior towards something like moose) or economic fig. 1. the proportion of full-length human dimension articles in relation to the total number of articles in alces 1974-2001, and in the proceedings of the international symposia 1984, 1990, 1998, and 2002. filled circles represent alces, filled squares represent proceedings 1984, 1990, and 1998, open square represents articles presented at the 2002 international symposium. fig. 2. the proportion (5-year moving average) of full-length human dimension articles in relation to the total number of articles in alces 1974-2001, and in the proceedings of the international symposia 1984, 1990, 1998, and 2002. filled circles represent alces, filled squares represent proceedings 1984, 1990, and 1998, open square represents articles presented at the 2002 international symposium. the solid lines represent the positive trend in hd publication rate 1974-1990, the broken line the trend 1991-1999. 0% 10% 20% 30% 40% 1973 1978 1983 1988 1993 1998 2003 publication year alces % h d a rt ic le s 0% 5% 10% 15% 20% 25% 1973 1978 1983 1988 1993 1998 2003 publication year alces % h d a rt ic le s alces vol. 39, 2003 ericsson human dimensions in moose research 17 values (brown et al. 2001, pierce et al. 2001). the economic value of moose can be classified into two major categories (steinhoff et al. 1987); exercised values (direct and indirect benefits of moose) and option values (to enjoy, to experience, or to use moose in the future). studies dealing with exercised value of moose dominate the literature. the first author to systematically identify, classify, and describe the potential role of moose as a recreational resource was david lime in 1975, although single topics had been touched upon earlier (cobus 1972). lime discussed several potential exercised and option values in his analysis of moose as a non-consumptive resource. in the years thereafter little new information was published aside from reviews by bisset (1987), wolfe (1987), and timmermann and buss (1998). despite this, little up-to-date hard facts exist today about moose as a nonconsumptive resource. timmermann and buss (1998) suggested that “attributing economic value to moose, specifically nonconsumptive expenditures, is difficult.” judging from published studies i do not fully agree. non-consumptive use, and especially valuation of wildlife, receive considerable attention worldwide (e.g., decker and goff 1987, mattsson and li 1993, brown et al. 2001). thus, the methods and techniques are there. i believe it is more a question of involving social science in moose research and management. studies focusing on exercised values of moose, with special reference to moose hunting and hunters, started to appear in the mid-1980s. at the second international moose conference 1984, bisset (1987), bluzma (1987), norling (1987), and wolfe (1987) provided comprehensive reviews from north america, sweden, and russia. shortly thereafter followed a series of [bio]economic studies (johansson et al. 1988; sødal 1989; mattsson 1989, 1990a, 1990b; mattson and li 1993; ericsson et al. 2000; storaas et al. 2001). mattsson’s studies are interand intradisciplinary important as they integrate ecological data with economic data, and deal with values beyond meat, license revenues, gadgets, and travel cost. central questions in mattsson’s work are determining the marginal benefit of a moose and how this changes when the number of moose varies over time. however, few recent studies exist in this field besides work from canada (newfoundland, condon and adamowicz 1995; saskatchewan, morton et al. 1995; ontario, sarker and surry 1998) and the usa. (maine, boyle and clark 1993). much work needs to be done in the field of economic values and valuation. especially needed are calculations of net economic benefit that take into account consumptive as well as non-consumptive use of moose in relation to damage, collisions, forestry management (euler 1975, boxall et al. 1996, boxall and macnab 2000, courtois et al. 2001), and cost of moose management in relation to policy implementation (ericsson et al. 2000). also desirable are studies that address option values of moose in relation to, for example, various biodiversity goals and policies. studies that target the asymmetric costs and benefits in moose management are also needed (ericsson et al. 2000, skonhoft 2002). new, updated national studies are required for all jurisdictions. attitudes. — a good example of both human attitudes towards moose and citizen participation in moose management was the reintroduction proposal for new york state, usa, 1992. new york state used several means (public meetings at various stages, eis, public comment period, public surveys, etc.) to ask the public if they should proceed with a moose reintroduction proposal or not (lauber and knuth 1997, 1998, 1999). this process of public involvement is becoming human dimensions in moose research ericsson alces vol. 39, 2003 18 increasingly important for moose managers and policy makers. two stimulating papers have applied solid social science theory, addressing both beliefs and attitudes towards moose hunting. both papers go beyond the limited question “do you support or oppose moose hunting?”. donnelly and vaske (1995) studied people’s beliefs and attitudes towards a proposed moose hunt in new hampshire, and whittaker et al. (2001) studied beliefs and attitudes about an urban moose hunt near anchorage, alaska. with respect to moose hunters, studies have addressed hunter satisfaction with management, such as the introduction of selective harvest regimes, but only for ontario and quebec, canada, and for maine, usa (ontario, rollins 1987, rollins and romano 1989, wedels et al. 1989, hansen et al. 1995; quebec, sigouin et al. 1999; maine, boyle et al. 1993). thus, there is little current knowledge worldwide about the perhaps most important moose manager, the hunter. man affects moose hunting. — studies, often ad-hoc to moose population studies, have addressed general topics about regulated or selective moose harvest (child 1983, euler 1983, gollat and timmermann 1983, pierce et al. 1985, timmermann and gollat 1986, cretê 1987, child and aitken 1989, boer 1991, heydon et al. 1992, timmermann and whitlaw 1992, hooper and wilton 1995, wilton 1995, ferguson and messier 1996, goudreault and milette 1999, lamoureux 1999, schwartz 2002), as well as more specific topics such as demographic consequences of selective harvest (e.g., baker 1975, schwartz et al. 1992, wilton 1992, timmermann and rempel 1998, ericsson 2001), trophy-management (smith et al. 1979), miscellaneous moose hunting “experiments” (crichton 1979, 1980), impact of native hunting (crichton 1981, feit 1987), moose alertness in relation to hunting (bangs et al. 1984, wilton and bisset 1988, garner et al. 1990, ericsson and wallin 1996, baskin et al. 2002), moose social structure in relation to hunting (timmermann and gollat 1994), and the impact of forest management practices (eason 1989). moose hunters. — some studies have systematically examined the characteristics of moose hunters and human aspects of moose hunting (e.g., timmermann 1977, euler 1985, norling 1987, child and aitken 1989, ferguson et al. 1989, redmond et al. 1997, ball et al. 1999, heikkilä and aarnio 2001, broman et al. 2002, gåsdal and rysstad 2002). however, we still know relatively little about the composition, demography, population dynamics, and socioeconomic characteristics of the human predator. infrastructure. – humans considerably affect moose through society’s infrastructure. collisions between moose and vehicles on roads and railroads have received attention in both alces and other peer-reviewed journals. the contributions have evolved, along with moose population increases, from quantifying the problem to evaluating the means and efforts to reduce the number of deadly impacts (e.g., björnstig et al. 1986, child and stuart 1987, andersen et al. 1991, becker and grauvogel 1991, child et al. 1991, del frate and spraker 1991, jaren et al. 1991, lavsund and sandegren 1991, mcdonald 1991, modafferi 1991, rattey and turner 1991, jolicoeur and crête 1994, belant 1995, farrel et al. 1996, garret and conway 1999, joyce and mahoney 2001, rea and child 2002). few studies have dealt with urban sprawl and the effect of human settlement (schneider and wasel 2000), hydroelectric projects (ballard et al. 1988), or potential barriers to moose movements, dispersal and migration, and habitat use (mcdonald 1991, forman alces vol. 39, 2003 ericsson human dimensions in moose research 19 and deblinger 2000, ball and dahlgren 2002). all would benefit from further research. response to other forms of human disturbance such as military activity (andersen et al. 1996) and snowmobile traffic (colescott and gillingham 1998) have received little attention relative to the amount of studies addressing human disturbance to other deer species. current studies usually describe moose behavioral response directly after the construction of potential barriers, but management and infrastructure planners need more than snapshots of reality, thus the need for longer term data collection. however, only recently have gps and gis-techniques been used to better understand the interaction of landscape and temporal features of moose and human infrastructure (e.g., gundersen et al. 1998). discussion the analysis and literature review indicates a declining proportion of human dimension publications in moose research and management over the last 10 years. with the exception of the 5th international symposium, years with an international symposium do not seem to stimulate a higher proportion of hd-contributions (fig. 2). wolfe’s (1987) prediction of an expanding hd-area was fulfilled up to 1991; thereafter the relative hd-publication rate in alces has declined (fig. 2). however, people have used outlets other than alces and the international symposia, but this does not fully compensate for the decreasing trend in alces since 1991; e.g., just 5 hd-moose papers appeared in the wildlife society bulletin and 2 in the journal of wildlife m a n a g e m e n t 1 9 8 4 2 0 0 2 . h o w e v e r , present data suggest that alternate journals are publishing hd-papers which are underrepresented in alces. recently moose papers written by social scientists (n = 3) have appeared in the journal “society and natural resources”. publishing moose papers in alternate journals is an exciting opportunity to strengthen the interdisciplinary interface among biologists, sociologists, economists, anthropologists, political scientists, law scholars, and others. the literature review suggests there is an increasing involvement in the moose world by these other disciplines. this undoubtedly has strengthened the quality of the hd-component in moose research since 1984, although the quantity of hd-papers has declined. it is important for the future that we continue to attract more professionals from other disciplines than natural science. a sure result of more attendees from a wider array of disciplines will be relatively more hd-contributions in alces that will benefit moose management worldwide. future directions the complexity of moose management has increased since 1984, as has the need for a better understanding of human dimensions. this is demonstrated by the presentations at the 5th international moose symposium, 2002. more scientific in-depth assessments of public attitudes toward moose and moose management are needed as moose populations expand into new areas, and move in closer to urban areas. hunting is a behavior, and hunters have strong attitudes about moose hunting (e.g., rollins and romano 1989, hansen et al. 1995). anti-hunting is, for most people, only an attitude a feeling and a set of beliefs about an object like moose hunting. attitudes are the broad structures that underlie behaviors. a behavior is like the tip of an iceberg, while the attitude is the part under water: one you can see, the other you must infer. social scientists often use surveys to infer people’s attitudes. attitudes become important when they are expressed in behavior. it is the behavioral expression of attitudes by the public that human dimensions in moose research ericsson alces vol. 39, 2003 20 often concerns managers and policy makers. it thus becomes crucial to understand the beliefs and dimensions of attitudes among consumptive and non-consumptive users of moose (whittaker et al. 2001). moose managers need to determine what policies are acceptable to the public, and the likely degree of opposition to existing and potential policies. from my data analysis and literature review i have identified several major trends specific to moose research and management that deserve top priority in the near future; 1. moose populations continue to increase in size, expand to new areas, and in some places are being re-introduced to former habitats. in some places, such as the baltic countries, moose populations are decreasing in size and distribution. how do people react to these changing moose populations? 2. with rebounding populations, moose will cause more nuisance and damage. what management alternatives are acceptable to handle conflicts between moose and humans? what makes them acceptable to hunters and the general public? 3. how important are moose to non-consumptive users in relation to other natural resources? 4. the consumptive use of moose will continue to be of central importance. what are the motivations behind the consumptive use of moose? how does the human predator function and think? what are the dynamics of the moose hunter population? conclusions despite the growing interest in how humans affect the ecological system in today’s society, there is still a gap between this awareness and the effort made to understand the human dimension of natural resource use. there is a recognized need in moose management to integrate traditional population ecology and the social sciences (e.g., sociology, psychology, and economics). among others, crichton, regelin, franzmann, and schwartz identified this as t h e “ t h e m a n a g e m e n t c h a l l e n g e ” (1998:659). to fulfill the two primary goals of moose management development of resources and optimization of public benefits we need more information about the human user. although humans play a significant and recognized role in moose management, the necessity of collecting and utilizing information about the human user has seldom been properly integrated into moose research. if we wish to manage moose optimally, we must better incorporate the human dimension. acknowledgements i gratefully acknowledge the fulbright commission and the swedish foundation for international cooperation in research and higher education (stint) for funding. i thank the department of rural sociology and the kemp natural resources station at the university of wisconsin-madison for providing facilities to collect and analyze these data. i thank the swedish research council for environment, agricultural sciences and spatial planning (formas) for support during write up and presentation of the results. constructive comments to earlier versions of the paper by john p. ball, lars edenius, ed telfer, and two anonymous reviewers greatly improved the quality of the paper. references andersen, r., j. d. c. linell, and r. langvatn. 1996. short term behavioural and physiological response of moose alces alces to military disturbance in norway. biological conservation 77:169-176. , b. wiseth, p. h. pedersen, and v. jaren. 1991. moose-train collisions: effects of environmental conditions. alces 27:79-84. alces vol. 39, 2003 ericsson human dimensions in moose research 21 baker, r. a. 1975. biological implications of a bull-only moose hunting regulation for ontario. proceedings of the north american moose conference and workshop 11: 464-476. ball, j. p., and j. dahlgren. 2002. browsing damage on pine (pinus sylvestris and p. contorta) by a migrating moose (alces alces) population in winter: relation to habitat composition and road barriers. scandinavian journal of forest research 17:427-435. , g. ericsson, and k. wallin. 1999. climate change, moose and their human predators. ecological bulletin 47:178187. ballard, w. b., a. f. cunning, and j. s. whitman. 1988. hypotheses of impacts on moose due to hydroelectric projects. alces 24:34-47. bangs, e. e., t. n. baily, and m. f. portner. 1984. bull moose behavior and movements in relation to harvest on the kenai national wildlife refuge. alces 20:187209. baskin. l., j. p. ball, and k. danell. 2002. moose escape behaviour in areas of high hunting pressure. presented at the 5th international moose symposium, hafjell, norway, august 2002. becker, e. f., and c. a. grauvogel. 1991. relationship of reduced train speed on moose-train collisions in alaska. alces 27:161-168. belant, j. l. 1995. moose collisions with vehicles and trains in northeastern minnesota. alces 31:45-52. bisset, a. r. 1987. the economic importance of moose (alces alces) in north america. swedish wildlife research supplement 1:677-698. björnstig, u., a. eriksson, j. thorson, and p.-o bylund. 1986. collisions with passenger cars and moose, sweden. american journal of physical health 76:460-462. bluzma, p. p. 1987. socio-economic significance of moose in the ussr. swedish wildlife research supplement 1:705723. boer, a. h. 1991. hunting: a product or a tool for wildlife managers? alces 27:7984. boxall, p. c., w. l. adamowicz, j. swat, m. williams, and j. louvered. 1996. a comparison of stated preference methods for environmental valuation. ecological economics 18:243-253. , and b. macnab. 2000. exploring the preferences of wildlife recreationists for features of boreal forest management: a choice experiment approach. canadian journal of forest research 30:1931-1941. boyle, k. j., and a.g. clark. 1993. does getting a bull significantly increase value? the net economic value of moose hunting in maine. alces 29:201211. , r. l. dressler, a. g. clark, and m. f. teisl. 1993. moose hunter preferences and setting season timings. wildlife society bulletin 21:498-504. broman, e., k. wallin, and m. broberg. 2002. shortcomings and the successes of the malawi principle in swedish moose management. presented at the 5th international moose symposium, hafjell, norway, august 2002. brown, d., and p. rothery. 1993. models in biology: mathematics, statistics and computing. john wiley and sons, london, england. brown, t. l., and d. j. decker 2001. evolution of human dimension interest. pages 23-38 in d. j. decker, t. l. brown, and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. , , and w. f. siemer. 2001. valuing wildlife: economic perspectives. human dimensions in moose research ericsson alces vol. 39, 2003 22 pages 57-73 in d. j. decker, t. l. brown, and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. brusewitz, g. 1992. from the middle ages to the present. pages 40-49 in r. bergström, h. huldt, and u. nilsson, editors. swedish game biology and management. svenska jägareförbundet, spånga, sweden. child, k. n. 1983. selective harvest of moose in the omineca: some preliminary results. alces 19:162-177. , and d. a. aitken. 1989. selective harvests, hunters and moose in central british columbia. alces 25:8197. , s. p. barry, and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27:41-49. , and k. m. stuart. 1987. vehicle and train collisions fatalities of moose: some management and socio-economic considerations. swedish wildlife research supplement 1:699-702. cobus, m. 1972. moose as an aesthetic resource and their summer feeding behavior. proceedings of the north american moose conference and workshop 8:244-275. colescott, j. h., and m. p. gillingham. 1998. reaction of moose (alces alces) to snowmobile traffic in the greysriver valley, wyoming. alces 34:329-338. condon, b., and w. adamowicz. 1995. the economic value of moose hunting in newfoundland. canadian journal of forest economics 25:319-328. courtois, r., j.-p. ouellet, and a. bugnet. 2001. moose hunters’ perception of forest harvesting. alces 37:19-34. cretê, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1:553-563. , f. p o t v i n , p . w a l s h , j . l . benedetti, m. a. lefebvre, j. p. weber, g. paillard, and j. gagnon. 1987. pattern of cadmium contamination in the liver and kidneys of moose and white-tailed deer in québec. science of the total environment 66:45-53. crichton, v. f. j. 1979. an experimental moose hunt on hecla island, manitoba. proceedings of the north american moose conference and workshop 15:245-279. . 1980. manitoba’s second experimental moose hunt on hecla island. proceedings of the north american moose conference and workshop 16:489-526. . 1981. the impact of treaty indian harvest on a manitoba moose herd. alces 17:56-63. , w. e. regelin, a. w. franzmann, and c. c. schwartz. 1998. the future of moose management and research. pages 655-663 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. danell, k., p. nelin, and g. wickman. 1989. cesium 137 in northern swedish moose: the first year after the chernobyl accident. ambio 108-111. decker, d. j., t. l. brown, and w. l. siemer. 2001. evolution of peoplewildlife relations. pages 3-21 in d. j. decker, t. l. brown, and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. , and g. r. goff, editors. 1987. valuing wildlife. economic and social perspectives. westview press, boulder, colorado, usa. del frate, g. g., and t. h. spraker. 1991. alces vol. 39, 2003 ericsson human dimensions in moose research 23 moose vehicle interactions and an associated public awareness program on the kenai peninsula, alaska. alces 27:1-7. donnelly, m. p., and j. j. vaske. 1995. predicting attitudes toward a proposed moose hunt. society and natural resources 8:307-319. eason, g. 1989. moose response to hunting and 1km2 block cutting. alces 25:6374. ericsson, g. 2001. reduced cost of reproduction in moose alces alces through human harvest. alces 37:61-69. , m. boman, and l. mattsson. 2000. selective versus random moose harvesting: does it pay to be a prudent predator? journal of bioeconomics 2:117-132. , and k. wallin. 1996. the impact of hunting on moose movements. alces 32:31-40. euler, d. 1975. the economic impact of prescribed burning on moose hunting. journal of environmental management 3:1-5. . 1983. selective harvest, compensatory mortality and moose in ontario. alces 19:148-161. . 1985. moose and man in northern ontario. forestry chronicle 61: 176179. farrel, t. m., j. e. sutton, d. e. clark, w. r. horner, k. i. morris, k. s. finison, g. e. menchen, and k. h. cohn. 1996. moose-motor vehicle collisions, an increasing hazard in new england. archives of surgery 131:377381. feit, h. a. 1987. north american native hunting and management of moose populations. swedish wildlife research supplement 1:25-42. ferguson, s. h., w. e. mercer, and s. m. oosenburg. 1989. the relationship between hunter accessibility and moose condition in newfoundland. alces 25:36-47. , and f. messier. 1996. can human predation of moose cause population cycles? alces 32:149-162. forman, r. t. t., and r. d. deblinger. 2000. the ecological road-effect zone of a massachusetts (u.s.a) suburban highway. conservation biology 14:3646. garner, d. l., m. l. wilton, and k. a. gustafson. 1990. importance of moose immigration into a heavily hunted area from an unhunted area. alces 26:3036. garret, l. c., and g. a. conway. 1999. characteristics of moose-vehicle collisions in anchorage, alaska, 1991-1995. journal of safety research 30:219-223. gåsdal, o., and s. rysstad. 2002. moose hunting as a recreational common pool good. presented at the 5th international moose symposium, hafjell, norway, august 2002. gollat, r., and h. r. timmermann. 1983. determining quotas for a moose selective harvest in northcentral ontario. alces 19:191-203. goudreault, f., and j. milette. 1999. hunting pressure and rate of increase of a moose population at a density below carrying capacity. alces 35:165176. gundersen, h., h. p. andreassen, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34:385-394. hansen, s., w. j. dalton, and t. stevens. 1995. on overview of a hunter opinion survey of satisfaction with the ontario moose management system. alces 31:247-254. heikkilä, r., and j. aarnio. 2001. forest owners as moose hunters in finland. alces 37:89-96. hendee, j. c. 1969. appreciative versus human dimensions in moose research ericsson alces vol. 39, 2003 24 consumptive uses of wildlife refuges: studies of who gets what and trends in use. transactions of the north american wildlife and natural resources conference 34:252-264. . 1974. a multiple satisfaction approach to game management. wildlife society bulletin 2:104-113. , and r. j. burdge. 1974. the substitutability concept: implication for recreation and management. journal of leisure research 6:157-162. , and d. r. potter. 1971. human behavior and wildlife management: needed research. transactions of the north american wildlife and natural resources conference 36:383-396. heydon, c., d. euler, h. smith, and a. bisset. 1992. modelling the selective moose harvest program in ontario. alces 28:111-122. hooper, c. a., and m. l. wilton. 1995. a selective moose hunt in south central ontario. alces 31:139-144. jaren, v., r. andersen, m. ulleberg, p. h. pedersen, and b. wiseth. 1991. moosetrain collisions: the effect of vegetation removal with a cost-benefit analysis. alces 27:93-99. johansson, p.-o., b. kriström, and l. mattsson. 1988. how is the willingness-to-pay for moose hunting affected by the stock of moose? an empirical study of moose-hunters in the county of västerbotten. journal of environmental management 26:163-171. jolicoeur, h., and m. crête. 1994. failure to reduce moose-vehicle accidents after a partial drainage of roadside salt pools in québec. alces 30:81-89. joyce, t. l., and s. p. mahoney. 2001. spatial and temporal distributions of moose-vehicle collisions in newfoundland. wildlife society bulletin 29:281291. karns, p. d. 1998. population distribution, density and trends. pages 125-140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. kim, c., m. c. hing, and o. receveur. 1998. risk assessment of cadmium exposure in fort resolution, northwest territories, canada. food additives and contaminants 15:307-317. lamoureux, j. 1999. effects of selective harvest on moose populations of the bas-saint-laurent region, québec. alces 35:191-203. lauber, t. b., and b. a. knuth. 1997. fairness in moose management decision-making: the citizens’ perspective. wildlife society bulletin 25:776-787. , and . 1998. refining our vision of citizen participation: lessons from a moose reintroduction proposal. society and natural resources 11:411424. , and . 1999. measuring fairness in citizen participation: a case study of moose management. society and natural resources 11:19-37. lavsund, s., and f. sandegren. 1991. moose-vehicle relations in sweden: a review. alces 27:118-126. mattsson, l. 1989. the economic value of wildlife for hunting. scandinavian forest economics 30:42-61. . 1990a. hunting in sweden: extent, economic values and structural problems. scandinavian journal of forest research 5:563-573. . 1990b. moose management and the economic value of hunting: towards bioeconomic analysis. scandinavian journal of forest research 5:575-581. , and c.-h. li. 1993. the nontimber value of northern swedish fore s t s . a n e c o n o m i c a n a l y s i s . scandinavian journal of forest realces vol. 39, 2003 ericsson human dimensions in moose research 25 search 8:426-434. mcdonald, m. g. 1991. moose movement and mortality associated with the glenn highway expansion, anchorage alaska. alces 27:208-219. modafferi, r. d. 1991. train moose-kill in alaska: characteristics and relationship with snowpack depth and moose distribution in lower susitna valley. alces 27:193-207. morton, k. m., w. l. adamowicz, and p. c. boxall. 1995. economic effects of environmental quality change on recreational hunting in northwestern saskatchewan: a contingent behaviour analysis. canadian journal of forest economics 25: 912-920. norling, i. 1987. moose hunting in sweden, description and evaluation. swedish wildlife research supplement 1:725733. pierce, c. l., m. j. manfredo, and j. j. vaske. 2001. social science theories in wildlife management. pages 39-56 in d. j. decker, t. l. brown, and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. pierce, d. j., b. w. ritcie, and l. kuck. 1985. an examination of unregulated harvest of shiras moose in idaho. alces 21:231-252. potter, d. r., j. c. hendee, and r. n. clark. 1973. hunting satisfaction: game, guns or nature? transactions of the north american wildlife and natural resources conference 38:220-229. rattey, t. e., and n. e. turner. 1991. vehicle moose accidents in newfoundland. journal of bone and joint surgery 73-a:1487-1491. rea, r. v., and k. child 2002. human dimension in new hampshire’s moose management. presented at the 5th international moose symposium, hafjell, norway, august 2002. redmond, g. w., a. arseneault, and c. lanteigne. 1997. using ivr technology to survey moose hunters in new brunswick. alces 33:75-84. rollins, r. 1987. hunter satisfaction with the selective harvest system for moose in northern ontario. alces 23:181-193. , and l. romano. 1989. hunter satisfaction with the selective harvest system for moose management in ontario. wildlife society bulletin 17:470475. sarker, r., and y. surry. 1998. economic value of big game hunting: the case of moose hunting in ontario. journal of forest economics 4:29-60. sas institute. 1989. sas/stat®. user’s guide. version 6. fourth edition. volumes 1 and 2. sas institute incorporated, cary, north carolina, usa. schneider, r. r., and s. wasel. 2000. the effect of human settlement on the density of moose in northern alberta. journal of wildlife management 64:513520. schwartz, c. c. 2002. carnivores, moose and humans: a changing paradigm of predator management. presented at the 5th international moose symposium, hafjell, norway, august 2002. , k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula alaska. alces 28:1-14. sigouin, d., s. st-onge, r. courtois, and j.-p. ouellet. 1999. change in hunting activity and hunters’ perception following the introduction of selective harvest in québec. alces 35:105-123. skonhoft, a. 2002. natural resource utilisation under asymmetric costs and benefits. moose management in norway. presented at the 5th international moose symposium, hafjell, norway, august 2002. human dimensions in moose research ericsson alces vol. 39, 2003 26 smith, c. a., j. b. faro, and n. c. steen. 1979. an evaluation of trophy moose management on the alaska peninsula. proceedings of the north american moose conference and workshop 15:280-302. sødal, d. p. 1989. the recreational value of moose hunting in norway: towards modelling optimal population density. scandinavian forest economics 30:6278. steinhoff, h. w., r. g. walsh, t. j. peterle, and j. m. petulla. 1987. evolution of the valuation of wildlife. pages 34-48 in d. j. decker and g. r. goffs, editors. valuing wildlife: economic and social perspectives. westview, boulder, colorado, usa. storaas, t., h. gundersen, h. henriksen, and h. p. andreassen. 2001. the economic value of moose in norway a review. alces 97-122. timmermann, h. r. 1977. the killing proficiency of moose hunters. proceedings of the north american moose conference and workshop 15:12-25. , and m. e. buss. 1998. population and harvest management. pages 559616 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , and r. gollat. 1986. selective moose harvest in northcentral ontario a progress report. alces 22:395-418. , and . 1994. early winter social structure of hunted vs. unhunted moose population in n. central ontario. alces 117-126. , and r. s. rempel. 1998. age and sex structure of hunter harvested moose u n d e r t w o h a r v e s t s t r a t e g i e s i n northcentral ontario. alces 34:21-30. , and h. a. whitlaw. 1992. selective moose harvest in north central ontario. alces 28:137-164. wedels, c. h. r., h. smith, and r. rollins, 1989. opinions of ontario moose hunters on changes to the selective harvest system. alces 25:15-24. whittaker. d., m. j. manfredo, p. j. fix, r. sinnot, s. miller, and j. j. vaske. 2001. understanding beliefs and attitudes about an urban wildlife hunt near anchorage, alaska. wildlife society bulletin 29:1114-1124. wilton, m. l. 1992. implications of harvesting moose during pre-rut and rut activity. alces 28: 31-34. . 1995. the case against calling and hunting dominant moose during the main rut period a viewpoint. alces 31:173-180. , and a. r. bisset. 1988. movement patterns of tagged moose from an unhunted area to a heavily hunted area. alces 24:62-68. wolfe, m. l. 1987. an overview of the socioeconomics of moose in north america. swedish wildlife research supplement 1:659-675. 4001.p65 alces vol. 40, 2004 todesco illegal moose kill in ontario 145 illegal moose kill in northeastern ontario: 1997 – 2002 charlie todesco ontario ministry of natural resources, p.o. box 1160, wawa, on, canada pos 1ko abstract: conservation officers found 793 illegally killed moose in the ontario ministry of natural resources northeast region during the period of 1997 – 2002. of these illegally killed moose, 365 were abandoned. the majority of abandoned moose were a result of illegal harvesting, as 68% of all abandoned moose had signs of positive human interaction with the carcass. three hundred and twenty moose (40%) spoiled and were unsuitable for human consumption. bulls were illegally killed at a significantly higher proportion, and calves at a significantly lower proportion, than their respective availability in the herd structure. cow moose are illegally killed proportional to their availability. illegal moose kills were positively and significantly correlated with moose populations, the number of applicants for adult validation tags, and the number of hunters checked by conservation officers. the illegal moose kill has both an immediate and a long-term impact on the regional herd population. an estimated 613 moose were not recruited into the regional herd as a result of illegal harvesting. moose watch, a program to reduce the region’s illegal moose kill was initiated in 2000, and was expanded province-wide in 2001. a toll-free 24-hour violation reporting line was established, and received 387 calls over 3 years regarding illegal hunting violations for a wide variety of wildlife species. during the 6 years, conservation officers in the region contacted over 108,000 hunters, issued 3,064 warnings, and laid 2,580 charges while conducting moose hunt enforcement duties. alces vol. 40: 145-159 (2004) key words: alces, illegal kill, moose watch, poaching based on a perceived increase in the number of illegally killed and abandoned moose (alces alces) in the mid 1990s, ontario ministry of natural resources (omnr) conservation officers began data collection in northeastern ontario. initial data collection was started in 1995, and became standardized across the northeast region (ner) in 1997. the objective of this initiative was to collect intelligence on the distribution and impact of the illegal harvest of moose, and to deliver a planned enforcement response to deal with the problem. in 2000, the ner developed and launched the “moose watch 2000” program. this program was aimed at reducing the illegal moose kill through public awareness, a 24-hour toll-free violation reporting line, and increased enforcement effort. the moose watch program was expanded provincially in 2001. the focus of this report is on the number of illegally killed and abandoned moose throughout the region from 1997 – 2002. study area the ner is a large (441,122 km2) and diverse land base, comprised of a variety of geographic and physiological features. it extends from the north shores of lake huron and lake superior to the james bay and hudson bay lowlands (fig. 1). there are 2 forest regions located within the ner, the boreal forest in the northern portion of the region and the great lakes – st. lawrence forest in the south (hosie 1979). boreal forest tree species are of fire origin, consisting of white spruce (picea glauca), illegal moose kill in ontario – todesco alces vol. 40, 2004 146 fig. 1. ontario ministry of natural resources northeast region and districts. black spruce (picea mariana), jack pine (pinus banksiana), balsam fir (abies b a l s a m e a ) , w h i t e b i r c h ( b e t u l a papyrifera), and trembling aspen (populus tremuloides). the great lakes – st. lawrence forest is comprised of conifers such as red pine (pinus resinosa), white pine (pinus strobus), eastern white cedar (thuja occidentalis), and tolerant hardwoods such as red maple (acer rubrum), sugar maple (acer saccharum), yellow birch (betula alleghaniensis), and red oak (quercus rubra). within the region, there are 9 omnr districts – chapleau, cochrane, hearst, kirkland lake, north bay, sault ste. marie, sudbury, timmins, and wawa (fig. 1). there are 5-10 conservation officers stationed in each district, and they patrol 28 wildlife management units (wmus), alces vol. 40, 2004 todesco illegal moose kill in ontario 147 including some that are extremely remote with little to no road access. most wmus have a moose archery season commencing from the saturday closest to september 17 to the third following friday, and a firearm season over the period from the saturday closest to october 8 to november 15. three wmus have extended firearms seasons until december 15. methods conservation officers patrolling in the region collected data on illegal moose kills from 1997 – 2002. all unlawfully killed and all abandoned moose encountered were classified as illegally killed moose. standardized data report sheets were completed for each illegal kill and included information on age/sex, whether the moose was seized or abandoned, date of kill, and general comments regarding the moose. when abandoned moose were encountered, indications of human interaction with the carcass were recorded in order to estimate the number of abandoned moose that may be a result of wounding mortality. human interaction with moose carcasses included gutting, concealment, and location of kill in relation to roads or waterways. wildlife management unit data have been collected for each illegally killed moose since 1998. only verified kill data were used, meaning that if a conservation officer “didn’t see it or didn’t touch it” the data were not included. the data in this report are considered to be a conservative estimate of illegally killed moose for the ner. in order to determine the impacts on herd recruitment in the ner, a population model was designed using the age/sex structure of the illegal kill, a low in-utero productivity rate (0.95), and an annual mortality rate of 10% (g. eason, v. crichton, personal communication). this model produces conservative estimates. moose population estimates in each wmu were obtained from the omnrs “ontario moose harvest planning system” computer program. the number of adult validation tags (avts) issued to hunters for harvesting bull and cow moose were obtained from copies of “ontario hunting regulation summary” for 1997-2002. enforcement statistics were derived from the omnrs “compliance activity violation reporting system” (cavrs) computer program. this program is utilized by all conservation officers to record their enforcement efforts, violation statistics, and violator information. a standard calculation of non-compliance rates was used (number of charges + number of warnings / field contacts). statistical analysis of illegal harvest data included correlation analysis and for availability – utilization analysis, a chi-square goodness-of-fit test and a bonferroni z-test were used (neu et al. 1974, byers et al. 1984). regression analysis was completed using the number of hunters contacted by conservation officers and the number of illegally killed moose to produce an estimate of the total number of illegally killed moose. the moose watch program promotional efforts were initiated through the production of posters, violation reporting cards, pens and pencils, and licence holders imprinted with the moose watch logo and the violation reporting line number (1 – 866 – 34 moose). in the 3 years of operation, 10,000 each of posters, contact cards, pens and licence holders were distributed across the province to hunters and the general public through businesses or places of employment, field contacts, outdoors shows, and presentations. moose watch was promoted in the “ontario hunting regulation summary” and all avt holders were shipped a moose watch promotion note with their tags. during the period of september 15 – december 15, media releases and interillegal moose kill in ontario – todesco alces vol. 40, 2004 148 t ab le 1 . i ll eg al ly k il le d m oo se in n or th ea st r eg io n 19 97 – 2 00 2. views were held with television, radio, and newspaper reporters promoting the program and releasing enforcement statistics. public service announcements were prepared and were run by radio and television stations. news releases dealing with convictions as a result of the moose watch program were released to all media sources. ministry of natural resources enforcement staff located at the provincial coordination centre (pcc) in sault ste. marie answered all calls to the moose watch violation line. all conservation officers in the province report at the start of their shift to the pcc, and pcc staff are able to pass violation complaints to on-duty officers or to the next officer starting their shift in the district from where the complaint was received. the traditional partnership with crime stoppers (a north american wide violation reporting system) has been maintained, and allows for anonymous violation reporting by those that choose to do so. results unlawful trends there were 793 verified illegally killed moose in the ner over the period of 1997 2002, of which 46% were abandoned (table 1). wawa district had the highest illegal kill and the highest rate of abandonment each year since 1997. six districts (chapleau, kirkland lake, north bay, sault ste. marie, and wawa) accounted for 78% of all illegal kills in the ner. abandonment rates were also highest in those districts as well. illegal moose kills in the ner peaked in the 1999 hunt, declined in 2000 and 2001, and rebounded in 2002. abandoned moose declined in the ner from 1997 – 1998, remained relatively constant from 1999 – 2001, and increased in 2002 to levels initially observed in 1997 (table 2). of the 365 abandoned moose, 251 (68%) showed positive signs of human interaction with the carcass. there was a a b a n . to ta l a b a n . to ta l a b a n . to ta l a b a n . to ta l a b a n . to ta l a b a n . to ta l a b a n . to ta l c ha pl ea u 10 12 6 11 6 13 5 11 4 10 14 21 45 78 c oc hr an e 5 7 2 10 0 2 3 5 2 5 3 10 15 39 h ea rs t 11 16 1 3 5 18 5 11 4 5 7 13 33 66 k irk la nd l ak e 8 14 7 19 3 10 7 13 10 14 9 13 44 83 n or th b ay 7 14 2 16 2 13 11 20 4 8 7 16 33 87 sa ul t s te . m ar ie 5 15 3 17 10 35 10 25 4 8 6 14 38 11 4 su db ur y 10 12 3 8 7 19 2 11 4 10 5 10 31 70 ti m m in s 4 8 4 13 8 17 8 16 12 16 7 11 43 81 w aw a 18 31 11 23 17 41 5 25 11 26 21 29 83 17 5 to ta l 78 12 9 39 12 0 58 16 8 56 13 7 55 10 2 79 13 7 36 5 79 3 20 01 20 02 to ta l 19 97 19 98 19 99 20 00 alces vol. 40, 2004 todesco illegal moose kill in ontario 149 strong, but not significant (p > 0.05), correlation between the number of abandoned moose and the number having human interaction (r = 0.78, tabular value = 0.811, 4 df) conservation officers located 114 abandoned moose with no signs of human interaction. a total of 320 abandoned moose spoiled, resulting in approximately 64,000 kg of meat becoming unsuitable for human consumption (assuming 200 kg meat/moose). wildlife management units the total estimated moose population in the 28 wmus steadily increased over the study period (fig. 2), although the moose populations in wmus 35 and 36 steadily declined over the 5 years. the illegal kill in each of the wmus was significantly correlated (p < 0.05, tabular value = 0.374, 26 df) to the population in the wmus for 1998 (r = 0.587), 2001 (r = 0.548), and 2002 (r = 0.413) (table 3). the illegal moose kill was significantly correlated to the number of hunter applicants (pool 1 – choice 1) who applied for avts in wmus in 1998 (r = 0.761), 2000 (r = 0.510), 2001 (r = 0.547), and 2002 (r = 0.503). the illegal moose kill was significantly correlated to the number of avts issued for only 2 years, 1998 (r = 0.554) and 2001 (r = 0.479). illegal kill data were not collected on a wmu basis in 1997. the illegal moose kill in 7 wmus (21b, 28, 29, 32, 35, 36, and 41) accounted for 50% of the ner verified illegal kill from 1998 – 2002. five of these wmus (28, 29, 35, 36, and 41) have the major urban centres of kirkland lake, timmins, sault ste. marie, sudbury, and north bay located in or adjacent to these respective units. four wmus (21b, 28, 29, 41) have the highest average moose populations, available avts, and avt applicants. of interest, wmu 32 had 49 illegal kills over the 5-year period. this wmu has 63% of its total area (11,424 km2) closed to hunting by the chapleau crown game preserve, resulting in a high illegal kill confined to a relatively small geographic area. in making avt decisions for each wmu, a 10% non-hunt mortality estimate is used to account for losses to predation, disease, accidents, lawful harvest by aboriginal people, and illegal kill. annual illegal kills do not exceed the 10% threshold in any wmu; however, these kills constitute the only verified component of the non-hunt mortality estimate. illegal kill structure based on aerial survey data for the ner over the study period, the regional moose herd structure is comprised of 33% bulls, 48% cows, and 19% calves. the structure of the illegal moose kill has remained constant and with 1 exception (cows 2001), has not fluctuated by more than 5% over the period of 1997 – 2002 (table 4). overall, cows constitute 49% of the illegal kill, bulls 40%, calves 7%, and moose that could not be identified 4% (fig. 3). the composition of the regional illegal kill is significantly different (p < 0.05) than t h e r e g i o n a l m o o s e h e r d s t r u c t u r e table 2. abandoned and spoiled moose observed in northeast region 1997 – 2002. 1997 1998 1999 2000 2001 2002 total abandoned 78 39 58 56 55 79 365 human interaction 52 29 31 23 48 68 251 spoiled 58 33 60 50 51 68 320 fig. 2. estimated northeast region moose population 1997 – 2002. 48603 50893 4744848522 45431 46344 40000 45000 50000 55000 1997 1998 1999 2000 2001 2002 illegal moose kill in ontario – todesco alces vol. 40, 2004 150 m o o se p o p . a v t a v t a p p . ill. m o o s e 1 8 b 5 2 8 1 0 9 1 8 6 2 1 9 1 6 9 0 2 9 1 7 4 2 0 2 1 a 3 2 0 5 7 9 0 1 9 9 4 9 2 1 b 3 1 0 5 8 4 5 3 2 1 6 1 1 2 2 2 6 0 0 1 3 0 8 0 4 9 2 3 1 3 0 0 1 7 6 7 3 0 6 2 4 2 0 7 4 1 6 4 1 1 8 4 5 2 5 1 6 5 6 1 3 0 1 3 2 1 2 6 6 7 5 4 5 4 3 4 0 2 7 1 8 3 4 1 4 0 8 7 3 2 2 8 2 7 6 2 5 6 1 4 4 8 5 6 2 9 1 6 2 9 3 5 7 2 4 6 4 1 6 3 0 2 8 2 7 2 6 5 1 8 7 2 1 3 1 1 8 7 7 1 5 1 1 4 9 1 3 3 2 1 1 8 7 3 0 4 6 6 9 3 3 1 2 6 6 4 7 3 3 7 6 3 4 8 0 1 2 5 2 2 0 2 3 5 1 4 3 1 1 5 9 1 5 3 1 1 8 3 6 1 6 0 1 1 9 1 1 3 3 8 1 3 3 7 9 9 3 5 9 5 6 4 9 3 8 2 5 6 4 3 7 5 2 9 8 9 1 4 3 9 1 2 0 0 1 7 5 1 7 1 4 4 4 0 2 6 9 3 2 9 5 2 9 6 2 9 4 1 2 3 2 2 4 9 0 4 1 9 8 4 4 2 2 8 4 4 2 4 1 1 9 7 3 3 4 6 3 4 9 4 8 4 1 3 0 4 7 3 8 9 2 3 0 2 1 3 8 3 4 8 1 1 5 0 3 4 5 2 8 6 1 2 t o ta l 4 8 5 5 2 6 8 6 4 4 4 3 1 1 1 6 7 w m u 1 9 9 9 m o o se p o p . a v t a v t a p p . ill. m o o se 1 8 b 5 2 8 1 0 9 1 7 8 1 1 9 1 6 9 0 3 3 8 9 0 3 1 2 1 a 3 2 0 5 7 9 0 2 3 4 5 0 2 1 b 3 1 0 5 7 3 5 3 0 5 8 7 2 2 2 6 0 0 1 5 9 8 4 4 8 2 3 1 8 0 3 2 0 0 9 0 0 2 2 4 2 0 9 0 1 8 7 1 1 2 6 4 2 5 1 6 5 6 1 3 0 1 4 5 1 2 6 1 4 0 4 4 5 4 0 3 2 2 7 1 1 1 3 1 4 0 1 0 5 4 1 2 8 2 7 6 2 5 1 1 4 4 5 4 1 0 2 9 1 6 2 9 3 5 7 2 7 9 3 7 3 0 1 5 8 6 2 2 0 1 8 0 5 4 3 1 1 9 6 4 1 7 0 1 4 3 3 9 3 2 1 1 8 7 3 7 4 7 9 8 3 3 1 2 6 6 4 5 3 5 2 2 3 4 8 6 9 2 5 2 1 9 0 3 5 1 4 3 1 1 5 9 1 4 5 8 6 3 6 1 0 8 1 9 2 1 1 6 9 1 3 3 7 9 9 3 5 9 5 7 5 1 0 3 8 2 5 6 4 3 7 5 2 5 5 7 5 3 9 1 2 0 0 1 7 5 1 7 7 5 2 4 0 2 6 9 3 2 9 5 2 8 8 5 8 4 1 2 3 2 2 4 9 0 4 2 8 1 1 6 4 2 2 8 4 4 2 4 1 2 1 6 8 3 4 6 3 0 2 4 8 4 3 0 0 4 7 4 1 1 2 3 0 2 1 3 8 0 4 8 1 1 5 0 1 3 3 2 6 3 7 7 t o t a l 4 7 4 4 8 6 4 9 5 4 4 5 6 4 1 3 7 w m u 2 0 0 0 m o o se p o p . a v t a v t a p p . ill. m o o se 1 8 b 2 8 5 7 2 1 6 1 0 1 9 1 6 9 0 3 6 4 1 0 4 4 0 2 1 a 3 2 2 0 7 9 0 2 3 7 1 3 2 1 b 3 1 0 5 7 3 5 2 7 2 6 1 1 2 2 2 3 0 0 2 3 1 1 0 8 6 5 2 3 1 7 5 5 1 8 6 7 9 1 1 2 4 2 0 9 0 2 0 1 1 1 5 6 2 2 5 1 6 5 6 1 4 7 1 6 5 0 2 6 1 4 0 4 5 0 4 4 2 2 2 7 1 1 1 3 1 4 0 9 9 0 1 2 8 2 9 7 0 5 0 0 4 3 4 4 1 1 2 9 1 7 2 7 3 0 2 2 1 6 0 7 3 0 1 5 8 6 2 2 0 1 7 6 3 4 3 1 1 9 6 4 1 9 2 1 5 5 6 1 0 3 2 1 7 0 6 1 3 9 7 5 2 1 1 3 3 1 3 3 0 4 5 3 0 0 0 3 4 8 6 9 2 5 1 9 6 0 3 5 1 4 3 1 1 4 4 1 3 5 0 5 3 6 1 0 8 1 9 2 1 0 2 9 5 3 7 9 9 3 6 6 6 6 4 1 3 8 2 1 3 2 2 9 0 2 3 0 1 1 3 9 1 2 0 0 1 2 5 1 6 6 3 1 4 0 3 2 3 6 4 0 7 3 1 3 3 4 4 1 2 9 9 8 5 4 1 4 3 1 2 7 4 2 2 8 4 4 2 7 8 2 4 6 3 6 4 6 3 0 2 5 8 4 2 2 1 4 7 9 2 5 2 6 2 2 1 7 9 3 4 8 6 9 1 1 3 3 2 3 4 8 0 t o t a l 4 8 6 0 3 6 7 3 5 4 3 8 6 7 1 0 2 w m u 2 0 0 1 table 3. moose populations, adult validation tags (avt), adult validation tag applicants, and illegal moose kills in northeast region 1998 – 2002. m o o se p o p . a v t a v t a p p . ill. m o o se 1 8 b 5 2 8 1 0 4 1 9 9 0 1 9 1 6 9 0 2 9 1 6 7 7 0 2 1 a 3 1 0 5 8 1 5 1 9 0 6 1 2 1 b 3 1 0 5 8 4 5 2 8 9 1 1 1 2 2 2 6 0 0 1 1 5 7 1 9 1 2 3 1 2 0 0 1 7 7 9 3 7 1 2 4 1 4 0 0 1 5 4 1 0 0 6 5 2 5 5 9 7 1 3 0 1 0 9 0 2 6 6 7 5 4 5 4 5 8 0 2 7 1 8 3 4 2 5 0 1 3 4 1 6 2 8 3 5 4 5 6 6 2 4 3 4 1 1 4 2 9 2 0 1 8 6 0 1 3 4 7 5 7 3 0 2 8 2 7 2 6 5 1 7 7 2 1 3 1 1 8 7 7 1 5 6 1 3 9 9 8 3 2 1 1 7 3 9 1 6 8 4 7 3 3 7 6 8 4 5 3 1 7 3 3 4 2 5 2 2 5 2 3 4 2 3 5 1 8 6 7 2 1 5 1 8 2 3 8 3 6 1 6 0 1 1 9 0 1 3 4 4 5 3 7 7 3 0 4 5 4 6 0 1 3 8 2 3 5 2 4 8 0 2 9 3 9 9 3 9 1 2 0 0 1 7 5 1 7 3 1 3 4 0 2 7 6 7 2 9 5 3 0 6 1 9 4 1 2 3 2 2 4 9 0 4 0 9 4 9 4 2 2 4 7 3 1 7 1 1 5 7 8 2 4 6 3 4 9 4 8 3 5 5 1 4 7 3 8 9 2 3 0 2 0 6 7 0 4 8 1 1 0 0 3 5 0 2 6 9 6 6 t o t a l 4 6 3 4 4 7 4 6 0 4 4 6 1 3 1 2 0 w m u 1 9 9 8 alces vol. 40, 2004 todesco illegal moose kill in ontario 151 table 3 (continued). moose populations, adult validation tags (avt), adult validation tag applicants, and illegal moose kills in northeast region 1998 – 2002. (table 5). using adjusted data for known age and sex ratios, moose were not killed proportional to their availability. bull moose were illegally killed at a higher proportion than they occurred in the herd structure, and calves were illegally killed at a lower proportion than they occurred in the herd. cow moose were illegally killed proportional to their availability. 129 120 168 137 102 137 17814611790 5428 0 100 200 300 400 1997 1998 1999 2000 2001 2002 illegal kill estimated annual recruitment loss fig. 4. northeast region illegal moose kill and estimated annual recruitment loss. m o o s e p o p . a v t a v t a p p . ill. m o o se 1 8 b 2 8 5 7 2 2 7 8 1 1 9 1 8 6 1 3 6 0 1 1 1 2 0 2 1 a 3 2 2 0 7 4 4 2 3 9 7 0 2 1 b 3 1 0 5 6 8 5 2 8 7 4 1 0 2 2 2 3 0 0 2 0 7 1 0 4 5 2 2 3 1 7 5 5 1 8 7 8 0 0 7 2 4 3 0 8 0 2 6 1 1 2 6 5 4 2 5 1 7 1 5 1 4 7 1 4 9 0 2 6 1 4 0 4 5 0 4 1 5 1 2 7 1 2 6 9 1 4 0 1 0 6 8 6 2 8 3 0 3 7 5 0 0 4 2 7 8 1 3 2 9 1 7 2 7 2 2 1 2 3 2 9 9 3 0 2 8 7 3 2 0 0 1 7 4 1 8 3 1 1 9 6 4 1 5 4 1 6 3 9 7 3 2 1 6 5 0 1 3 2 8 9 9 1 4 3 3 1 3 3 0 4 0 2 9 1 2 3 4 8 6 9 2 5 1 8 7 1 3 5 1 4 3 1 8 8 1 2 8 2 7 3 6 1 0 8 1 5 0 8 7 9 6 3 7 8 9 9 3 5 6 3 2 6 3 8 2 1 3 2 1 8 4 2 2 8 6 1 0 3 9 1 2 0 0 1 1 0 1 6 1 4 1 4 0 3 2 3 6 3 8 6 3 4 5 2 4 4 1 2 9 9 8 2 6 2 4 0 7 6 9 4 2 2 4 3 0 1 4 5 2 4 9 9 7 4 6 4 2 6 5 0 4 4 3 0 4 7 9 2 5 1 4 5 2 2 6 3 1 4 8 6 9 1 9 1 4 1 3 1 t o ta l 5 0 8 9 3 5 5 8 9 4 3 6 0 6 1 3 7 w m u 2 0 0 2 recruitment loss the illegal moose kill has an immediate impact on the ner moose herd, as well as a long-term impact regarding potential recruitment that did not occur. using a basic but conservative population model, an estimated 613 moose were not recruited into the regional herd over the 6-year period. this would result in a total loss of 1,406 moose in the ner from 1997 – 2002 (fig. 4). enforcement effort moose enforcement effort by ner conservation officers steadily increased from 1997 – 1999 and peaked in 2000 (the first year of the moose watch program) (table 6). conservation officers checked over 108,000 hunters, with the highest number of hunters being checked in 1999. the overall non-compliance rate was 5.2%, with the highest non-compliance rate (6%) occurring in 2000. conservation officers issued 3,064 warnings and laid 2,580 charges while completing field moose enforcement duties during the 6-year period. penalties assessed through tickets or by the courts as a result of trials amounted to $822,186. fines are paid into the “fish and wildlife special 1997 1998 1999 2000 2001 2002 total bulls 53 47 65 57 39 59 320 cows 60 59 83 69 58 64 393 calves 12 11 14 6 4 10 57 unknown 4 3 6 5 1 4 23 total 129 120 168 137 102 137 793 table 4. northeast region illegal moose kill structure. fig. 3. northeast region illegal moose kill age and sex structure. 40% 49% 7% 4% bulls cows calves unkn. illegal moose kill in ontario – todesco alces vol. 40, 2004 152 table 5. occurrence of observed and expected illegally killed moose in the northeast region. of hunters contacted per illegal moose was variable, but remained close to the 6-year average of 139 hunters contacted per illegal moose. enforcement efficiency appeared to be steadily improving over the period of 1997 – 2001 as a result of increased violation detection, but declined sharply in 2002. using the number of illegally killed moose and the number of hunters contacted by conservation officers, a regression equation (r2 = 0.85, y = 77.517 + 0.01158 x) was derived. this equation was used with the estimated number of moose hunters in the ner obtained through postcard surveys (1999 – 2002 data only available, p. davis, personal communication). the annual illegal moose kill estimate was calculated to range from 557 to 577 (table 7). using the (chi-square = 77.43, tabular value, p < 0.05, 2 df = 5.99). 126 152 93 116 106 83 97 82 136 124 151141 0 20 40 60 80 100 120 140 160 1997 1998 1999 2000 2001 2002 enforcement effort (hours) hunter contacts fig. 5. hunter contacts and enforcement effort per illegal moose in northeast region 1997 – 2002. ner proportion observed expected proportion legally killed bonferroni intervals preference bull 0.33 319 253 0.417 0.362 < p1 < 0.472 + cow 0.48 393 367 0.514 0.459 < p2 < 0.570 0 calf 0.19 53 145 0.069 0.0409 < p3 < 0.098 -total 765 765 purpose account” and, along with hunting and fishing licence revenues, are used to fund fish and wildlife management and enforcement programs. the number of illegally killed moose was positively, but not significantly (p > 0.05, tabular value = 0.811, 4 df) correlated to the number of hours spent in the field by conservation officers (r = 0.599). the number of illegally killed moose was positively and significantly (p < 0.05, tabular value = 0.811, 4 df) correlated to the number of hunters contacted by conservation officers (r = 0.919). enforcement effectiveness was assessed by examining the number of hours of enforcement effort and the number of hunters contacted per illegal moose over the 6 years (fig. 5). the number alces vol. 40, 2004 todesco illegal moose kill in ontario 153 average of 139 hunters contacted per illegal moose and the ner projected numbers of moose hunters, the annual illegal kill estimate ranges from 394 to 406. based on these estimates, conservation officers may only be locating 20 – 40% of all illegally killed moose in the ner. moose watch program the moose watch violation reporting line received the highest number of calls in 2001, the year when the program was expanded province-wide (table 8). the ner accounted for 50% of all calls to the violation reporting line in 2001 and 54% of all calls in 2002. wawa and sault ste. marie districts received the most moose watch violation calls in the region. over the 3 years that the moose watch violation-reporting line was in operation, a total of 392 calls were received (table 9). violation reports of illegal or abandoned moose constituted 57% of all calls received, and overall, only 4% of calls received were from hunters reporting that they had killed an animal that they were not licenced for. calls about illegal night hunting accounted for 7% of violations reported. in the absence of a general omnr violation reporting line, calls were also received regarding illegal poaching of deer, elk, fish, and turkeys, as well as other resource related infractions. calls to the moose watch violation reporting line were consistent over the 3 years table 6. northeast region enforcement efforts 1997 – 2002.1 where the majority of calls provided violation information that did not require an immediate enforcement response. less than 20% of the calls received were of a nature requiring an immediate response by conservation officers. discussion the number of illegally killed moose in the ner is of concern to enforcement staff, wildlife managers, and stakeholders, especially as the numbers in this report are considered to be minimum estimates. wolfe (1987) broadly defined illegal harvest as the “taking of protected wildlife contrary to conditions prescribed by provincial / state / territorial or federal wildlife statutes”, and that most wildlife agencies consider reports of illegal kill by enforcement staff as a minimum estimate. furthermore, the illegal moose kill has a direct socio-economic impact through reduction of hunting opportunities and lost licensing revenue. table 7. northeast region verified and estimated illegal moose kill 1999 2002. 1 (period september 1 – december 31). 1999 2000 2001 2002 ve rifie d ille g a l m o o s e k ill 167 137 102 137 pro je c t e d ille g a l m o o s e k ill – fro m re g re s s io n eq u a t io n 573 557 577 560 pro je c t e d ille g a l m o o s e k ill – fro m co co n t a c t 404 394 406 396 pro je c t e d n er m o o s e h u n t e rs 56,152 54,790 56,491 55,026 1997 1998 1999 2000 2001 2002 total enforcement effort (hrs) 10,588 11,635 13,892 14,569 11,835 12,780 75,299 hunter contacts 18,167 18,132 20,805 17,369 15,471 18,674 108,618 charges 322 458 559 454 381 406 2580 warnings 565 562 493 586 459 399 3064 non compliance rate (%) 4.9 5.6 5.1 6 5.4 4.3 5.2 penalties ($) 93,290 163,176 203,120 157,215 100,055 105,330 822,186 illegal kills 129 120 168 137 102 137 793 illegal moose kill in ontario – todesco alces vol. 40, 2004 154 table 8. moose watch violation report line calls 2000 – 2002. 2000 2001 2002 chapleau 3 2 9 cochrane 1 3 0 hearst 2 4 2 kirkland lake 7 7 10 north bay 8 13 10 sault ste. marie 9 21 7 sudbury 10 18 6 timmins 6 15 10 wawa 13 17 17 total northeast region calls 61 100 71 total provincial calls 61 200 131 the reduced illegal moose kill in 2000 and 2001 appeared to be a response to the moose watch program; however, the 35% increase in verified illegal kills in 2002 indicates an apparent decrease in hunter compliance. a similar trend was observed in north central ontario in the late 1970s – early 1980s by timmermann and gollat (1984) when hunting regulations were changed to prohibit party hunting. charges laid by conservation officers during the moose hunting season declined following the first 2 years of regulation change, and sharply increased in the third year as a result of enforcement efforts and a less cautious approach taken by hunters. ontario reported high non-compliance during the first year of the moose selective harvest system in 1983 (wolfe 1987), and the illegal harvest of moose continues to be a serious compliance issue in northeastern ontario 20 years later. the high number of abandoned moose is of great concern, and the majority of the ner abandoned moose were killed with unlawful intent, rather than by accident, based on human interaction. in a similar study, beattie et al. (1980, citing hardin and roseberry 1975) reported that 20% of abandoned deer carcasses on the crab orchard national wildlife refuge in illinois had been intentionally abandoned. high abandonment rates of moose offend law abiding hunters and the general public, and resulted in 85 complaints to the moose watch violation reporting line. increased promotion of hunter ethics and increased enforcement effort are required to deter this behavior. approximately 1/3 of the ner abandoned moose showed no sign of human interaction, and may be a result of wounding mortality. moose hunter shooting proficiency was studied by timmermann (1977) and buss et al. (1989) and they estimated that potential wounding loss based on shooting exercises at life-sized moose targets by ontario hunters could be in the magnitude of 25 – 30%, and close to 40% on moving targets. wolfe (1987) reported that crippling loss of moose in ontario was considered to be of moderate concern. it is also possible that these moose were killed as a result of illegal activity. pursley (1977) observed a 25% wounding mortality rate in unlawfully killed deer in new mexico. higher wounding mortality rates have been observed for night hunted deer, ranging from 27% in manitoba (bessey 1984) to as high as 50% on manitoulin island, ontario (i. anderson, personal communication). little information exists on wounding rates in unlawfully hunted moose; however, it would be reasonable to assume that similar rates as observed in deer would apply to moose. illegal harvesting of moose in the ner is a function of moose populations and hunter pressure. the 7 wmus that comprise 50% of the regional kill are located near urban centres, have high hunter preference, and high competition for available avts. perceived availability of animals is a primary consideration for those that are predicated to unlawfully taking wildlife (bessey 1984, alces vol. 40, 2004 todesco illegal moose kill in ontario 155 table 9. violations reported to moose watch line 2000 – 2002. 1 ner only 2000, province-wide 2001-02. violation reported 1 2000 2001 2002 total moose poaching 17 60 60 137 abandoned moose 15 37 33 85 turning self in 6 5 5 16 night hunting 5 16 5 26 aircraft hunting 5 2 0 7 deer poaching 3 37 13 53 fish poaching 1 10 1 12 elk poaching 0 3 1 4 turkey poaching 0 1 0 1 non-violations 9 29 13 51 total 61 200 131 392 glover 1982 as cited by bessey 1984, gregorich 1992). strategic enforcement effort needs to be focussed on these 7 wmus. competition for declining levels of avts and opportunities to harvest adult moose may be pressuring some hunters to violate moose hunting regulations. for the 1999 and 2000 hunting seasons, the ner had the highest number of moose hunters in the province, and the second lowest number of avts (bisset 2002). hunter satisfaction is influenced by the ability to harvest an animal, and in areas with high hunter densities, hunting techniques are selected to avoid losing preferred hunting locations to other hunters (crête 1987). hall et al. (1990) found a perception among migratory bird hunting violators that the temptation to violate was enhanced by the belief that more waterfowl were being killed elsewhere along the flyway. these situations lead to increased competition among hunters, and may influence some individuals to take increased risks in order to harvest a moose. benson (2000) stated that in hunting “opportunity elicits actions that sometimes would not be considered”. in sweden, where high moose populations and strictly regulated hunting occur (cederlund and markgren 1987), losses to poaching are considered to be negligible (boer 1991). if moose populations and the availability of avts increase, the incidence of illegal harvesting of moose may decrease based on an improvement in hunter satisfaction. the relatively stable sex structure of the ner illegal kill indicates that there is differential vulnerability to poaching. it is reasonable that bull moose constitute a higher than expected percentage of the illegal kill based on their increased availability to hunters resulting from rut and post-rut activity. calves are under-represented in the illegal kill as all licenced hunters in ontario can lawfully harvest them in any wmu with an open moose season. crête (1987) observed that when hunters can choose, vulnerability is determined by hunter preference for “large bulls, small bulls, large cows, small cows, and calves” in decreasing preference. moose are being killed opportunistically as they become available to poachers. bessey (1984) assumed that the majority of deer poachers were opportunists who violated hunting legislation when opportunities were presented. the ner data suggest that moose are being killed opportunistically as they are encountered and abandoned if an adult validation tag is not affixed in a reasonable period of time. if the ner illegal moose kill were solely based on cow moose being mistaken for calves, the proportion of cows in the illegal kill would be significantly higher. an assumption may be made that the verified annual illegal kill does not constitute a sustainability issue, as it does not exceed the 10% nonillegal moose kill in ontario – todesco alces vol. 40, 2004 156 hunt mortality estimate in any wmu. of all of the contributing non-hunt mortality factors, the only verified data that exists for any of these factors is the illegal moose kill data. poaching is the only non-hunt mortality factor that is actively managed or only factor that can be reasonably controlled at this time in the ner. individual wmu moose population models and harvest levels may have to be adjusted in wmus where known estimates approach or exceed the 10% non-hunt mortality estimate. there is little information on the impact of illegal kills on moose populations in north america (wolfe 1987). illegal kills may have a significant impact on moose populations when combined with other nonhunt mortality factors. illegal harvest and brainworm (parelaphostrongylus tenuis) are cited as contributing factors in moose p o p u l a t i o n d e c l i n e s i n n o v a s c o t i a (timmermann 1987). poaching and collisions were the highest cause of all known moose non-hunt mortality in maine, minnesota, nova scotia, and new brunswick in 1970, and the illegal kill accounted for an average of 31% of the non-hunt mortality in these jurisdictions (karns et al. 1974). the illegal kill has a recruitment impact on the ner moose herd, which will reduce the availability of moose for law-abiding hunters. there are 2 ways of looking at this loss in terms of hunting opportunities. one assumption would be that all illegally harvested moose and recruitment loss constitutes the total loss of opportunities to the hunting community. wolfe (1987) stated that every illegally killed moose could support an additional 6 resident hunters or 3 non-resident hunters in north america. using the verified illegal kill and estimated recruitment loss of 1,404 moose being unavailable for lawful harvesting, this would represent a loss of opportunities to 8,424 resident hunters or 4,212 non-resident hunters to hunt within the ner. the other viewpoint would be that the illegal kill and recruitment would have accrued into the ner moose population and been apportioned to hunters using the current allocation methods. in this example, the verified illegal kill and estimated recruitment loss would represent approximately 240 avts over the 6-year period (assuming a planned harvest level of 10% and a 50% success rate of filling an avt). each avt has a multiplier value in terms of hunter opportunities. provincially, 58% of all hunters apply for avts in groups (average group size 4.25), and 42% apply as individuals. these 240 avts would have permitted an opportunity for a minimum of 643 individuals to legally hunt an adult moose. as all hunters can harvest calf moose in ontario, the loss of approximately 175 calves has an extremely high multiplier effect. regardless of the viewpoint taken, any reduction in illegal harvest would have a compensatory value in reducing the non-hunt mortality estimate for wildlife managers, and allow for an increase in moose hunting opportunities. estimates of illegal moose kill fluctuate across north america, ranging from 5100% of the legal harvest level, with a mean of 30% (wolfe 1987). wolfe (1987) reported that ontario’s estimated illegal moose kill was 10% of the legal harvest based on a 1983 questionnaire. mercer and manuel (1974) estimated that the illegal moose kill accounted for 5 – 10% of the moose population in accessible areas of newfoundland in the early 1970s. violation simulation studies indicate low detection rates (< 1%) of violations by enforcement staff and low violation reporting rates (< 10%) by the public to enforcement agencies (vilkitis 1971, pursley 1977, bessey 1984, boxall and smith 1987). while estimates of the number of violators and the number of illegally killed wildlife using these violation simulations have poor statistical precision, alces vol. 40, 2004 todesco illegal moose kill in ontario 157 the estimates are useful in that they suggest a higher incidence of illegal kill than previously assumed (wolfe 1987). illegal moose kill estimates documented here are likely overestimated using the regression analysis, as the relationship is probably more curvilinear than linear. however, in the absence of any violation simulation exercises or other substantive estimates in ontario for illegal moose harvesting, the 2 estimates derived in this report on annual illegal kill provide a baseline on which further testing can be made. wolfe (1987) states “additional research is necessary to improve means of quantifying the magnitude of illegal kill and of separating out the relative contribution of various components”. efficiency and effectiveness of wildlife enforcement programs are difficult matters to assess and enhance to ensure violation deterrence and compliance with legislation (cowles et al. 1979, bessey 1984, hall et al. 1990, gregorich 1992). hunter compliance with legislation is directly related to favorability of attitude towards the legislation (bessey 1984). while overall moose hunting non-compliance rates of hunters checked by conservation officers are less than 10% in the ner, there are limitations on the relevance of simple compliance estimations (cowles et al. 1979). this is best illustrated by the 2002 statistics which had the lowest overall non-compliance rate, and a 35% increase in illegal moose kills from the previous year. condonation of illegal wildlife harvesting occurs in many jurisdictions across north america (vilkitis 1971, bessey 1984, hall et al. 1990, gregorich 1992), and can limit the effectiveness of wildlife enforcement. the elimination of public acceptance of illegal wildlife harvesting and the imposition of penalties that are severe enough to provide deterrence are required to reduce the illegal harvest of moose. one of the primary purposes of the moose watch program was to increase public and stakeholder awareness, and to provide a general deterrence through their involvement in compliance monitoring and violation reporting. the moose watch program has been effective in dealing with illegal harvesting activities and has received close to 400 calls since its inception, especially as other jurisdictions have noted low rates of violation reporting by the public (vilkitis 1971, pursley 1977, bessey 1984, boxall and smith 1987). despite calls that deal with non-enforcement matters or lead officers to investigate occurrences in which no charges are laid, the favorable public response to the program indicates the effectiveness of the promotional program and acceptance by the hunting and non-hunting community. efficient enforcement action can be initiated by conservation officers investigating timely and accurate complaints. in 2000, the moose watch program had a $25,000 budget for initial start up costs and promotional materials. one call to the violation reporting line in october 2000 regarding 2 illegal moose led to the discovery of a third illegal moose, and resulted in the conviction of 6 poachers, fines totaling $34,500 and 29 years of hunting suspensions. this case alone paid for the entire moose watch program. the moose watch program is not the panacea for enforcement in the ner, but rather another tool available to conservation officers. the illegal moose kill is not uniformly distributed across the ner wmus, and strategic enforcement initiatives are required in problem districts, including enhanced promotion and education, increased uniformed officer presence, and special investigations. hall et al. (1990) state that “actions to reduce violations of recreational hunting regulations can be as effective as those that limited commercial hunting”. continued hunter and public support, and adequate and efficient law enillegal moose kill in ontario – todesco alces vol. 40, 2004 158 forcement presence will be required to reduce the illegal harvesting of moose in the ner. conclusions illegal harvesting of moose in the ner is an issue affecting the general public, hunters, wildlife managers, and conservation officers. the verified illegal kill of 793 moose from 1997 – 2002 represents a loss of viewing and hunting opportunities, a recruitment loss to the moose herd, and tarnishes the image of lawful hunters. these verified kills represent a bare minimum number of illegally killed moose and demonstrate a non-compliance issue in localized areas within the ner. the actual level of illegal harvest is not known and modeling systems to determine the appropriate level of enforcement effort to suppress and deter this activity have not been developed. in order to continue reducing the illegal moose kill, stakeholder involvement and effective enforcement actions need to continue in the ner, using a blend of education, promotion, field enforcement, and appropriate penalties for those few that choose to violate. references beattie, k. h, c. j. cowles, and r.h. giles jr. 1980. estimating illegal kill of deer. pages 65 – 71 in r.l. hine and s. nehls, editors. white-tailed deer population management in the north central states. proceedings of a symposium held at the 41st midwest fish and wildlife conference, urbana, illinois, usa, 10 december 1979. benson, d. e. 2000. hunting ethics and the 6 r’s: relevance, reasoning, resources, respect, restraint and responsibility. pages 78 – 84 in w.d. mansell, editor. proceedings of the 2000 premier’s symposium on north america’s hunting heritage, ottawa, ontario, canada. bessey, k. m. 1984. analysis of the illegal harvest of white-tailed deer in agromanitoba: implications for program planning and management. m.sc. thesis, university of manitoba, winnipeg, manitoba, canada. bisset, a. 2002. 1999 and 2000 moose harvest in ontario. ontario ministry of natural resources, northwest science and information. nwsi technical report tr-131. thunder bay, ontario, canada. boer, a. 1991. hunting: a product or tool for wildlife managers? alces 27:74– 78. boxall, p. c, and l. c. smith. 1987. estimates of the illegal harvest of deer in alberta: a violation simulation study. occasional paper #2. alberta, forestry, lands, and wildlife, fish and wildlife division, resource economics and assessment section. edmonton, alberta, canada. b u s s , m. e., r. g o l l a t , and h. r. timmermann. 1989. moose hunter shooting proficiency in ontario. alces 25:98-103. byers, c. r., r. k. steinhorst, and p. r. krausman. 1984. clarification of a technique for analysis of utilization – availability data. journal of wildlife management 48:1050-1053. cederlund, g. n., and g. markgren. 1987. the development of the swedish moose population 1970-1983. swedish wildlife research supplement 1:55–62. cowles, c. j., k. h. beattie, and r. h. giles jr. 1979. limitations of wildlife law compliance estimators. wildlife society bulletin 7:188–191. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1:553–563. glover, r. l. 1982. characteristics of deer poachers and poaching in misalces vol. 40, 2004 todesco illegal moose kill in ontario 159 souri. m.sc. thesis, university of missouri, columbia, missouri,usa. gregorich, l. j. 1992. poaching and the illegal trade in wildlife parts in canada. canadian wildlife federation, ottawa, ontario, canada. hall, d. l., j. g bonnaffons, and r. m. jackson. 1990. the relationship of enforcement, courts and sentencing to compliance with waterfowl hunting regulations. journal of wildlife law enforcement 2:1-15. hardin, j. w., and j. l. roseberry. 1975. estimates of unreported deer loss resulting from a special deer hunt on crab orchard national wildlife refuge. conference proceedings of the southeast association of fish and wildlife agencies 29:460–466. hosie, r. c. 1979. native trees of canada. fitzhenry & whiteside limited. don mills, ontario, canada. karns, p. d., h. haswell, f. f. gilbert, and a. e. patton. 1974. moose management in the coniferous-deciduous ecotone of north america. naturaliste canadien 101:643–646. mercer, w. e., and f. manuel. 1974. some aspects of moose management in newfoundland. naturaliste canadien 101:657-671. neu, c. w., c. r. beyer, and j. m. peek. 1974. a technique for analysis of utilization – availability data. journal of wildlife management 38:541-545. pursley, d. 1977. illegal harvest of big game during closed season. new mexico department of game and fish, santa fe, new mexico, usa. timmermann, h. r. 1977. the killing proficiency of moose hunters. proceedings of the north american moose conference and workshop 13:13-25. _____. 1987. moose harvest strategies in north america. swedish wildlife research supplement 1:565–579. _____, and r. gollat. 1984. sharing a moose in north central ontario. alces 20:161-186. vilkitis, j. r. 1971. the violation simulation proves as reliable as field research in estimating closed-season illegal big game kill in maine. transactions of the northeast section of the wildlife society 28:141–144. wolfe, m. l. 1987. an overview of the socioeconomics of moose in north america. swedish wildlife research supplement 1:659–675. alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 41 differential habitat selection by moose and elk in the besa-prophet area of northern british columbia michael p. gillingham and katherine l. parker natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia, canada v2n 4z9, email: michael@unbc.ca abstract: elk (cervus elaphus) populations are increasing in the besa-prophet area of northern british columbia, coinciding with the use of prescribed burns to increase quality of habitat for ungulates. moose (alces alces) and elk are now the 2 large-biomass species in this multi-ungulate, multipredator system. using global positioning satellite (gps) collars on 14 female moose and 13 female elk, remote-sensing imagery of vegetation, and assessments of predation risk for wolves (canis lupus) and grizzly bears (ursus arctos), we examined habitat use and selection. seasonal ranges were typically smallest for moose during calving and for elk during winter and late winter. both species used largest ranges in summer. moose and elk moved to lower elevations from winter to late winter, but subsequent calving strategies differed. during calving, moose moved to lowest elevations of the year, whereas elk moved back to higher elevations. moose generally selected for mid-elevations and against steep slopes; for stunted spruce habitat in late winter; for pine-spruce in summer; and for subalpine during fall and winter. most recorded moose locations were in pine-spruce during late winter, calving, and summer, and in subalpine during fall and winter. elk selected for mid-elevations except in summer and for steep slopes in late winter. use and selection of 3 habitat classes were prominent for elk: deciduous and elymus burns, and subalpine. highest overlap between moose and elk occurred during fall and winter when both species used and strongly selected for subalpine habitat. neither elk nor moose selected areas to minimize the risk of wolf predation, but elk selected areas with lower risk of predation by grizzly bears and higher vegetation quality during calving and summer. alces vol. 44: 41-63 (2008) key words: alces alces, cervus elaphus, elevation, habitat selection, habitat use, home range, individual variation, movement rates, resource selection moose (alces alces) and elk (cervus elaphus) often provide the majority of prey biomass for large predators in complex predatorprey systems of north america. as examples, elk support wolf (canis lupus) populations in yellowstone and yukon (hayes and harestad 2000, smith et al. 2003), moose are common prey for wolves in northern coniferous forests (e.g., post et al. 2002, vucetich et al. 2002), and both moose and elk provide a prey base for wolves and grizzly bears (ursus arctos) in northern british columbia (bergerud et. al. 1983, bergerud and elliott 1998, parker and milakovic 2007). moose and elk are relatively profitable prey types in comparison to smaller ungulates or alternative prey, and given sufficient densities, can sustain large predator populations. hence, moose and elk are keystone species in the functioning of large-scale large-mammal systems. in multiungulate, multi-predator systems, however, they are not studied as commonly as other species because they are less susceptible to disturbance than some species (e.g., stone’s sheep, ovis dalli stonei), use smaller areas and, therefore, are not as subjected to landscape disturbance as other species (e.g., woodland caribou, rangifer tarandus caribou), or have large populations that are less vulnerable to, and can better accommodate change. in addition, the requirements of moose and elk are assumed to be relatively well known (e.g., habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 42 franzman and schwartz 1998, toweill and thomas 2002). because moose and elk are highly visible species with strong social and ecological values, and have the benefit of being high-profile game species, they are often managed to maintain or increase numbers. yet there are relatively few published studies that have examined the concurrent resource use by these 2 species (e.g., jenkins and wright 1988). moose have long occurred in northern british columbia (kelsall 1987), but elk herds are expanding into new areas in response to habitat fragmentation and management, and in some cases translocations (spalding 1992). prescribed burning has traditionally been used to create and maintain elk habitat in portions of northeastern british columbia (peck and peek 1991). fires temporarily result in shruband herb-dominated communities and increases in forage biomass, often with higher nutritional value. burning and its impacts on vegetative communities have been linked to the increase and expansion of elk herds (e.g., luckhurst 1973, silver 1976, parminter 1983). following fire, elk winter primarily on younger post-burn vegetation dominated by grasses or shrubs, except during severe winter conditions when there is higher use of conifer stands (peck and peek 1991). moose also frequent fire-associated habitats (peek 1998), but their use of burned habitats can depend on their past exposure to burned areas (gasaway et al. 1989). throughout most of their range, moose are primarily browsers and are associated with habitats containing a high proportion of preferred shrubs (boer 1998). elk on the other hand, are classified as grazers (cook 2002, stewart et al. 2002), but their food selection can shift in response to food availability and they are successful using browsing strategies (houston 1982). because of their flexible foraging habits, ability to use a wide variety of terrain types and high fecundity, elk could compete with moose for food (flook 1964) although competition between the 2 species is thought to be low (miller 2002). in complex predator-prey systems, high numbers of high-biomass ungulates may alter predator populations, and in turn other species in the same system. the overall goal of this study was to provide an initial description of habitat use and selection by moose and an expanding elk population in the besa-prophet area of northern british columbia. specifically, we asked whether there was overlap in use and selection by moose and elk that may have implications to the system, and if there might be potential impacts through predation on other species. to do this we used global positioning satellite (gps) radio-locations, remote-sensing imagery of vegetative communities, assessments of predation risk from concurrent studies on grizzly bears and wolves in the same area, and habitat selection modeling. these data and analyses help characterize the ungulate-predator landscape of the besaprophet watershed and contribute to better land-use planning. study area the muskwa-kechika management area (mkma) in northern british columbia is known for its abundance of large ungulates (moose, elk, caribou, stone’s sheep, a few mountain goats (oreamnos americanus) and deer (odocoileus spp.)) and large predators (wolves, grizzly bears, black bears (u. americanus), coyotes (canis latrans), wolverines (gulo gulo), and a few cougars (puma concolor)). the besa-prophet study area is within the besa-prophet pre-tenure planning area (fig. 1), one of several pre-tenure areas within the mkma requiring specific wildlife planning prior to resource extraction or development. it is a highly diverse area including the besa river south of the prophet river, and covering ~740,887 ha. located within the muskwa ranges and rocky mountain foothills between 57°11’ and 57°15’ n, and 121°51’ and 124°31’ w, the besa-prophet is characterized alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 43 by numerous east-west drainages and southfacing slopes (fig. 1) that provide benefits to wintering species because they are often blown free of deep snows. other than several permanent outfitter camps and 1 governmentdesignated all-terrain vehicle trail, there is relatively little access into the besa-prophet. this activity occurs mostly during late summer and fall during hunting seasons, with some snowmobile use in winter. valleys at ~800-1300 m are commonly lined with white spruce (picea glauca), some lodgepole pine (pinus contorta) and trembling aspen (populus tremuloides) on dry sites, and black spruce (p. mariana), willow-birch (salix spp., betula glandulosa) communities on poorly drained sites (meidinger and pojar 1991). there also are slopes that have been burned by the british columbia ministry of environment and local guide outfitters to enhance ungulate populations, especially stone’s sheep. the subalpine area is characterized by an abundance of willow and scrub birch, as well as some balsam fir (abies lasiocarpa) and white spruce often in krummholz form, and various grasses, sedges and fescues (festuca spp.). treeline occurs between ~1,450-1,600 m. alpine tundra above ~1600 m consists of permanent snowfields, rock, mat vegetation, and grasslands (demarchi 1996). anecdotal evidence in the besa-prophet suggests that elk populations are expanding, enabled by prescribed burns that are conducted primarily for the enhancement of stone’s sheep populations. in this area, prescribed fire has been officially managed since the early 1980s, although there also have been natural burns and locally initiated fires before and since that time. concerns regarding the implications of a rapidly increasing elk population on ecosystem dynamics were the impetus for the comparisons in our study. densities of moose and elk populations in the besa-prophet are not well documented; very rough estimates are approximately 2000 moose and 500 elk (j.p. elliott, bc ministry of environment, fort st john, personal communication). it is important to note, however, that habitat use and selection will vary as a consequence of population density (boyce et al. 2003). methods fifteen adult female moose and 14 adult female elk were fitted with gps (global positioning satellite) collars (gtx, advanced telemetry systems, isanti, mn) between march 2003 and january 2005. collars were programmed to record locations 4 times daily at 6-h intervals with a range of start times between midnight and 0500 hr. data were retrieved when collars were recovered at the end of a 1-year sampling period. we defined 5 seasons distinguished by biological and ecological characteristics for our analyses of range use and movements, and habitat use and selection: winter (1 november – 28 february) corresponding with the formation fig. 1. location of the study area within the besa-prophet pre-tenure planning area (inset) in northeastern british columbia. contour lines (200-m intervals) illustrate the predominance of east-west valleys within the study area. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 44 of sex-specific groups following rut; late winter (1 march – 15 may) when movement rates were lowest; calving (16 may – 15 june) during which parturient females became solitary and the onset of plant greening occurred; summer (16 june – 15 august) from plant green-up through peak vegetation biomass to the start of plant senescence; and fall (16 august – 31 october) when senescence of vegetation occurred, males and females formed mixed sex groups, and females came into estrus. seasonal movements and home ranges to identify seasonal movement rates of moose and elk, we averaged movement rates (m/h) per individual using gps locations obtained from consecutive 6-h fixes by season, and then averaged across individuals for each season by species (n = number of individuals). we compared movement rates between moose and elk within seasons using a repeated measures analysis of variance (anova), with differences identified following bonferroni adjustment of confidence intervals. we used the same approach to compare seasonal elevations used by both species. we determined sizes of annual and seasonal home ranges using 100% minimum convex polygons (mcp, jennrich and turner 1969) around gps locations, as well as by fixed-kernel analysis (worton 1989) for each individual. the mcps, calculated by connecting the outer locations of all animal-use points, tend to overestimate range sizes for animals that have infrequent movements away from a centralized area, but they provide a relative comparison of the extents of the landscape used by moose and elk. in fixed-kernel analyses, kernels are calculated from the 95% probability density of all locations and delineate areas of higher use (core areas) within the home range. depending on the arrangement of animal locations, the fixed-kernel analysis may identify multiple core areas. hereafter, we use the term ‘kernel area’ to refer to the total area identified by the fixed-kernel analysis. because kernels are mathematical interpolations, however, they may exclude some areas where movements take place between core areas, and may include substantial ‘buffer’ areas with no animal locations around highdensity locations, particularly with small numbers of locations (seaman et al. 1999). for comparison with other studies on moose and elk, however, we used the animal movement extension (hooge and eichenlaub 2000) in arcview (esri 2002) to calculate both mcp and fixed-kernel range sizes. we used anova to compare annual and seasonal range sizes between species within season. we used spatialviewer (m. gillingham, unpublished visual basic program) to examine movement patterns of individual animals. habitat use and availability to compare seasonal use of habitat classes between moose and elk, we determined the proportion of gps locations within each class by individual within season. to index resources available to each collared individual, we defined availability at the scale of seasonal movement, within johnson’s (1980) third order of selection. seasonal movement is an animal’s movement potential within a season (e.g., gustine et al. 2006b), or a circle around each use point with a radius equivalent to the distance traveled at each individual’s 95th percentile movement rate from 6-h gps fixes. within that area of movement potential, we selected 5 random points for availability locations. we used a raster geographic information system (gis; imageworks xpace; pci geomatics corp. 2001) to query habitat classes for used and available points. we ensured that no 2 points were used twice and that there was no overlap between used and available points (manly et al. 2002). for both moose and elk within season, we then averaged the proportions of habitat classes that were used by, and available to, each individual to eliminate effects of uneven sample sizes among individuals (se was based on number of collared alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 45 individuals). we visually compared use to availability of different habitat classes, but then determined resource selection for combinations of additional variables because habitat use occurs in response to multiple variables and not to vegetation class alone. we defined 10 habitat classes for the besaprophet based on a vegetation classification system with 25-m resolution for the area, developed using landsat tm and enhanced thematic mapper (etm) remote-sensing imagery (lay 2005) (table 1). two burn classes (elymus burn and deciduous burn) may have included small, but unknown amounts of other disturbed areas such as avalanche chutes, which could not be distinguished separately with remote-sensing imagery. avalanche chutes were relatively rare in the areas used by moose and elk. resource selection we used a suite of gis layers to extract the attributes for defining resource selection by moose and elk by season, using all gps use locations and available locations as defined above. these layers included habitat class, vegetation biomass, vegetation quality, slope, aspect, elevation, and risk of predation by grizzly bears and wolves. habitat class and vegetation indices – in addition to defining habitat class as we did in analyses for habitat use and availability, we used the same tm (4 june and 22 july 2001) and etm (15 august 2001) imagery to index vegetation biomass during june, july, and august using a normalized difference vegetation index (ndvi) that is related to leaf area and plant biomass (tucker and sellers 1986, ruimy et al. 1994). we also developed an index to vegetation quality during the calving and summer seasons by calculating the change in ndvi (subtraction of individual pixel values) between june and july, and july and august images. a positive change in ndvi during the growing season corresponds with growth of new tissue (groten and ocatre 2002) and highest rates of green-up are likely the most digestible forage (griffith et al. 2002, oindo 2002). habitat class description non-vegetated rock and rock habitats; permanent snowfields or glaciers and water bodies. elymus burn recently burned and open disturbed sites dominated by elymus innovatus. deciduous burn older burned and disturbed areas containing populus tremuloides and populus balsamifera shrubs (<2 m) and trees (≥2 m). subalpine deciduous shrubs ≥1600 m in elevation; and spruce-shrub transition zone at middle to upper elevations with white and hybrid spruce (picea glauca and p. glauca x engelmanni), and dominated by birch and willow. stunted spruce low productivity sites often on north-facing slopes with picea glauca of limited tree height and percent cover. pine-spruce white and hybrid spruce-dominated communities; and lodgepole pine dominated communities. riparian low-elevation, wet areas with black (picea mariana) and hybrid spruce; often with standing water in spring and summer; exposed gravel bars adjacent to rivers and creeks. alpine dry alpine tundra habitat ≥1600 m characterized by dryas spp.; and wet alpine tundra habitat ≥1600 m dominated by cassiope spp. and sedge (carex spp.) meadows. low shrub deciduous shrubs <1600 m dominated by birch and willow. carex wetland meadows dominated by sedges (carex spp.) at elevations <1600 m, with intermittent salix shrubs. table 1. description of the 10 habitat classes, derived from landsat tm and enhanced thematic mapper remote-sensing imagery, used to describe habitat use and selection by moose and elk in the besa-prophet area of northern british columbia. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 46 slope, aspect and elevation – we obtained elevation, slope, and aspect from a digital elevation model (dem) in the 1:20,000 british columbia terrain and resource inventory management program (british columbia ministry of crown lands 1990). we modeled aspect as 2 continuous variables (i.e., northness and eastness; roberts 1986) to avoid introducing additional categorical variables. northness (the cosine of aspect) values of 1.00 and -1.00 suggest selection for north and south aspects, respectively, whereas values near 0.00 suggest selection for east and west aspects. eastness (the sine of aspect) values show selection for east (i.e., 1.00) and west (i.e., -1.00) aspects; values of 0.00 show selection for northern/southern exposures (palmer 1993). we did not assign an aspect to pixels with a slope ≤1°. predation risk – we defined potential risk of predation using logistic regression models by season from gps-collared wolves and grizzly bears, which are assumed to be the most significant large mammal predators in the besa-prophet area (bergerud and elliott 1998). details of these predator models are in gustine et al. (2006a, b) and walker et al. (2007). the predation-risk models included slope, aspect, elevation, habitat class, fragmentation (an index of vegetation diversity), and distance to linear features (e.g., seismic lines). from these models, we generated a risk surface as a gis layer that defined selection value to grizzly bears or wolves in each season by applying the coefficients from models to each 25 x 25-m pixel, based on its topographic and vegetation features. we scaled values from 0 to 1 to standardize selection surfaces, and then assumed that the risk of predation to moose and elk by grizzly bears and wolves was directly related to selection values of the predators. model calving summer fall winter late winter elevation1+aspect+habitat2      elevation+slope+aspect+habitat      wolf 3+habitat      elevation+slope+aspect+wolf+bear3+biomass+habitat    elevation+slope+aspect+wolf+bear+quality+habitat   aspect+wolf+bear+biomass+habitat    aspect+wolf+bear+quality+habitat   elevation+slope+aspect+wolf+biomass+habitat    elevation+slope+aspect+wolf+quality+habitat   aspect+wolf+biomass+habitat    aspect+wolf+quality+habitat   table 2. candidate models developed a priori to describe resource selection by moose and elk by season in the besa-prophet area of northern british columbia. vegetation biomass during calving, summer, and fall was based on ndvi measures for june, july, and august, respectively. vegetation quality, assessed by the change in ndvi between summer months, was only used in calving and summer models. no risk of predation by grizzly bears was included during hibernation (winter and late winter seasons). 1 elevation was modeled as a quadratic with both a linear and squared term. 2 habitat as defined by classes in table 1. 3 wolf and bear represent risk of predation by wolves and grizzly bears, respectively; see text for details. alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 47 selection models – we developed 11 a priori, ecologically plausible models (table 2) from the previously described attributes to define resource selection by moose and elk by season. we used logistic regression with these parameters (k) to characterize differences between use and availability for each individual, and ranked the suite of models using akaike’s information criterion (aic) values corrected for small sample size (aicc) when n/k < 40 (burnham and anderson 2002). deviation coding was used for categorical variables (hendrickx 1999). to avoid inflated coefficients and inflated error terms in the models, we used tolerance scores to assess model inputs for collinearity and multicollinearity. if tolerance scores were <0.20, covariates were not included in the same model (menard 2002). because logistic regression models do not provide reliable estimates if there is complete or near-complete separation in levels of categorical variables (menard 2002), we dropped both used and available points for those habitats in which either use or available points in a habitat were ≤4 (gillingham and parker 2008). we validated the top models using kfold, cross-validation (boyce et al. 2002), and an averaged spearman’s rank correlation coefficient. within each model set (i.e., by individual and season), we calculated akaike weights (wi). if there was not a single model for which this relative weight of evidence, wi, was ≥0.95, we considered competing models to be those for which the sum of wi was ≥0.95 (burnham and anderson 2002). for each model set for each individual animal, we averaged the selection coefficients (β) in competing models based on their relative wi. we evaluated the importance of specific resources to moose and elk in general after developing pooled models from these averaged individual models by averaging coefficients from the individual models within species during each of the 5 seasons. we assumed significance of all tests at α = 0.05. we used stata for all statistical and modeling procedures (version 9.2; statacorp 2007). all means are presented as x ± 1 se unless noted otherwise. results we retrieved 14,534 gps locations from 14 of the collared moose and 14,870 locations from 13 collared elk. the number of gps locations recorded as a percentage of the number of attempted gps locations was 76.7 ± 0.03% (x ± se) for moose and 82 ± 1.6% for elk. seasonal movements and home ranges distances moved between consecutive 6-h gps fixes ranged from <1 m to 14.5 km (straight-line distance) by moose and from <1 m to 10.3 km by elk. both species moved at lowest rates during winter and late winter (35-41 m/h), and then increased movements to highest rates in summer (>100 m/h) (fig. 2). moose usually tended to move at rates slightly lower than elk (repeated-measures anova, p = 0.049), but these rates were significantly lower only during the calving season (moose = 59 ± 21 m/h, elk = 93 ± 27 m/h). the range sizes estimated by fixed-kernel analyses were always smaller, as expected, than those determined by mcp for both fig. 2. comparative differences in the movement rates (m/hr, x ± se) of adult female moose (n = 14) and elk (n = 13), averaged by individual and then averaged across individuals, by season in the besa-prophet area of northern british columbia. seasonl ate w inte r ca lvin g su mm er fal l wi nte r m ov em en t r at e (m /h ) 0 20 40 60 80 100 120 140 moose elk habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 48 species (table 3). kernel areas were most comparable to mcps (~85-90 % of mcp size) during the calving season for moose, and during calving and fall for elk. they were less than half the size of mcps during fall for moose, and during winter and late winter for elk. annual home ranges by mcp for moose averaged 195 km2, but were highly variable among individuals, ranging from a minimum of 39 to a maximum of 899 km2 (kernel area = 14-124 km2). seasonal ranges for moose were typically smallest during the calving season (18 km2), and more than 7 times larger during summer. annual ranges of mcp for elk averaged 191 km2 and were not different than those of moose (table 3), and also were highly variable among individuals (mcp range = 50-1000 km2; kernel area = 10107 km2). excluding 1 animal that made a large circular excursion in july away from its other seasonal use areas, annual ranges for elk averaged 123 ± 19 km2 (range = 50-250 km2; kernel area = 10-107 km2). in contrast to moose, seasonal ranges for elk were smallest during the winter and late winter seasons (16-20 km2). similar to moose, seasonal ranges for elk were largest during summer (table 3). compared to elk, the seasonal ranges of moose (by mcp) were more than twice as large during winter and late winter, but less than half as large during calving. habitat use and availability moose and elk used elevations on the landscape differently among seasons (repeated-measures anova, p < 0.001) (fig. 3). during calving and summer (may–august) and winter (november–february), moose used lower elevations than elk (moose = 1333 ± 81 m in calving, 1397 ± 53 in summer, 1519 ± 88 in winter; elk = 1551 ± 49 m in calving, 1671 ± 40 m in summer, 1624 ± 51 m in winter) (all p <0.016). both species moved to lower elevations from winter to late winter, however, calving strategies appeared to differ between the 2 species. in june, moose were at lowest elevations of the year, and after the calving season moved gradually upslope durseason estimate moose1 elk p mean se mean se annual mcp 195.28 59.71 190.81 69.61 0.961 kernel 57.04 8.81 45.64 8.19 0.355 calving mcp 17.59 4.09 38.00 4.05 0.002 kernel 15.05 4.28 34.09 5.82 0.013 summer mcp 132.80 60.25 118.38 63.66 0.871 kernel 99.42 52.81 57.76 28.70 0.504 fall mcp 46.85 10.96 60.22 12.86 0.434 kernel 26.86 6.69 51.02 13.16 0.107 winter mcp 45.91 9.83 20.40 2.24 0.015 kernel 32.16 8.88 8.82 1.14 0.012 late winter mcp 30.49 4.60 15.52 4.60 0.030 kernel 25.49 6.39 6.04 1.60 0.009 table 3. comparison of annual and seasonal home-range sizes (km2) for 14 female moose and 13 female elk based on 100% minimum convex polygons (mcp) and 95% fixed-kernel (kernel) estimates. p-values are from one-way anovas comparing home-range sizes between species for a given technique and season. 1 only 12 animals were used in the winter models because of collar failure. alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 49 ing summer and fall. elk, in contrast, moved from their late wintering areas to higher elevations to calve in june, continuing upslope in july, and descending to the same elevations used by moose in the fall (fig. 3). predominant use of specific habitat classes differed between moose and elk, coinciding with some of the elevational differences between the species. during late winter, calving, and summer, most locations for moose (2836% across animals) were in the pine-spruce habitat class (fig. 4). this contrasts to the fall and winter periods, when most locations were in subalpine vegetation (33-39%). there was relatively little use by moose during any season of riparian (<6 % of locations; distinct from low shrub and carex), alpine (<4%), or non-vegetated (<1%) habitat classes. low shrub vegetation was used least (10 %) by moose during the calving season, and most during winter (22%). use of the deciduous burn class by moose was relatively consistent at 13-16% throughout the year. for elk, the use of 3 habitat classes was prominent: elymus burn, deciduous burn, and subalpine (fig. 5). seasonally, the 3 classes always totaled between 59 and 83% of use locations. highest use by elk occurred in the subalpine in all seasons (33.0% of locations during calving; 29.5% during fall, and 40.5% in winter) except late winter (13.1% of locations), when they increased use of both burned habitat classes (~70% of locations). during summer when elk used the subalpine more than any other season (64% of locations), they spent less time in elymus and deciduous burn habitats. compared to moose, elk used the pine-spruce habitat class very little (<5% of locations) in any season except fall (17 ± 2%). resource selection there were relatively few variables that were selected for or against by moose because of variation in individual selection strategies (table 4). in general, moose selected for mid-elevations in all seasons but late winter. during calving, summer, and fall they avoided non-vegetated areas and steep slopes. in the calving season, most moose rarely used (<4 locations) subalpine (10/14 animals), alpine (13/14), or carex (12/14) habitat classes even though avoidance of these classes was not indicated by the pooled calving model. moose selected strongly for pine-spruce habitat in summer, subalpine and low shrub in fall, subalpine in winter, and stunted spruce in late winter. in no seasons were vegetation biomass and quality important to overall resource selection (although some individual moose selected for or against these variables). risk of predation was an important factor for some moose, but was only important in the pooled selection models during fall, when moose locations were in areas of relatively higher predation risk than that present in the area around them. elk exhibited selection by season for and against more variables than did moose (table 5). mid-elevations were selected in all seasons but summer, when higher elevations were more important. elk selected against steep slopes from summer through winter, but for them in late winter. they consistently selected against northness and alpine habitat in all seasons, fig. 3. comparative differences in elevations (x ± se) used by gps-collared female moose (n = 14) and elk (n = 13), averaged by individual and then averaged across individuals, by season in the besa-prophet area, northern british columbia. seasonl ate w inte r ca lvin g su mm er fal l wi nte r e le va tio n (m ) 0 1200 1400 1600 1800 elk moose habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 50 and non-vegetated habitat in all seasons but calving. during calving, elk selected elymus burns and stunted spruce (although <2% of locations across animals were in this class) and avoided the subalpine habitat class (even though it comprised 32 ± 5% of use locations). in summer, they selected for subalpine as well as stunted spruce. in both calving and summer, elk selected for areas with high vegetation quality (as measured by change in ndvi) even though the risk of wolf predation was relatively high in those areas, and against the risk of bear predation (although coefficients were very small; table 5) and low shrub habitat. from calving through fall, elk selected against areas with high vegetation biomass (as indexed by ndvi). from fall through late winter, there was strong selection for elymus fig. 4. comparison of proportional use versus availability (+ se) of habitat classes for female moose in the besa-prophet area of northern british columbia. standard errors were determined from averages for each individual by season. p ro po rt io n summer p ro po rt io n 0.0 0.1 0.2 0.3 0.4 0.5 winter late winter 0.0 0.1 0.2 0.3 0.4 0.5 no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x p ro po rt io n 0.0 0.1 0.2 0.3 0.4 0.5 no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x calving 0.0 0.1 0.2 0.3 0.4 0.5 p ro po rt io n no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x p ro po rt io n fall 0.0 0.1 0.2 0.3 0.4 0.5 used available no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 51 fig. 5. comparison of proportional use versus availability (+ se) of habitat classes for female elk in the besa-prophet area of northern british columbia. standard errors were determined from averages for each individual by season. late winter 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 summer 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 winter 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 p ro po rt io n no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x p ro po rt io n p ro po rt io n no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x calving 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 fall 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 used available p ro po rt io n p ro po rt io n no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x no nve ge ta te d el ym us b ur n de cid uo us b ur n su ba lp in e st un te d sp ru ce pi ne s pr uc e ri pa ria n al pi ne lo w sh ru b ca re x and deciduous burns and subalpine habitat. the pooled selection models indicated that the carex habitat class was selected by elk in all seasons, but <1% of used locations were in this class and the majority of individuals (n = 7-11, depending on season) rarely or never used the class. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 52 parameter calving summer coef se p + coef se p + elevation 78.94 40.01 0.048 7 1 40.74 15.75 0.010 10 0 elevation2 -30.0 15.19 0.048 1 7 -14.4 5.59 0.010 0 10 slope -0.05 0.02 0.029 1 6 -0.04 0.01 0.007 1 8 northness <0.01 0.18 0.994 0 1 -0.06 0.14 0.672 0 1 eastness 0.01 0.19 0.962 1 0 -0.01 0.13 0.968 0 0 wolf risk -1.91 2.11 0.365 1 2 0.46 0.97 0.639 2 0 bear risk 0.06 1.30 0.962 1 1 0.48 0.99 0.624 0 0 biomass -0.49 0.88 0.574 0 1 -1.27 1.00 0.206 1 3 quality 1.28 1.61 0.428 2 1 0.78 0.77 0.308 3 0 nonveg -1.22 0.15 <0.001 0 1 -2.47 0.50 <0.001 0 4 elymus burn 0.25 0.37 0.487 1 1 0.23 0.51 0.648 2 1 deciduous burn 0.24 0.45 0.594 3 1 0.62 0.35 0.077 5 0 subalpine -0.17 0.28 0.554 0 0 0.59 0.50 0.234 5 1 stunt 0.07 0.39 0.854 1 0 0.57 0.41 0.167 5 0 pine/spruce -0.19 0.51 0.711 2 2 0.80 0.33 0.016 5 0 riparian -0.45 0.53 0.398 1 1 0.50 0.51 0.328 3 1 alpine 0.18 0.09 0.055 0 0 -0.13 0.59 0.831 0 0 low shrub 0.51 0.51 0.321 1 0 0.08 0.40 0.834 3 4 carex 0.11 0.14 0.434 0 0 -0.80 0.37 0.032 1 1 parameter fall winter1 late winter coef se p + coef se p + coef se p + elevation 54.99 23.93 0.022 7 0 37.65 13.67 0.006 6 2 58.26 33.70 0.084 7 0 elevation2 -18.6 8.15 0.023 0 7 -12.3 4.41 0.005 1 7 -21.7 12.84 0.092 0 7 slope -0.03 0.01 0.005 0 8 -0.01 0.01 0.235 2 4 -0.01 0.01 0.305 2 4 northness 0.01 0.12 0.946 1 0 <0.01 0.08 0.987 0 1 0.03 0.10 0.748 0 0 eastness -0.02 0.13 0.879 0 1 -0.02 0.08 0.838 0 0 -0.02 0.11 0.836 0 1 wolf risk 0.94 0.44 0.033 4 0 0.47 0.62 0.443 2 0 -0.21 0.76 0.787 1 2 bear risk -0.06 1.35 0.966 2 2 biomass -1.16 1.25 0.351 1 5 quality <0.01 0 0 nonveg -1.39 0.51 0.006 1 3 -0.74 0.46 0.107 0 4 -0.46 0.47 0.330 0 2 elymus burn 0.30 0.37 0.413 4 0 -0.10 0.34 0.779 1 1 0.00 0.36 0.989 2 0 deciduous burn 0.65 0.36 0.068 8 0 0.35 0.18 0.055 6 1 0.29 0.24 0.220 2 0 subalpine 1.12 0.30 <0.001 11 0 0.53 0.23 0.021 9 0 0.09 0.27 0.738 2 0 stunt -0.36 0.44 0.404 0 1 0.06 0.23 0.782 1 0 0.42 0.22 0.050 6 1 pine/spruce 0.01 0.32 0.966 1 3 -0.22 0.20 0.275 2 6 0.08 0.22 0.733 2 1 riparian -0.30 0.38 0.430 0 2 0.08 0.26 0.774 1 0 -0.09 0.41 0.822 1 0 alpine -0.94 0.67 0.157 0 5 -0.40 0.40 0.310 0 4 -0.27 0.33 0.414 0 1 low shrub 0.87 0.35 0.012 9 0 0.44 0.23 0.054 6 0 -0.06 0.37 0.882 2 0 carex 0.03 0.21 0.888 0 0 -0.01 0.25 0.970 1 0 -0.01 0.20 0.979 0 0 table 4: comparison of significant selection coefficients by season from averaged resource selection models for 14 female moose in the besa-prophet area of northern british columbia. for each season, the number under the + indicates the number of individual final models that showed selection for that parameter; the number under the – indicates the number of individuals that avoided that attribute. coefficients, se, and p-values refer to the pooled models. 1 only 12 animals were used in the winter models because of collar failure. alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 53 parameter calving summer coef se p + coef se p + elevation 37.31 4.82 <0.001 8 7.81 3.65 0.032 7 3 elevation2 -11.9 1.55 <0.001 1 9 -2.02 1.09 0.065 3 7 slope 0.00 0.01 0.350 4 5 -0.02 0.00 <0.001 2 9 northness -0.31 0.07 <0.001 4 8 0.20 0.05 <0.001 8 3 eastness -1.74 0.12 <0.001 13 -0.71 0.06 <0.001 12 wolf risk 0.98 0.48 0.042 6 2 1.22 0.35 <0.001 6 4 bearrisk -0.00 0.00 <0.001 10 2 -0.00 0.00 <0.001 4 5 biomass -2.68 0.30 <0.001 1 7 -1.51 0.13 <0.001 9 quality 5.46 0.88 <0.001 10 1 5.62 0.67 <0.001 12 nonveg 0.06 0.28 0.824 1 6 -0.91 0.29 0.002 2 6 elymus burn 0.34 0.16 0.036 6 4 -0.09 0.20 0.660 3 4 deciduous burn 0.14 0.17 0.415 8 3 0.25 0.19 0.200 6 3 subalpine -0.63 0.21 0.003 4 7 0.39 0.20 0.045 6 2 stunt 0.48 0.18 0.006 4 2 0.59 0.18 0.001 5 1 pine/spruce 0.05 0.23 0.814 3 6 0.11 0.24 0.655 6 4 riparian 0.35 0.21 0.097 4 3 0.56 0.18 0.002 5 1 alpine -0.84 0.19 <0.001 1 7 -1.25 0.18 <0.001 11 low shrub -1.02 0.24 <0.001 3 6 -0.89 0.23 <0.001 2 10 carex 1.07 0.14 <0.001 4 1.23 0.17 <0.001 5 parameter fall winter late winter coef se p + coef se p + coef se p + elevation 8.58 2.46 <0.001 7 4 11.50 2.63 <0.001 5 6 54.31 6.19 <0.001 12 1 elevation2 -2.55 0.85 0.003 5 7 -2.49 0.83 0.003 6 5 -18.1 2.13 <0.001 1 12 slope -0.02 0.00 <0.001 2 9 -0.04 0.00 <0.001 12 0.02 0.01 <0.001 8 3 northness 0.02 0.04 0.649 6 5 0.06 0.03 0.059 6 4 -0.09 0.07 0.197 3 6 eastness -0.45 0.06 <0.001 1 11 -0.81 0.05 <0.001 13 -1.90 0.16 <0.001 13 wolf risk -0.03 0.14 0.841 3 3 0.58 0.29 0.047 5 5 -0.22 0.40 0.587 3 6 bearrisk -0.28 0.44 0.526 4 6 biomass -2.92 0.47 <0.001 2 11 quality nonveg -2.63 0.33 <0.001 12 -1.52 0.18 <0.001 12 -0.84 0.23 <0.001 1 7 elymus burn 0.91 0.15 <0.001 12 1 0.24 0.08 0.005 8 2 0.45 0.12 <0.001 8 deciduous burn 1.39 0.13 <0.001 13 0.37 0.07 <0.001 10 2 0.59 0.11 <0.001 11 2 subalpine 1.01 0.10 <0.001 11 1 0.51 0.07 <0.001 8 1 0.34 0.16 0.034 6 2 stunt -0.40 0.19 0.032 4 6 0.10 0.11 0.386 5 3 -0.69 0.29 0.018 7 pine/spruce 0.00 0.09 0.985 5 4 0.01 0.10 0.940 5 5 -0.66 0.11 <0.001 7 riparian -0.52 0.17 0.002 4 6 0.23 0.18 0.183 6 2 0.26 0.16 0.111 4 1 alpine -0.71 0.14 <0.001 1 9 -0.30 0.09 0.001 2 7 -0.32 0.15 0.040 2 6 low shrub 0.51 0.14 <0.001 7 1 0.20 0.11 0.077 6 4 0.37 0.17 0.029 5 1 carex 0.44 0.16 0.007 3 0.18 0.03 <0.001 2 0.50 0.11 <0.001 3 table 5: comparison of significant selection coefficients by season from averaged resource selection models for 13 female elk in the besa-prophet area of northern british columbia. for each season, the number under the + indicates the number of individual final models that showed selection for that parameter; the number under the – indicates the number of individuals that avoided that attribute. coefficients, se, and pvalues refer to the pooled models. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 54 discussion resource partitioning by ungulates typically occurs relative to habitat, dietary, and special niche requirements (bowyer and kie 2004). both moose (miquelle et al. 1992, bowyer 2004) and elk (peek and lovaas 1968, weckerly 1998) sexually segregate to a greater or lesser extent throughout their range. consequently, we make inferences only to female moose and elk in the besa-prophet, with the recognition that our findings provide preliminary comparative insights based on only 1 year of data for each species. further, although locations of animals were based on 4 fixes per day distributed throughout noctural and diurnal periods, these fixes that were 6 h apart do not allow us to examine habitat selection at finer feeding-patch scales. seasonal movements and home ranges home-range estimates are difficult to compare among studies because of differences in methodologies (e.g., definitions of seasons), analyses (lawson et al. 1997), and available resources. in this study, most fixed-kernel home ranges had multiple core areas and, particularly for moose, often did not include large portions of the valley bottoms that were used regularly for moving back and forth between core areas within a season. consequently, defining availability of resources within just the core areas would have greatly underestimated available habitat during a given season. we observed large seasonal variation in moose and elk home ranges whether we compared the extent of use (mcp) or kernel areas. sizes of annual home ranges for moose in the besa-prophet probably reflect the relatively open, mountainous terrain of the study area and were more similar in size to mcp home ranges of female moose reported for the kenai peninsula of alaska (hundertmark 1998: 128 km2), south-central alaska (ballard et al. 1991: 290 km2), and the mackenzie valley of northwest territories (stenhouse et al. 1994: 174 km2) than to the smaller ranges from more southern latitudes (hundertmark 1998). in summer, when moose exhibited the longest average movements between 6-h fixes, the extent of their ranges (mcp) was also comparable to the large home ranges reported for south-central alaska (hundertmark 1998). in winter, the seasonal ranges of moose were comparable in size to mcp home ranges at more similar latitudes (e.g., 42-47 km2 in north-central alberta; lynch and morgantini 1984). cederlund and sand (1994) observed that female moose with calves had larger home ranges than those without calves. although we do not know the reproductive status of the animals in our study, home ranges for moose during calving were the smallest of all seasons, an observation that would be consistent with an ungulate with a hiding reproductive strategy. elk also hide their young, and their home ranges during calving were small, although they were not the smallest by season. home-range size can be related to the abundance of important resources because a seasonal home range must meet an animal’s energy and nutritional requirements (anderson et al. 2005). for example, the home-range sizes of elk in summer and winter in alberta and wisconsin were inversely related to mean forage biomass; elk increased home-range sizes in winter when quality of forage was reduced and snow cover reduced forage biomass (anderson et al. 2005). in our study, elk did not have larger range sizes in winter or late winter than summer. rather, home ranges were smallest during winter and late winter. this may indicate that snow was more limiting in our system, but perhaps more likely that food was not limiting, particularly on the wind-blown south-facing slopes in the besa-prophet. elk home ranges during summer are reported to be highly variable even within the same study area (strohmeyer and peek 1996: mcp = 79-593 km2) because of the juxtaposition of habitat components within alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 55 the area. summer was also the season with the most variability in home ranges for elk in the besa-prophet. in our study, elk exhibited the highest 6-h movement rates during summer (as much as 10 km in 6 h) and summer home ranges based on fixed-kernel estimates were comparable to those reported for largeherd, migratory elk in yellowstone national park (boyce 1991). we recorded a very long movement of a minimum of 138 km (mcp area around those locations was ~800 km2) over 20 days in july by 1 animal that returned to its original starting point. edge et al. (1986) found that only 2-3% of marked female elk dispersed and most other studies have reported strong philopatry by female elk for seasonal and annual home ranges (e.g., craighead et al. 1972). resource selection models calculations of use, based on proportions of gps locations, give some indication of areas that are most frequented, and therefore, important to managers. selection of locations by animals on the landscape, however, is usually a response to multiple variables and not simply to elevation or habitat alone. selection models allow the quantification of tradeoffs that animals make, for example, in relation to predation risk or nutritional value. at the same time, selection models require that numerous assumptions be met (e.g., thomas and taylor 2006), sometimes resulting in the exclusion of habitats that are rarely used or completely avoided (gillingham and parker 2008). in our analyses, we presented seasonal selection models for moose and elk after pooling individual models instead of developing a seasonal global model across individuals. rather than unequally weight individual animals with more gps locations, delete known observations from some animals to equalize sample sizes (thomas and taylor 2006), or use a random-effects approach to account for unequal sample sizes (gillies et al. 2006), we chose to develop models for each individual in each season, dropping those habitats that had complete or near-complete separation (<4 used points). further, our approach for considering which competing models were averaged was conservative (sum wi ≥ 0.95) (burnham and anderson 2002). we then determined a pooled model for each season by averaging the final individual models within each season. there are 2 different approaches that can be used when averaging competing models for which estimates of all coefficients are not found in each model (e.g., predation risk was not in all competing models; burnham and anderson 2002). we chose the more conservative approach of assuming that missing coefficients had the value of 0 in a given model. our approach ensures that the final models are not unduly weighted by individual animals (burnham and anderson 2002), but it has the effect of reducing the magnitude of coefficients that occur in only some models and likely results in models with fewer significant coefficients when compared to the alternative approach of simply combining all individual data to develop seasonal, global models. it is important to understand (but not over-emphasize) the influence of individual variation. in our study, other than elk selecting against eastness (i.e., for western exposures) during calving and for deciduous burns in fall, all animals did not show the same selection for any attribute on the landscape. our approach incorporates into each seasonal pooled model any zero selection value by an animal for an attribute with a weighting equal to significant selection coefficients by other animals. as such, the final pooled models do not overemphasize the importance of attributes selected by one or a few individuals. we also recommend presenting selection analyses with observations of use to identify habitat classes (or other categorical variables) that are avoided or rarely used, and to assess the relative magnitude of use for highly selected classes. findings in our study may be limited by bihabitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 56 ases attributed to relatively low fix rates (moen et al. 1996, frair et al. 2004), particularly for moose. most locations within each season (55-65%), however, were obtained from habitat classes with forest cover (pine-spruce, stunted spruce, subalpine) that probably had the poorest signal attenuation. therefore, although our interpretation may underestimate the magnitude of use and selection of those classes, the classes were nonetheless noted as important in our analyses despite potential fix bias. habitat use and selection moose were often at lower elevations than elk on the besa-prophet landscape. both species, however, moved down in elevation between winter and late winter. in other areas of british columbia with high topographical diversity, the greatest single determinant of late winter habitat use by moose was decreasing elevation, which may be a surrogate for snow depth (poole and stuart-smith 2005, 2006). snow depth is a primary factor affecting late winter distribution of moose populations (peek 1998), and moose in interior mountainous areas typically move to lower elevations throughout the winter (e.g., pierce and peek 1984, van dyke et al. 1995). elk also move to lower elevations during the winter (e.g., unsworth et al. 1998, boyce et al. 2003), possibly to take advantage of increased food availability. because of the abundance of south-facing, wind-swept slopes in the besaprophet, moose and elk may not be as affected by snow depth as they are in other areas. moose frequently inhabit shrub communities throughout their range whenever snow depths do not exceed 100 cm (peek 1974), and select coniferous forests in regions of deeper snow, provided that browse is available (bunnell and eastman 1976, peek et al. 1982, pierce and peek 1984). in the mountainous interior regions of alaska, shrub-dominated communities above timberline are important moose habitat (peek 1998) and those dominated by willow species appear to be the most important to moose (risenhoover 1989). moose in our study used subalpine habitats most in fall and winter, and the pine-spruce habitat class most during late winter, calving, and summer. elk often occupy south-facing, seral brushfields (irwin and peek 1983) or windswept, grass-dominated slopes (knight 1970; houston 1982) during winter, except when deep or crusted snow causes them to seek timber (houston 1982). peck and peek (1991) reported that elk in northeastern british columbia wintered primarily in post-fire grass and shrub communities, except during severe weather when conifer stands were used. timbered habitats also have been reported to be important in other seasons. for example, elk in idaho shifted from using a high proportion of shrub and open timber habitats in spring to using more timbered habitats in fall (unsworth et al. 1998). the availability of forest cover may help reduce thermal stress and predation risk (anderson et al. 2005). habitats most used by elk in the besa-prophet were deciduous and elymus burns during late winter, and subalpine during all other seasons. according to pooled selection models, when elk selected areas with relatively high wolf risk, they did not select significantly for forest cover to help minimize risk. in yellowstone national park, when wolf activity was centered around dens and rendezvous sites in summer, elk apparently avoided wolves by selecting higher elevations, less open habitat, burned forest, and, in areas of high wolf density, steeper slopes (mao et al. 2005). elk did not spatially separate themselves from wolves in winter, but relied on behavioral anti-predatory strategies such as grouping (mao et al. 2005). although we have no data on group sizes, anecdotally elk in the besa-prophet tended to group together more frequently than moose during summer through winter. elk typically select habitats characterized by early seral stage (thomas et al. 1979, alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 57 irwin and peek 1983, grover and thompson 1986), which may be facilitated by burns. in the besa-prophet, the relative selection by elk for deciduous and elymus burns varied by season. elk showed greater selection for the more open elymus burns during calving, and for deciduous burns during fall, winter, and late winter. moose response to burns also is generally positive throughout their range, although gasaway et al. (1989) observed that traditional movement patterns by moose apparently prevented animals without pre-fire use from finding burns. in the besa-prophet, moose did not specifically select for burned areas, but frequented the older deciduous burns as a small, but consistent part of their habitat use. potentially related to predation risk and forage quality, the calving strategies of moose and elk appeared to differ in our study. moose used the lowest elevations of the year (1333 ± 81 m) during calving and had the smallest seasonal use areas. these locations were not typically on valley bottoms (~800-1300 m) per se, but rather on the coniferous side slopes. use and selection of pine-spruce by moose was greatest during the subsequent summer season. poole et al. (2007) reported that moose in southern british columbia showed 2 elevational strategies during calving related to predation risk. they described ‘climber’ moose, which moved to higher elevations to calve in areas with lower forage quality and quantity and, therefore, farther from grizzly bears found at lower elevations. in contrast, ‘non-climber’ moose calved at low elevations with much higher forage values, but potentially at increased risk of predation. given that grizzly bears in the besa-prophet tend to remain in higher areas during spring (parker and milakovic 2007), moose at lower elevations would avoid bears and have access to early green-up of shrub vegetation. this calving strategy, however, would come with the potential risk of wolf predation, given than wolves select for shrub vegetation in spring (parker and milakovic 2007). although the pooled selection models did not indicate that moose selected locations to specifically avoid predation risk during calving or summer, they did, however, avoid risk of wolf predation during the fall. in contrast to moose, elk moved upslope from their late winter locations to calve ~220 m higher than moose. during calving and summer, the higher elevations corresponded with the high use of the subalpine habitat class. their selection for vegetation quality was probably facilitated by access to elymus and deciduous burns, which typically green up earlier in spring, and then by the spruce-shrub transition zone of the subalpine in summer. unlike moose, elk appeared to select calving and summer areas on the landscape that minimized some predation risk by bears given that they used similar elevations during these seasons (parker and milakovic 2007). management implications and recommendations combining both use and selection information from this study, it appears that the highest potential for overlap between moose and elk may be during fall and winter, when both species used the subalpine habitat class more than other classes and selected strongly for it. in winter, there may be some elevational separation between the 2 species, given that elk locations were on average 100 m higher than those of moose. in fall, however, the elevations used by moose and elk were not different. both species also selected for low shrub habitat during fall, where elk undoubtedly consume higher amounts of forbs and grass that are not yet senescent (stevens 1970) compared to the more browse-dominated diets of moose. elk are generalist feeders that maximize their food intake through mechanisms of habitat selection rather than selection of specific foods (irwin and peek 1983). consequently, they can successfully shift from herbaceous species to browse (houston 1982), and may habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 58 be efficient competitors with moose when resources are limited (flook 1964). during periods when resources are limited, overlap in resource use between the 2 species could result in temporary interspecific competition (jenkins and wright 1988). if prescribed burns that are conducted primarily to enhance stone’s sheep populations are enabling increases in non-target elk populations in the besa-prophet, there also may be potential for competition between elk and stone’s sheep during some times of the year (walker et al. 2007). further, it is likely that with an expanding elk population, predator numbers will increase in response to the increased prey base. higher wolf numbers would be expected to affect predator-prey dynamics by expanding into adjacent areas, particularly via burned slopes to upper elevations, and by increasingly preying on stone’s sheep and caribou. prescribed burning also may provide additional opportunities for grizzly bears that select for burned vegetation classes (i.e., elymus burn and deciduous burn) throughout the non-denning period (parker and milakovic 2007), thereby augmenting predation risks to moose and elk. the management action of prescribed fire may help to sustain some of the diversity and abundance of large mammals for which the besa-prophet area is known, but it could also shift the prey base for predators and change the dynamics of the system. additional studies involving population estimates and animal distributions should be specifically designed to determine how intensity, frequency, and locations of prescribed burns affect habitat use by ungulates (principally elk, moose, and stone’s sheep) and subsequently predators (e.g., wolves and grizzly bears) and their movements relative to ungulate prey. our findings suggest that there is currently some seasonal overlap between elk and moose in the besa-prophet, and that expanding elk numbers will affect other species in the system. acknowledgements we thank the muskwa-kechika trust fund, the british columbia ministry of environment and the university of northern british columbia for their support. g. williams helped familiarize us with the besa-prophet landscape. r. b. woods, with the bc ministry of environment, captured and collared all of the animals monitored in this study. we also acknowledge j. b. ayotte, s. g. emmons, d. d. gustine, r. j. lay, b. milakovic, and a. b. d. walker who assisted in aspects of data visualization and analyses. references anderson, d. p., j. d. forester, m. g. turner, j. l. frair, e. h. merrill, d. fortin, j. s. mao, and m. s. boyce. 2005. factors influencing female home range sizes in elk (cervus elaphus) in north american landscapes. landscape ecology 20:257271. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monograph 114. bergerud, a. t., and j. p. elliott. 1998. wolf predation in a multiple-ungulate system in northern british columbia. canadian journal of zoology. 76:1551-1569. _____, w. wyett, and j. b. snider. 1983. the role of a wolf population in limiting a moose population. journal of wildlife management 47:977-988. boer, a. h. 1998. interspecific relationships. pages 337-349 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d.c., u.s.a. bowyer, r. t. 2004. sexual segregation in ruminants: definitions, hypotheses, and implications for conservation and management. journal of mammalogy 85:1039-1052. _____, and j. g. kie. 2004. effects of foraging alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 59 activity on sexual segregation in mule deer. journal of mammalogy 85:498-504. boyce, m. s. 1991. migratory behavior and management of elk (cervus elaphus). applied animal behaviour science 29:239-250. _____, p. r. vernier, s. e. nielsen, and f. k. schmiegelow. 2002. evaluating resource selection functions. ecological modelling 157:281-300. _____, m. s., j. s. mao, e. h. merrill, d. fortin, m. g. turner, j. fryxell, and p. turchin. 2003. scale and heterogeneity in habitat selection by elk in yellowstone national park. ecoscience 10:421-423. british columbia ministry of crown lands. 1990. terrain and resource inventory management in british columbia specifications and guidelines for geomatics: digital baseline mapping at 1:20,000. ministry of crown lands for the government of british columbia, victoria, british columbia, canada. burnham, k. p., and d. r. anderson. 2002. model selection and multi-model inference: a practical information–theoretic approach. second edition. springer-verlag, new york, new york, u.s.a. bunnell, f. l., and d. s. eastman. 1976. effects of forest management practices on wildlife in the forests of british columbia. pages 631-676 in proceedings xvi iufro world congress, oslo, norway. cederlund, g., and h. sand. 1994. homerange size in relation to age and sex in moose. journal of mammalogy 75:10051012. cook, j. g. 2002. nutrition and food. pages 259-349 in d. e. toweill, and j. w. thomas, editors. north american elk: ecology and management. smithsonian institution press, washington, d.c., u.s.a. craighead, j. j., g. atwell, and b. w. o’gara. 1972. elk migrations in and near yellowstone national park. wildlife monograph 29. demarchi, d. a. 1996. introduction to the ecoregions of british columbia. british columbia wildlife branch, ministry of environment, lands and parks, victoria, british columbia, canada. edge, w. d., c. l. marcum, s. l. olson, and j. f. lehmkuhl. 1986. nonmigratory cow elk herd ranges as management units. journal of wildlife management 50:660-663. esri 2002. arcview gis, version 3.3. environmental systems research institute, redlands, california, u.s.a. flook, d. r. 1964. range relationships of some ungulates native to banff and jasper national parks, alberta. pages 119-128 in d. j. crip, editor. grazing in terrestrial and marine environments. blackwell scientific publications, oxford, u.k. frair, j. l., s. e. nielsen, e. h. merrill, s. r. lele, m s. boyce, r. h. m. munro, g. b. stenhouse, and h. l. beyer. 2004. removing gps collar bias in habitat selection studies. journal of applied ecology 41:201-212. franzman, a. w., and c. c. schwartz. 1998. ecology and management of the north american moose. smithsonian institute press, washington, d.c., u.s.a. gasaway, w. c., s. d. dubois, r. d. boertje, d. j. reed, and d. t. simpson. 1989. response of radio-collared moose to a large burn in central alaska. canadian journal of zoology 67:325-329. gillies, c. s., m. hebblewhite, s. e. nielsen, m. a. krawchuk, c. l. aldridge, j. l. frair, d. j. saher, c. e. stevens, and c. l. jerde. 2006. application of random effects to the study of resource selection by animals. journal of animal ecology 75:887-898. gillingham, m. p., and k. l. parker. 2008. the importance of individual variation in defining habitat selection by moose in northern british columbia. alces 44: 7-20. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 60 griffith, b., d. c. douglas, n. e. walsh, d. d. young, t. r. mccabe, d. e. russell, r. g. white, r. d. cameron, and k. r. whitten. 2002. the porcupine caribou herd. pages 8-37 in arctic refuge coastal plain terrestrial wildlife research summaries. u.s. geological survey, biological resources division, biological science report 2002-0001, anchorage, alaska, u.s.a. groten, s. m. e., and r. ocatre. 2002. monitoring the length of the growing season with noaa. international journal of remote sensing 23:2797-2815. grover, k. e., and m. j. thompson. 1986. factors influencing spring feeding site selection by elk in the elkhorn mountains, montana. journal of wildlife management 50:466-470. gustine, d. d., k. l. parker, r. j. lay, m. p. gillingham, and d. c. heard. 2006a. calf survival of woodland caribou in a multi-predator ecosystem. wildlife monograph 165. _____, _____, _____, _____, and _____. 2006b. interpreting resource selection between scales among individual woodland caribou in winter. journal of wildlife management 70:1601-1614. hayes, r. d., and a. s. harestad. 2000. wolf functional response and regulation of moose in the yukon. canadian journal of zoology 78:60-66. hendrickx, j. 1999. using categorical variables in stata. stata technical bulletin 52:2-8. hooge, p. n., and b. eichenlaub. 2000. animal movement extension to arcview. version 2.0. alaska science center, biological sciences office, u.s. geological survey, anchorage, alaska, usa. houston, d. b. 1982. the northern yellowstone elk: ecology and management. macmillan publishing, new york, new york, u.s.a. hundertmark, k. j. 1998. home range, dispersal and migration. pages 303-335 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d.c., u.s.a. irwin, l. l., and j. m. peek. 1983. elk habitat use relative to forest succession in idaho. journal of wildlife management 47:664-672. jenkins, k. j., and r. g. wright. 1988. resource partitioning and competition among cervids in the northern rocky mountains. journal of applied ecology 25:11-24. jennrich, r. i., and f. b. turner. 1969. measurement of non-circular home range. journal of theoretical biology 22:227-37. johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61:65-71. kelsall, j. p. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research (supplement) 1:1-10. knight, r. r. 1970. the sun river elk herd. wildlife monograph 23. lawson, e. j. gallerani, and a. r. rodgers. 1997. differences in home-range size computed in commonly used software programs. wildlife society bulletin 25:721-729. lay, r. j. 2005. use of landsat tm and etm+ to describe intra-season change in vegetation with consideration for wildlife management. m.sc. thesis, university of northern british columbia, prince george, british columbia, canada. luckhurst, a. j. 1973. stone sheep and their habitat in the northern rocky mountains foothills of british columbia. m.sc. thesis, university of british columbia, vancouver, british columbia, canada. lynch, g. m., and l. e. morgantini. 1984. sex and age differential in seasonal home alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 61 range of moose in northwestern alberta. alces 20:61-78. manly, b. f., l. l. mcdonald, and d. l. thomas. 2002. resource selection by animals: statistical design and analysis for field studies. second edition. chapman-hall, london, u.k. mao, j. s., m. s. boyce, d. w. smith, f. j. singer, d. j. vales, j. m. vore, and e. h. merrill. 2005. habitat selection by elk before and after wolf reintroduction in yellowstone national park, wyoming. journal of wildlife management 69:16911707. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. british columbia ministry of forests, victoria, british columbia, canada. menard, s. 2002. applied logistic regression analysis. second edition. sage, thousand oaks, california, u.s.a. miller, w. 2002. elk interactions with other ungulates. pages 259-349 in d. e. toweill, and j. w. thomas, editors. north american elk: ecology and management. smithsonian institution press, washington, d.c., u.s.a. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monograph 122. moen, r., j. pastor, y. cohen, and c. c. schwartz. 1996. effects of moose movement and habitat use on gps collar performance. journal of wildlife management 60:659-668. oindo, b. 2002. predicting mammal species richness and abundance using multi-temporal ndvi. photogrammetric engineering and remote sensing 68:623-629. palmer, m. w. 1993. putting things in even better order: the advances of canonical correspondence analysis. ecology 74:22152230. parker, k. l., and b. milakovic. 2007. defining the predator landscape in the besa-prophet. part 3 of an ecosystem approach to habitat capability modelling and cumulative effects management. final report submitted to the muskwakechika advisory board, fort st. john, british columbia, canada. parminter, j. 1983. fire-ecological relationships for the biogeoclimatic zones and subzones of the fort nelson timber supply area. northern fire ecology project, british columbia ministry of forests, victoria, british columbia, canada. pci geomatics corporation. 2001. pci works version 7.0. richmond hill, ontario, canada. peck, v. r., and j. m. peek. 1991. elk, cervus elaphus, habitat use related to prescribed fire, tuchodi river, british columbia. canadian field-naturalist 105:354-362. peek, j. m. 1974. on the winter habitats of shiras moose. canadian naturalist 101:131-141. _____. 1998. habitat relationships. pages 351-375 in a.w. franzman, and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., u.s.a. _____, and a. l. lovaas. 1968. differential distribution of elk by sex and age on the gallatin winter range, montana. journal of wildlife management 32:553-557. _____, m. d. scott, l. j. nelson, d. j. pierce, and l. l. irwin. 1982. role of cover in habitat management for big game in northwestern united states. transactions of the north american wildlife and natural resources conference 47:363-373. pierce, d. j., and j. m. peek. 1984. moose habitat use and selection patterns in north-central idaho. journal of wildlife management 48:1335-1343. poole, k. g., and k. stuart-smith. 2005. fine-scale winter habitat selection by moose in interior montane forests. alces 41:1-8. habitat selection by moose and elk – gillingham and parker alces vol. 44, 2008 62 _____, and _____. 2006. winter habitat selection by female moose in western interior montane forests. canadian journal of zoology 84:1823-1832. _____, r. serrouya, and k. stuart-smith. 2007. moose calving strategies in interior montane ecosystems. journal of mammalogy 88:139-150. post e., n. c. stenseth, r. o. peterson, j. a. vucetich, and a. m. ellis. 2002. phase dependence and population cycles in a large-mammal predator-prey system. ecology 83:2997-3002. risenhoover, k. l. 1989. composition and quality of moose winter diets in interior alaska. journal of wildlife management 53:568-577. roberts, d. w. 1986. ordination on the basis of fuzzy set theory. vegetatio 66:123-131. ruimy, a., g. dedieu, and b. saugier. 1994. methodology for the estimation of terrestrial net primary production from remotely sensed data. journal of geophysical research 99:5263-5283. seaman, d. e., j. j. millspaugh, b. j. kernohan, g. c. brundige, k. j. raedeke, and r. a. gitzen. 1999. effects of sample size on kernel range size estimates. journal of wildlife management 63:739-747. silver, r. s. 1976. ecological features of moose (alces alces andersoni) winter habitat in the boreal white and black spruce zone of northeastern british columbia. m.sc. thesis, university of british columbia, vancouver, british columbia, canada. smith, d.w., r.o. peterson, and d.b. houston. 2003. yellowstone after wolves. bioscience 53:330-340. spalding, d. j. 1992. the history of elk (cervus elaphus) in british columbia. publication number 18, pp. 1-27. royal british columbia museum, victoria, british columbia, canada. statacorp. 2007. stata version 9.1. college station, texas, u.s.a. stenhouse, g. b., p. b. latour, l. kutny, n. maclean, and g. glover. 1994. productivity, survival, and movements of female moose in a low-density population, northwest territories, canada. arctic 48:57-62. stevens, d. r. 1970. range relationships of elk and livestock, crow creek drainage, montana. journal of wildlife management 30:349-363. stewart, k. m., r. t. bowyer, j. g. kie, n. j. cimon, and b. k. johnson. 2002. temporospatial distributions of elk, mule deer, and cattle: resource partitioning and competitive displacement. journal of mammalogy 83:229-244. strohmeyer, d. c., and j. m. peek. 1996. wapiti home range and movement patterns in a sagebrush desert. northwest science 70:79-87. thomas, d. l., and e. j. taylor. 2006. study designs and tests for comparing resource use and availability ii. journal of wildlife management 70:324-336. thomas, j. w., h. j. black, r. j. scherzinger, and r. j. pedersen. 1979. deer and elk. pages 104–127 in j. w. thomas, technical editor. wildlife habitats in managed forests: the blue mountains of oregon and washington. u.s. department of agriculture agricultural handbook 533, washington, d.c., u.s.a. toweill, d. e., and j. w. thomas. 2002. north american elk: ecology and management. smithsonian institution press, washington d.c., u.s.a. tucker, c. j., and p. j. sellers. 1986. satellite remote sensing of primary production. international journal of remote sensing 7:1395-1416. unsworth, j. w., l. kuck, e. o. garton, and b. b. butterfield. 1998. elk habitat selection on the clearwater national forest, idaho. journal of wildlife management 62:1255-1263. alces vol. 44, 2008 gillingham and parker – habitat selection by moose and elk 63 van dyke, f., b. l. probert, and g. m. van veek. 1995. seasonal habitat use characteristics of moose in south-central montana. alces 31:15-26. vucetich, j. a., r. o. peterson, and c. l. schaefer. 2002. the effect of prey and predator densities on wolf predation. ecology 83:3003-3013. walker, a. b. d, k. l. parker, m. p. gillingham, d. d. gustine, and r. j. lay. 2007. habitat selection and movements of stone’s sheep in relation to vegetation, topography and risk of predation. ecoscience 14:55-70. weckerly, f. w. 1998. sexual segregation and competition in roosevelt elk. northwestern naturalist 79:113-118. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70:164168. p85-120_4101.pdf alces vol. 41, 2005 timmermann and rodgers moose values 85 moose: competing and complementary values h. r. timmermann1 and a. r. rodgers2 1r r #2 nolalu, on, canada p0t 2k0; 2centre for northern forest ecosystem research, ontario ministry of natural resources, 955 oliver road, thunder bay, on, canada p7b 5e1 abstract: moose (alces spp.) are the largest and one of the most widespread land mammals in boreal and mixed-wood forests of the northern hemisphere. they provide essential food for carnivores and human subsistence users in remote areas. during the 20th century, their increasing densities and distribution in north america and fennoscandia have provided added recreation through licensed hunting, viewing, a variety of cultural, spiritual, and commercial activities, and numerous scientific/ecological studies. in some areas, their high densities are considered unacceptable due to damage caused by moose/vehicle collisions, to commercially valuable trees, to agricultural crops, and added management costs. most management agencies attempt to manage populations through an allocation process which includes provisions for native harvest, licensed hunter harvest, and control of illegal harvests. moose benefits or worth, both competing and complementary, are discussed, based on a wide array of monetary and non-monetary values reported in the literature. alces vol. 41: 85-120 (2005) key words: aesthetics, allocation, collision damage, commercial value, cultural/spiritual value, forest damage, hide and antler value, meat value, moose, recreational value, scientific/ecological value moose (alces alces) are the largest and one of the most widely distributed land mammal species across canada, portions of the usa, sweden, norway, and finland (peterson 1955). an important part of natural communities, moose are an essential food source for wolves (canis lupus) and bears (ursus americanus, u. arctos), as well as a host of scavengers including red fox (vulpes vulpes) and ravens (corvus covax). in addition to their intrinsic ecological role, moose have been exploited by man over the centuries, providing food, clothing, and tools for native people and early settlers in northern latitudes (reeves and mccabe 1998, rodgers 2001). moose populations generally expanded and increased during the latter half of the 20th century (karns 1998, timmermann and buss 1998, lavsund et al. 2003). considered a renewable resource, they continue to provide many benefits and socio-economic advantages if managed in a sustainable manner (eagen et al. 1989). we believe the current “value” of moose has never been as diverse or extensive. however, “how do we value moose and which values compete or complement one another?” in common terms, a “value” is thought of as being usually more or less desirable, useful, important, worthwhile, worthy of esteem for its own sake, a monetary worth, or a thing or quality having intrinsic worth (webster 1967). however, values are perceived in many different ways and are acquired by a host of interactions and experience over time (kellert 1980, 1987). consequently, values may change over time and circumstances, and differ among individuals (kellert 1980, 1987). all values are not necessarily shared by everyone, nor should they be, as some values conflict to moose values – timmermann and rodgers alces vol. 41, 2005 86 some extent or may be harmonious to some extent, depending on the individual point of view. not surprisingly then, sociologists, economists, and wildlife managers have developed fundamentally different notions of value. whereas social values might be categorized as cultural, societal, psychological, and physiological, upon which relative importance or worth may be assigned (brown and manfredo 1987), economists may prefer to divide wildlife assets into, for example, use and nonuse values (bishop 1987). wildlife managers, on the other hand, tend to employ a mixture of socio-economic values. steinhoff (1978), for example, classified big game values as either recreational (related to sports or hobbies), aesthetic (related to beauty, inspiration, or art), educational (added knowledge), biological (ecosystem component), social (non-monetary related to quality of life), or commercial (meat and trophy). differences among disciplines have made it very difficult to integrate their various perspectives into a generally accepted value system. for instance, wildlife managers may value wildlife using recreation expenditures based on the monetary amount hunters or others may spend in pursuit of their activities, while economists may find that approach objectionable because it may only help in understanding the local economic impacts of wildlife-related activities and does not provide a good measure of the economic value of wildlife to society as a whole (bishop 1987). nor can economists and wildlife managers agree entirely on their respective definitions of competing and complementary values. whereas economists may determine the extent to which values compete or complement one another on the basis of production, such that one value may be excluded at the expense of another, wildlife managers tend to accept the co-existence of multiple values with varying levels of importance determined by competing and complementary interactions through time. in the absence of general agreement on terms and definitions, we have chosen, for the purposes of this discussion, to categorize moose values into two components: those we consider competing and those we consider complementary based on a literature search and personal communication with wildlife managers. although broadly consistent with the definitions of economists, we have attempted to allow for the wider mixture of socio-economic perspectives commonly held by wildlife managers. thus, complementary values are simply those we consider to agree with each other or those that come together, while competing values we believe to be incompatible, or which disagree with each other. these categories provide a framework within which to assess the positive and negative values of moose. attempting to maximize meat production (a positive value), for example, will undoubtedly impact on a manager’s ability to maximize trophy production (also a positive value), or to minimize forest or vehicle collision damage (negative values) at the same time. the first is an example of complementary values: both meat and trophy production are positive values that coexist but vary in their predominance through time. in this case, the production of one value increases (complements) the production of the other value. the latter is a case of competing values: maximizing the positive value of meat production by increasing the size of a moose population will also increase the negative aspects of forest or moose/vehicle collision (mvc) damage. in this way, the production of one value detracts from (competes with) the ability of wildlife managers to promote another value. evaluation methods moose have substantial recreational alces vol. 41, 2005 timmermann and rodgers moose values 87 and economic values. each moose has some intrinsic economic value associated with both consumptive and non-consumptive use according to schwartz and bartley (1991). economists use indirect valuation approaches including travel cost, contingent valuation, and hedonic methods, or variations thereof (steinhoff et al. 1987, sarker and surry 1998). others, such as legg and kennedy (2000), have developed socio-economic impact models which factor in gross output, value added employment, labor income, and taxes. however, the most common approach taken by wildlife managers is to measure direct and indirect monetary costs and benefits associated with moose. direct costs or expenditures related to hunting commonly include costs of equipment, transportation, accommodation, food and beverage, and license fees (ruhr and crichton 1985). evaluation of wildlife viewing activities is assessed by a survey sample to determine the number of days spent viewing, trips taken to participate, and associated expenditures (duwors et al. 1999). a sampling of methods used to evaluate the benefits of moose include those from; british columbia (reid 1997), saskatchewan (ross 1975, ross and paul 1976), manitoba (capel and pandy 1973, ruhr and crichton 1985), ontario (bisset 1987, legg 1995, umali 1997, sarker and surry 1998, legg and kennedy 2000), newfoundland (condon and adamowicz 1995), maine (boyle and clark 1993), idaho (loomis et al. 1985), sweden (kriström 1987, johansson et al. 1988, mattsson 1990, sylvén 1995), and norway (storaas et al. 2001). as it is not the intent of this paper to undertake a thorough socio-economic analysis of the value of moose, we have taken the simplest approach of providing descriptive indices of monetary costs and benefits associated with the positive and negative values of moose that are readily understood by most wildlife managers and may contribute to future indepth analyses. this review is restricted to studies of moose values conducted and reported in north america and fennoscandia. cost estimates are given in u.s. dollars using a conversion rate of $1.00 us to $1.30 canadian, unless otherwise stated. moose densities an understanding of moose densities and distribution is fundamental to any discussion concerning values. north american moose populations in 2000 were estimated at about 1 million distributed in 28 jurisdictions (timmermann 2003). populations occur in 11 canadian provinces or territories and in at least 17 u.s. states with continuing range expansion in new england and several western u.s. states. in the early 1990s, north american moose densities in 2 where bear and wolves are unexploited (messier 1994) to as high as 6.0/km2 for specific areas on isle royale and >12.0 moose/km2 for management area 17 in newfoundland (brandner et al. 1990, thompson and curran 1993, mercer 1995, mercer and mclaren 2002). crête (1987) believed the north american carrying capacity of moose in a predator-free environment to be 2.0 moose/km2. target densities in forested habitats of newfoundland are currently 2 (mclaren et al. 2004). several herd reduction programs have occurred in elk island national park, alberta since 1960 in an effort to reduce winter densities to 2.0/km2 and reduce over-winter mortality (lynch et al. 1995). n e w f o u n d l a n d a n d f e n n o s c a n d i a have consistently had the highest reported moose densities in the world (hörnberg 1995, harkonen 1999, storaas et al. 2001, mclaren et al. 2004) due to little natural predation, modern forestry producing large areas of good habitat, and a closely managed human harvest system. densities have also increased close to many urban / suburban moose values – timmermann and rodgers alces vol. 41, 2005 88 areas across north america, contributing to added conflicts between moose and humans (karns 1998). high densities in fennoscandia, primarily on winter ranges have led to many competing values, including high annual hunter harvests, significant forest damage, and loss of life, injury, and property damage from moose/vehicle collisions (lavsund and sandegren 1991, heikkilä and aarnio 2001, storaas et al. 2001). in 2000, about 200,000 moose were harvested by hunters in all 3 nordic countries, from a post hunt winter population estimated at 450,000 (lavsund et al. 2003). swedish moose densities were estimated at 0.7 2.2/ km2 in local areas during the early 1980s and a record harvest of 175,000 occurred in 1982 in an effort to reduce populations (lavsund and sandegren 1989, cederlund 1996). hunter harvest was again increased in the late 1990s to reduce densities and subsequent forest damage (carlestål 2000). annual harvests became stabilized at about 90,000 per year in the early 1990s by targeting a harvest of 0.3-0.6 moose/km2. populations since 1995 have again increased and in 1998/99, 101,930 moose were harvested (faber 1999). norwegian moose densities and harvests increased during the 1970s and 1980s. in the late 1990s densities averaged 1.0-2.0/km2 and were as high as 5.0-6.0/km2 during winter in some valleys (solberg et al. 1998, 2003). harvests fluctuated between 34,000 in 1995 and 36,000 in 1997 and peaked at 39,309 in 1999 (storaas et al. 2001, lavsund et al. 2003). such high densities prompted preparation of an action plan to help reduce damage to timber plantations and mvcs to acceptable levels (jaren et al. 1995). in finland, moose harvests peaked at 68,843 in 1984 and averaged 27,750 per year (range 22,836-32,484) between 1995 and 1999 (harkonen 1999, lavsund et al. 2003). finnish managers increased the number of moose licenses from 17,060 in 1997 to 38,134 in 1999 to target a harvest of 50,000 and reduce damage to commercially valuable timber. in some regions of norway moose are considered a benefit for both landowners and hunters because of consistent high densities over time (storaas et al. 2001). large forest companies owning land in sweden and finland regard moose as a recreation benefit rather than an economical resource for the local community. timber companies typically want to reduce moose densities and their resultant damage while hunters desire densities near the maximum sustainable yield (angelstam et al. 2000). complementary values recreational hunting reasons or motivations given for recreational hunting of moose include nature appreciation, companionship, meat and trophy value, stress release, and to practice outdoor skills (rollins 1987, rollins and romano 1989, bottan 1999). providing recreational licensed hunting opportunities has been the primary focus of most management agencies. we consider such regulated hunting as complementary to other values including cultural/spiritual, aesthetic, commercial, and scientific/ecological. others may view these as competing values. in 2000-01, 23 north american jurisdictions managed a licensed moose hunt and an estimated 385,569 hunters harvested 82,466 moose (timmermann 2003). moose hunting provides a significant annual economic impact in some jurisdictions. bisset (1987) provided a detailed review of value determination, economic concepts, techniques, and problems of evaluation. his review represents the only comprehensive attempt to estimate the total gross value of north american moose hunting related expenditures ($375.8 million in 1982) based on limited data. total expenditures per moose averaged $1,688 for non-residents and $1,285 for resident moose hunters in 1982. alces vol. 41, 2005 timmermann and rodgers moose values 89 some jurisdictions have since attempted to quantify the value of moose hunting. reid (1997) estimated a contribution of $12.2 million for resident hunters in british columbia in 1995. legg and kennedy (2000) reported an impact of $59.2 million to the gross ontario income and 1,645.8 years of employment in 1996. during that year, ontario resident hunters spent an estimated total of $154.46 million. regelin and franzmann (1998) estimated the annual economic impact of 33,000 resident and 1,000 non-resident alaskan hunters to represent $32.6 million in the late 1990s. in ontario, sarker and surry (1998) indicated a decline in recreational moose hunting demand with higher travel costs and lower income. in fennoscandia, nearly 400,000 moose hunters harvested about 200,000 moose in 2000 (lavsund et al. 2003). storaas et al. (2001) placed the 2000 value of moose hunting at $172-257 million in norway. licenses in north america, all hunters are required to pay a license fee for the opportunity to hunt and harvest a moose. resident licenses in canada in 2001 averaged $27.54 (range $7.70-$44.28) and non-residents $157.40 (range $15.40 $354.20, n = 11, timmermann 2003). resident fees in the u.s. averaged $106.00 (range $20.00 $310.00), while non-resident licenses averaged $727.70 (range $80-$1,643, n = 12). export permits or trophy fees were required in addition to license fees to transport an animal from alaska, northwest territories, alberta, and ontario. legg and kennedy (2000) reported license revenue from ontario moose hunting was approximately $3.12 million. umali (1997) reported that current ontario moose license fees do not reflect the benefit hunters receive from moose hunting and that the true value is measured by hunter’s willingness to pay over and above the price of a license. to illustrate this point, some hunters have paid as high as $19,000 (range $7,000$19,000) for a single moose license at a special u.s. governors auction (table 1). moose license revenues in finland have ranged from $1-3 million (heikkilä and aarnio 2001). hunting fees in norway were estimated at $30 $52 per animal in 2000 according to storaas et al. (2001). in addition to basic license fees, additional revenue is generated by a price system based on carcass weight that was introduced in 1960 to encourage hunters to select young and small moose (jon lykke, personal communication 2004). meat moose are highly regarded for their large size, low fat (>20 times less than lean beef), and nutritious meat quality (hansson and malmfors 1978, crichton 1998a, crichton and redmond 1998). moose meat contains relatively high levels of eicosapentaenoic acid (epa), a protective fatty acid not found in domestic animals. rowland (1989) reported that epa can protect against heart attacks, hardening of the arteries, and certain types of arthritis. midkiff (2004) has suggested moose meat is a natural product and therefore a healthy alternative to most feedlot raised beef which is driven by volume, efficiency, uniformity, and profit. many newfoundlanders depend on moose as a source of food (condon and adamowicz 1995). the monetary value of moose meat in north america is often determined by comparison to an equivalent retail cost of a kilogram or pound of domestic beef. average yields of processed moose meat have been reported as varying between 160 kg in ontario (hamilton 1981) and 180 kg in alberta (renecker et al. 1987), whereas hansson and malmfors (1978) assumed an average carcass weight of 130 kg for alces alces, the smaller european moose. bisset (1987) used an average of moose values – timmermann and rodgers alces vol. 41, 2005 90 170 kg per moose and a retail beef price of $6.28/kg, or an average of $1,067.60 per animal. he applied this value to estimate the total north american value of moose meat taken by licensed hunters at $76 million based on a 1982 harvest of 71,000. oosenbrug et al. (1991) reported the meat value of each newfoundland moose at $1,016 in 1990. more recently crichton (1998a) estimated a conservative value of $1,084/moose using a yield of 160 kg/moose by applying a cost of $6.78/kg ($3.08/lb) used by hamilton (1981). if one assumes a current value of $8.80/kg ($4.00/lb), the total value of moose meat taken by licensed hunters in north america could exceed $116 million using a 2001 harvest of 82,466 as reported by timmermann (2003). in fennoscandia, as opposed to north america, moose (alces alces) can be sold on the free market. storaas et al. (2001) reported moose meat in norway has a potential value well above $42 million based on an annual harvest of up to 40,000 in 2000. likewise the swedish moose harvest (105,000) contributed 13.65 million kg of meat or close to 10% of cattle production in 2001 (sylvén 2003). the total value of moose meat taken in finland in 1997 was about $13 million (heikkilä and aarnio 2001). only about 10% of finnish hunters gave meat as the primary reason for hunting, compared to a much higher value placed on meat by swedish and norwegian moose hunters (mattsson and kriström 1987, sødal 1989). antlers and hides trophies are considered objects preserved or mounted as a memorial (webster 1967). trophy moose antlers provide evidence of population status (bubenik 1989) and are highly prized and often prominently displayed on mounted heads or with a portion of the skull intact. their value lies in a lasting symbol of a successful hunting experience since no tender or monetary value is legally permitted in north america. trophy antlers are rated by size and conformation and are usually from prime bulls 6.5-10.5 years of age (timmermann 1971, gasaway 1975). trophies are cumulatively scored by specific measurements, the number of points and conformity between right and left palms (boone and crockett club 1988). in north america 4 types of recognition are listed. the boone and crockett club initiated the first standards in the early 1930s for all legally taken moose (crichton 1998a). they designated 3 categories: the canada moose (a. a. americana and a. a. andersoni), the alaskan/yukon moose (a. a. gigas) and the shiras or wyoming moose (a. a. shirasi). trophy antlers taken by archers can be certified by the pope and young club, those taken by muzzle loaders by the longhunter association, and shed antlers by the north american shed hunters club (crichton 1998a, b). management strategies designed to maintain trophy class bulls in the population have been reported for alaska (smith et al. 1979, schwartz et al. 1992) and for sweden by sylvén (1995). antlers are also valued by artists for carving and items such as buttons, picture frames, furniture, and lamps (crichton 1998a). prices paid for shed antlers in alaska vary from $1-$6 per pound according to their condition (tom cooper, personal communication 2004). once processed, items can be legally sold in most jurisdictions and often fetch a high value. moose hides can provide a valuable source of leather. reeves and mccabe (1998:27-32) provide an extensive review of the traditional value of moose hides used by natives for a wide variety of clothing and footwear. some agencies have offered incentives to hunters to submit raw moose hides which are tanned and given or sold to native peoples for further processing to traditional clothing (crichton 1998a). unfortunately, today many hunters discard alces vol. 41, 2005 timmermann and rodgers moose values 91 state 1996 1997 1998 1999 2003 wy n/a $12,500 $7,500 $8,250 ($12,500 & $14,000)1 co $19,000 $9,750 $7,000 $7,500 ut n/a n/a n/a $7,250 me $8,000 $7,200 n/a n/a table1. cost of a single moose license received at u.s. governors moose license auction 19962003 (brimeyer 1999). 1 n. a. moose foundation sun valley id convention december 2003 (m. orwig, personal communication 2004) — note: previous convention yielded $10,500 & $12,500 (wy). moose hides after processing the meat. in norway, moose hides are valued at between $17 and $32 (sødal 1985). potential current value of hides in norway based on a harvest of 40,000 could be as high as $1.3 million according to storaas et al. (2001). cultural/spiritual brownlee et al. (2002) reported finding a single moose antler along with human bones from a cree burial site near the manitoba/ontario/minnesota border. the bones, which help provide knowledge of native heritage, were dated to 6750 b.p., the oldest yet found in manitoba. evidence of moose in native culture is found in indian pictographs dating back 500 years or more along interconnected lakes west of lake superior (dewdney and kidd 1962). moose are depicted on several sites including those on lac la croix, hegman, blindfold, crooked, and darkey lakes in the quetico park-boundary waters canoe area of ontario and minnesota. moose petroglyphs have also been reported in kejimkujik national park, nova scotia, and in sweden and norway (hallström 1960, jansson et al. 1989). such rock carvings and paintings in fennoscandia are commonly interpreted as the art of prehistoric man and some date back to the early 1600s (jansson et al. 1989, skogsstyrelsen 2002). this art bears witness to a mobile hunting and trapping culture and moose (the crucial winter prey) represents the dominant type of figure, suggesting it was a prime target. “flesh foods”, primarily caribou (rangifer spp.) and moose were “the only significant form of sustenance” for many north american natives in the 1600s according to gillespie (1981:15). the entire carcass including meat, internal organs, hide, and skeleton were used. hunting, fishing, and trapping provides over half the total income for some of canada’s native populations and, for some, moose accounts for a high percentage of dietary needs (hamilton 1981, eagan et al. 1989). native peoples often viewed wildlife including moose as their spiritual kin where hunting success was obtained by following prescribed rituals and atonement after the kill (feit 1987). the “rabbit and the frog” is one of several sacred legends of the sandy lake cree of northwestern ontario (stevens 1971). the frog attacked the “windigo-moose”, by crawling up its rectum and biting a vital organ that killed the moose. both the frog and her husband, the rabbit, ate on the moose until full. wolves came along, and while the frog escaped by leaping into a bloody hole in the snow, the rabbit, who crawled into the carcass, was eaten by the wolves. moose are also part of ojibway legend in the lake nipigon region of northern ontario, as related by morriseau (1977). of all the ungulates in north america, kay (1997) believes moose were the easiest to kill. he proposed that moose biogeography in western north america moose values – timmermann and rodgers alces vol. 41, 2005 92 was controlled primarily by native hunting and presented evidence supporting an aboriginal overkill hypothesis. crichton (1981) reported unregulated hunting by treaty indians in manitoba was a prime factor responsible for reduced moose populations. crichton (1987, 2001) and nepinak and payne (1988) reported many first nations peoples are concerned over moose conservation and wish to be involved in active management programs. feit (1987) suggested the licensed sport or recreational hunt or extensive forestry practices (clearcutting and road access) could disrupt native management practices in some areas. he further believes that conflicts develop when such groups of resource users do not share a common cultural heritage. aesthetics the intrinsic value of wildlife often exceeds utilitarian values normally associated with monetary measures (omnr 1991). moose provide a challenging and rewarding subject for naturalists, photographers, painters, and outdoor recreationists (fraser 1978). wildlife viewing of large mammals including moose and whales were reported by 43.3% of participants in a 1996 environment canada study (duwors et al. 1999). non-consumptive use and appreciation may include direct observation, photography, painting/sketching, and exploring habitat for sign (droppings, tracks, browse, sound, shed antlers, beds, tree damage). most management agencies recognize non-consumptive values and have developed policies to provide moose viewing opportunities. viewing areas can be especially fruitful in locations such as provincial, state, and federal parks where densities are generally higher and hunting is prohibited (cobus 1972, timmermann and buss 1998). indeed, hunting and viewing could be considered an incompatible activity in some areas, as hunting tends to reduce densities and conditions moose to avoid people (yukon renewable resources 1996). moose are native to at least 35 north american national parks in 16 jurisdictions and are highly prized by residents and visiting tourists (timmermann 2003). in alaska, regelin and franzmann (1998) report moose to be a favored viewing animal for over 1 million visiting tourists each year. many agencies promote the development of moose viewing areas along specific road corridors, rivers, and lake shores, where expectations of observing moose are high. yukon renewable resources (1999) for example lists 6 popular summer and 4 fall/winter moose viewing sites. vermont features moose viewing opportunities around selected roadside wetlands and salt-licks in the “vermont watchable wildlife guide” (alexander 1993). in wyoming, moose are viewed by thousands, especially in grand teton and yellowstone national parks (hnilicka and zornes 1994). ontario targeted the development of specific interim (1985, 1995, and 2000) moose viewing opportunities (omnr 1980). however, little effort was made to identify moose viewing sites and it was unclear how much progress was achieved in reaching this goal (timmermann et al. 2002). moose are maintained in captive conditions for display and education, scientific research, and commercial breeding (schwartz 1992, monska 2001). although difficult to keep and expensive to feed, at least 29 facilities kept captive or semicaptive moose in north america in 1990 (schwartz 1992). filion et al. (1983) suggested that total expenditures on north american wildlife related activities other than hunting were 3.5 times direct expenditures by hunters. bisset (1987) used this value to speculate that the total value of moose in wildlife appreciation including non-consumptive use could be as much as $1,315 million in 1982. in saskatchewan, an ecotourism alces vol. 41, 2005 timmermann and rodgers moose values 93 outfitter’s license is required to conduct moose viewing tours (arsenault 2000). in norway, storaas et al. (2001) reported current moose viewing enterprises that offer a “moose viewing safari” are small and unstable, but that the potential exists to provide added economic benefits to landowners or local forest companies. commercial the commercial value of moose is often associated with the economic impact of commercial outfitters who market a hunt. outfitters generally provide a package which may include food and accommodations, a remote fly-in experience, a guide, and occasionally a license. in 2003 such a packaged hunt (1bull tag between 4 u.s. non-residents) typically sold for $6,000 in ontario (peter davis, personal communication 2003). assuming a tag fill rate of 50%, each harvested bull could generate $11,600 $15,400 based on the 2003 currency ex-change rate. hunter outfitting packages for moose in the northwest territories average $6,500 depending on travel mode, (e.g., horse, helicopter, backpacking, or use of packers; veitch and simmons 2002). in northern ontario, the outfitting industry is the third largest industry, behind only forestry and mining (omnr 1986). competing values have resulted from increased commercial logging activities and associated roads penetrating remote areas (mckercher 1992). tourist outfitters who provide a traditional fly-in “wilderness” hunting experience are negatively affected, while timber contractors benefit from the newly accessed wood supply. the economic benefits and costs associated with habitat management, such as controlled burns, vegetative crushing, and scarification may generate significant employment and expenditures of equipment in some areas (oldemeyer and regelin 1987, al franzmann, personal communication 2004). moose are used to market and sell a wide variety of products throughout north america (appendix). moosehead, canada’s oldest independent brewery is the leading canadian beer imported to the u.s. “moose, mountains and mounties” were tourism canada’s marketing focus in the early 1980s. bars, restaurants, gift shops, motels, lodges, rv parks, children’s toys and books, and a host of products feature moose as a marketing tool. the distinctive shape and size of moose, and its symbolic identification with canadian wilderness as the “monarch of the north”, lends itself to an image (an icon of the virtues of strength and independence) which attracts attention and sells products. the commercial value of such products is unknown, but thought to be significant and growing. moose are the national animal of norway (storaas et al. 2001), and the official state animal of maine (morris and elowe 1993). a detailed study of “moose-marketing” and its impact on the economy is needed. scientific/ecological scientific studies of moose have been carried out by many agencies as summarized in 19 chapters in ecology and management of the north american moose (franzmann and schwartz 1998). the alaskan moose research center (mrc) located on the kenai peninsula has, since its inception in 1968, provided data useful in evaluating the capability of land to produce moose (schwartz and hundertmark 1994, franzmann 1996). the mrc produced 90 professional journal publications, 32 publication proceedings, 12 book chapters, 8 dissertations/theses, and numerous reports up to 1996. isle royale boasts the longest running ecological study of wolves and moose beginning in 1959 (peterson and vucetich 2002). published research on moose-related studies on the island (150+) was summarized by jordan et al. (2000). moose values – timmermann and rodgers alces vol. 41, 2005 94 future research efforts in north america will likely concentrate on new knowledge concerning the role of predators, habitat quality, and how hunting influences moose population dynamics (crichton et al. 1998). forest ecosystem management has replaced featured species management in north america during the late 1990s, but moose will remain a crucial management species in eastern canada (hénault et al. 1999). the impact of moose on forest ecology can be used as an indicator of forest health and biodiversity crichton (1998c). in sweden, the effects of a dense moose population on forest biodiversity and harvest regulation have been a research priority at the grimsö wildlife research station in riddarhyttan and the swedish university of agricultural sciences in uppsala (persson et al. 2000, edenius et al. 2002, sylvén 2003). the norwegian institute for nature research (nina) in trondheim and the finnish game and fish research institute in ilomantsi finland also conduct moose research on similar topics (andersen and saether 1992, jaren 1992, härkönen et al. 1998, angelstam et al. 2000, heikkilä and härkönen 2000, danielsen 2001). habitat utilization deciduous shrubs and trees may compete with coniferous plantation growth, but they also provide seasonal food for moose. intense moose browsing on deciduous shrubs and trees, particularly in winter, is considered beneficial by providing a release effect on adjacent commercially more valuable coniferous tree species (lavsund 1987, andrén and angelstem 1993, posner and jordan 2002). summer leaf-stripping opens the canopy to more light, which helps accelerate growth and forest succession (bubenik 1989). in addition, moderate browsing temporarily retards terminal shoot development and helps reinforce the root system of trees and saplings. thompson and curran (1993) reported heavy moose browsing in some areas of newfoundland may improve the commercial value of forests by thinning balsam dominated stands and increasing the relative spruce component. connor et al. (2000) suggest moose have altered species composition, the quantity of remaining available browse, and influenced forest successional patterns in gros morne national park, newfoundland. hänninen (1994) speculated one moose in finland could provide forest cultivation worth $400.00 per year and that density dictated the degree of cultivation. on the other hand, silvicultural practices, including use of herbicides following logging, may reduce the period of vegetative succession, thus reducing the overall value of added browse production (kennedy and jordan 1985, hjeljord and gronvold 1988, cumming et al. 1995, eschholz et al. 1996). competing values resource allocation moose have traditionally provided an important source of food and clothing for native people inhabiting north american moose range (crête 1987). moose are publicly owned in most jurisdictions and held in trust by government (schwartz et al. 2003). in ontario, the heritage hunting and fishing act (2002; http://www.e-laws. gov.on.ca/dblaws/source/statutes/english/2002/s02010_e.htm#toc) recognizes the right of all persons, native and non-native, to hunt and fish in accordance with provincial laws. however, subsistence use by native people as provided under treaty or other legal agreements is given priority in harvest allocation by at least 10 of 23 agencies that manage a harvest. native people under treaty, followed by resident hunters, are typically favored over nonresidents in allocating harvest opportunities (franzmann and schwartz 1983, timmermann 2003). in canada, status indians are alces vol. 41, 2005 timmermann and rodgers moose values 95 permitted under treaty to hunt for food for personal use without provincial licenses year round. in september 2003, the supreme court of canada ruled that metis people of mixed native and european descent, who can prove a historic link to surviving metis communities and customs, can also claim hunting and other aboriginal rights (canadian press 2003). potentially 600,000 status indians across canada will have to share available wildlife resources with about 300,000 metis people. non-native ontario moose hunters are fearful that some moose stocks could be overharvested and that tag allocations to non-native residents will be reduced as a higher percentage of the sustainable moose harvest is allocated to native and metis people (john kaplanis, personal communication, 2003). non-resident moose hunting opportunities have been significantly reduced or eliminated in many jurisdictions (timmermann and buss 1998). non-resident hunter numbers in 1982 were one-third those of 1972 due in part to increased license fees, resident-only seasons, guide and registration requirements, and limited permits. such reductions can be considered competing to the extent that they diminish the potential economic return, as non-residents typically spend an average of 25% or more per moose than resident hunters (bisset 1987). illegal harvest the illegal kill probably comprised at least 30% of the total north american licensed harvest in the early 1980s (wolfe 1987). this would be an illegal kill of approximately 27,000 moose based on an estimated licensed harvest of 70,300 in 1982 (timmermann 1987). other authors since then believe the illegal harvest in some jurisdictions may approach or exceed the annual legal harvest (timmermann 2003). illegal harvests and their resultant value losses (reduced populations and economic losses to the licensing jurisdiction) are considered by many to compete with values generated by licensed legal harvests, although such harvests may also be considered positive in helping lower high moose densities and by providing a meat supply. todesco (2004) and pilgrim (2000) have recently reported on the impact of illegal moose kills in north-eastern ontario and the great northern peninsula of newfoundland. reliable estimates of severity of illegal loss are lacking and are further compounded in some jurisdictions by unregulated removals of unknown magnitude by native people under treaty and by a recent decision by the canadian supreme court to grant metis people specific hunting rights (crichton 1981, canadian press 2003). fine schedules and resultant monetary penalties for illegal hunting convictions can constitute an indirect value. most jurisdictions have established a minimum or base fine schedule for hunting infractions. under some circumstances, the court may assess considerably higher fines or penalties. in the u.s. for example, moose liquidation or replacement costs are used to set a minimum fine schedule for illegal harvests. replacement costs set by 13 western/mid-western states in 2002 averaged $1,789.00 (range $262.50 in wisconsin to $6,000 in montana) and were second only to bighorn sheep (jensen 2002). replacement costs double in minnesota and montana if a moose is classified as a trophy. in ontario, the current set fine for a resident hunting moose without a license is $234.85. this fine would be considered a minimum penalty or out of court settlement if a ticket is issued. it does not reflect the true value of a moose (charlie todesco, personal communication 2003). penalties can range from $19,250 $77,000 under the fish and wildlife conservation act (1997; http://www.e-laws.gov.on.ca:81/ isysquery/irlf22d.tmp/16/doc) if a moose values – timmermann and rodgers alces vol. 41, 2005 96 summons is issued instead of a ticket, and the actual fine is determined in court. on december 5th, 2002, 4 ontario hunters were fined $21,560 in a thunder bay ontario court for several hunting violations (bob stewart, personal communication 2003). two of the moose hunters were fined $7,700 each for hunting moose with an aircraft and banned from hunting for 3 years. one of this party was fined $2,310 for shooting a firearm from a motor boat, while other fines of $770 were levied for giving false statements. some jurisdictions including alaska may confiscate equipment used for illegal harvest, including aircraft, boats, and firearms, worth $100,000 in some cases (al franzmann, personal communication 2004). lacasse (1986) proposed a formula to determine an appropriate fine for moose poaching in quebec by considering 3 variables: animal biomass, management costs, and socio-economic losses. two biomass values (70 kg for calves and 190 kg for all lent beef value ($4.20/pound). management costs included summations of expenditures for research, conservation, inventory, data processing, and administration of the hunt. a proportional factor (58-day firearm and archery season x 365 days = 15.9%) expressed the relative significance of moose hunting activities circa 1985. this percentage was applied to estimate management costs and then divided by the estimated quebec moose population to derive an annual expenditure per moose. socio-economic losses considered the average number of days to kill a moose in 1984 (70) multiplied by the average daily expenses per moose hunter ($61.60). this method proposed a lost value of $4,620 for a calf and $5,390 for an adult animal. forest damage moose are a dominant or keystone species throughout much of the boreal forests of north america (bergerud and manuel 1968, paine 1988) and in pine/spruce forests of sweden, norway, and finland (lavsund 1987). similarities between the impact of 2) on commercial stands in newfoundland and fennoscandia are remarkable (thompson and curran 1989). balsam fir (abies balsamea) stands that dominate newfoundland commercial forests provide similar early successional moose habitats following logging as do scots pine (pinus sylvestris) in fennoscandia (lavsund 1981, 1987). in north america, damage by moose browsing, primarily in balsam-dominated stands has been reported by krefting (1951, 1974) on isle royale and in newfoundland (pimlott 1963; bergerud and manuel 1968; thompson and curran 1989, 1993; mclaren et al. 2000). concern is centered on the reduction of available food leading to declining densities and significant damage to growing commercial trees leading to serious financial costs and short-term wood supply deficits. several moose dieoffs have been reported on isle royale due to reduction of available browse by high density populations (risenhoover and maass 1987, peterson 1997). after each reduction, moose numbers have gradually increased following habitat recovery. forest damage has increased, concurrent with a moose population increase in newfoundland (mclaren et al. 2004). moose are especially attracted to pre-commercial thinned stands (thompson 1988) and, as such, the management challenge is to reduce commercial damage by moose in such stands. regenerating balsam fir are commonly mechanically thinned at 10-12 years post cut to reduce stem density and enhance tree growth (thompson and curran 1989). heavy moose browsing in thinned stands can remove >50% of current growth, especially in high production areas. newfoundland currently spends >$3 alces vol. 41, 2005 timmermann and rodgers moose values 97 million per year in a pre-commercial thinning silvicultural program, which yields a supply >400,000 m3 of wood per year (lingard 1997). managers can help reduce stand damage by lowering moose densities (increasing hunter harvest), retaining hardwood competition to help re-direct browsing pressure, and delaying thinning 3m (i.e., 13-15 years post cut; thompson 1988, mclaren et al. 2000). some of the highest densities of moose in the world are found in fennoscandia. low natural predation, evolution of modern forestry practices creating a checkerboard of good habitat, and a closely regulated human selective harvest in norway (myrberget 1979), sweden (cederlund and markgren 1987), and finland (heikkilä and aarnio 2001) are believed responsible. forestry is essential to the economy of all 3 countries and much of the moose habitat is used for commercial wood production (lavsund and sandegren 1989). forest products represented 25% and 40% of export markets for sweden and finland, respectively, in the early 1980s (lavsund 1987). damage to economically important scots pine, the principle winter forage species in young plantations, is common and directly related to moose density (ahlén 1975, angelstam et al. 2000). breakage of pine tops of saplings up to 3-4 m, and bark stripping by moose results in reduced future growth and wood quality (sandgren 1980, faber and edenius 1998). in addition, retention of deciduous forest biodiversity is reduced (angelstam et al. 2000). special survey methods were recently introduced in sweden to measure level of damage to commercially valuable trees as well as species important to biological diversity (skogsstyrelsen 2002). in norway, lavsund (1987) reported that a density of 2.0 moose/km2 would produce 25% damage in thinned stands. a lower density of moose in sweden (1.7/km2) caused up to 57% damage in pine-dominated stands according to angelstam et al. (2000). damage in sweden alone was believed to be in the range of $200 to $500 million dollars per year, or $1,000 per moose shot in the early 1980s (lavsund 1989). estimated damage to swedish forests in 2003 could equal $60 – $175 million (roger bergström, personal communication 2004). compensation paid to finnish landowners varied from $1-4 million (heikkilä and aarnio 2001), while estimated browsing damage to young pine in norway was $20 – $40 million after 80 years (solbraa 1998). many studies (50+) in fennoscandia have resulted in attempts to quantify and reduce conflicts between forest owners who wish to lower moose populations and hunters who wish to retain or increase moose densities (ball and dahlgren 2002). the cost of browsing on commercial tree species in norway according to solbraa (1998) exceeds the present income many forest owners derive from the sale of hunting rights. due to moose damage, densities were reduced in all 3 countries by increased harvests in the 1940s and 1950s (dahl 1979, myrberget 1979, jaren 1992). moose harvests in sweden increased from around 35,000 per year in the mid-1960s to 132,000 in 1980 (lavsund 1981). populations were purposely reduced by increasing harvests from a high of 174,000 in 1982 to around 90,000 per year thereafter due to excessive damage to commercial stands (lavsund 1987). in sweden and norway, lavsund et al. (2003) reported a growing trend to replant spruce (picea abies) which is much less attractive to moose than pine. this effort to reduce damage to commercial tree species and subsequent reduction in pine browse may result in lower moose densities in the future. agricultural crop and garden damage moose occasionally damage agriculmoose values – timmermann and rodgers alces vol. 41, 2005 98 tural crops. in newfoundland, high moose densities have impacted a growing agrifoods sector (wicks 2002). a 2-year project tested the effectiveness of electrobraid fencing tm to protect 27 acres of cabbage, resulting in zero crop losses in fields that had previously suffered large crop losses. in fairbanks, the matanuska valley, and kenai peninsula of alaska, moose routinely forage in vegetable gardens and on ornamental trees. preventative measures used include erecting 10-foot-high fences, placing dried bloodmeal around plants, and hanging pieces of bear hide to keep moose away (anonymous 1998). increasing moose populations in idaho have caused substantial concerns to private landowners from damage to crops and breaking of fences in the late 1980s (idfg 1990). damage to agricultural crops (baled alfalfa, sugar beets, sunflowers, and small grains) has occurred in northwestern minnesota and nuisance permits are occasionally issued to deal with such damage (mark lenarz, personal communication 2004). in fennoscandia, governments pay compensation damage for moose damage to cultivated crops. such costs in norway were estimated at $64,500 – 165,000 in 1999 (storaas et al. 2001). vehicle collisions moose/vehicle collisions are a serious concern in some areas of north america and fennoscandia (almkvist et al. 1980; damas and smith 1983; sanderson 1983; nilsson 1987; child et al. 1991; del frate and spraker 1991; lavsund and sandegren 1991; mcdonald 1991; oosenbrug et al. 1991; schwartz and bartley 1991; child 1998; seiler 1999, 2003; joyce and mahoney 2001; redmond et al. 2004). moose killed by vehicles rather than hunters represent an economic expense for automobile repair and human health care as well as potential economic loss from recreation. moose are particularly attracted to some roadsides by new growth along right-of-way edges as well as high salt concentrations resulting from winter de-icing operations (child 1998). mvcs have a far greater average and overall economic impact than do collisions with other wildlife (humphrey 2002). most estimates of collision costs are conservative, as most agencies do not maintain accurate records (child and stuart 1987, romin and bissonette 1996, sullivan and messmer 2003). mvcs in newfoundland varied considerably, depending on reporting sources (table 2), and those by joyce and mahoney (2001) were considerably higher than those of oosenbrug et al. (1991) and rattey and turner (1991). in sweden, seiler (2003) believes mvcs may be at least double those suggested by official statistics. in british columbia, child et al. (1991) reported that the number of reported mvcs may be underestimated by 2 6 times the actual number of moose killed, especially as an unknown number are also killed or crippled on logging, mining, and rural roads. in newfoundland, as moose densities and traffic volumes increased in the 1980s, collision deaths nearly doubled (oosenbrug et al. 1986, 1991). property damage during the 1980s averaged $1,155 per accident in ontario and $1,848 in newfoundland (table 3). a later study by joyce and mahoney (2001) estimated the following annual losses based on 750 mvcs/year: $269,500 in initial health costs, $1,212,750 in vehicle damage, $431,200 in consumable moose meat, and $161,700 in losses to the outfitting and related industries or $2,767 per moose. in alaska, property damage averaged $4,000 per vehicle at one auto body shop after 366 mvcs on the kenai peninsula during the winter of 1989/90. a decade later, nearly 300 moose fatalities were reported up to mid-february in alaska (anonymous 1999a). human deaths (no year given) resulting from mvcs were reported by child and stuart (1987) for newfoundland alces vol. 41, 2005 timmermann and rodgers moose values 99 1 kenai peninsula, severe deep-snow winter (human deaths occurred in summer). 2 average # of dead moose reported 1999-2003 (range 177-215). 3 first human death (kilpatrick pers. com. 2004). 4 one human death over 20 years ago. table 2. estimated numbers of moose/vehicle collisions in fennoscandia and north america. jurisdiction year # collisions # human deaths reference sweden 1980 6,000 15 lavsund and sandegren (1991) 1996 4,000 n/a groot-bruinderink and hazebroek (1996) 2003 10,000+ 10-15 seiler (2003) norway 1996 1,500 n/a groot-bruinderink and hazebroek (1996) 2002/03 2,600 n/a lykke (pers. com. 2004) 2003 1,200 n/a seiler (2003) finland 1996 150 n/a groot-bruinderink and hazebroek (1996) 2001 3,000 n/a rajamäki and mänttäri (2002), heikkilä (pers. com. 2004) north america early 90's 3,500+ n/a child (1998) nl 87-88 661 3 rattey and turner (1991) 1989 897 4 joyce and mahoney (2001) 1990 460 4 oosenbrug et al. (1991) 1990 867 4 joyce and mahoney (2001) 1994 616 0 joyce and mahoney (2001) 1997 595 0 joyce (pers. com. 2004) me 90-92 600/yr n/a morris and elowe (1993) 96-98 2,126 8 humphrey (2002) 99-01 2,068 8 morris (pers. com. 2004) ak 89-901 366 2 del frate and spraker (1991) 89/90 665 n/a schwartz and bartley (1991) 98/991 300 n/a anonymous (1999a) 2000 689 n/a sullivan and messmer (2003) nh 2003 225 n/a bontaites (pers. com. 2004) vt 1997 165 n/a alexander et al. (1998) 99-03 166/yr n/a alexander (pers. com. 2004) wy 79-93 206 n/a hnilicka and zornes (1994) pq early 90s 265 n/a child (1998) 99-032 188 2-3 lamontagne (pers. com. 2004) ns 2003 20 1/10yrs power (pers. com. 2004) bc 1990 400-1200/yr n/a child et al. (1991) 99/03 713 n/a child (1998) ma 88-923 10 1 (2003) vecellio et al. (1993) yt 20034 <12 0 ward (pers. com. 2004) ct 2003 2 0 kilpatrick (pers. com. 2004) nwt 99/03 4 0 veitch (pers. com. 2004) moose values – timmermann and rodgers alces vol. 41, 2005 100 (6), quebec (5), british columbia (2), and new brunswick (1). economic and human impacts of mvcs in maine have more than doubled in 10 years (1988-98). during a 3year period (1996-98), humphrey (2002) reported 2,126 collisions resulted in 8 human fatalities (table 2) and 637 injuries with an estimated economic impact of $23,600 per collision, totaling $50.2 million. a similar pattern held for the period 1999-2001 when 2,068 mvcs caused 583 human injuries, 8 human deaths and an economic impact of $48.59 million (karen morris, personal communication 2004). mvc damage costs include; material loss to vehicles, human injuries (ambulances and medical expenses), human fatalities (life insurance, funeral expenses), call-out costs for police, veterinarians, and wildlife officials to deal with injured or dead moose, loss of meat and hunting opportunities, and the societal costs of traffic delays (seiler 2003). fennoscandia experiences the highest number of mvcs (table 2). in sweden, 500 human injuries and between 5 and 20 deaths were reported each year during the 1980s (lavsund and sandegren 1991). in recent years, sweden leads norway and finland with an estimated 10,000+ mvcs per year (seiler 2003). the swedish national road administration calculated an year. similarly, in norway, mvc costs were believed to be between $11 and $17 million 3,500 moose were killed annually on north american roadways in the early 1990s. this could represent a direct economic loss of $10 million or more using average meat value and vehicle damage losses assumed by oosenbrug et al. (1991). wildlife managers have historically been in conflict with transportation agencies in placing management emphasis on increasing moose populations, which cause increasing mvcs on modern high-speed, high-traffic, highways. in the future, managers must balance the positive aspects of higher moose populations against potential negative impacts when formulating population objectives (sullivan and messmer 2003). train collisions losses of moose to train collisions are considered socially unacceptable and economically costly (rausch 1958, child 1983). the frequency of moose/train collisions varies from year to year and is closely linked to winter snow levels, especially on winter range in mountain valleys of alaska, british columbia, and norway (rausch 1958, becker and grauvogel 1991, child 1998, gundersen et al. 1998). the majority of reported kills occur in winter when seasonally migrating moose move to valley bottoms where transportation corridors are located (jaren et al. 1991, modafferi 1991). in british columbia, losses have exceeded 1,000 animals in years of above average snowfall (child et al. 1991). similarly, 725 moose died in alaska in the deep snow winter of 1989-90 (modafferi 1991). an average of 500 moose/year were killed by trains in norway during the 1980s (jaren et al. 1991), and rose to 1,000 in 1993 (lavsund et al. 2003). although poorly documented, seiler (2003) believes swedish trains killed at least 900 moose each year since 2000. schwartz and bartley (1991) believe such losses can be biologically significant to some populations. unfortunately, most jurisdictions only subjectively report moose/train collisions and damage costs. these are often considered underestimates and are largely unspecified (child and stuart 1987, schwartz and bartley 1991). collisions with moose may damage trains (including derailments), cause passenger delay and physical strain on train personnel, hold population levels below potential, and reduce income to landownalces vol. 41, 2005 timmermann and rodgers moose values 101 ers with hunting rights according to child (1983), child and stuart (1987), and andersen et al. (1991). vegetative clearing of railway right-of-ways to reduce available forage appears to hold promise in reducing collision frequency (andersen et al. 1991, jaren et al. 1991). schwartz and bartley (1991) suggested conducting a special “train hunt” during severe winters on the premise that a human harvest is a wiser use of moose than moose killed by trains. habitat enhancement near railroad tracks and clearing escape routes away from the railbeds has been used in alaska to reduce moose-train collisions (al franzmann, personal communication 2004). aircraft collision/damage although rare, moose/aircraft collisions have occurred at several alaskan airports including those at soldotna, kenai municipal, and anchorage international (child 1998). near-collisions of aircraft with moose have necessitated expensive airport fencing. moose breaching these fences have been killed in special hunts within the 1 total expenditures per moose. 2 average for 13 u.s. states (range $ 262.50 in wi to $ 6,000 in mt). 3 average per vehicle at one body shop. 4 estimated economic impact per mvc. 5 one bull tag per 4 non-resident hunters ($6,000) & a tag-fill rate of 50%. 6 appropriate pq fine formula (animal biomass, mgmt. costs, & socio-economic losses). 7 average for 750 mvc (health costs+vehicle damage+loss of meat+outfitting losses). 8 average direct cost per mvc (swedish national road administration). jurisdiction year $value/ moose subject reference north america 1982 $1,285 / res. hunter1 bisset (1987) 1982 $1,688 / non-res. hunter1 bisset (1987) 1982 $1,068 meat bisset (1987) early 90s $1,082 meat crichton (1998a) united states 2002 $1,789 replacement2 jensen (2002) ak 1990 $4,000 mvc3 del frate and spraker (1991) me 2000 $23,000 mvc4 humphrey (2002) canada on 2003 $11,600-15,400 / harvested bull5 peter davis (pers.com. 2003) 1980s $1,155 mvc fraser and hristienko (1982) pq 1986 $5,390-adult poaching loss6 lacasse (1986) $4,620-calf nl 1980s $1,848 mvc oosenbrug et al. (1986) 1990 $1,016 meat oosenbrug et al. (1991) 2000 $2,767 mvc7 joyce and mahoney (2001) norway 1980s $17-32 / hide sødal (1985) 2000 $1,050 meat storaas et al. (2001) sweden 2001 $1,143 meat sylvén (2003) 2003 mvc8 seiler (2003) finland 1994 $400/yr forest cultivation hänninen (1994) moose values – timmermann and rodgers alces vol. 41, 2005 102 airport boundaries (al franzmann, personal communication 2004). moose believed to be in rut caused thousands of dollars of damage to 18 parked floatplane aircraft in 1999 (anonymous 1999b). in february, 1997, moose damaged 7 small aircraft in one week. pre-antler cast rubbing was the suspected cause. kastdalen (1998) reported on measures taken in norway to minimize moose collisions with aircraft and connecting road and rail traffic to the new gardermoen national airport near oslo. complete fencing in combination with overor underpasses were employed to prevent collisions and allow seasonal moose migration. prey value moose provide a source of food for large carnivores including wolves, black bears, and grizzly bears, as well as an array of birds and small mammals that scavenge moose carcasses. ballard and van ballenberghe (1998) provide an extensive review of moose/predator-prey relationships and suggest predation may act as a major limiting factor in many moose populations. moose densities are generally higher in newfoundland and fennoscandia where predation of adult moose is low (mercer and mclaren 2002, lavsund et al. 2003). one might argue that predators are valuable in helping to keep densities below carrying capacity (k) by removal of less fit members of the population (petersen et al. 1984, ballard et al. 1987). on the other hand, gasaway et al. (1983) and schwartz and franzmann (1989) proposed management actions to reduce predation and increase moose densities when prey populations are depressed (predator pit) and habitat is adequate. predator control programs are extremely controversial and present managers with competing values; to reduce predation and increase moose populations or to allow moose to remain at low densities. in fennoscandia and yellowstone national park, wolves have been allowed to re-establish former ranges after receiving protection (sweden 1966, norway 1972, yellowstone national park, wyoming 1995) (andrén et al. 1999, white et al. 2003). in 1998, wabakken et al. (2001) reported 50-72 wolves in 6 reproducing packs had been established on the scandinavian peninsula. these increased to an estimated 97-107 wolves in the winter of 2002 (olof liberg, personal communication 2004). wolves were re-introduced to yellowstone national park in 1995 and 1996, and by 2004 an estimated 170-180 wolves use the park at least part of the year (rolf peterson, personal communication 2004). white et al. (2003) and gundersen (2003) believe hunting harvest and wolf predation to be largely additive and recommend managers reduce hunting harvest in yellowstone and in the koppang area of norway’s hedmark county, respectively. some woody plants, including willows (salix spp.), preferred by moose have been released from suppression in yellowstone national park following wolf establishment according to ripple et al. (2001). in the future, the major challenge will be to set population targets and maintain and improve public understanding and local acceptance of wolves as the population increases. interspecific values moose and woodland caribou (rangifer tarandus) are sympatric over much of the boreal coniferous forests of north america (boer 1998). as logging proceeds northwards, the habitat requirements of both species must be considered in formulating logging plans. in the past, habitat for both moose and caribou in ontario were managed on a species-specific basis through guidelines designed to favor each species (omnr 1988, 1999). the crown forest sustainability act (1996; http://www.e-laws.gov.on.ca/ dblaws/statutes/english/94c25_e.htm), alces vol. 41, 2005 timmermann and rodgers moose values 103 directs forest managers to maintain biodiversity on all managed forests, by emulating natural disturbance and landscape patterns (omnr 2001). woodland caribou in ontario are currently classified as threatened and a recovery strategy is being developed (omnr 2003). the objective of this strategy is to maintain caribou in northern parts of their historical range by managing for larger cuts and reducing edge density, both parameters which favor caribou but not moose (john mcnicol, personal communication 2004). hence managers are forced to weigh competing values and decide whether to manage for moose or woodland caribou. habitat management costs certain types of new habitat created by logging (i.e., edge, early succession, residual cover) will result in increased moose reproduction, survival, and population densities. consequently, agencies have developed habitat management guidelines aimed at sustaining or increasing moose densities (thompson and stewart 1998). management of moose habitats to provide sustained moose populations will impose additional constraints and higher investment costs on timber production (racey et al. 1989, sarker and surry 1998). identification of moose habitat values may limit the time during which all or part of an area will be logged (chamberlin 1981). with such competing values, managers must weigh the benefits from hunting and non-consumptive use against the timber values. summary the majestic moose, “monarch of the north” and a symbol of wilderness, is a much valued species by native indians, metis people, recreational hunters, and a host of non-consumptive users. they also provide food for a variety of mammalian and avian carnivores, including scavengers. moose represent a multitude of values, each representing varying levels of importance determined by compatible and incompatible interactions through time. we have attempted to identify and categorize such values as those considered positive or complementary and those which we believe to be largely negative or to compete. our list comprises 22 values based on a literature search of known and expressed assets of moose in north america and fennoscandia. identified values, many of which interact, do not necessarily follow accepted economic value systems and nomenclature. our objective was to assemble an array of recognized values that will aid economists and decision makers to better understand moose as a valuable renewable resource. we believe such knowledge should lead to a more comprehensive basis for improved evaluation and species management in an ecological context. knowledge of current moose densities is considered fundamental to an understanding of competing and complementary values which likely vary among individuals. twelve complementary values are identified, including recreational hunting, license revenues, and a discussion of the worth of meat, antlers, and hides for which a monetary value can be more readily identified (table 3). licensed recreational harvest, for example, can help control and sustain moose densities and promote economic benefits to local economies valued in the hundreds of million dollars annually. current license fees in north america are modest (table 1) and do not reflect the recreational and meat value of moose harvested by licensed hunters (table 3). values more difficult to quantify, are identified, including those related to cultural/spiritual, aesthetics, commercial, scientific/ecological, and habitat utilization. ten competing values largely associated with moderate to high moose densities are identified. high densities, (>2.0/km2), moose values – timmermann and rodgers alces vol. 41, 2005 104 although perceived desirable by many, may lead to extensive property/forest damage and loss of human life and injury. conflicts between moose and humans have increased, especially in proximity to urban/suburban areas (karns 1998). consequently their presence in many areas has become more socially unacceptable. values considered to compete include damage caused to commercial forests and agricultural crops, collisions with vehicles (mvcs see table 3), trains, and aircraft, as well as resource allocation conflicts, added habitat management costs, predator/prey control costs, illegal harvests, and setting interspecific priorities. in future, some aspects of moose management may need to be more decentralized, so local knowledge and values can be used more effectively in estimating local moose densities, assessing and mitigating forestry and property damage, and taking appropriate measures in sustaining healthy populations (lavsund and sandegren 1989, jaren et al. 1995). moose viewing opportunities need to be better identified and funded to promote use of prime viewing areas in a variety of locations. expansion and development of moose ecotourism to provide added benefits, especially in northern economically depressed areas, should receive attention. a detailed analysis of moose marketing, used to sell a growing variety of products (appendix), and its economic impact is needed. illegal harvests, thought to approach or exceed current licensed harvests, are believed to constitute a major competing value that begs resolution. co-management of the resource is dependent on effective joint participation between government mandated and native wildlife managers. mvcs appear to be significant and increasing in some jurisdictions and more research is needed to identify mitigating measures. mvc data can provide an inexpensive index to population change and managers should consider its use in improving the quality of associated data (mccaffery1973, case 1978, hicks 1993, alexander et al. 1998). harvest quotas in some jurisdictions may need adjustment, due to increased predation losses, following re-introductions of wolves. agencies need to re-examine their moose policies and objectives. a more holistic approach to management, that recognizes a rich variety of values, is needed. production of higher densities should be balanced with added damage costs accrued. more moose are not necessarily always beneficial. agencies need to recognize that the past policy of managing moose primarily for recreational harvest may no longer be appropriate or receive majority public support. acknowledgements we would like to thank the following individuals who provided information and advice: al franzmann and tom cooper, soldotna, ak; rolf peterson, michigan technological university, houghton, mi; marty orwig, north american moose foundation, mackay, id; bill peterson, libby, mt; mark lenarz, minnesota department of natural resources, grand rapids, mn; karen morris, maine department of inland fisheries and wildlife, bangor, me; kristine bontaites, new hampshire fish and game, new hampton, nh; cedric alexander, vermont fish and wildlife, st. johnsbury, vt; howard kilpatrick, north franklin, ct; al hicks, new york state department of environmental conservation, delmar, ny; rick ward, yukon renewable resources, whitehorse, yt; alasdair veitch, department of renewable resources, norman wells, nwt; ed telfer, edmonton, ab; vince crichton, wildlife and ecosystem management, manitoba conservation, winnipeg, mb; gilles lamontagne, quebec ministère de l’environment et de la faune, quebec city, pq; vince powers, alces vol. 41, 2005 timmermann and rodgers moose values 105 nova scotia wildlife division, kentville, ns; brianmclaren, newfoundland and labrador department of tourism, culture and recreation, inland fish and wildlife division, cornerbrook, nf; bruce ranta, ted armstrong, bob stewart , john mcnicol, bill dalton, charlie todesco, and peter davis, ontario ministry of natural resources, kenora, thunder bay, sault ste marie, and wawa, on; john kaplanis, thunder bay, on; mike buss, dwight, on; alex cringan, ft. collins, co; jon lykke, verdal, norway; sten lavsund, uppsala, sweden; olof liberg, grimso wildlife research station, riddarhyttan, sweden; and risto heikkilä, forest research institute, vantaa, finland. we are especially indebted to ed addison, al franzmann, ed telfer, mike buss, john mcnicol, alex cringan, michael wolfe, and nicole mccoy for providing constructive comments on an earlier draft, and to susan rodgers for preparation of the manuscript. references ahlén, i. 1975. winter habitats of moose and deer in relation to land use in scandinavia. viltrevy 9:45-192. alexander, c. e. 1993. the status and management of moose in vermont. alces 29:187-195. ____, p. fink, l. garland, and f. hammond. 1998. moose management plan for the state of vermont, 1998-2007. agency of natural resources, department of fish and wildlife, waterbury vermont, usa. almkvist, b., t. andré, s. ekblom, and s. a. rempler. 1980. slutrapport viltolycksporjekt. (final report of the game accident project). swedish national road administration. tu146:1980-05, borlänge, sweden. (in swedish with english summary). andersen, r., and b-e. saether. 1992. interactions between a generalist herbivore, the moose, and its winter food resources: a study of behavioural variation. alces supplement 1:101-104. _____, b. wiseth, p. h. pedersen, and v. jaren. 1991. moose-train collisions: effects of environmental conditions. alces 27:79-84. andrén, h., and p. angelstem. 1993. moose browsing on scots pine in relation to stand size and distance to forest edge. journal of applied ecology 30:133-142. _____, o.liberg, and h. sand. 1999. de stora rovdjurens inverkan på de vilda bytesstammarna i sverige. sou 1999:146. angelstam, p., p. e. wikberg, p. danilov, w. e. faber, and k. nygrén. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland, and russian karelia. alces 36:133-145. anonymous. 1998. out damn moose. the moose call 8:7. anonymous. 1999a. tough winter on the kenai peninsula, alaska. the moose call 9:29-30. anonymous. 1999b. plane bashing moose wearing out welcome. the moose call 10:27. arsenault, a. a. 2000. status and management of moose (alces alces) in saskatchewan. saskatchewan environment and resource management, fish and wildlife branch. fish and wildlife technical report 00-1. saskatoon, saskatchewan, canada. ball, j. p., and j. dahlgren. 2002. browsing damage on pine (pinus sylvestris and p. contorta) by migrating moose (alces alces) population in winter: relation to habitat composition and road barriers. scandinavian journal of forest research 17:427-435. ballard, w. b., and v. van ballenberghe. moose values – timmermann and rodgers alces vol. 41, 2005 106 1998. predator/prey relationships. pages 247-273 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, j. s. whitman, and c. l. gardner. 1987. ecology of an exploited wolf population in south-central alaska. wildlife monographs 114. becker, e. f., and c. a. grauvogel. 1991. relationship of reduced train speed on moose-train collisions. alces 27:161168. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. journal of wildlife management 32:729-746. bishop, r. c. 1987. economic values defined. pages 24–33 in d. j. decker and g. r. goff, editors. valuing wildlife: economic and social perspectives. westview press, boulder, colorado, usa. bisset, a. r. 1987. the economic importance of moose (alces alces) in north america. swedish wildlife research supplement 1:677-698. boer, a. h. 1998. interspecific relationships. pages 337349 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. boone and crockett club. 1988. records of north american big game. 9th edition. boone and crockett club, dumfries, virginia, usa. bottan, b. j. 1999. exploring the human dimension of thunder bay moose hunters with focus on choice behaviour and environmental preferences. m.sc. thesis, faculty of forestry and the forest environment, lakehead university, thunder bay, ontario, canada. boyle, k. j., and a. g. clark. 1993. does getting a bull significantly increase value? the net economic value of moose hunting in maine. alces 29:201-211. brandner, t. a., r. o. peterson, and k. l. risenhoover. 1990. balsam fir on isle royale: effects of moose herbivory and population density. ecology 71:155-164. brimeyer, d. 1999. reno auction-governor’s licences. the moose call 9:28. brown, p. j., and m. j. manfredo. 1987. social values defined. pages 12–23 in d. j. decker and g. r. goff, editors. valuing wildlife: economic and social perspectives. westview press, boulder, colorado, usa. brownlee, k., e. l. syms, v. crichton, and b. j. smith. 2002. an ancient burial site and an ancient moose antler. the moose call 15:1. bubenik, a. b. 1989. coevolution of forest ecosystems, as a basis for coexistence of wildlife and forest, with special respect to pecoran. unpublished report, forestry/wildlife workshop, ontario ministry of natural resources, algonquin region, huntsville, ontario, canada. canadian press. 2003. metis hail ‘major’ win, supreme court affirms their hunting rights. the canadian press, september 20, 2003. capel, r. e., and r. k. pandey. 1973. demand estimation in planning for intensive resource management: deer and moose hunting in manitoba. north american wildlife conference 38:389403. carlestål, b., editor. 2000. är älgen ett hinder för att nå de skogspolitiska målen? (is moose an obstacle to reach the goals of the forestry policy). kungl. skogs-och lantbruksakademiens tidskrift 139(2):1-97. case, r. m. 1978. interstate highway alces vol. 41, 2005 timmermann and rodgers moose values 107 road-killed animals: a data source for biologists. wildlife society bulletin 66: 6-13. cederlund, g. 1996. highest moose density in the world? the moose call 4:11. _____, and g. markgren. 1987. the development of the swedish moose population, 1970-1983. swedish wildlife research supplement 1:55-62. chamberlin, l. c. 1981. managing for fish and wildlife values within the forest management planning process. alces 17:193-228. child, k. n. 1983. railways and moose in the central interior of british columbia: a recurrent management problem. alces 19:118-135. _____, 1998. incidental mortality. pages 275-301 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, s. p. barry, and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27:41-49. _____, and k. m. stuart. 1987. vehicle and train collision fatalities of moose: some management and socio-economic considerations. swedish wildlife research supplement 1:699-703. cobus, m. w. 1972. moose (alces alces) and campers in sibley provincial park: a study of wildlife aesthetics. m.sc. thesis, university of guelph. guelph, ontario, canada. condon, b., and w. adamowicz. 1995. the economic value of moose hunting in newfoundland. canadian journal of forest research 25:319-328. connor, k. j., w. b. ballard, t. dilworth, s. mahoney, and d. anions. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 36:111-132. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1:553-563. crichton, v. f. j. 1981. the impact of treaty indian harvest on a manitoba moose herd. alces 17:56-63. _____. 1987. moose management in north america. swedish wildlife research supplement 1:541-551. _____. 1998a. hunting. pages 617-653 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____. 1998b. shed antler records of north american big game. the moose call 8:28-29. _____. 1998c. moose and ecosystem management in the 21st century does the king have a place? a canadian perspective. alces 34:467-477. _____. 2001. co-management the manitoba experience. alces 37:163-173. _____, and g. redmond. 1998. did you know? the moose call 8:18. _____, w. l. regelin, a. w. franzmann, and c. c. schwartz. 1998. the future of moose management and research. pages 655-663 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. cumming, h. g., c. p. kelly, r. a. lautenschlager, and s. thapa. 1995. effects of conifer release with vision® (glyphosate) on moose forage quality. alces 31:221-232. dahl, e. 1979. historical aspects of swedish moose management. meddelelser fra norsk viltforskning 3(8):49-59. damas and smith company. 1983. wildlife mortality in transportation corridors in moose values – timmermann and rodgers alces vol. 41, 2005 108 canada’s national parks. impact and mitigation. consultants report to parks canada, ottawa, canada. danielsen, j. 2001. local community based moose management plans in norway. alces 37:55-60. del frate, g. g., and t. h. spraker. 1991. moose-vehicle interactions and an associated public awareness program on the kenai peninsula, alaska. alces 27:1-7. dewdney, s., and k. e. kidd. 1962. indian rock paintings of the great lakes. university of toronto press, toronto, ontario, canada. duwors, e., m. villeneuve, f. filion, r. reid, p. bouchard, d. legg, p. boxall, t. williamson, a. bath, and s. meis. 1999. the importance of nature to canadians: survey highlights. environment canada, ottawa, ontario, canada. eagan, m., p.logan, and e. duwors. 1989. the benefits of wildlife. ministry of environment, canadian wildlife service, ottawa, ontario, canada. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forest. silva fennica 36:57-67. eschholz, w., f. a. servello, k. s. rayond, and w. b. krohn. 1996. winter use of glyphosate-treated clearcuts by moose in maine. journal of wildlife management 60:764-769. faber, w. e. 1999. the 1998/99 moose harvest in sweden. the moose call 10:31. _____, and l. edenius. 1998. bark stripping by moose in commercial forests of fennoscandia a review. alces 34:261-268. feit, h. a. 1987. north american native hunting and management of moose populations. swedish wildlife research supplement 1:25-42. filion, f. l., s. w. james, j-l. ducharm, w. pepper, r. reid, p. boxall, and d. teillet. 1983. the importance of wildlife to canadians: highlights of the 1981 national survey. canadian wildlife service, environment canada, ottawa, ontario, canada. franzmann, a. w. 1996. alaska’s moose research center: the concept, construction and early history. the moose call 3:10-11. _____, and c. c. schwartz. 1983. management of north american moose populations. pages 517-522 in c. m. wemmer, editor. biology and management of the cervidae. smithsonian institution press, washington, d.c., usa. _____, and _____, editors. 1998. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. fraser, d. g. 1978. moose watching in sibley provincial park. ontario fish and wildlife review 17(4):13-18. _____, and h. hristienko. 1982. moosevehicle accidents in ontario: a repugnant solution. wildlife society bulletin 10:266-269. gasaway, w. 1975. moose antlers: how fast do they grow? alaska department of fish and game, fairbanks, alaska, usa. _____, r. o. stephenson, j. l. davis, p. e. k. sheperd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. gillespie, b. c. 1981. major fauna in the traditional economy. pages 15-18 in j. helm, editor. subarctic handbook of north american indians. volume 16. smithsonian institution press, washington, d.c., usa. groot-bruinderink, g. w. t. a., and e. hazebroek. 1996. ungulate traffic collisions in europe. conservation biology 10:1059-1067. alces vol. 41, 2005 timmermann and rodgers moose values 109 gundersen, h. 2003. vehicle collisions and wolf predation: challenges in the management of a migrating moose population in southeast norway. ph.d. dissertation, university of olso, olso, norway. _____, h. p. andreassen, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34:385-394. hallström, g. 1960. monumental art of northern sweden from the stone age. almqvist & wiksell, stockholm, sweden. hamilton, g. d. 1981. practical importance of moose and other wild foods to natives in a remote northern ontario community. alces 17:44-55. hänninen, p. 1994. koll på elgskadorna genom vettig skogsvård. jägaren 3:34-35. hansson, i., and g. malmfors. 1978. meat production from moose, alces alces (l). swedish journal of agricultural research 9:155-159. härkönen, s. 1999. moose hunting in finland. the moose call 10:7. _____, r. heikkilä, w. e. faber, and å. pehrson. 1998. the influence of silvicultural cleaning on moose browsing in young scots pine stands in finland. alces 34:409-422. heikkilä, r., and j. aarnio. 2001. forest owners as moose hunters in finland. alces 37:89-95. _____, and s. härkönen. 2000. thinning residues as a source of browse for moose in managed forests in finland. alces 36:85-92. hénault, m., l. bélanger, a. r. rodgers, g. redmond, k. morris, f. potvin, r. courtois, s. morel, and m. mongeon. 1999. moose and forest ecosystem management: the biggest beast but not the best. alces 35:213-225. hicks, a. c. 1993. using road kills as an index to moose population change. alces 29:243-248. hjeljord, o., and s. gronvold. 1988. glyphosate application in forest-ecological aspects. vi. browsing by moose (alces alces) in relation to chemical and mechanical brush control. scandinavian journal of forest research 3:115-121. hnilicka, p., and m. zornes. 1994. status and management of moose in wyoming. alces 30:101-107. hörnberg, s. 1995. moose density related to occurrence and consumption of different forage species in sweden. report 58. department of forest surveys, swedish university of agricultural sciences, umeå, sweden. humphrey, b. 2002. interesting moose facts from maine. the moose call 14:11. (idfg) idaho department of fish and game. 1990. species management plan 1991-1995: moose. idaho department of fish and game, boise, idaho, usa. jansson, s., e. b. lundberg, and u. bertilsson. 1989. hällristningar och hällmålningar i sverige. (rock carvings and rock paintings in sweden). bokförlaget forum, stockholm, sweden. jaren, v. 1992. monitoring norwegian moose populations for management purposes. alces supplement 1:105-111. _____, r. andersen, m. ulleberg, p. h. pedersen, and b. wiseth. 1991. moose-train collisions: the effects of vegetation removal with a cost-benefit analysis. alces 27:93-99. _____, a. b. leifseth, a. hoel, s. brainerd, t. borgos, p. h. pedersen, v. holthe, t. punsvik, b. kristiansen, j. hageland, and c. a. libach. 1995. management of cervids towards the year 2000: action plan, directorate of nature conservation, report 1995-1. directorate for nature management, moose values – timmermann and rodgers alces vol. 41, 2005 110 trondheim, norway. (in norwegian with english summary). jensen, b. 2002. summary of liquidation or replacement costs for big game species native to north dakota. the moose call 15:9-10. johansson, p-o., b. kriström, and l. mattsson. 1988. how is the willingness to pay for moose affected by the stock of moose? an empirical study of moose-hunters in the county of västerbotten. journal of environmental management 26:163-171. jordan, p. a., b. e. mclaren, and s. m. sell. 2000. a summary of research on moose and related ecological topics at isle royale, u.s.a. alces 36:233267. joyce, t. l., and s. p. mahoney. 2001. spatial and temporal distributions of moose-vehicle collisions in newfoundland. wildlife society bulletin 29:281-291. karns, p. d. 1998. population, distribution, density and trends. pages 125-140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. kastdalen, l., 1998. summary from a report: “the consequences for moose of building a new national airport in norway”. the moose call 7:13-16. kay, c. e. 1997. aboriginal overkill and the biogeography of moose in western north america. alces 33:141-164. kellert, s. r. 1980. american’s attitudes and knowledge of animals. transactions of the north american wildlife and natural resource conference 45:111-124. _____, 1987. the contributions of wildlife to human quality of life. pages 222-229 in d. j. decker and g. r. goff, editors. valuing wildlife: economic and social perspectives. westview press, boulder, colorado, usa. kennedy, e. r., and p. a. jordan. 1985. glyphosate and 2,4-d: the impact of two herbicides on moose browse in forest plantations. alces 21:149-160. krefting, l. w. 1951. what is the future of the isle royale moose herd? transactions of the north american wildlife conference 16: 461-472. _____. 1974. moose distribution and habitat selection in north central north america. naturaliste canadien 101:80-100. kriström, b. 1987. the value of a hunting permit under rationing; an application to moose hunting in sweden. report 58. department of forest economy, swedish university of agricultural sciences, s-901 83 umeå, sweden lacasse, m. 1986. un orignal, ca vaut combien? sentier chasse-pêche. juillet 1986:28-32. lavsund, s. 1981. moose as a problem in swedish forestry. alces 17:165-179. _____. 1987. moose relationships to forestry in finland, norway and sweden. swedish wildlife research, supplement 1:229-244. _____. 1989. älgskador vårt största skogsskyddsproblem f.n. k. skogs-o. lantbr. akad. tidskr. 128:111-116. _____, t. nygren, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39:109-130. _____, and f. sandegren. 1989. swedish moose management and harvest during the period 1964-1989. alces 25:58-62. _____, and _____. 1991. moose-vehicle relations in sweden: a review. alces 27:118-126. legg, d. 1995. the economic impact of hunting for large mammals in ontario, 1993. social and economic research alces vol. 41, 2005 timmermann and rodgers moose values 111 and analysis section, resource stewardship and development branch, ontario ministry of natural resources., peterborough, ontario, canada. _____, and m. kennedy. 2000. the economic impact of moose hunting in ontario, 1996. analysis and planning section, land use planning branch, ontario ministry of natural resources, peterborough, ontario, canada. lingard, m. 1997. aggressive silviculture program planned by abitibi-price in 1997. pages 29-30 in focus on forestry. robinson blackmore publication, st. john’s, newfoundland, canada. loomis, j. b., d. m. donnelly, c. f. sorg, and l. oldenburg. 1985. net economic value of hunting unique species in idaho: bighorn sheep, mountain goat, moose, and antelope. resource bulletin rm-10, u.s. forest service, rocky mountain forest and range experiment station, fort collins, colorado, usa. lynch, g. m., b. lajeunesse, j. willman, and e. telfer. 1995. moose weights and measurements from elk island national park, canada. alces 31:199-207. mattsson, l. 1990. moose management and economic value of hunting: toward bioeconomic analysis. scandinavian journal of forest research 5:575-581. _____, and b. kriström. 1987. älgens jaktvärde: en naturresursekonomisk analys baserad på 1985 års älgjakt i västerbotten län. arbetsrapport 60, sveriges landbruksuniversitet, institutionen for skogekonomi, umeå, sweden. mccaffery, k. r. 1973. road-kills show trends in wisconsin deer populations. j o u r n a l o f wi l d l i f e m a n a g e m e n t 37:212-216. mcdonald, m. g. 1991. moose movement and mortality associated with the glenn highway expansion, anchorage alaska. alces 27:208-219. mckercher, b. 1992. tourism as a conflicting land use: northern ontario’s outfitting industry. annals of tourism research 19:467-481. mclaren, b. e., s. p. mahoney, t. s. porter, and s. m. oosenbrug. 2000. spatial and temporal patterns of use by moose of pre-commercially thinned, naturallyregenerating stands of balsam fir in central newfoundland. forest ecology and management 133:179-196. _____, b. a. roberts, n. djan-chekar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40:45-59. mercer, w. e. 1995. moose management plan for newfoundland. report on file with wildlife division, newfoundland and labrador, st. john’s, newfoundland, canada. _____, and b. e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose. alces 38:123-141. messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75:478-488. midkiff, k. 2004. the meat you eat. how corporate farming has endangered america’s food supply. st. martan’s press, new york, new york, usa. modafferi, r. d. 1991. train moose-kill in alaska: characteristics and relationship with snowpack depth and moose distribution in lower susitna valley. alces 27:193-207. monska, l. 2001. moose husbandry at the columbus zoo: nutritional aspect. alces 37:35-41. morris, k., and k. elowe. 1993. the status of moose and their management in maine. alces 29:91-97. morriseau, n. 1977. legends of my people: the great ojibway. mcgraw-hill, moose values – timmermann and rodgers alces vol. 41, 2005 112 toronto, ontario, canada. myrberget, s. 1979. the norwegian moose population 1945-1977. meddeleslser fra norsk viltforskning 3(8):18-33. nepinak, h., and h. payne. 1988. hunting rights of indian people in manitoba: an historical overview and a contemporary explication toward enhanced conservation through joint management. alces 24:195-200. nilsson, j. 1987. effekter av vitstängsel. viltolyckor. nordiska trafiksäkerhetsrådet, rapport 45. linköping, sweden. oldemeyer, j. l., and w. l. regelin. 1987. forest succession, habitat management, and moose on the kenai national wildlife refuge. swedish wildlife research supplement 1:163-179. (omnr) ontario ministry of natural resources. 1980. moose management policy. queen’s printer for ontario, toronto, ontario, canada. _____. 1986. timber management guidelines for the protection of tourism values. ontario ministry of natural resources, toronto, ontario, canada. _____. 1988. timber management guidelines for the provision of moose habitat. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. _____. 1991. looking ahead: a wildlife strategy for ontario. ontario ministry of natural resources, wildlife policy branch, queens printer for ontario, toronto, ontario, canada. _____. 1999. a management framework for woodland caribou conservation in northwestern ontario. ms report, ontario ministry of natural resources, thunder bay, ontario, canada. _____. 2001. forest management guide for natural disturbance pattern emulation. version 3.1. ontario ministry of natural resources, queen’s printer for ontario, toronto, ontario, canada. _____. 2003. recovery strategy for the forest-dwelling woodland caribou (rangifer tarandus caribou) in ontario. ontario ministry of natural resources, peterborough, ontario, canada. oosenbrug, s. m., r. w. mcneily, e. w. mercer, and j. f. folinsbee. 1986. some aspects of moose-vehicle collisions in eastern newfoundland, 19731985. alces 22:377-393. _____, e. w. mercer, and s. h. ferguson. 1991. moose-vehicle collisions in newfoundland-management considerations for the 1990’s. alces 27:220-225. paine, r. t. 1988. food webs: road maps of interactions or grist for theoretical development? ecology 69:1648-1654. persson, i.-l., k. danell, and r. bergström. 2000. disturbance by large herbivores in boreal forests with special reference to moose. annales zoologici fennici 37:251-263. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. peterson, r. o. 1997. the crash of isle royale moose, 1996. the moose call 5:4-5. _____, and j. a. vucetich. 2002. ecological studies of wolves on isle royaleannual report 2001-2002. the moose call 15:13-16. _____, j. d. woolington, and t. n. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88. pilgrim, e. p. 2000. blood on the hills. flanker press, st. john’s, newfoundland, canada. pimlott, d. h. 1963. influence of deer and moose on boreal forest vegetation in two areas of eastern canada. transactions of the international union of game biologists 6:106-116. posner, s. d., and p. a. jordan. 2002. competitive effects on plantation white alces vol. 41, 2005 timmermann and rodgers moose values 113 spruce saplings from shrubs that are important browse for moose. forest science 48:283-289. racey, g. d., j. mcnicol, and h. r. timmermann. 1989. application of the moose and deer habitat guidelines on investment. pages 119131 in forest investment: a critical look. canadian forest research centre symposium proceedings. o-p-17. canadian forest service, sault ste. marie, ontario, canada. rajamäki, r., and j. mänttäri. 2002. hirvieläinonnettomuudet yleisillä teillä vuonna 2001. tiehallinto (finnish road administration), helsinki, finland. (in finnish). rattey, t., and n. e. turner. 1991. vehicle-moose accidents in newfoundland. journal of bone and joint surgery 73-a, 10:1487-1491. rausch, r. a. 1958. the problem of railroad-moose conflicts in the susitna valley. alaska department of fish and game. federal aid in wildlife research. final report. project w-3. job 1-4. juneau, alaska, usa. redmond, g., m. phillips, d. bryson, and j. noel. 2004. mitigating moose-vehicle accidents in northeast new brunswick: an experimental electrobraid fence project. final report. new brunswick department of transportation, planning and land management branch, p.o. box 6000, fredericton, new brunswick, canada. reeves, h. m., and r. e. mccabe. 1998. of moose and man. pages 1-75 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. regelin, l., and a. w. franzmann. 1998. past, present, and future moose management and research in alaska. alces 34:279-286. reid, r. 1997. the economic value of resident hunting in british columbia. 1995. wildlife branch, ministry of environment, lands and parks, victoria, british columbia, canada. renecker, l. a., r. j. hudson, and g. w. lynch. 1987. moose husbandry in alberta, canada. swedish wildlife research supplement 1:775-780. ripple, w. j., e. j. larsen, r. a. renkin, and d. w. smith. 2001. trophic cascades among wolves, elk, and aspen on yellowstone national park’s northern range. biological conservation 102:227-234. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17:357-364. rodgers, a. 2001. moose. world life library, voyageur press, stillwater, minnesota, usa. rollins, r. 1987. hunter satisfaction with the selective harvest system for moose in northern ontario. alces 23:181-193. _____, and l. romano. 1989. hunter satisfaction with the selective harvest system for moose management in ontario. wildlife society bulletin 17:470-475. romin, l. a., and j. a. bissonette. 1996. deer-vehicle collisions: status of state monitoring activities and mitigation efforts. wildlife society bulletin 24:276-283. ross, c. 1975. economic evaluation of resident big game hunting in saskatchewan. research, planning and policy branch, saskatchewan tourism and renewable resources, research report number 1. regina, saskatchewan, canada. _____, and h. j. paul. 1976. economic evaluation of non-resident big game hunting in saskatchewan. research planning and policy branch, sasmoose values – timmermann and rodgers alces vol. 41, 2005 114 katchewan tourism and renewable resources, research report number 2. regina, saskatchewan, canada. rowland, d. 1989. outdoor health and nutrition. eating wild game is no trivial pursuit. courier (spring):18. ruhr, r. d. c., and v. f. j. crichton. 1985. a methodology for evaluating the benefits of moose. alces 21:299-320. sanderson, k. 1983. wildlife road-kills and potential mitigation in alberta. environmental council of alberta report eca 83-st/1. edmonton, alberta, canada. sandgren, m. 1980. produktionsförluster och kvalitetsnedsättningar i en älgbetad tallkultur. examensarbete, inst f skogsskötsel 1980-5. sarker, r., and y. surry. 1998. economic value of big game hunting: the case of moose hunting in ontario. journal of forest economics 4:1-29-60. schwartz, c. c. 1992. techniques of moose husbandry in north america. alces supplement 1:177-192. _____, and b. bartley. 1991. reducing incidental moose mortality: considerations for management. alces 27:227231. _____, and a. w. franzmann. 1989. bears, wolves, moose and forest succession, some management considerations on the kenai peninsula, alaska. alces 25:1-10. _____, and k. hundertmark. 1994. the moose research center. the moose call 1:9-10. _____, _____, and t. h. spraker. 1992. an evaluation of selective moose harvest on the kenai peninsula, alaska. alces 28:1-13. _____, j. e. swenson, and s. d. miller. 2003. large carnivores, moose, and humans: a changing paradigm of predator management in the 21st century. alces 39:41-63. seiler, a. 1999. uppföljningsstude höga kusten: barriäreffekt pålg. pages 57-60 in b. iuell, editor. proceedings of the nordic conference on roads, road traffic and habitat fragmentation. statens vegvesen, oslo, norway. (in swedish). _____. 2003. the toll of the automobile: wildlife and roads in sweden. ph.d. thesis, swedish university of agricultural sciences, uppsala, sweden. s k o g s s t y r e l s e n . 2002. enkel älgbetningsinventering äbin. skogsstyrelsen http://www.svo.se. (in swedish). smith, c. a., j. b. faro, and n. c. steen. 1979. an evaluation of trophy moose management on the alaskan peninsula. proceedings of the north american moose conference and workshop 15:280-302. sødal, d. p. 1985. elg-økonomi. rapport fra et forprosjekt. institutt for skogøkonomi, rapport 1. norges landbrukshøyshøyskole, ås, norway. (in norwegian). _____. 1989. økonomisk verdsetting av elgjakt. (economic valuation of moose hunting). ph.d. thesis, department of forest economics, agricultural university of norway. ås, norway. (in norwegian with english summary). solberg, e. j., m. heim, and b-e. saether. 1998. moose harvest in norway. the moose call 7:3. _____, h. sand, j. linnell, s. brainerd, r. andersen, j. odden, h. bröseth, j. swenson, o. strand, and p. wabakken. 2003. store rovdyrs innvirking på hjorteviltet i norge: ökologiske prosesser og konsekvenser for jaktuttak og jaktutövelse. norwegian institute for nature research fagrapport 63. trondheim, norway. (in norwegian). solbraa, k. 1998. elg og skogsbruk, -biologi, økonomi, beite, taksering foralces vol. 41, 2005 timmermann and rodgers moose values 115 valtning. skogsbrukets kursinstitutt, biri, norway. steinhoff, h. w. 1978. big game values. pages 271-282 in j. l. schmidt and d. l. gilbert, editors. big game of north america. stackpole books, harrisburg, pennsylvania, usa. _____, r. g. walsh, t. j. peterle, and j. m. petulla. 1987. evolution of the valuation of wildlife. pages 34–48 in d. j. decker and g. r. goff, editors. valuing wildlife: economic and social perspectives. westview press, boulder, colorado, usa. stevens, j. 1971. sacred legends of the sandy lake cree. mclelland and stewart limited, toronto, ontario, canada. storaas, t., h. gundersen, h. henriksen, and h. p. andreassen. 2001. the economic value of moose in norway a review. alces 37:97-107. sullivan, t. l., and t. a. messmer. 2003. perceptions of deer-vehicle collision management by state wildlife agency and department of transportation administrators. wildlife society bulletin 31:163-173. sylvén, s. 1995. moose harvest strategy to maximize yield value for multiple goal management-a simulation study. agricultural systems 49:277-298. _____. 2003. management and regulated harvest of moose (alces alces) in sweden. ph.d. thesis, swedish university of agricultural sciences, uppsala, sweden. thompson, i. d. 1988. moose browsing damage in pre-commercially thinned balsam fir stands in central newfoundland. alces 28:56-61. _____, and w. j. curran. 1989. moose damage to pre-commercially thinned balsam fir stands: review of research and management implications. information report n-x-272. canadian forest service publication, st. john’s, newfoundland, canada. _____, and _____. 1993. a reexamination of moose damage to balsam fir-white birch forests in central newfoundland: 27 years later. canadian journal of forest research 23:1388-1395. _____, and r. w. stewart. 1998. management of moose habitat. pages 377-401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. timmermann, h. r. 1971. the antlers of the moose development related to age. ontario fish and wildlife review 10:11-18. _____. 1987. moose harvest strategies in north america. swedish wildlife research supplement 1:565-579. _____. 2003. the status and management of moose in north america — circa 2000-01. alces 39:131-151. _____, and m. e. buss. 1998. population and harvest management. pages 559-615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, r. gollat, and h. a. whitlaw. 2002. reviewing ontario’s moose management policy-1980-2000targets achieved, lessons learned. alces 38:11-45. todesco, c. 2004. illegal moose kill in northeastern ontario: 1997-2002. alces 40:145-159. umali, g. m. 1997. licenced-based valuation method: a travel cost approach to valuing moose hunting. analysis and planning services section, land use planning branch, ontario ministry of natural resources, peterborough, ontario, canada. vecellio, g. m., r. d. deblinger, and j. moose values – timmermann and rodgers alces vol. 41, 2005 116 e. cardoza. 1993. status and management of moose in massachusetts. alces 29:1-7. veitch, a., and e. simmons. 2002. mackenzie mountain non-resident and nonresident alien hunter harvest summary. report number 137. department of resources, wildlife and economic development, norman wells, northwest territories, canada. wabakken, p., h. sand, o. liberg, and a. bjärvall. 2001. the recovery, distribution, and population dynamics of wolves on the scandinavian peninsula, 1978-1998. canadian journal of zoology 79:710-725. webster. 1967. webster’s seventh new collegiate dictionary. g. and c. merriam company, springfield, massachusetts, usa. white, p. j., r. a. garrott, and l. l. eberhardt. 2003. evaluating the consequences of wolf recovery on northern yellowstone elk. unpublished report dated october 2003, on file at the yellowstone center for resources, yellowstone national park, mammoth, wyoming, usa. wicks, g. 2002. moose exclusion with electrobraid tm fence, 2000-2001. the moose call 14:9. wolfe, m. l. 1987. an overview of the socioeconomics of moose in north america. swedish wildlife r e s e a r c h s u p p l e m e n t 1 : 6 5 9 6 7 5 . yu k o n re n e wa b l e re s o u r c e s. 1996. moose management guidelines. fish and wildlife branch, department of renewable resources, whitehorse, yukon territory, canada. _____. 1999. yukon moose. fish and wildlife branch, department of renewable resources, whitehorse, yukon territory, canada. appendix a sampling of moose related items or products sold commercially. antique storeflying moose antique mall, wichita, kansas, usa auto dealersmoose motors incorporated —112 lindgren road west, huntsville ontario, canada, p1h 1yz bars and restaurantsl o o s e m o o s e , to r o n t o , o n t a r i o , canada lonely moose bar, anchorage, alaska, usa moose tooth pub & pizza, anchorage, alaska, usa the moose is loose, sterling, alaska, usa moosquitos bar, sterling, alaska, usa moose’s saloon— kalispell, montana, usa moose winooski’s— brantford & kitchener, ontario canada moose delaney’s sports grill— 3 cann street, huntsville ontario, canada. p1h 1h3 r u s t i c m o o s e — k e t c h u m , i d a h o , usa moosehead lounge, glenwood, newfoundland, canada. bedroom itemsberber throw pillow (with bull moose illustration)— woolrich company— $19.99 us flannel sheet set-(illustrated with moose, geese, & trees)— woolrich company — $24.99 us beermoosehead beer— moosehead breweralces vol. 41, 2005 timmermann and rodgers moose values 117 ies, st. john new brunswick canada moose drool beer— big sky brewing co., missoula montana, usa moose tooth beer— moose tooth brewing co., anchorage alaska, usa candymaple moose pops, bread & chocolate incorporated, wells, vermont, usa childrens booksdeneki, an alaskan moose. by w.d. berry. 1965. macmillan publishing company, new york. usa thidwick the moose— dr. seuss moose for kids— j. fair a moose for jessica—1987— p. a wakefield and l. carrara, e.p. dutton, new york mickey moose by bob reese — aro publishing, provo, utah 1986 chocolatesharry and davids moose munch bar— www.harryanddavid.com the chocolate moose — 2839 bathurst street, north york, ontario, canada m6b 3a4 moose droppings— chocolates shaped like moose droppings christmas ornaments 30 inch holiday standing moose— safeways ($12.99 us) — nov. 11/ 03 flyer coffee table booksmoose: giant of the northern forest. bill silliker, jr., key porter books, 1998; $28.95 cnd (hardcover), $19.95 us (softcover) canning moose, by richard e. mccabe. rusty rock east press, p.o. box 34646, washington, d.c. 20034. usa— $10.95 + $1.50 shipping (us) moose by daniel wood., whitecap books limited., 351 lynn avenue, north vancouver, b.c., canada v7j 2c4 wild moose country, by paul i.v. strong. northwood press, incorporated. minnetonka, minnesota., usa ($39.00 us; $55.00 cnd— hardcover or $19.95 us & $27.95 cnd softcover moose, by art rodgers 2001, world life library, voyageur press, stillwater, minnesota, usa welcome to the world of moose. by diane swanson. 27pp. ($6.95 cnd, $5.95 us) compact discsmoose music— v. crichton, 1046, mcivor avenue, winnipeg, manitoba, canada, r2g 2j9 computer consultantsmoose eagle computers — 416-422 0505 cookbooksmoosewood restaurant (new recipes) — 1996 vegetable kingdom, incorporated moose in the pot— tim lundt, matsu alternative school, 1775 west parks highway, wasilla, alaska, usa low bush moose and other alaskan recipes 1978. alaska northwest publishing company the moose— national meat institute, montreal quebec, canada. 1970. fruits & vegetablesr e d m o o s e — g o u r m e t to m a toes on the vine—great northern hydroponics, ruthven, ontario, canada gift shopsblue moose— grand marais, minnesota, usa the purple moose—skagway, alaska, usa moose values – timmermann and rodgers alces vol. 41, 2005 118 moose creek antler lighting and decor— libby, montana, usa moosetrack quilts — whitefish, montana, usa mostly moose’s— kalispell, montana, usa moose crossing, incorporated, marion, montana, usa golf coursemoose run golf course, anchorage, alaska, usa hockey teamsmanitoba moose, american hockey league, winnipeg, manitoba, canada halifax mooseheads, halifax, nova scotia, canada insignias, emblems, coats of arms, state sealshudsons bay company coats of arms — ontario, canada, maine, usa, michigan, usa hawker hurricane fighter squadron #242 canadian & 503rd bombardment — ww 2 royal canadian airforce squadron # 419, badge: moose attacking, motto: moosea aswayita (beware of the moose), authority: king george 6th, june 1944 community crests— ft. resolution, ft. simpson, kakisa, ft. simpson, northwest territories, canada internet providermoose web corporation, kalispell, montana, usa jewelery-moose beads— d a l e p e t e r s o n , l i b b y, m t, u s a limited edition printsrobert batemanautumn overture— bull moose1981 stefen lyman1988 location namesmoosejaw, saskatchewan, canada, mooseonee, ontario, canada, moosomin saskatchewan, canada, moose factory, ontario, canada, moose lake, minnesota, usa, moosic village, 6 miles south of scranton pennsylvania, usa. moose pass, alaska, moose, wyoming, usa, mooseup, connecticut, usa, moosehide, klondike, yukon, moosehorn, manitoba, canada, moose creek, ontario, canada, moosejaw creek, saskatchewan canada 79 based on moose in maine, usa including moosehead lake 81 based on moose in alaska mascotsmoose mascot— seattle mariners baseball team, seattle, washington, usa moose marketing— a samplingwww.mooseworld.com moose visitor centregould co, usa— co state parks — the moose viewing capital of co motelsthe moose head inn, kenosee lake sask., canada moose river rv, sterling, alaska, usa moose lane b&b, anchorage, alaska, usa moose hollow, b&b, soldotna, alaska, usa moose creek lodge, soldotna, alaska, usa moose motel 226 highway #11, smooth rock falls, ontario, canada, p0l 2b0 movies brother bear— walt disney— 2003, rutt and tuke— canadian moose brothers, alces vol. 41, 2005 timmermann and rodgers moose values 119 inspired by bob & doug mckenzie “the moose are hilarious, and phil colin’s music is terrific!”— leonard matlin, maclean’s november, 24, 2003:66 ‘rocky and bullwinkle’ national animalnorway (storaas et al. 2001) naval shipmoose— commissioned quebec city, sept. 8th, 1939 & assigned to halifax ns local defence force until may 1942, transferred to sydney & employed as a training ship until july 1945 organizationsloyal order of moose— an international fraternal & benevolent organization north american moose foundation— (610 w. custer, po box 30, mackay, idaho, usa 83251 outdoor clothingmoose creek brand, usa. (made in china)— shirts, vests, jackets shirts @ moose figures— cabella’s & l.l. bean, coldwater creek, abercrombie’s outdoor equipmentthe moose hunter & outfitter. rr# 2 hwy 11-17, thunder bay ontario, canada outfittersmoose point camps — portage maine, usa pastam o o s e p a s t a ( ’ p â t e s d ” o r i g n a l ) — gourmet du village, morin heights canada, jor 1ho politicsbull moose progressive party— teddy roosevelt’s unsuccessful bid for a 3rd term 1912 postage stamps— canada post’s new $5.00 stamp issued december 19, 2003, by wildlife artist david preston-smith radio stationmoose radio muskoka— 50 balls drive, bracebridge ontario, canada, p1l 1ti real estate brokersmoose realty limited, 877 jane street, york, ontario, canada, m6n 4c4 soapmoose drool soap—montana, usa moose spit soap— british columbia, canada songgotta get me moose, b'y— written by kevin blackmore, wayne chaulk and ray johnson state animalmaine (morris and elowe (1993) state parksmoose lake state park, moose lake, minnesota, usa moose river resort, sterling, alaska, usa tattoosrub-on moose tattoos theater companymooseberry theater company, moosomin, saskatchewan, canadatourist advertizing“moose, mountains and mounties”— tourism canada’s marketing focus— early 1980s (noto 1985) moose values – timmermann and rodgers alces vol. 41, 2005 120 “moose in the city”— toronto, ontario, canada’s millennium event (2000)—100’s of life-sized moose sculptures gracing city streets (the moose call 2001, 12:25) toys / ornamentschocolate moose— (fun puppet & animated plush— chocolate scented)—($9.99 us)— dandie international limited, 106 harbor drive, jersey city usa—walgreens.com— nov. 2003 2-piece pet and people moose antlers ($9.99 us)— petsmart— moose bookends— cabellas videosmoose hunt, a guide to success— interesting services incorporated. 1989 moose close-up— v. crichton, 1046, mcivor avenue, winnipeg, manitoba, canada in the company of moose. gisele benoit,films franc-sud—cbc, quebec, canada the high season of the moose wood carvingthe canadian carver, hwy. # 17, pancake bay, lake superior, ontario, canada << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice alces16_338.pdf alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alces vol. 16, 1980 alcessupp1_22.pdf i in memoriam dr. vince crichton (doc moose), 1942–2020 it was a very sad day for alces and the north american moose conference and workshop group when we learned that vince, a near-inaugural member, a regular scientific contributor, a relentless inspirational leader, and treasured friend of so many, was taken from us at 78 years old on december 3, 2020. his presence at every meeting since 1972 was hugely important to the scientific and social development of our association. first and foremost, he will always be the beloved husband of kim, dad to scott (vita) and susan (craig), and grampy to julia. dr. vincent frederick joseph crichton was born november 7, 1942 in the small northern town of chapleau, ontario. his love of the outdoors and wildlife undoubtedly developed during his early years travelling in the bush with his father vince crichton sr. who was fish and wildlife supervisor of the chapleau region. vince earned his bachelor and masters of science degrees at the university of manitoba and his doctorate at the university of guelph in the field of wildlife diseases. his 40-year-career with the province of manitoba began in 1972 as eastern region wildlife biologist and culminated in 2012 as manager of game, fur and problem wildlife, manitoba conservation, wildlife and ecosystem branch. an annual opportunity to exchange scientific information with colleagues and a journal in which to publish it, doesn’t happen on its own, especially in the absence of a formal organization. but for moose biologists, this was the case for more than 50 years, in large part due to vince crichton’s influence. in the early days of alces and the namcw, vince was a key member of a small steering group, affectionately called the “moose mafia”, who never allowed the ball to be dropped. and he continued until his last days to provide that key leadership. he was a regular scientific contributor to alces, an issue editor and a permanent associate editor. vince originated and co-sponsored the antler dmb award carved each year since 1981 by tom copper, ak, and now presented 36 times. he was instrumental with others in producing the acclaimed “ecology and management of the north american moose” in which he has two chapters. as an aside, he initiated and coedited the “moose call” newsletter intending to bring moose research and interesting stories to readers who might otherwise never pick up a journal. assisted by kim, they produced 27 issues. vince organized and hosted three namcws, one of which (50th), was combined with the 8th international moose symposium. in 1988 he received the distinguished moose biologist award, and in 2016 at the 50th, was presented with the first namcw professional ii commitment and appreciation award. this was his 45th consecutive annual meeting, including all 8 international symposia, the only “mooser” to have done so. vince served as associate editor of the wildlife society bulletin and was the canadian vice-president of the north american moose foundation. for these reasons he will be remembered by many colleagues from around the northern hemisphere wherever moose roam. vince’s unique character and persuasiveness came from his never failing pursuit of good wildlife management, not to mention his tall stature. one might presume that his outspokenness survived in the civil service only because of the principled and absolute conviction of his positions. his objective he said was to get people “into the same canoe, paddling in the same direction”. dr. crichton was given an award of merit from the province of manitoba for 40 years of service and was the recipient of the 2014 conservation award from the manitoba chapter of the wildlife society for support of the conservation and management of wildlife and their habitats. in retirement he became an expert on the spread of cwd in north america and continued to pursue his passion for moose and caribou management as a consultant, public speaker, environmentalist, conservationist, hunter and writer. he was always ready and willing to give a talk on his beloved moose, whether to an international scientific conference, the gynecological association of montreal, the idea city symposium, toronto, his local university, the local chapter of the wildlife society, or to a class of local elementary students. he was a recorder for the boone and crockett club for many decades, measuring trophy heads at big game nights all over manitoba, and along with others, founded the manitoba big game trophy association. in november while hospitalized, vince was recognized in the manitoba legislative assembly “for being one of the first advocates for co-management of moose by first nations and government and for his passion, dedication and commitment to moose management provincially, and around the world”; he was awarded the first ever honorary manitoba moose hunting license which brought a smile to his face. riding mountain national park was vince’s second home. in the spring and fall he iii spent much of his spare time, cameras in hand and infant grand-daughter julia in tow behind his bike searching for his “rubber-nosed swamp donkeys”. he was an accomplished photographer and was involved in the production of several videos for naturalists and hunters. on his 78th birthday, the cbc aired the documentary “giants of the boreal forest” documenting his never ending passion for moose. his work has been featured on discovery channel (champions of the wild) and animal planet (the man who would be moose). his private entrepreneurial interests ranged from telonics canada, to speculating on the price of winter-dried moose pellets; that is until the bottom fell out of the souvenir stick-creature market leaving him with a garage full of pellets needing to be turned into his garden. to paraphrase an old physicist (newton), if we have all moved a little further in our views of moose management and of how to communicate those ideas to the public, in part, it is because we have sat with a genuine giant. vince’s absence from future meetings of the north american moose conference and workshop will leave a hole in our hearts but he will always be there to remind us of the much needed moose research left to be accomplished and shared. dewdney (1957) sketch of pictographs, darkly lake, quetico pk. alces vol. 34 (1), (1998) ii foreword international moose symposiums have been held every 5-10 years since the first in québec city, québec, canada in 1973. subsequent meetings occurred in uppsala, sweden in 1984, syktyvkar, russia in 1990, fairbanks, alaska, usa in 1997, and hafjell, norway in 2002. russia became the first country to repeat as host when the 6th international moose symposium was held in yakutsk, russia where moose biologists from throughout the world gathered on 13-23 august 2008. although each volume of alces is typically international in content, an international symposium provides an opportunity to focus a substantial portion of a volume on the biology and management of moose in the host country. thus, volume 45 of alces features 8 papers that provide regional and national perspectives on the history and status of moose and moose management in russia and the former soviet union. both contemporary and well known russian moose biologists contributed to this volume, and they have provided a unique collection of papers concerning regional histories and management of moose. the publication of soviet union-authored papers from the 3rd international moose symposium at syktyvkar 18 years ago was fraught with translation difficulties that reflected, in part, political circumstances that prevented adequate communication among scientists. despite this, alces editors arthur rodgers and kris hundertmark strove to ensure that the work of such authors would be published in alces, and eventually supplement no. 2 contained 30 related papers. the relative ease of submitting, translating, editing, and publishing the current set of russian-authored papers not only demonstrates improved global communication among moose biologists and researchers, it reaffirms the critical importance and advantages of our international cooperation. dr. vince crichton was instrumental in establishing communication with many authors and served as the initial “clearinghouse” for many papers submitted for volume 45. although i was unable to contact each author individually (some papers were submitted by colleagues), each paper was reviewed by a russian author to approve the final edited version. certain papers underwent the standard review process by peer scientists, but i handled singly most papers that focused on descriptive history and status of moose. i particularly want to thank dr. leonid baskin and dr. taras sipko for their help in contacting other authors and checking the accuracy of my translations during the editing process. volume 45 is organized such that the first 8 papers are history-status-review papers about eurasian moose from authors of russia and ukraine. the remaining papers were authored by international researchers who traveled and participated in the 6th international moose symposium. the editorial staff believes that volume 45 truly reflects alces as a premier international scientific journal dedicated to the research and management of moose. peter j. pekins, chief editor editorial policy alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented at the annual north american moose conference and workshop, but works may be submitted directly to the editors at any time. reviewers judge submitted manuscripts on data originality, ideas, analyses, interpretation, accuracy, conciseness, clarity, appropriate subject matter, and on their contribution to existing knowledge. page charges current policies and charges are explained in a covering letter and invoice sent to authors with galley proofs. manuscript preparation authors should follow “manuscript guidelines for contributors to alces”, by rodgers et al. appearing in alces, vol. 34 (1): 1998 (available from the co-editors and associate editors). updates are posted on the alces web page; http://alcesjournal.org/ publicdocs/manuscriptguidelines.pdf. copy – please provide an electronic copy of the manuscript in ms word to the submissions editor. this copy should maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. double-space and left-justify all text. except for the first page, number all pages consecutively, including tables and figure captions. revisions should be handled similarly. corresponding author do not use a title page. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlespaced in the upper left corner of the first page. if available, the author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (<10 words) begins left justified on the next line. type the title in upper-case bold letters. do not use abbreviations or scientific names in the title. abstract & key words following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. do not use abbreviations or literature citations. type alces vol. 00: 000 000 (0000), right justified on the line following the abstract. after leaving a single blank line, provide 6-12 key words in alphabetical order. footnotes -use only in tables and at the bottom of the first page to provide the present address of an author when it differs from the address at the time of the study. style -accompany the first mention of a common name with its scientific name. do not use scientific names for the names of domesticated animals or cultivated plants. use syst me international d’unités (si) units and symbols. use digits for numbers unless the number is the first word of a sentence, in which case it is spelled out. italics should only be used in the text for scientific names and statistical symbols. use the name-and-year system to cite published literature. cite references chronologically in the text. references – use large and small capitals for author’s last names and initials. do not use any abbreviations in the references. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. titles and all parts of tables must be typed double-spaced. tables must be constructed to fit the width of the page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use numerical superscripts to identify footnotes or asterisks for probabilities. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table columns must be generated with tab settings or a table editor. do not use spaces (i.e., the space bar). illustrations type figure captions on a separate page. identify each illustration by printing the author’s name and the figure number on the back in soft pencil. if necessary, also indicate the orientation of the illustration on the back. each illustration (either a photograph or line-drawn figure), must be of professional graphics quality, and reduced to fit into the area of either 1 (67 mm) or 2 (138 mm) columns of text by the author(s). letters and numbers on reduced figures must remain legible and be no less than 1.5 mm high after reduction. the same size and font of lettering should be used for all figures in the manuscript. photographs must be of high contrast and printed with a matte finish. typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original electronic graphics files or bitmap images (preferably as tagged image file format files) in an ibm-compatible format on 9-cm (3.5-inch) diskette or cd-rom. please submit manuscripts online at: http://alcesjournal.org if a problem is encountered, please contact: roy rea, submissions editor natural resources and environmental studies institute university of northern british columbia 3333 university way prince george, british columbia canada v2n 4z9 e-mail: reav@unbc.ca telephone: (250) 960 5833 instructions for contributors to alces f:\alces\supp2\pagema~1\rus 17s alces suppl. 2, 2002 lopatin and rosolovsky – leslie matrix analyses 77 evaluation of the state and productivity of moose populations using leslie matrix analyses v. n. lopatin and s. v. rosolovsky institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: we examined the use of a leslie matrix analysis for estimating moose (alces alces) population parameters and allowable harvests from a moose population near leningrad, russia, during 1959 – 1975. leslie matrix analysis indicated that moose fecundity and mortality exhibited cyclic fluctuations. alces supplement 2: 77-80 (2002) key words: alces, fecundity, harvests, leningrad, leslie matrix, mortality, populations, productivity evaluation of population status is a traditional problem of modern ecology. closely associated with it are the practical problems of determining management strategy. many population models have been used to mimic populations, but the question remains if these models are representative of nature. many quantitative methods have been used, but solution of many mathematical and ecological problems necessitates expansion of the traditional methods of examining ecological problems (watt 1971, jeffers 1981). one method of expressing population dynamics is through the use of a leslie matrix model (leslie 1945), which considers age–specific birth and survival rates. rusakov (1979) used a leslie matrix model to examine a moose population in northwestern russia. also, peterson (1977) used a leslie matrix model to examine moose populations on isle royale, michigan, usa. based on these and similar analyses (aivazyan 1968, pesaran and slater 1984), we refined these methods for this study as provided by lopatin and rosolovsky (1988). these modifications result in rates of fecundity and mortality of individuals dependent on changes in density. for example, knowing the fecundity and death rate, one can determine the relative change in numbers over time and can estimate death rate. the purpose of this paper is to examine the dynamics of a moose population near leningrad, russia for the period 1959 – 1975. methods we used a leslie matrix analysis to examine moose population fluctuations within the leningrad region, russia, during 1959 – 1975 (rusakov 1979). under natural conditions, fecundity rate is often difficult to obtain under field conditions. often, field managers use the average number of calves per cow as an estimate of fecundity rate. these ratios can be represented as number of calves per female (x axis) and death rate (fig. 1). this indicates that population numbers are more sensitive to changes in death rate than fecundity. if the numbers of calves per female and total number in the moose population are determined accurately, equations derived from fig. 1 can be helpful in managing the moose population. leslie matrix analyses – lopatin and rosolovsky alces suppl. 2, 2002 80 for evaluation of the state of populations. pages 24–39 in population studies of animals in reserves. nauka, moscow, russia. (in russian). pesaran, m., and l. slater. 1984. dynamic regressions: theory and algorithms. finances and statistics. moscow, russia. (in russian). peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. u.s. national park service science monograph 11:1–200. rusakov, o. s. 1979. moose. pages 174– 271 in ungulates in northwestern ussr. nauka, leningrad, russia. (in russian). smith, d. m. 1975. models in ecology. mir, moscow, russia. (in russian). watt, c. w. 1971. ecology and management of resources. mir, moscow, russia. (in russian). alcessupp1_77.pdf f:\alces\supp2\pagema~1\rus 11s alces suppl. 2, 2002 gaidar et. al. collective moose hunting 53 efficiency of collective moose hunting in a forest-taiga zone of russia alexander a. gaidar, nikolai n. grakov, and b. m. zhitkov all-union research institute of game management and fur farming of the ussr centrosoyuz, kirov, russia abstract: we discuss results of hunting moose by coordinated drives by various-sized hunting groups in the kirov region. we report the optimum size of a hunting brigade, its advantages over hunting by individuals or by small groups, and the opportunity for selective harvest by sex and age classes. alces supplement 2: 53-55 (2002) key words: moose hunting, forest-taiga, hunting brigade, hunting season, licensed harvest, selective harvest in the forest-taiga zone, the major problem with moose management is the selective harvest of the optimum number of moose rather than the population as a whole. vast forested hunting grounds in the northern regions of russia and, especially, roadless portions of siberia make it extremely difficult to manage moose for sustained yield. access by helicopters to remote moose hunting grounds has been considered in the press by conservation and game management organizations, but is cost-prohibitive. the kirov region is situated in central and southern taiga and broad-leaved forest zones. only small islands of forests occur in southern portions of the kirov. moose hunting is more successful there than in the northern areas, in part, because of better habitat and forest roads. in the northern regions, moose numbers are greater, but access is difficult and harvest quotas are not achieved. in the forest and forest-taiga zone of russia, moose hunting is one of the most prestigious and favorite activities. besides the sporting interest, hunting provides an opportunity to make money and harvest quality meat. moose are hunted under sporting and commercial licenses. a sporting license costs 150 rubles (1 us$ ≈ 29 russian rubles) for hunting a bellowing male and 75 rubles for any moose after 1 october. a commercial license costs 40 rubles for an adult moose and 20 rubles for a calf. all age groups of animals are hunted and calves compose 20% of the harvest. meat of moose taken under sporting licenses belongs to the hunter. under a commercial license, moose meat is delivered to a trading network at a price of 1.5– 1.7 rubles per kg. in certain regions, a hunter may sell meat at the retail price of 3.5 rubles per kg. those terms are very attractive to hunters, especially for urban ones because meat is scarce in russia. a single hunter or small groups of 2–3 hunters get 1–2 licenses for moose hunting. it appears that such an approach makes moose hunting accessible to most hunters but leads to undesirable harvest selection. most hunters with 1–2 licenses attempt to take large animals, especially healthy adult females. if there is no license to take calves, then 1 or even 2 orphaned calves may not survive. glushkov (1985) estimated that every year after a hunting season in the kirov, at least collective moose hunting -gaidar et. al. alces suppl. 2, 2002 54 400 moose calves were orphaned and most had died by the end of winter. hunters use private or rented vehicles for transportation to moose hunting areas. all-terrain vehicles are necessary because it is often impossible to reach hunting grounds by car. hunters usually use vehicles like “niva,” “luaz,” “uaz,” or “gaz66” and “ural” for brigades. when traveling along snowy areas, hunters use “buran” and “icar” snowmobiles. study area our study area was located at the research-experimental hunting grounds (65,800 ha) of the all-union research institute of game management and fur farming in kirov. there are 60 members of our institute’s hunters collective. during the hunting season, an average of 300–400 moose inhabit our hunting ground. an average of 61.7 animals are taken per season; approximately 1 moose per 1,000 ha. on our hunting ground there are 40,000 ha of forested areas suitable as moose habitat and approximately 5,000 ha of brushwoods in the river flood plain. a forest area is divided into quarters of 1 x 2 and 2 x 2 km in size. the hunting grounds are crossed by 2 highways: one of them for timber trucks is covered with ferroconcrete slabs, the other one a partially asphalted earth road that becomes almost impassable in autumn. dirt roads and roads for timber trucks that branch from these highways become impassable for motor transport in deep snow. hunting seasons typically ran from 1 october to 15 january (188 days). over 9 years, 556 moose were taken (range 46–80 per year). hunters were transported by a “gaz-66” truck and by tractor. they hunted in 1–2 brigades, which ranged in size from 9 to 55 persons. methods hunting is carried out on permanent sites. most hunters are set up as shooters and several people (4–6) drive moose to the shooters. as a rule, shooting lines are invariable. hunters use smooth-bore guns or carbines of 7.62 and 9 mm caliber. shooters with guns stand near overgrown moose paths with little field of vision. hunters with carbines hunt in open sites, clearings, glades, small meadows, and fields. they may shoot only in the fixed sectors following strict accident prevention rules. moose drivers wear orange waistcoats. of course, not every drive is successful. sometimes there are no moose. at other times, moose pass through open areas in the drivers’ chain, rush between shooters, or run where there are no hunters. in view of this, it seems that the larger the brigade, the greater the probability of shooting a moose. however, the best results were in brigades of 9–15 (an average of 12) hunters. when hunter numbers approached the maximum, moose harvest fell 1.2–1.5 times because a small brigade is more mobile and can make more drives per day. control of a large brigade is a problem and takes more time for every drive because the size of the area covered increases. the advantages of dressing and loading harvested animals do not compensate for lost hunting time. since daylight is limited, hunters in smaller brigades may use twilight hours for dressing, loading, and transporting carcasses. the highest number of moose is taken in november (198 moose per 38 days of hunting). it takes 11 man-days to shoot 1 animal. although the indices of man-days for taking moose do not greatly differ during the other months (13.3–14.1), november is the preferred month for hunting moose in the kirov region. shallow snow depth and frozen ground permit hunters to use transport vehicles successfully. in october huntalces suppl. 2, 2002 gaidar et. al. collective moose hunting 55 ing is hampered by the lack of moose tracks, and in december to january by deep snow, hard frosts, and disturbance from previous hunting. discussion a brigade is preferred over individual hunters or hunting by small groups in the forest-taiga part of the country. using a brigade makes it possible to join the hunters’ collective, strengthens cooperative spirit among hunters in the forest, serves to acquire hunting skills, regulates the harvest of age and sex groups, and makes renting transport vehicles easier. with this style of hunting, we harvest 22–32% calves, compared to a 20% average in the kirov region. that permits us to maintain post-harvest moose numbers at a level of 300–400 individuals and a steady high yield of moose per area. references glushkov, v. m. 1985. moose population management: biological prerequisites and practical possibilities. pages 5-13 in management of wild ungulate populations. collected papers of rsfsr cnil glavokhota, moscow, russia. (in russian). alces 48 (2012) contents in memoriam: harold greenfield cumming ............................................... i using pelvis morphology to identify sex in moose skeletal remains ............................................. jason a. duetsch and rolf o. peterson 1 seasonal variation of phenols, nitrogen, fiber, and in vitro digestibility in swedish moose .................................................................... ................ r. thomas palo, peter a. jordan, åke pehrson and hans staaland 7 effects of essential oils on the feeding choice by moose ........... ...................................................................... sabine edlich and caroline stolter 17 distribution and prevalence of elaeophora schneideri in moose in wyoming ............. john c. henningsen, amy l. williams, cynthia m. tate, steve a. kilpatrick, and w. david walter 35 effective temperature differences among cover types in northeast minnesota ....................... amanda m. mcgraw, ron moen, and lance overland 45 ecothermic responses of moose (alces alces) to thermo regulatory stress on mainland nova scotia ......... hugh g. broders, andrea b. coombs, and j. r. mccarron 53 history and status of moose in oregon .............. patrick e. matthews 63 potvin double-count aerial surveys in new brunswick: are results reliable for moose? .................... roderick e. cumberland 67 comparing stratification schemes for aerial moose surveys .... ..................................................................... john r. fieberg and mark s. lenarz 79 visibility of moose in a temperate rainforest ..................................... ......................................... susan a. oehlers, r. terry bowyer, falk huettmann, david k. person, and winifred b. kessler 89 age, sex, and seasonal differentials of carcass weights of moose from the central interior of british columbia: a comparative analysis ........................ daniel a. aitken, kenneth n. child, roy v. rea, and olav g. hjeljord 105 perceptions of moose-human conflicts in an urban environment ..................................... alaina marie h. mcdonald, roy v. rea, and gayle hesse 123 (continued on inside back cover) 46th north american moose conference and workshop ................ 131 previous meeting sites................................................................................. 133 distinguished moose biologist kjell danell .................................... 135 distinguished moose biologist past recipients............................... 136 distinguished moose biologist award criteria ............................... 138 editorial review committee........................................................................ 139 additional copies additional copies and back issues of alces (issn 0835-5851 called proceedings of the north american moose conference and workshop up to vol. 16, 1980), most international moose symposia, and special symposia can be purchased online through the lakehead university alumni bookstore http://bookstore.lakeheadu.ca/esolution/course.php. all past publications are priced at cdn or us $49.99 each plus applicable tax (gst or hst). price includes mailing and handling costs. prices are subject to change. for more information contact the alces business editor. make cheques, money orders or purchase orders payable to lakehead university bookstore. acknowledgements brooke pilley worked long hours formatting and typesetting manuscripts. alces home page further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our website: http://bolt.lakeheadu.ca/~alceswww/alces.html 1 in memoriam victor van ballenberghe, 1943–2022 victor van ballenberghe was born in bayshore, long island, new york and grew up on a dairy farm in upstate new york. he received a biology degree from state university of new york (suny) at oneonta and then entered the university of minnesota in 1967 where he earned ms and phd degrees. vic’s phd (1972) research focused on wolves as part of a larger moose study in northeastern minnesota. his research occurred in the early days of radiotelemetry and was of the first to employ the pioneering technology. vic’s first job was as an extension trapper at south dakota state university. two years later, he was hired by the alaska department of fish & game to investigate the effect of the trans-alaska pipeline on moose migration. in 1980, he joined the research branch of the u.s. forest service and conducted moose and wolf research in denali national park and the copper river delta. he initiated long-term, well known investigations into moose biology and behavior that extended over 40 years. a truly unique aspect of this research was vic’s documentation of the ecology, breeding activity, and mortality across the lifespan of multiple bull moose. although vic retired in 2000, he continued these studies until parkinson’s disease limited his activity. vic was a “boots on the ground” field biologist and believed that one had to spend time in the field and patiently observe animals in their natural habitat to truly understand selection processes affecting behavior and survival. he was well known for his consummate field skills and knowledge of natural history. vic was a collaborative researcher, advising and supervising numerous graduate students including those researching moose and wolves in the copper river and denali. vic authored and co-authored over one hundred technical journal articles, book chapters, and symposium papers. he published many popular articles and wrote numerous newspaper opinion pieces on controversial wildlife management issues. he worked with multiple film crews including the bbc, national geographic, and animal planet, while accommodating national park service naturalists, private photographers, and so many others seeking his expertise and commentary. his exemplary professional efforts and leadership were recognized by his receiving the distinguished moose biologist award (1996) from the international alces working group. vic was appointed to the alaska board of game by governor bill sheffield in 1985, serving one full term and two subsequent partial terms in 1996 and 2002. he was a strong 2 advocate of scientific, evidence-based wildlife management, including support for bears and wolves. fittingly, his last publication in alces addressed large carnivore management in alaska, advocating for stronger scientific versus political approaches. vic was an avid and extremely skilled photographer. his beautiful book in the company of moose (2004) contains 120 photographs from denali national park and elsewhere, set to vic’s descriptive and passionate writing portraying the year-round ecology and behavior of moose. the final chapter “death of a warrior” won awards for creative nonfiction. vic is survived by his wife linda masterson, daughter andrea bradford, son jonathan van ballenberghe, and several nieces and nephews. he was truly a pioneer and stalwart of the moose world, with a professional legacy for future moosers to recognize and appreciate. _goback alces39_153.pdf alces vol. 39, 2003 silverberg et al. – moose response to wildlife viewing 153 moose responses to wildlife viewing and traffic stimuli judith k. silverberg1, peter j. pekins2, and robert a. robertson3 1new hampshire fish and game department, 2 hazen drive, concord, nh 03301, usa; 2department of natural resources, university of new hampshire, james hall, durham, nh 03824, usa; 3department of resources and economics, university of new hampshire, james hall, durham, nh 03824, usa abstract: we examined behavioral response of moose to wildlife viewers and traffic stimuli at a moose viewing blind located on a roadside salt lick in northern new hampshire during summer, 1997-1999. feeding, fleeing, alertness, looking, grooming, and moving were measured relative to a standard viewer and a variety of stimuli associated with viewers and traffic. in general, moose were reasonably tolerant of most stimuli as moose never fled the lick > 15% of the time. educational material likely influenced viewer behavior. stimuli that caused a reduction in feeding and increased fleeing were loud viewers, cars stopped, and trucks passing, as well as combinations of stimuli including these factors. viewing satisfaction and impacts can be addressed by considering these findings at moose viewing sites. alces vol. 39: 153-160 (2003) key words: alces alces, behavior, wildlife viewing wildlife managers should understand and minimize the often poorly understood and measured impacts of nonconsumptive wildlife users on species and habitats (duffus and dearden 1993). the behavioral response of moose (alces alces) to viewing has been explored in a few park situations. m c m i l l a n ( 1 9 5 4 ) s t u d i e d m o o s e i n yellowstone national park that were subjected to heavy tourist pressure and often were photographed at close range. by comparing moose in a heavily used tourist area to moose in a lesser visited area, he found that: (1) the closeness of approach permitted was dependent on the manner of approach; (2) some moose were able to recognize an individual; and (3) their awareness of a person was dependent on visibility, not the individual. moose eventually reduced their wariness to human approach, with approach distance dependent upon the moose’s activity. cobus (1972b) also found that the reactions of moose to humans indicated a developed tolerance in sibley provincial park, ontario, canada. mcmillan (1954) examined the response of moose to sounds and found that moose in yellowstone reacted to the snapping of twigs or rustling through brush. the metallic click of a field notebook brought a quick response, whereas shouting, a sharp whistle, automobile horns, and other sounds from the highway failed to produce any response. cobus (1972b) found that voices frequently scared moose that seemed relatively unaffected by the sight and scent of viewers at a lake. he also noted that the noise of traffic passing the lake caused no reaction, but a sudden car horn or slam of a door frequently disturbed moose 457 m (500 yards) away. the effect of road traffic from 1973-1983 was examined in denali national park, alaska, where there was a 50% increase in daily vehicular traffic on the main park road. this elevated volume correlated with a 72% decrease in moose sightings (signer and beattie 1986). a study of moose reactions to snowmobile moose response to wildlife viewing – silverberg et al. alces vol. 39, 2003 154 traffic in the greys river valley in wyoming showed that moose bedding within 300 m and feeding within 150 m of passing snowmachines altered their behavior in response to disturbance (colescott and gillingham 1998). moose often appear unalert because they can be approached closely without eliciting a visible reaction. however, devos (1958) found that ear position was a good indicator of the level of alertness, and moose extended their ears upward at a 45 degree angle to the head when alert. he also found that flight, flushing distance, and the relative sign of alarm varied among moose. in yellowstone national park, wyoming, altmann (1958) found that flight distance varied by month and situation. for example, during the fall hunting season moose fled at 183-274 m (200-300 yards), whereas a cow with a new calf could be approached within 27-64 m (30-70 yards) in may and june. in new hampshire, moose are viewed commonly along major roadways where salt licks are created by runoff of road salt. moose are observed primarily from cars, although a substantial number of viewers exit their vehicles at many salt licks. given the popularity of moose viewing, its direct relationship to tourism in northern new hampshire (silverberg 2000), and the concern for viewer safety and minimizing impacts of viewers, the new hampshire fish and game department constructed a moose viewing blind on route 26 in dixville notch, new hampshire, approximately 16 miles east of errol, new hampshire and 16 miles west of colebrook, new hampshire. the viewing site provided viewers an opportunity to view moose out of their vehicle off the roadway, thereby reducing traffic congestion, road safety concerns, and direct human-moose interactions. the site had the potential to change how people viewed moose and how moose responded to viewing. specifically, people can park their cars away from the lick, walk a short pathway with educational signs, and view moose from within the blind. the planning phase provided the opportunity to design a research project that would explore behavioral responses of moose to viewer-caused stimuli. three factors at the dixville notch site distinguish it from previous research in parks: (1) visitors were encouraged to leave their cars and walk to a blind; (2) educational information was available; and (3), the viewing location was on a well-traveled highway. this study was designed to categorize moose reaction to stimuli caused by wildlife viewers and vehicular traffic in order to determine whether there were predictable and measurable behavioral responses. study area a 4-hectare study site that incorporated the viewing area was located just east of dixville notch state park, in the township of dixville notch, new hampshire on route 26. the area was harvested (clearcut) in 1991 and is characterized by a regenerating northern hardwood/spruce-fir forest community. on the north side of the road was a substantial road run-off salt lick about 175 m long, with a smaller 70 m lick on the south side. the site included a 6-car parking lot, trail, and viewing blind built in december 1996. the trail was approximately 125 m long with educational signs, and led to the viewing blind that could accommodate up to 20 people. the blind afforded a view across the roadway, and had viewing slits that faced the lick and a moose trail entering the lick from the east. methods we recorded reactions of moose to viewer and traffic stimuli during june and july, 1997-1999. we recorded time, viewer numbers, and moose behavior on a data grid (lehner 1979). most observation periods alces vol. 39, 2003 silverberg et al. – moose response to wildlife viewing 155 occurred during the early evening when moose were most likely to visit the lick (silverberg et al. 2002). typically, multiple moose behaviors and stimuli were recorded during each observation. seven specific stimuli were categorized: car passing, truck passing, car stopped, car stopped with human outside of vehicle, viewer walking to or from blind, viewer in the blind talking, and viewer talking loudly. moose response was defined as one of 6 behaviors: feeding, looking, alert, moving, fleeing, and grooming. the number of moose in the lick and their sex, if determinable, were recorded during each observation period. a moose was considered feeding if it was actively feeding or licking mud. looking was defined as when a moose appeared to stare at the stimulus. alertness was defined as when a moose stopped its previous behavior, stared, and had its ears in a 45 degree position (devos 1958). a moose was regarded as moving if it took several steps and resumed its previous behavior. fleeing meant a moose rapidly moved from the lick to cover. grooming was defined as licking or moving to repel insects. an observation period was defined as the elapsed time when a moose entered the lick to the time it left, or it was too dark to continue observation. we recorded all moose behavior and stimuli that occurred every other minute. because moose were not marked, and moose have affinity for specific salt licks, the same moose was probably observed on different days. multiple observations occurred each observation period. these two facts meant that observations were not independent. the researcher hereafter referred to as the "perfect viewer", set the standard of behavior to which the behavior of other wildlife viewers was compared. the perfect viewer approached the blind quietly, was quiet in the blind, and usually was in the blind before moose visited the lick. presumably, moose rarely detected the presence of the perfect viewer or, at the very least, showed no reaction to the perfect viewer. baseline moose behavior was recorded only when the perfect viewer was present and there were no other human stimuli. the recording sheets and other written comments of the researcher were used to construct a narrative of each period to provide further description of the interactions. analysis of single and multiple combinations (2-4) of stimuli were necessary because multiple stimuli often occurred simultaneously (e.g., car stopped and truck passing). moose response was quantified by totaling the number of observed responses and calculating the percentage of each response that was exhibited for individual and combinations of stimuli. a chisquare test (p = 0.05) of independence (zar 1996) was used to compare the patterns of behavioral responses to different stimuli to the pattern of responses associated with the perfect viewer. emphasis was placed on interpreting the change in feeding and fleeing because reduced feeding and increased fleeing are negative responses for both moose and viewers. results a total of 48 observation periods occurred: 9 in 1997, 19 in 1998, and 20 in 1999. without the moose being marked it is difficult to determine the exact number of moose observed, however, because of antler development, multiple moose in the lick at one time, and the number of days between observations, it is possible to make a realistic estimate of the number of moose observed in each year: 1997, 5 males, 3 females, and 2 calves; 1998, 9 males, 4 females, and 2 calves; 1999, 11 males, 9 females, and 3 calves. observation periods ranged from 593 minutes, averaging 22 minutes. these observation periods occurred only when moose response to wildlife viewing – silverberg et al. alces vol. 39, 2003 156 moose were in the lick and the length of the observation period depended upon the amount of time the moose were present. an average of 6.4 cars passed, 1.6 trucks passed, 3.2 cars stopped, and 0.9 humans were out of their car during an observation period. no observation period consisted of only viewers in the blind and moose in the lick. during the 342 minutes of observation when only the perfect viewer was present, feeding, looking, and alertness were the most common behaviors (> 20%); grooming and fleeing were observed < 5% of the time (fig. 1). a difference in behavioral response pattern relative to that of the perfect viewer was found when a truck passed (χ2 = 26.5, df = 5, p = 0.000) or a car stopped (χ2 = 18.8, df = 5, p = 0.002) (table 1). when trucks passed, moose fled 14.5% of the time, or > 3 times as often as with the perfect viewer, and feeding declined > 25% (fig. 2). when cars stopped, moose fled 12% of the time, or nearly 3 times more than with the perfect viewer, and feeding behavior declined by > 30% (table 1, fig. 2). moose were most alert (> 29%) when a truck or car passed the lick. cars passing had minimal effect on feeding, as did visitors talking in a normal voice, or walking to the viewing blind (table 1, fig.1). conversely, although only 20 minutes of loud viewers were recorded, they caused the largest reduction in feeding (> 46%, fig. 2). trucks passing caused moose to flee 14.5% of the time (fig. 2). analysis of combinations of stimuli (24; table 2) indicated that a change in behavior, relative to the standard visitor, occurred only if a truck passed or a car stopped. chi-square values were within the same ranges, indicating no additive effects. eight combinations were significant (χ2 > 12, p < 0.05), including truck passing-car fig. 1. behavioral responses of moose when only the researcher was present at the dixville notch wildlife viewing area, summer 1997-1999. these data were used to compare all other response patterns to individual and combined stimuli. alces vol. 39, 2003 silverberg et al. – moose response to wildlife viewing 157 table 1. chi-square analysis results of single stimuli and behavioral responses of moose, and percent time feeding and fleeing as observed from the viewing blind, dixville notch viewing area, summer 1997-1999. stimulus number of chidf p-value % time % time observations square fled feeding perfect viewer 246 4.2 33.6 car passing 267 3.84 5 0.572 7.1 31.3 truck passing 72 26.5 5 0.000 14.5 24.2 car stopped 117 18.5 5 0.002 12.0 23.3 viewer walking 37 5.08 5 0.406 9.0 35.2 viewer talking 128 2.81 5 0.779 3.8 31.6 viewer loud 20 4.54 5 0.475 7.4 18.5 stopped, viewer walking-truck passing, viewer walking-car stopped, viewer walking-truck passing-car stopped, truck passing-car stopped-human out of car, viewer talking-visitor walking-car stopped, viewer talking-viewer walking-trucks passing-car stopped, viewer walking-car passing-truck passing-car stopped, and viewer walkingtruck passing-car stopped-human out of car (fig. 2). the narratives indicated that if a moose didn’t flee when a car stopped, it generally fled when a person approached within 5 m. no moose showed aggression towards people. discussion the primary purpose of a wildlife viewing site is to provide a satisfactory viewing fig. 2. moose feeding and fleeing response to various stimuli and combination of stimuli at the dixville notch wildlife viewing area, summer 1997-1999. stars represent stimuli that caused significant change in behavior. moose response to wildlife viewing – silverberg et al. alces vol. 39, 2003 158 opportunity with minimal impact. consequently, it was necessary to determine whether the act of viewing may reduce the opportunity to view moose. in general, reactions of moose to humans at the dixville notch wildlife viewing area indicated a high tolerance of human stimuli. the presence of quiet, well-behaved viewers had minimal effect on feeding activities and fleeing occurred < 4% of the time. in no situation did moose flee the lick > 15% of the time or feeding occur < 20% of the time, except when visitors were loud, but results were not significant (fig. 2). similar tolerance was found in park situations by mcmillan (1954), devos(1958), and cobus (1972b). although the incidence of loud viewers was low, feeding declined to its lowest level and looking increased measurably, although not significantly (fig. 2). conversely, moose showed little reaction when viewers walked to or from the site, talked in normal tones, or viewed quietly from the blind. educational signs placed along the trail may have had a positive impact on most viewing behavior, and/or viewer behavior was affected by the presence of the researchers. the signs provided tips for viewers like visiting the area at dawn and dusk, being patient, keeping a respectful distance, and being quiet. it is highly probable that impacts can be reduced by on-site education of wildlife viewers. while there was minimal change in moose response to viewers in the blind, responses to trucks passing and cars stopping were measurable and pronounced as moose fled at > 3 times the rate relative to response to the standard visitor. although observers in some parks found little response to vehicular traffic (mcmillan 1954, cobus 1972a), moose sightings declined in denali national park when traffic increased measurably (signer and beattie 1986). in addition, changes in feeding behavior were observed in wyoming with snowmobile traffic (colescott and gillingham 1998). it should be emphasized that local summer traffic at dixville notch was > 3,000 cars table 2. chi square analysis results of two simultaneous stimuli and behavioral responses of moose, and percent time feeding and fleeing as observed from the viewing blind, dixville notch viewing area, summer 1997-1999. stimuli number of chidf p-value % time % time observations square fled feeding perfect viewer 4.2 33.6 car stoppedhuman-out-of-car 47 5.48 5 0.360 6.6 27.2 car passing-truck passing 304 2.36 5 0.79 7.8 27.2 car passing-car stopped 357 6.71 5 0.242 7.5 28.6 truck passing-car stopped 236 15.3 5 0.002 11.1 25.2 viewer walking-truck passing 102 12.12 5 0.033 13.6 26.7 viewer walkingcar passing 289 3.96 5 0.055 6.9 29.3 viewer walkingcar stopped 207 18.9 5 0.002 10.9 25.3 viewer talkingviewer walking 149 1.59 5 0.901 4.7 32.1 viewer talking-viewer loud 56 8.32 5 0.138 8.5 30.8 alces vol. 39, 2003 silverberg et al. – moose response to wildlife viewing 159 daily, with a speed limit of 89 km/h (55 mph), unlike parks with slow moving traffic. logging and semi-tractor trailer trucks were audible at considerable distances as they gained speed entering and leaving the notch and moose responded to such noise. each summer one or more moose were killed at the site, and the obvious relationship between vehicle collisions and roadside saltlicks has implications for positive moose viewing. the incidence of wildlife viewing is greater in parks than at the dixville notch study site, and moose subjected continuously to viewing presumably become habituated to stopped cars. given that the dixville notch wildlife viewing area was established in 1997 and the site is on a welltraveled highway, the ratio of stopped cars to cars passing is relatively small. consequently, local moose were probably not habituated to stopped cars and responded with reduced feeding and increased fleeing. the negative influence of stopped cars on moose behavior and viewing opportunities could be alleviated with road signs prohibiting such activity. presumably the increased fleeing response attributed to a combination of stimuli was indicative of the single strongest stimuli, that is a truck passing or car stopped. there appeared to be additive effects with certain combinations, for example, moose fled twice as often when a car stopped and a viewer was walking versus the single stimuli of a viewer walking. when viewers were talking, walking, and a car stopped, moose fled twice as often as when viewers were walking or talking. one exception was the combination of viewer talking, walking, car stopped, and humans-out-of-cars, as moose fled only 5.3% of the time. one particular moose represented the majority of these observations, and relative to other moose, appeared extremely tolerant of all stimuli. it should be recognized that moose less tolerant of people could use the site predominantly at night. most human visitation occurred during midday and early evening when moose visitation was relatively low; moose visitation was highest at 2200-2400 h and 0400-0600 h (silverberg et. al. 2002). on the few occasions when loud viewers were present, the decline in feeding behavior probably had minimal impact because the incidents were short, lasting less than 5 minutes. substantial impact on feeding behavior could influence use of salt licks on a daily or long term scale. if disturbances were more frequent and of longer duration, moose may alter their visitation time and duration, or conversely, become habituated to the presence of noisy visitors. individual moose could be monitored to determine their frequency and time of visitation, and whether individual, age, or gender patterns exist. although certain behavioral changes occurred, the overall effect may not be meaningful in the context of time spent to fulfill nutritional requirements. wildlife viewers have a potentially negative influence on moose behavior and their own viewing opportunities. specifically, this study documented that cars stopped adjacent to the lick and viewers out of their cars increased fleeing behavior and ultimately reduced viewing opportunity and satisfaction. further, combinations of stimuli often had additive impact. several points relevant for managing moose viewing sites included: (1) viewing can alter moose behavior; (2) quiet viewers had no measurable impact on moose; (3) education of viewers should reduce potential disturbance of moose; and (4) viewing sites on heavily trafficked roads introduce stimuli not easily controlled. consideration of these findings should help ensure satisfactory, low-impact viewing opportunities throughout the northe a s t e r n u n i t e d s t a t e s w h e r e m o o s e populations and public interest in viewing moose are expanding. moose response to wildlife viewing – silverberg et al. alces vol. 39, 2003 160 references altmann, m. 1958. the flight distance in free-ranging big game. journal of wildlife management 22:207-209. cobus, m. w. 1972a. moose as an aesthetic resource and their summer feeding behavior. proceedings of the north american moose conference and workshop 8:244-275. . 1972b. moose (alces alces) and campers in sibley provincial park: a study of wildlife aesthetics. m.s. thesis, university of guelph, guelph, ontario, canada. colescott, j. h., and m. p. gillingham. 1998. reactions of moose (alces alces) to snowomobile traffic in the greys river valley, wyoming. alces 34:329338 de vos, a. 1958. summer observations on moose behavior in ontario. journal of mammalogy 9:128-139. duffus, d. a., and p. dearden. 1993. recreational use, valuation and management of killer whales (orcinus orca) on canada’s pacific coast. environmental conservation 20:149-156. lehner, p. n. 1979. handbook of ethological methods. garland stpm press. new york, new york, usa. mcmillan, j. f. 1954. some observations on moose in yellowstone park. american midland naturalist 52:392-399. signer, f. j., and j. b. beattie. 1986. the controlled traffic system and associated wildlife responses in denali national park. arctic 39:195-203. silverberg, j. k. 2000. impacts of wildlife viewing: a case study of dixville notch wildlife viewing area. ph.d. dissertation, university of new hampshire, durham, new hampshire, usa. , p. j. pekins, and r. a. robertson. 2002. impacts of wildlife viewing on moose use of a roadside salt lick. alces 38:205-211. zar, j. h. 1996. biostatistical analysis. third edition. prentice hall incorporated, englewood cliffs, new jersey, usa. 140 distinguished moose biologist past recipients 1991 charles c. schwartz, alaska dept. of fish and game, soldotna, alaska. 1990 rolf peterson, michigan technological university, houghton, michigan. 1989 warren b. ballard, alaska dept. of fish and game, nome, alaska. 1988 vince f. j. crichton, manitoba dept. of natural resources, winnipeg manitoba. and michel crête, ministère du loisir, de la chasse et de la péche, service de la faune terrestre, québec, pq. 1987 w. c. (bill) gasaway, alaska dept. of fish and game, fairbanks, alaska. 1986 h. r. (tim) timmermann, ontario ministry of natural resources, thunder bay, ontario. 1985 ralph ritcey, fish and wildlife branch, kamloops, british columbia. 1984 edmund telfer, canadian wildlife service, edmonton, alberta. 1983 albert w. franzmann, alaska division of fish and game, soldotna, alaska. 1982 a. (tony) bubenik, ontario ministry of natural resources, maple, ontario. 1981 patrick d. karns, minnesota division of fish and wildlife, grand rapids, minnesota. and al elsey, ontario ministry of natural resources, thunder bay, ontario. in 1974, prior to the establishment of the distinguished moose biologist award, the group recognized the pioneering moose research of the late laurits (larry) krefting, u.s. fish and wildlife service, with an individual award. 2007 kris j. hundertmark, university of alaska fairbanks, fairbanks, alaska. 2006 kristine m. rines, new hampshire fish and game department, new hampton, new hampshire. 2005 w. m. (bill) samuel, university of alberta, edmonton, alberta. 2004 w. eugene mercer, wildlife division, st. john's, newfoundland. 2003 arthur r. rodgers, ontario ministry of natural resources, thunder bay, ontario. 2002 bernt-erik sæther, norwegian university of science and technology, trondheim, norway. 2001 r. terry bowyer, university of alaska, fairbanks, alaska. 2000 gerry m. lynch, alberta environmental protection, edmonton, alberta. 1999 william j. peterson, minnesota department of natural resources, grand marais, minnesota. 1998 peter a. jordan, university of minnesota, st. paul, minnesota. 1997 margareta stéen, swedish university of agricultural sciences, uppsala, sweden. 1996 vic van ballenberghe, u.s. forest service, anchorage, alaska. 1995 not presented 1994 james m. peek, university of idaho, moscow, idaho. 1993 murray w. lankester, lakehead university, thunder bay, ontario. 1992 not presented 131 season of detachment of winter ticks (dermacentor albipictus) from southern ontario moose (alces alces) edward m. addison1,2, r. f. mclaughlin3, and d. j. h. fraser4 1wildlife research and development section, ontario ministry of natural resources and forestry, peterborough, ontario, canada k9j 7b8; 226 moorecraig road, peterborough, ontario canada k9j 6v7; 3r.r. #3, penetanguishene, ontario, canada l0k 1p0; 4344 wessex lane, nanaimo, british columbia, canada v9r 6h5 abstract: detachment of engorged female winter ticks (dermacentor albipictus) from captive moose (alces alces) was studied in ontario during march and april, 1981–1984. the earliest detached engorged female was observed on 15 march, and for 9 of 15 moose, on 25–26 march. detachment increased in early to mid-april with most adult ticks remaining on captive moose in late april. few ticks were observed on wild cow moose by midto late may, 1981–1984, and detachment was considered complete in late may. more ticks dropped from moose at night than during daylight hours. the primary period of detachment was considered mid-april to mid-may during all 4 years of the study. prediction of relative infestation the following autumn may be possible by considering the drop-off time and ground conditions that influence survival of gravid adult female ticks. alces vol. 57: 131–138 (2021) key words: alces alces, dermacentor albipictus, moose, winter ticks, detachment of ticks. the winter tick (dermacentor albipictus) has one of the broadest geographic distributions of north american ticks (gregson 1956) and displays range-wide variation in life history traits including the duration of its parasitic phase. time from infestation with larvae to detachment of engorged females varied from 22 to 30 days in california (howell 1939) and texas (drummond et al. 1969) to 175 days in ontario (addison and mclaughlin 1988). there are innumerable incidental reports of time of winter tick detachment, but only drew and samuel (1989) working in edmonton, alberta described the complete period when engorged female winter ticks detach from moose. our study augments their earlier findings by describing this period of detachment from captive moose in ontario in 1981–1984. importantly, the portion of an infested moose’s home range that is occupied during the period of detachment will become the potential source of infestation with larval winter ticks the following autumn, with overlapping habitat use common in these seasons (healy et al. 2018). it also delineates the period of greatest energetic demands on calves and gestating females in spring, which are most negatively affected by the amount of blood extracted during severe infestations (see pekins 2020). methods moose were raised in captivity in 1980–1983 in algonquin provincial park (algonquin park), ontario (45° 33’n, 78° 35’w) as described by addison et al. (1983). they were held in outdoor pens (29.6 × 16.5 m) ticks on moose in spring –addison et al. alces vol. 57, 2021 132 ticks on moose in spring –addison et al. with ~50% summer canopy cover of white pine (pinus strobus), white birch (betula papyrifera), trembling aspen (populus tremuloides), and largetooth aspen (p. grandidentata). with the exception of red-berried elder (sambucus pubens), ground vegetation in the pens disappeared rapidly during the first summer of occupancy from browsing, grazing, and trampling. crushed stone (5 cm diameter) was added to the feeding area and perimeter of the pens to increase hoof wear. straw-lined sheds (2.9 × 2.9 × 2.7 m) were available but moose seldom chose to bed in them. larvae for infesting moose were collected by flagging vegetation in the wild in algonquin park. in an initial trial in 1980 one moose was infested on 11 november. subsequently, 14 moose were administered larvae between 17 september and 12 october 1981–1984; the mean date of infestation was 30 september (table 1). the 15 captive moose used in this study were part of other tick-related experiments and varied in sex, age, time of infestation, size of the infesting dose, and number of moose per pen. one male calf was infested with 8000 larvae in 1980; two male calves were infested with 21,000–22,500 larvae each in 1981; two male and two female calves were infested with 21,000 larvae each, and two male and two female calves with 42,000 larvae each in 1982; one female and three male yearlings were infested with 21,000 larvae each in 1983. three of the yearlings infested in 1983 had been infested as calves in 1982 (table 1). beginning in early march each year, the total area of each pen was surveyed for detached ticks by walking parallel 1.0 m wide transects. some observers changed between years and new staff received training in the established methods. each pen was checked for ticks for 30–45 min during mornings (0700–0830 h) and evenings (1700– 1800 h), with the number of engorged females (efs) recorded as morning and evening collections. when other behavioral studies occurred, surveys for detached engorged female winter ticks sometimes occurred up to 1 h earlier in the morning and 1 h later in the evening. in 1981–1984, pens were checked for efs through 22, 18, 17, and 27 april, respectively, when moose were euthanized upon direction from our animal care inspector and/or to prepare pens for introduction of new moose. hair and hides of moose euthanized in 1982 (n = 2) and 1983 (n = 8) were dissolved and ticks counted as detailed in addison et al. (1979). adult ticks collected from hides in 1983 were tallied by sex. in table 1. infestation and first detachment of dermacentor albipictus from captive moose in southern ontario, canada, 1981–1984. infestation date age of moose (months) number of moose prior infestation with d. albipictus current infestation with d. albipictus first date of detachment nov 11, 1980 4 1 0 8000 march 25 sept 25, 1981 4 1 0 21,000 march 18 sept 27, 1981 4 1 0 22,500 march 18 sept 23–oct 2, 1982 4 4 0 21,000 march 25 sept 23–oct 2, 1982 4 4 0 42,000 march 24 sept 30, 1983 16 1 19,000 21,000 march 19 sept 30, 1983 16 1 21,000 21,000 march 16 sept 30, 1983 16 1 0 21,000 march 21 sept 30,1983 16 1 42,000 21,000 march 15 alces vol. 57, 2021 ticks on moose in spring –addison et al. 133 1982, the perianal region of one moose was examined for 14 consecutive days and the number, location, and degree of engorgement of each tick was recorded. additionally, presence of ticks was noted but not quantified on 10 radio-collared wild cows in algonquin park during calving behavior studies in late may 1981–1984 (see addison et al. 1993). results detachment time of efs was independent of time and dose of infestation (table 1). the earliest date of detachment was 15 march, and for 9 of 15 moose occurred on 25 or 26 march. the median date of first detachment for all animals was 25 march. detachment of efs in march was similar among years, although in 1982 there were 6 days when the daily count of efs was ≥ 25% of the maximum daily ef count for that year (fig. 1). of winter ticks that were collected as detached efs and others that remained on hides of moose when euthanized, only 11.8%, 33.4%, 6.7%, and 6.8% had detached as efs in march fig. 1. early season daily detachment of dermacentor albipictus from captive southern ontario moose (alces alces), 1981–1984 (expressed as the percentage of a day’s detachment relative to the maximum daily number of detached engorged females in that year). the maximum number of detached engorged females collected in one day in each year was: 1981–101, 1982–39, 1983 – 294, and 1984 – 271. the total number of engorged detached females counted each year was: 1981–1489, 1982–230, 1983–2455, and 1984 –2410. ticks on moose in spring –addison et al. alces vol. 57, 2021 134 ticks on moose in spring –addison et al. of 1981, 1982, 1983, and 1984, respectively. generally, the daily counts of detached efs increased within the first few days of april. the single exception was in 1984 when the daily total count was < 25% of the maximum daily count until well into the second week of april (fig. 1). tick collections in the morning represented approximately a 13–14 h crepuscular and nocturnal period since the previous evening collection. the evening collection represented ticks dropping in the prior 10–11 h of daylight. morning and evening collections yielded 4399 and 2051 efs, respectively, suggestive of a higher rate of detachment during nocturnal and/or crepuscular hours. number, location, and degree of engorgement of female ticks in the perianal area of one moose varied daily from 0 to 20 ticks. between 20 march and 2 april, ≤ 3 female ticks were observed engorging and increasing in size rapidly between consecutive days. in 1982, 419 detached efs were collected from the 2 moose prior to euthanasia; 7565 adult ticks were counted on the hides during the post-mortem inspections. in 1983, the 2481 detached efs collected from the 8 moose prior to death comprised 16.2% of the 15,272 adult female ticks remaining on the hides postmortem. of the 922–3900 (μ = 1909) adult female ticks/moose on the eight moose euthanized in late april 1983, 0.5–2.8% (μ = 1.8%) per moose were fully engorged or near fully engorged on the day of death. few engorging winter ticks remained on the 10 radio-collared wild cows in late may (1981–1984) in algonquin park (see addison et al. 1993). discussion in edmonton, alberta, detachment of efs from moose was first observed on 27 february and >50% had detached by the end of march (drew 1984), with peak detachment occurring in late march in 1982 and 1983 (drew and samuel 1989). this was at least 2 and up to 4 weeks earlier than documented in this study in ontario. drew and samuel (1989) referenced the possible influence of photoperiod on synchrony of detachment as proposed by patrick and hair (1977) for another ixodid tick, amblyomma americanum. the similar detachment times for engorged female winter ticks, independent of year and time and dose of infestation in the 4 years of our study, is consistent with winter ticks adapting their parasitic phase to local prevailing seasonal conditions. on captive moose in 1982, 5–22% (μ = 12.4%) of the infesting dose of ticks remained on the hides when moose were euthanized in midto late april 1983 (addison et al. 2019). presumably, many of the remaining adult female ticks may have engorged and detached during late april into early may. few engorging winter ticks remained on wild algonquin moose captured in midto late may (e. addison, pers. comm.). the period of detachment of efs in our study lasted 4–6 weeks beginning in late march through to mid-may, and the peak period of 2–4 weeks was similar to that observed by drew and samuel (1989) and consistent with the variability in engorgement and detachment of some other ixodid ticks (usually 5–21 days). for example, adult female d. albipictus have become replete 5–14 d after attaching to the host (howell 1939), amblyomma americanum in 7–15 days (μ = 9.8 d; sauer and hair 1972), and dermacentor variabilis in 7–10 d (μ = 8.7 d; sonenshine 1967). these times can be extended if mating of females is delayed after their initial period of feeding (sonenshine 1967). although variation exists in the timing of final detachment of winter ticks among studies and areas, detachment is complete by alces vol. 57, 2021 ticks on moose in spring –addison et al. 135 the end of may consistently. in alberta, drew (1984) reported all efs having detached by 1 may, and glines and samuel (1989) reported all winter ticks detached from experimentally infested moose by 14 may. only 32, 52, and 233 ticks were recovered from digested hides of 3 moose killed on 20 may in 1963–1965 in the chapleau crown game preserve of northern ontario (addison et al. 1979). the latest that an ef has been reported on a host was on 31 may, and this was from a horse maintained at 5500 ft above sea level in pony, montana (bishopp and wood 1913). ritcey and edwards (1958) working with moose in wells gray park, british columbia, were the first to note the apparent synchrony between presence of efs on moose and the appearance of bare ground and proposed that presence of snow increased mortality of detached ticks. drew and samuel (1986) tested this hypothesis and demonstrated higher mortality of efs dropped on snow as compared to efs placed on bare ground. edmonton, alberta has relatively little snowfall with a mean mid-winter snowpack of 18 cm and a mean of 16.9 d in march with at least 5 cm of ground snow (1981–2010; canadian climate normals). in contrast, snow stations near the end of march in algonquin park, ontario (45.58°n, 78.3°w) and chapleau, ontario (45.58°n, 83.45°w) in 1961–1986 had mean ground snow depths of 42.6 cm (± 24.9 sd) and 65.2 cm (± 17.76 sd), respectively; near the end of april, mean ground snow depth was 5.1 and 14.6 cm, respectively (warren et al. 1998, ontario ministry of natural resources and forestry 2020). ticks detaching from moose in algonquin and in chapleau at the time of detachment reported for edmonton would have had lower chance of survival, as well as chapleau ticks during the early algonquin detachment period of late march-early april. these data suggest that detachment times of the adult stage of winter ticks are adaptive to local weather conditions and coincident with appearance of bare ground in spring, as proposed by ritcey and edwards (1958). a period of detachment of winter ticks prolonged over 4–6 weeks has adaptive advantages as compared to a more compressed period. during spring in moose range, remaining ground snow and/or fresh snow varies annually and, in the case of fresh snow, varies daily among habitats within a local area and year, and daily in the case of fresh snow. a prolonged period of detachment of d. albipictus increases the probability of some winter ticks detaching onto substrates optimal for survival and egg laying. additionally, detachment of efs over a prolonged period may spread their distribution over a larger portion of the spring home range; subsequently, more of the home range would be a source of infesting larvae the next autumn. extraction of blood from moose by winter ticks can be extensive and energetically demanding (see samuel 2004, musante et al. 2007, pekins 2020). extending the period of repletion and detachment of efs over numerous weeks as compared to a more compressed period may result in increased survival of some infested moose. the rapidity of final engorgement and detachment of individual winter ticks was apparent from the rapid increase in size of engorging efs on the perianal region of one moose followed for 14 consecutive days. rapid engorgement of individuals is also observed in the ixodid, boophilus microplus, where partially engorged females doubled in length and increased in weight 5–25 × during their last night of feeding before detaching from the host (wharton and utech 1970). the energetic demands of extraction of large amounts of blood just prior to detachment may be ameliorated by asynchronous final engorgement of adult female winter ticks. ticks on moose in spring –addison et al. alces vol. 57, 2021 136 ticks on moose in spring –addison et al. the small proportion (0.5–2.8, μ = 1.8%) of adult female ticks/moose fully or near fully engorged on the day of death on the 8 moose euthanized in late april, 1983 suggests a degree of asynchrony in the final engorgement and detachment of efs. numerous variables made it more likely that a greater proportion of detached efs were recovered in the alberta study than in the ontario study. drew and samuel (1989) maintained their moose in smaller pens with concrete floors and solid wood partitions between pens, unlike the forest floor in the ontario study. during our behavioral studies in march and april, we observed efs detaching from moose, yet a few hours later, were unable to locate all efs known to have detached. it was likely more efs were recovered from our pens before all snow was gone due to the contrasting appearance of ticks and snow and the inability of efs to rapidly burrow into the duff layer or through snow. it was more difficult to locate recently detached efs during days of fresh heavy snowfall than on snow-free days. ravens (corvus corax) and canada jays (perisoreus canadensis) were observed feeding on detached engorged efs during the observation periods (addison et al. 1989). a daily rhythm to the time of detachment has been observed for many species of ixodid ticks (belozerov 1982), and the efs detaching at night may be less vulnerable to predation by corvids. our data are not indicative of absolute numbers, but instead reflect trends over time in early detachment of efs. much current emphasis is placed on the potential impacts of climate change on biotic and abiotic components of ecosystems, including ecosystems with moose. many die-offs of moose associated with winter ticks occurred in the first half of the 20th century (seton 1909, samuel 2004) and again more recently in the northeastern united states (jones et al. 2017, 2019). the extent to which changes in climate can be attributed as a primary cause of these dieoffs both past and relatively current remains conjectural. however, it is clear that unpredictability in weather patterns has increased greatly in recent decades and continues to increase. if climate changes and associated weather patterns lead to increased snow depth during the expected local period of detachment of efs, production of winter tick larvae will be suppressed and this will be advantageous to moose populations. conversely, warming weather that decreases the extent of ground snow cover during the period of detachment of efs may negatively impact moose populations. summer weather conditions also influence recruitment of the non-parasitic stages of winter ticks and those influences vary among habitats (addison et al. 2016). additionally, a longer questing period in autumn may elevate infestations (pekins 2020). in summary, the main period of adult feeding and detachment of winter ticks on captive moose in algonquin park, ontario was 4–6 weeks in length between early april and mid-may. although generally similar, this was approximately one month later than documented in central alberta. differences in detachment time of efs among areas appears consistent with local variation of bare ground in spring. knowledge of local detachment times of efs combined with snow cover information in a specific spring might provide insight and predictability of the annual tick burdens on moose the following autumn. local differences in moose density and weather affecting larval survival and questing will influence the relative accuracy of these predictions. acknowledgements we thank s. fraser, s. gadawaski, a. jones, s. mcdowell, l. berejikian, k. long, k. paterson, l. smith, d. bouchard, v. ewing, alces vol. 57, 2021 ticks on moose in spring –addison et al. 137 j. jefferson, m. van schie, a. macmillan, a. rynard, n. wilson, c. pirie, m. mclaughlin, d. carlson, and others for their strong commitment to some or all of capturing, raising, and maintaining of moose calves and collection of winter ticks. we appreciate the assistance of r. marotte and e. newton of the wildlife research and development section, ontario ministry of natural resources and forestry for assistance in preparing a figure and provision of data from the ontario snow network data set, respectively. special thanks to r. addison who has assisted throughout our moose studies including editorial assistance with manuscripts. field work was conducted at the wildlife research station in algonquin park, ontario. references addison, e. m., d. j. h. fraser, and r. f. mclaughlin. 2019. grooming and rubbing behavior by moose experimentally infested with winter ticks (dermacentor albipictus) alces 55: 23–35. _____, f. j. johnson, and a. fyvie. 1979. dermacentor albipictus on moose (alces alces) in ontario. journal of wildlife diseases 15: 281–284. doi: 10.7589/0090-3558-15.2.281 _____, and r. f. mclaughlin. 1988. growth and development of winter ticks, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670–678. doi: 10.2307/ 3282188 _____, _____, p. a. addison, and j. d. smith. 2016. recruitment of winter ticks (dermacentor albipictus) in contrasting habitats, ontario, canada. alces 52: 29–40. _____, _____, and d. j. h. fraser. 1983. raising moose calves in ontario. alces 19: 246–270. _____, _____, _____, 1993. observations of preand post-partum behaviour of moose in central ontario. alces 29: 27–33. _____, r. d. strickland, and d. j. h. fraser. 1989. gray jays, perisoreus canadensis, and common ravens, corvus corax, as predators of winter ticks, dermacentor albipictus. canadian fieldnaturalist 103: 406–408. belozerov, v. n. 1982. diapause and biological rhythms in ticks. pages 469–500 in f. d. obenchain and r. galun, editors. the physiology of ticks. pergamon press, oxford, united kingdom. bishopp, f. c., and h. p. wood. 1913. the biology of some north american ticks of the genus dermacentor. parasitology 6: 153–187. doi: 10.1017/s0031182 000003012 canadian climate normals. meteorological service of canada. environment canada. . (accessed july 2020.) drew, m. l. 1984. reproduction and transmission of the winter tick, dermacentor albipictus (packard) in central alberta. m. s. thesis, university of alberta, edmonton, alberta, canada. _____, and w. m. samuel. 1986. reproduction of the winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64: 714–721. _____, and _____. 1989. instar development and disengagement rate of engorged female winter ticks, derma centor albipictus, (acari: ixodidae), following singleand trickle-exposure of moose (alces alces). experimental and applied acarology 6: 189–196. drummond, r. o., t. m. whetstone, s. e. ernst, and w. j. gladney. 1969. biology and colonization of the winter tick in the laboratory. journal of economic entomology 62: 235–238. doi: 10.1093/ jee/62.1.235 glines, m. v., and w. m. samuel. 1989. effect of dermacentor albipictus (acari: ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. environmental & applied acarology 6: 197–213. http://climate.weather.gc.ca/climate_normals/index_e.html http://climate.weather.gc.ca/climate_normals/index_e.html ticks on moose in spring –addison et al. alces vol. 57, 2021 138 ticks on moose in spring –addison et al. gregson, j. d. 1956. the ixodoidea of canada. publication 930. canada department of agriculture, ottawa, canada. healy, c., p. j. pekins, l. kantar, r. g. congalton, and s. atallah. 2018. selective habitat use by moose during critical periods in the winter tick life cycle. alces 54: 85–100. howell, d. e. 1939. the ecology of dermacentor albipictus (packard). proceedings of the sixth pacific science congress 4: 439–458. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. _____, _____, _____, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi: 10.1139/ cjz-2018-0140 musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. ontario ministry of natural resources and forestry. 2020. the snow network for ontario wildlife database: 1952–2020. unpublished database. ontario ministry of natural resources and forestry, wildlife research and monitoring section, peterborough, ontario, canada. patrick, c. d., and j. a. hair. 1977. seasonal abundance of lone star ticks on white-tailed deer. environmental entomology 6: 263–269. doi: 10.1093/ ee/6.2.263 pekins, p. j. 2020. metabolic and population effects of winter tick infestations on moose: unique evolutionary circumstances? frontiers in ecology and evolution 8: 176. doi: 10.3389/fevo. 2020.00176. ritcey, r. w., and r. y. edwards. 1958. parasites and diseases of the wells gray moose herd. journal of mammalogy 30: 139–145. doi: 10.2307/1376619 samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series volume 1. federation of alberta naturalists, edmonton, alberta, canada. sauer, j. r., and j. a. hair. 1972. the quantity of blood ingested by the lone star tick (acarina: ixodidae). annals of the entomological society of america 65: 1065–1068. seton, e. t. 1909. life-histories of northern mammals. volume 1. grasseaters. charles scribner’s sons, new york city, new york, usa. sonenshine, d. e. 1967. feeding time and oviposition of dermacentor variabilis, (acarina: ixodidae) as affected by delayed mating. annals of the entomological society of america 60: 489–490. doi: 10.1093/aesa/60.2.489 warren, r., a. r. bisset, b. a. pond, and d. voigt. 1998. the snow network for ontario wildlife: the why, when, what and how of winter severity assessment in ontario. ontario ministry of natural resources, peterborough, ontario, canada. wharton, r. h., and k. b. w. utech. 1970. the relation between engorgement and dropping of boophilus microplus (canestrini) (ixodidae) to the assessment of tick numbers on cattle. journal of the australian entomological society 9: 171–182. doi: 10.1111/j.1440-6055.1970. tb00788.x f:\alces\vol_39\p65\3918.pdf alces vol. 39, 2003 rolandsen et al. detectability of moose 79 sustainable moose (alces alces) management depends on regular information about population size and structure to determine the annual number of hunting permits. in norway, the two most frequently used methods are aerial censuses and observational data obtained by hunters during the hunting season (solberg and sæther 1999). aerial census of the population may be a relatively precise method given the choice of an appropriate sampling design (caughley 1974, tärnhuvud 1988), but the method is expensive, and therefore of restricted local use. moreover, because of the long sea2present address: gnd naturkart as, n-7898 limingen, norway. sonal migrations that occur in many f e n n o s c a n d i a n m o o s e p o p u l a t i o n s (cederlund et al. 1987, sweanor and sandegren 1988, sæther et al. 1992, andersen and sæther 1996), the size and distribution of the population during winter, when censuses normally occur, may differ extensively from the population during the autumn hunting season. in practice, the winter estimates may therefore be of limited use for local moose managers. a much less costly method than aerial censuses is to estimate the change in moose density and structure based on moose obfactors affecting detectability of moose alces alces during the hunting season in northern norway christer moe rolandsen1,2, erling johan solberg3, jarle tufto4, bernt-erik sæther1, and morten heim3 1department of zoology, norwegian university of science and technology, n-7491 trondheim, norway; 3norwegian institute for nature research, n-7485 trondheim, norway; 4department of mathematical sciences, norwegian university of science and technology, n-7491 trondheim, norway abstract: the use of hunter observations of moose (alces alces) to index variation in population size and structure is based on the assumption that there is a monotonic relationship between moose seen per hunter-day and population size of moose. for this relationship to also be proportional, the probability of detecting a given moose should increase proportionally with the number of hunters and days hunting; i.e., a doubling of the number of hunter-days should double the probability of detecting a moose. moreover, to obtain a precise index, the index should be independent of moose reproductive status, date of hunting season, weather conditions, and hunting area. we examined the influence of these factors on the probability of detecting individually radio-collared moose in a population in northern norway. our results support a proportional relationship between the number of moose seen and the number of hunters observing them. moreover, we found no difference in observation rate among female moose in relation to the number of calves following them. however, large variation existed in the proportion of possible moose observed by different hunting teams. this can result in varying observation rates between years in situations where large annual variation exists in the number of days hunted by each hunting team. we therefore recommend that the pooling of observation data should be performed over a more carefully selected period of the hunting season (e.g., the first hunting week) rather than over the whole hunting season. alces vol. 39: 79-88 (2003) key words: alces alces, hunter observations, hunting, management, moose, norway detectability of moose rolandsen et al. alces vol. 39, 2003 80 servations obtained and reported by hunters during the hunting season (crichton 1993, andersen and sæther 1996, ericsson and wallin 1999, solberg and sæther 1999). this method has been in regular use as a management tool in norway since the mid1980s and in one area since the late-1960s (solberg and sæther 1999). the data obtained from these reports includes total number of observed (corrected for known duplications) and shot moose by sex and age, as well as the hunting effort (andersen and sæther 1996, solberg and sæther 1999). such data are frequently used to estimate the change in population density (moose seen per hunter-day), sex structure (male per female), and recruitment rate (calves per female). studies that have examined the relationship between moose observations by hunters and independent estimates of population size obtained through other methods all found support for a general monotonic relationship between the equivalent measurements (fryxell et al. 1988, ericsson and wallin 1999, solberg and sæther 1999). however, the observational indices did not always show the same direction of population change as other independent estimates of population size or structure (ericsson and wallin 1999, solberg and sæther 1999), and did not increase proportional with the population size. for instance, based on the relationship between hunter observations and independent measures of population size among different populations, ericsson and wallin (1999) found a diminishing increase in the moose seen per hunter-day with increasing moose densities. similarly, solberg and sæther (1999) found that the moose seen per hunter-day tended to overestimate population size in years with high hunting success, indicating that the probability of detecting a given moose co-varies with conditions that lead to high hunting success (e.g., weather conditions). in turn, this variation in the probability of detecting a moose among years reduces the precision of observation indices as a predictive management tool. one suggested factor that may affect the probability to observe a moose is the number of hunters participating in the hunt (ericsson and wallin 1994). although more hunters are likely to observe more moose, the number of moose observed may not necessarily increase in proportion to hunting effort; i.e., because observation efficiency may decrease with the number of hunters. alternatively, if more hunters lead to more intense hunting, and subsequently greater movement of moose, moose may expose themselves more often to the hunters. different sex or age groups may also expose themselves with different probability, or groups of animals may be more easily detected than singletons. as a consequence the moose seen per hunter effort and recruitment indices (e.g., calves per female) may vary in relation to variation in population structure. these effects may be of minor importance for estimating the general trend over years for a population; i.e., the relationship may still be monotonic (sensu williams et al. 2002), but may be a major impediment for developing moose observation indices into a more precise management tool. in this study, we used hunter observations of individually radio-collared moose to examine how the detection probabilities of moose vary within and among hunting areas. more specifically, we tested to what extent variation in hunting effort, moose movement, and weather conditions were associated with the probability of detecting a moose. study area the study was carried out in the municipality of bardu (69οn) in the county of troms, northern norway. the area is situalces vol. 39, 2003 rolandsen et al. detectability of moose 81 ated within the medium boreal and the northern boreal vegetation zones (moen 1998). dominating tree species on the mountain sides are birch (betula pubescens) and scots pine (pinus sylvestris), interspersed with rowan (sorbus aucuparia), aspen (populus tremula), grey alder (alnus incana), bird cherry (prunus padus), and willow species (salix spp.) along the rivers (sæther and heim 1993, solberg et al. 1999). during the summer there is a high production of herbaceous plants, including many important browse species like cicerbita alpina, equisetum fluviatile, athyrium filix-femina, matteuccia struthiopteris, and dryopteris expansa (sæther and heim 1993). the area has cold winters (mean january temperature -10.4οc), cool summers (mean july temperature 13.0οc), and a mean yearly precipitation of 652 mm. the municipality of bardu is currently divided into 29 hunting zones (x = 25.27 km2, sd = 24.31). eleven of them were included in the present study (x = 19.45km2, sd = 4.31). each hunting zone had 1 team of hunters, with an average of 6.4 (sd = 2.1) hunters per team during the study period. methods data collection and measurements adult moose were captured by darting from a helicopter during february/march in 1996 and 1997. the animals were subsequently ear-tagged and fitted with 5 cm wide radio-collars. before the 1997 hunting season the radio-collared females were approached on foot and the number of calves with the females was recorded. during the period september 25 october 18, radio-collared moose within the selected hunting zones were triangulated with a precision level of + 100 m once per day, at approximately the same time (i.e., + 2 hours) each day. the interval between september 25 and october 18 included 2 periods with hunting (september 25 october 1 and october 10 october 18). no hunting occurred between october 2 9. each team of hunters in the study area recorded the number of radio-collared moose observed daily and the number of calves present with the radio-collared females. they also recorded the locality of the observation, and whether any radiocollared individuals or calves following them were shot. by triangulating radio-collared moose each day we also knew how many of the radio-collared moose were present in areas with hunters. the probability of detecting a moose was defined as the proportion of still living radio-collared moose within hunting zones with active hunting that were observed each day. if a radio-collared moose was with certainty observed twice or more by the same team, these observations were not recorded. this is in accordance with the standard procedure for recording moose observed by hunters on the observation form (see below). on one occasion the same individual moose was observed in 2 different hunting zones on the same day, and was counted as 2 observations in the data analysis. from the moose observation forms completed each year by all hunting teams (andersen and sæther 1996, solberg and sæther 1999), we calculated the daily mean number of hunters in a team and the daily number of teams that were hunting. the daily distance moved by individual moose was calculated as the linear distance in meters between the positions on consecutive days (m/day). because the radiocollared moose were not followed continuously, the estimates of movement underestimated the actual distance moved by the moose. the proportion of radio-collared moose that crossed borders between hunting zones each day was used as an index of interzonal movement. this variable was calculated as detectability of moose rolandsen et al. alces vol. 39, 2003 82 the number of radio-collared moose found in a different hunting zone than the preceding day divided by the total number of radiocollared moose present in the study area the same day. the proportion of radio-collared females with calf/calves in hunting zones with active hunters was used to reflect the structural composition of the moose population. all climatic variables were measured at bardufoss meteorological station and provided by the norwegian meteorological institute in oslo. both temperature and the amount of precipitation decreased during the study period. predictions and data analysis we tested several assumptions regarding the use of 'moose seen per hunter-day' as an index of variation in population density. this involved (1) the variation in the number and spatial distribution of the observers, the hunters; (2) the climatic conditions that may influence the visibility in the forest; and (3) the structural composition of the moose population. first, the number of moose observed may vary with the number of hunters. the basic assumption behind the use of moose seen per hunter-day as an index of density is that the number of moose observed increases proportionally (that is, with a coefficient not different from 1) with the number of hunters observing. thus, given a fixed population density within a given area, the moose seen per hunter-day will be independent of the number of hunters in the area. this is not trivial as the moose seen per hunter-day has been suggested to decrease with increasing number of hunters because of decreasing observation efficiency (ericsson and wallin 1999). however, increasing number of hunters may also observe more moose per hunter-day because more hunters lead to higher disturbance and movement of moose, thus increasing chances for moose to be observed (ericsson and wallin 1996). accordingly, the use of moose seen per hunter-day may be a poor index of population density if the number of hunters varies. another possibility is that the observation rates vary with the area used for hunting independent of the number of hunters, as hunting over large areas may be expected to decrease the number of 'sanctuaries' where moose may hide. in the study area, each hunting team hunted exclusively within fixed hunting areas, suggesting that the disturbance and movement may be expected to increase with the number of hunting teams hunting at a given time. the observation rate may also vary with the variation in climate and the progress of the season (i.e., date within hunting season) as this may affect the visibility in the forest, hence the chance to observe a moose. here, we test whether the observation rate varies with the level of precipitation and with the progress of the hunting season. high level of precipitation is usually associated with low visibility. similarly, the transparency of the forest is assumed to increase with the progress of the hunting season because of proceeding leaf fall. because a high proportion (> 50%) of the study area is covered with deciduous forest (birch, willow, rowan, and alder), progression of leaf fall is likely to have a significant effect on the observation rate. finally, we examined whether the observation rate varied with the structural composition of the population. for instance, females with calf/calves may be easier to detect because of the larger group size, or alternatively less easy to detect if they are more elusive to protect their calf/calves. thus, the proportion of moose observed may either decrease or increase with the proportion of females with calves in the area. here, we examined to what extent the observation rate varied among radiocollared females with different reproducalces vol. 39, 2003 rolandsen et al. detectability of moose 83 tive status and tested to what extent the observation rate changed as the proportion of radio-collared females with calf/calves decreased during the hunting season. the factors affecting the number of moose observed each day were examined using poisson regression with a log link function (proc genmod, sas institute 1996), as the observed number of radiocollared moose was expected to have a poisson probability distribution. in general we modelled the expectation in the poisson distribution λ as a curve linear function of covariates of interest such that; 87654321 bbbbbbbb zdptsinxam=λ (1) on log scale this model takes the form; )ln()ln()ln()ln()log( 21 nbxbma +++=λ (2) to adjust for the daily variation in the number of radio-collared moose within hunting areas, we used the logarithm of all radiocollared moose within hunting areas as an offset variable; i.e., the exponent of the parameter (m) equal to 1 (see equation 1). in addition, we included the total number of hunters (mean number of hunters in a team (ln), number of hunting teams (lnn)) as offset, as we assumed the observed number of radio-collared moose to increase in direct proportion to the number of hunters (the null hypothesis). we then examined alternative models by including one or a combination of the different explanatory variables (mean distance moved by radio-collared moose (d), interzonal movement of moose (i), temperature (t), precipitation (p), date within the hunting season (z), and the structural composition of the moose population (s) as covariates (equation 2), and finally by testing models where ln and/or lnn were included as covariates rather than offset variables. in this case we tested to what extent the variation in the observed number of radio-collared moose changed with the number of hunters with a coefficient different from 1, and/or to what extent variation in number of hunting teams influenced the number of observations. because of the generally small data set (low power), we only tested models with 1 or 2 covariates included at a time. the statistical significance of the different models was tested using the likelihood ratio test based on the change in deviance (sas institute 1996). the change in deviance between 2 nested models, e.g. d(h 0 ) d(h), is approximately chi-square distributed, with p -p 0 degrees of freedom where p and p 0 are the number of parameters under the models h and h 0 , respectively. if an independent variable significantly (p < 0.05) reduced the error deviance (d), we rejected h 0 (no effect, a proportional relationship between moose observed and number of hunters) in favor of the more general model h. results during the study period, 23 radio-collared moose were located within the selected hunting zones. of these, 6 were females without calves, 8 single-calf females, 8 females with twins, and 1 male. the timing of the harvest of radio-collared moose or calves following radio-collared females indicated that most moose were shot during the first week of hunting (table 1). the proportion of radio-collared moose observed varied among days (0.16 ± 0.11), and generally decreased during the hunting season (r = -0.63, n = 16, p = 0.009). during the same period of time, the mean number of hunters, the number of hunting teams, temperature, precipitation, and interzonal )ln()ln()ln()ln( ptsi +++++ )ln()ln( zd ++ detectability of moose rolandsen et al. alces vol. 39, 2003 84 movement of moose decreased (table 2). the proportion of radio-collared females with calf/calves in hunting zones with active hunters increased (table 2), while the distance moved by radio-collared individuals was unrelated to time (table 2). comparing the different models indicated that no combination of covariates better explained the variation in number of moose observed than the null hypothesis (table 3, fig. 1). an alternative model with interzonal movement of moose as a covariate was the best alternative model, but not significantly different from h 0 (table 3). the model including the mean number of hunters hunting and the number of hunting teams as covariates also indicated a lower deviance, but not significantly different from h 0 (table 3, fig. 1). indeed, regressing the number of observed moose (with moose present as offset) on the total number of hunters (mean number of hunters, number of hunting teams), revealed that the slope did not significantly deviate from 1 (fig. 1). other combinations of covariates such as date within season, temperature, precipitation, proportion of females with calves, and distance moved by moose (m/day), did not significantly contribute when the number of hunters simultaneously was kept as an offset variable. thus, given the present power, we cannot reject the null hypothesis that moose observed is a direct proportional function of the moose present in the area table 1. the number of radio-collared moose or calves following radio-collared females shot during the hunting season in bardu in 1997. the number in parentheses gives the % shot each week. number of shot animals hunting week period adults calves total 1 september 25 october 1 5 9 14 (60.9) 2 october 10 october 16 2 2 4 (17.4) 3 october 17 october 23 1 1 (4.3) 41 october 24 october 31 2 2 4 (17.4) 1 the fourth hunting week includes one extra day (i.e., 8 days). table 2. matrix of pearson correlation coefficients for independent variables, where z = date within the hunting season, x = mean number of hunters in a hunting team, n = number of hunting teams, t = temperature, p = precipitation, s = proportion of radio-collared females with calves, d = mean daily distance moved by radio-collared moose, and i = interzonal movement of moose. independent variables z x n t p s i x -0.76*** n -0.82*** 0.84*** t -0.85*** 0.87*** 0.82*** p -0.67** 0.44 0.50* 0.65** s 0.87*** -0.72** -0.87*** -0.70** -0.52* i -0.69** 0.52* 0.67** 0.69** 0.41 d -0.41 0.33 0.33 0.18 0.11 -0.64** 0.49 * p ? 0.05, **p ? 0.01, ***p ? 0.001. * p < 0.05, **p < 0.01, ***p < 0.001. alces vol. 39, 2003 rolandsen et al. detectability of moose 85 total number of hunters 0 20 40 60 80 100 n u m b e r o f m o o se o b se rv e d 0 1 2 3 4 5 p ro p o rt io n o f m o o se o b se rv e d 0.0 0.1 0.2 0.3 0.4 and the number of hunters observing. in order to examine whether group size affects the probability of observing a moose, we compared the probability of observing female moose in relation to the number of calves following them. no significant difference was found among the categories (x + se; 0.16 + 0.04, 0.20 + 0.07, 0.10 + 0.04 for females without calves, single calf females, and females with twins, respectively; χ2 = 2.62, df = 2, p = 0.27). similarly, no significant relationship was found between the rate of movement and whether the females were without calves, single calf females, or females with twins (x + se; 1728 + 549 m/day, 1076 + 154 m/day, 1645 + 471 m/day, for females without calves, single calf females, and females with twins, respectively; f = 0.75, df = 2, p = 0.49) or precipitation (r = 0.12, n = 15, p = 0.67). table 3. the best models explaining daily variation in the number of radio-collared moose observed (dependent variable) during the hunting season. model 1 is the null hypothesis (h 0 ) and models 2 and 3 are the 2 best alternative models. d(h 0 ) – d(h) is the change in deviance between h 0 and the alternative models. m = the number of radio-collared moose present in hunting zones with active hunters, x = mean number of hunters in a hunting team, n = number of hunting teams, and i = interzonal movement of moose. fig. 1. the number of radio-collared moose observed given a fixed number (16) of radio-collared moose present (left axis, stippled line) and the proportion radio-collared moose observed (right axis, filled circles) in relation to total number of hunters. the solid line represents a proportional relationship between number of moose observed and number of hunters. model deviance d(h 0 ) d(h) df p 1 λ = -5.89 ± 0.13mx n 12,14 2 λ = -6.38 ± 0.55mx ni1.88±2.00 11,06 1.08 1 > 0.1 3 λ = -7.96 ± 2.82mx 0.173±0.797n2.916±1.640 10,44 1.70 2 > 0.1 detectability of moose rolandsen et al. alces vol. 39, 2003 86 moreover, comparing the rate of movement before and after females lost their calf/ calves by hunting, 4 females decreased their rate, whereas 3 females increased their rate of movement (x difference = 171 m/day; paired t-test, t = 0.38, n = 7, p = 0.72), indicating no general difference in movement before or after the calves were killed. thus, no difference in observation frequency or behavior was found in relation to whether females had calves or not. discussion the use of observational data to monitor changes in population size and structure is based on several assumptions (ericsson and wallin 1999, solberg and sæther 1999, hochachka et al. 2000), among which is the assumption that the number of moose observed increases proportionally with the number of hunters. previous analyses, however, have indicated that the moose seen per hunter-day increases with population size, but with a slope less than 1 (ericsson and wallin 1999; solberg and sæther 1999). as the number of hunters also tends to increase with population size, we therefore expected the lack of proportional increase to be due to an increasing saturation of the number of observations as the number of hunters increases (ericsson and wallin 1999). also the practice in norway of not reporting moose that with certainty were observed by other hunting team members the same day is likely to generate a disproportional relationship between the number of moose observed and number of hunters. although we observed a tendency of a disproportional slope (fig. 1), we could not reject the hypothesis of a proportional relationship between the number of moose observed by hunters and the number of moose present. thus, other compensatory mechanisms may also influence the number of moose observed; e.g., the movement pattern of moose as the number of hunters increase. in the present study, interzonal movement (i.e., the proportion of moose that cross borders between hunting zones) of moose decreased with decreasing number of active hunting teams (table 2), with a possible response being that relatively fewer moose are seen, and registered, by more than 1 team of hunters on the same day. also, due to the low sample size, we cannot completely exclude the possibility that the lack of significant deviance from a proportional relationship could be due to low statistical power. another factor that may affect the moose observed per hunter-day and the recruitment indices is different detectability of moose depending on sex, age, or number of calves in company with female moose. in two recent contrasting studies, solberg and sæther (1999) and ericsson and wallin (1999) suggested that managers could either underestimate or overestimate recruitment rate based on hunter observations because of different observation rates of females with calves and females without calves. in the present study, however, the probability of detecting a female moose did not differ significantly in relation to her reproductive status. gustafsson and cederlund (1994) indicated similar results in a study with 16 radio-collared female moose during the hunting season. hence, we now have similar results from 2 independent studies indicating that reproductive status does not affect observation rate. accordingly, the number of moose seen per hunter day and the observed calves per female should not be affected by annual variation in the proportion of calf-rearing females in the population. a third factor that may affect the observation rate is the distribution of observations among hunting teams and to what extent hunting teams vary in the time spent hunting among years. for instance, ericsson and wallin(1999) found that the slope bealces vol. 39, 2003 rolandsen et al. detectability of moose 87 tween the moose seen per hunter-day and moose density differed among hunting areas, indicating that hunting teams differed in their efficiency in finding and recording moose; i.e., because observation skills or observation conditions varied among the different teams and hunting areas. variation among years in the number of days the different teams are hunting may therefore introduce variation in the observation index that is not caused by variation in moose density. indeed, in the present study, the 3 hunting teams that finished the hunt earliest had the highest proportion of observations (58%) compared to other teams (17%). because the practice in norway is to pool moose observation data over the whole hunting season (solberg et al. 1997), variation in the time spent hunting among teams and years may generate variation in the moose seen per hunter day even if the population density is stable. this may also explain why solberg and sæther (1999) found that moose seen per hunter-day tended to underestimate population size in years with low hunting success. during such years, teams with relatively low observation frequency may spend more time hunting to fulfill their quota, hence contribute many hunter-days and few observations to the index. because most moose in norway are harvested during the first week (solberg et al. 1997), prolonged hunting effort may lead to an index that more resembles the post-harvest population density than during years when the effective hunting season is shorter (solberg and sæther 1999). we therefore suggest that pooling data over more carefully selected periods of the hunting season (e.g., the first hunting week) may provide better indices of variation in population size and structure. by so doing, we would expect less variation among years in the period observations are collected in different zones, and therefore more similar conditions for observing moose among hunting teams. in turn, these indices may give more precise estimates of changes in population density and structure. acknowledgements we wish to thank all the hunters in bardu who reported their observations and harvest of radio-collared moose. we are also grateful to karl johan kjæreng for help with the collection of data. jon swenson gave valuable comments on earlier drafts of the manuscript. the study was funded by the norwegian research council ('changing landscapes'), the directorate for nature management (dn), and the norwegian institute for nature research (nina). references andersen, r., and b.-e. sæther. 1996. elg i norge. n.w. damm & søn a.s., teknologisk forlag. (in norwegian). caughley, g. 1974. bias in aerial surveys. journal of wildlife management 38:921933. cederlund, g., f. sandegren, and k. larsson. 1987. summer movement of female moose and dispersal of their offspring. journal of wildlife management 51:342-352. crichton, v. 1993. hunter effort and observations the potential for monitoring trends of moose populations a review. alces 29:181-185. ericsson, g., and k. wallin. 1994. antallet älgar som ses – bare en fråga om hur månge som finns. swedish university of agricultural sciences, department of animal ecology, umeå, sweden. (in swedish). , and . 1996. the impact of hunting on moose movements. alces 32:31-40. , and . 1999. hunter observations as an index of moose alces alces population parameters. wildlife biology 5:177-185. detectability of moose rolandsen et al. alces vol. 39, 2003 88 fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52:14-21. gustafsson, l., and g. cederlund. 1994. observerbarhet och förflyttningar hos älgar i samband med jakt. grimsö research station, swedish university of agricultural sciences, riddarhyttan, sweden. (in swedish). hochachka, w. m., k. martin, f. doyle, and c. j. krebs. 2000. monitoring vertebrate populations using observational data. canadian journal of zoology 78:521-529. moen, a. 1998: nasjonalatlas for norge. v e g e t a s j o n . s t a t e n s k a r t v e r k , hønefoss. (in norwegian). sæther, b.-e., and m. heim. 1993. ecological correlates of individual variation in age at maturity in female moose (alces alces): the effects of environmental variability. journal of animal ecology 62:482–489. , k. solbraa, d. p. sødal, and o. hjeljord. 1992. sluttrapport elg-skogsamfunn. the final report from the project “moose-forest-society”. nina forskningsrapport 28. (in norwegian with english summary). sas 1996. sas/stat software. changes and enhancements. sas institute, cary, north carolina, usa. solberg, e. j., m. heim, b.-e. sæther, and f. holmstrøm. 1997. oppsummeringsrapport. overvåk-ningsprogram for hjortevilt. elg 1991-95. nina fagrapport 30. (in norwegian with english summary). , and b.-e. sæther. 1999. hunter observations of moose alces alces as a management tool. wildlife biology 5:107-117. , , m . h e i m , a n d t . stubsjøen. 1999. elgbeiteregistreringer i bardu og målselv vinteren 1997/98. assessment of moose browse in bardu and målselv during the winter 1997/98). nina oppdragsmelding 590. (in norwegian with english summary). sweanor, p. y., and f. sandegren. 1988. migratory behaviour of related moose. holarctic ecology 11:190-193. tärnhuvud, t. 1988: utveckling av metoder for älginventering. flyginventering. – slutrapport. uppsala, 25 pp. (in swedish). williams, b. k., j. d. nichols, and m. j. conroy. 2002. analysis and management of animal populations. academic press, san diego, california, usa. persistent organic pollutants in the livers of moose harvested in the southern northwest territories, canada nicholas c. larter1, derek muir2, xiaowa wang2, danny g. allaire1, allicia kelly3, and karl cox3 1department of environment & natural resources, government of the northwest territories, po box 240, fort simpson, northwest territories, canada x0e 0n0; 2aquatic contaminants research division, environment and climate change canada, burlington, ontario, canada l7s 1a1; 3department of environment & natural resources, government of the northwest territories, po box 900, fort smith, northwest territories, canada x0e 0p0. alces vol. 53: e1–e33 (2017) analytical methods pcbs, ocos, and bfrs samples were thawed and thoroughly homogenized in a small stainless steel blender. subsamples (3.5 to 5 g ww) were soxhletextracted with dichloromethane; all were spiked with 1,3,5-tribromobenzene (trbb) prior to extraction. recovery of brominated compounds was monitored using bde-71, d16-ghbcd, and c13-bde-209. 13c12-pcb133 was added as a recovery standard for gpc performance for samples analyzed by als global. a laboratory blank consisting of all reagents and 2 nist reference materials (cod liver srm 1588b and fish muscle srm 1946) were also analyzed with the 14 samples. the extracts were then rotary evaporated under vacuum, exchanged into dcm: hexane (1:1) and applied to a gel permeation chromatography (gpc) column (60 g biobeads sx3) to remove lipids and other biogenic materials. the gpc eluate was reduced to 1 ml under vacuum. percent lipid was determined gravimetrically on a subsample of the extract or by evaporating the first gpc fraction. extracts were cleaned up on a silica gel column. nlet utilized activated silica gel (8 g, 1.1 cm i.d. column), and eluted with hexane followed by n-hexane/dcm (1:1) to separate pcbs from most of the pbdes and ocps. bde-209 was quantitatively eluted in the silica fraction 1. the als global split the gpc elution into separate ocp/oco fractions that were chromatographed on a 2% deactivated silica gel column and then reduced to 0.05 ml for analysis. the pcb fraction was cleaned up on an acid-silica gel column (45% w/w h2so4 on silica gel topped with neutral silica gel) then reduced to 0.04 ml for anlaysis. gc-ecd analysis was conducted on 7 dehcho samples using a gc-ecd (agilent 6890 gas chromatograph with a 63ni-electron capture detector [ecd]) using a 30 m x 0.25 mm (i.d.) db-5 column (internal film thickness 0.25 mm; j&w scientific, folsom, california, usa) with h2 carrier gas (constant flow rate 0.91 ml min-1). ultra-pure n2 was used as the makeup gas for the ecd (detector temperature: 325 c). the gc-ecd quantification of ocs in each sample was performed using a 4-point external standard calibration curve. calibration standards were quantified after every 10 samples. toxaphene-related compounds, including 22 polychlorinated bornane congeners as well as αand β-endosulfan and endosulfan sulfate, were quantified by gc-electron capture-negative ion mode (ecni) mass spectrometry using an agilent 6890 gc-5975 1 ms system as described by hoekstra et al. (2002). toxaphene and homologues were quantified using a “hercules” technical standard as described by glassmeyer et al. (1999). individual polychlorinated bornane congeners were quantified using a series of ex‐ ternal calibration standards (dr. ehrenstofer, augsburg, germany). alphaand betaendosulfan, and endosulfan sulfate were quantified by external standards using the characteristic fragment ions m/z 406 and m/z 273. analyses of all pbdes and other brominated flame retardants (bfrs) was carried out by gc-ecnims on an agilent 68905975 ms using an hp5-ms capillary column (30 m x 0.25 mm x 0.25 um film thickness). helium was the carrier gas, and separation was performed at a constant flow of 1.2 ml/min (muir et al. 2006). the mass spectrometer was operated in the nci mode, methane was the buffer gas, and temperature was 106, 150, and 300 °c for the quadrupole, the ion source, and the interface, respectively. the analytes were monitored at m/z 79/81 using an external standard calibration, except for c13-bde-209 which was monitored at m/z 493/495 and native bde-209 at m/z 487/ 485. any βand γ-hbcdd residues in the samples were most likely thermally isomerized to α-hbcdd in the gc injection port, thus, results represent total hbcdd (muir et al. 2006). pfass an internal standard mixture of 13c-pfass was added to every sample and extracted by shaking twice with acetonitrile. the extract was evaporated under nitrogen to dryness and reconstituted with 1 ml of methanol. the extract was cleaned with a graphite carbon solid phase cartridge (supelco). cleaned up extracts were analyzed for pfcas as well as pfsas. the analyses were performed by liquid chromatography with negative electrospray tandem mass spectrometry (lc-ms/ms). analytes were detected using an api 4000 q trap (applied biosystems, carlsbad, california, usa) after chromatographic separation with an agilent 1100 lc. chromatography was performed using an ace c18 column (50 mm x 2.1 mm, 3 µm particle size; aberdeen, united kingdom), preceded by a c18 guard column (4.0 x 2.0 mm, phenomenex) and the column oven was set to 30 °c. samples were quantified with a 6 point calibration curve and isotopic dilution method. quality assurance recoveries of internal standards ranged from 79% for δ-hch to 120% for pcb-204 (table s4) in dehcho samples analysed by gc-ecd, and from 59% for endrin ketone to 138% for 1245-ttbb in samples analysed by gc-hrms/lrms. a recovery spike demonstrated good recoveries of 32 ocp/ ocos (55-101%) and 18 pbde/bfrs (71– 133%) (table s5). slight losses of more volatile compounds (e.g., hexachlorobutadiene) and recovery enhancement due to contri‐ bution from laboratory reagent blanks (bde 47) explain the recovery variation. no corrections for recovery were made based on this information. analysis of the reference materials (nist srms 1588b and 1946) showed good agreement with all analytes quantified to within ±25% of certified values of ocp/ocos (17 compounds) and pcbs (29 congeners). 2 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 1 . l is t o f m o o se sa m p le s, co ll ec ti o n y ea r, ag e, se x, b io lo g ic al ch ar ac te ri st ic s, h ar v es t d at e an d lo ca ti o n . r eg io n id a g e s ex c o n d it io n e /g /f 1 h u n te r es t. ag e d at e h ar v es te d l at it u d e o n l o n g it u d e o w s o u th s la v e s s r -m o -1 1 -4 < 1 m f c al f 2 0 -f eb -1 0 6 0 .2 1 11 2 .1 3 s s r -m o -1 1 -7 1 m e a d u lt 1 0 -j an -1 0 6 0 .6 4 11 2 .2 7 s s r -m o -1 1 -8 4 m n /a a d u lt 4 -o ct -1 0 6 0 .7 6 11 2 .1 9 s s r -m o -1 1 -9 4 m g a d u lt 1 3 -d ec -1 0 6 0 .7 9 11 6 .2 2 s s r -m o -1 1 -1 0 < 1 m g y ea rl in g 3 0 -s ep -1 0 6 0 .8 5 11 4 .4 9 s s r -m o -1 1 -1 2 < 1 f g ca lf 8 -f eb -1 0 6 0 .7 8 11 2 .7 8 s s r -m o -1 1 -1 3 9 f n /a a d u lt 8 -f eb -1 0 6 0 .7 8 11 2 .7 8 d eh ch o g e n -1 4 3 m g a d u lt 2 1 -f eb -0 6 6 1 .5 5 1 2 1 .1 8 g e n -1 5 6 f g a d u lt 0 2 -m ar -0 6 6 1 .6 4 1 2 1 .0 6 g e n -1 6 1 2 f g a d u lt 0 6 -m ar -0 6 6 1 .6 0 1 2 1 .1 4 g e n -1 7 < 1 f g c al f 0 6 -m ar -0 6 6 1 .6 0 1 2 1 .1 4 g e n -1 8 < 1 m g c al f 2 4 -m ar -0 6 6 1 .5 0 1 2 0 .6 2 g e n -1 9 2 m e a d u lt 1 8 -m ar -0 6 6 1 .4 8 1 2 0 .6 2 jm r -4 5 f e a d u lt 2 4 -m ar -0 6 6 1 .4 9 1 2 0 .6 4 1 h u n te r co n d it io n ev al u at io n . e = ex ce ll en t, g = g o o d , f = fa ir . n /a = n o t av ai la b le . alces vol. 53, 2017 larter et al. – pops in moose livers 3 t ab le s 2 . l is t o f in d iv id u al o rg an oh al o ge n an al yt es al o ng w it h th ei r m d l s (n g /g w w ) u si ng ei th er g c w it h h ig h o r lo w re so lu ti o n m s , g c -n c im s , o r g c -e c d . m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d o c o h ex ac h lo ro b u ta d ie n e g c -h r m s , g c -e c d < 0 .0 0 5 0 .0 3 2 o c o 1 ,2 ,4 ,5 -t et ra ch lo ro b en ze n e 1 ,2 ,4 ,5 -t et ra ch lo ro b en ze n e g c -h r m s , g c -e c d 0 .0 9 9 0 .0 0 7 o c o 1 ,2 ,3 ,4 -t et ra ch lo ro b en ze n e 1 ,2 ,3 ,4 -t et ra ch lo ro b en ze n e g c -h r m s , g c -e c d 0 .1 0 8 0 .0 1 5 o c p p en ta ch lo ro b en ze n e p en ta ch lo ro b en ze n e g c -h r m s , g c -e c d 0 .0 7 7 < 0 .0 0 2 o c o 3 4 5 -t ri ch lo ro v er at ro le g c -h r m s , g c -e c d n a 0 .0 1 2 o c p p en ta ch lo ro an is o le g c -h r m s , g c -e c d < 0 .0 0 8 0 .0 2 4 o c o h ex ac h lo ro b en ze n e g c -h r m s , g c -e c d < 0 .0 2 6 < 0 .0 0 2 o c o 3 ,4 ,5 ,6 -t et ra ch lo ro v er at ro le g c -h r m s , g c -e c d < 0 .0 1 6 < 0 .0 11 o c p αh c h αh ex ac h lo ro cy cl o h ex an e g c -h r m s , g c -e c d < 0 .0 5 9 0 .0 1 5 o c p βh c h βh ex ac h lo ro cy cl o h ex an e g c -h r m s , g c -e c d < 0 .1 0 .0 1 3 o c p γh c h li n d an e g c -h r m s , g c -e c d < 0 .0 7 5 < 0 .0 0 2 o c p h ep ta ch lo r g c -h r m s , g c -e c d < 0 .0 1 8 0 .0 8 o c p p en ta ch lo ro n it ro b en ze n e g c -h r m s , g c -e c d < 0 .0 7 5 n a o c p a ld ri n g c -h r m s , g c -e c d < 0 .0 11 0 .0 2 6 o c p d ac th al g c -h r m s , g c -e c d < 0 .0 5 6 n a o c p o ct ac h lo ro st y re n e g c -h r m s , g c -e c d < 0 .0 2 5 0 .0 1 6 o c p h ep ta ch lo re p o x id e g c -h r m s , g c -e c d < 0 .0 6 9 0 .0 4 7 o c p o x y ch lo rd an e g c -h r m s , g c -e c d < 0 .0 1 9 < 0 .0 0 2 o c p tr an sch lo rd an e g c -h r m s , g c -e c d < 0 .0 3 7 0 .0 0 5 o c p ci sch lo rd an e g c -h r m s , g c -e c d < 0 .0 3 2 0 .0 1 2 o c p tr an sn o n ac h lo r g c -h r m s , g c -e c d < 0 .0 3 5 0 .0 6 2 o c p d ie ld ri n g c -h r m s , g c -e c d < 0 .0 1 8 0 .0 8 8 o c p ci sn o n ac h lo r g c -h r m s , g c -e c d < 0 .0 3 9 0 .0 2 4 o c p e n d ri n g c -h r m s , g c -e c d < 0 .3 8 7 0 .0 2 3 t ab le s 2 co n ti n u ed . . . . 4 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d o c p 2 4 'd d e g c -h r m s , g c -e c d < 0 .0 1 0 .0 0 1 o c p 4 4 'd d e g c -h r m s , g c -e c d < 0 .0 1 0 .0 2 1 o c p 2 4 'd d d g c -h r m s , g c -e c d < 0 .0 1 0 .0 1 4 o c p 4 4 'd d d g c -h r m s , g c -e c d < 0 .0 1 0 .0 2 o c p 2 4 'd d t g c -h r m s , g c -e c d < 0 .0 1 0 .0 2 o c p 4 4 'd d t g c -h r m s , g c -e c d < 0 .0 1 0 .0 2 o c p m et h o x y ch lo r g c -h r m s , g c -e c d < 0 .0 1 0 .0 1 9 o c p m ir ex g c -h r m s , g c -e c d 0 .0 6 8 0 .0 1 p c b s p c b -1 m o n o ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b -3 m o n o ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 4 /1 0 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 6 6 p c b s p c b 7 /9 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 5 8 p c b s p c b 6 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 4 2 p c b s p c b 8 /5 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 7 8 p c b s p c b 1 2 /1 3 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 1 .1 1 p c b s p c b 1 5 d ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 3 6 p c b s p c b 1 9 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 < 0 .0 0 2 p c b s p c b 1 8 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 11 p c b s p c b 1 7 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 2 7 /2 4 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 3 6 p c b s p c b 1 6 /3 2 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 1 9 p c b s p c b 2 6 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 1 2 p c b s p c b 2 5 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 9 t ab le s 2 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 5 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d p c b s p c b 3 1 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 4 5 p c b s p c b 2 8 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 5 p c b s p c b 2 0 /3 3 /2 1 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 5 5 p c b s p c b 2 2 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 9 7 p c b s p c b 3 7 tr ic h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 p c b s p c b 5 3 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 1 7 p c b s p c b 4 5 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 3 6 p c b s p c b 4 6 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 5 8 p c b s p c b 7 3 /5 2 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .9 1 3 p c b s p c b 4 3 /4 9 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 4 6 p c b s p c b 4 8 /4 7 /7 5 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 5 .1 5 2 p c b s p c b 4 4 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .5 5 4 p c b s p c b 5 9 /4 2 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 1 p c b s p c b 7 1 /4 1 /6 8 /6 4 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 1 7 p c b s p c b 1 0 0 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 5 2 p c b s p c b 6 3 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 4 p c b s p c b 7 4 /6 1 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 4 1 p c b s p c b 7 0 /7 6 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .6 6 3 p c b s p c b 8 0 /6 6 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .3 7 5 p c b s p c b 5 6 /6 0 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 2 9 p c b s p c b 8 1 te tr ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 1 3 p c b s p c b 9 5 /9 3 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .6 4 1 p c b s p c b 9 1 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 4 6 t ab le s 2 co n ti n u ed . . . . 6 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d p c b s p c b 9 2 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 5 6 p c b s p c b 8 4 /9 0 /1 0 1 /8 9 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 6 p c b s p c b 8 9 -1 0 1 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .6 9 2 p c b s p c b 9 9 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .3 1 9 p c b s p c b 11 9 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 2 p c b s p c b 8 3 /1 0 8 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 8 9 p c b s p c b 9 7 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 7 2 p c b s p c b 8 6 /1 11 /1 2 5 /1 1 7 /8 7 /1 1 6 /1 1 5 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 2 0 /8 5 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d 0 .0 5 5 0 .0 9 2 p c b s p c b 11 0 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 6 1 p c b s p c b 1 3 6 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 1 4 p c b s p c b 8 2 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 4 9 p c b s p c b 1 0 7 /1 0 9 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 5 9 p c b s p c b 1 2 3 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .4 0 6 p c b s p c b 11 8 /1 0 6 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 9 8 p c b s p c b 11 4 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 4 p c b s p c b 1 0 5 /1 2 7 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 4 1 p c b s p c b 1 2 6 p en ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 5 1 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 7 3 p c b s p c b 1 3 5 /1 4 4 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .6 5 1 p c b s p c b 1 3 9 /1 4 9 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 3 1 /1 6 5 /1 4 2 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 1 p c b s p c b 1 4 6 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 5 6 t ab le s 2 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 7 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d p c b s p c b 1 5 3 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 9 3 p c b s p c b 1 3 2 /1 6 8 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 3 5 p c b s p c b 1 4 1 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 6 8 p c b s p c b 1 3 7 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 3 .9 7 1 p c b s p c b 1 6 3 /1 6 4 /1 3 8 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .2 8 3 p c b s p c b 1 5 8 /1 6 0 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 4 4 p c b s p c b 1 2 9 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 8 p c b s p c b 1 5 9 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 2 8 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 8 p c b s p c b 1 6 7 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 p c b s p c b 1 5 6 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 3 p c b s p c b 1 5 7 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 < 0 .0 0 1 p c b s p c b 1 6 9 h ex ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 8 2 /1 8 7 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 0 9 p c b s p c b 1 8 3 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 4 6 p c b s p c b 1 7 4 /1 8 1 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 7 2 p c b s p c b 1 7 7 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 4 3 p c b s p c b 1 7 1 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 4 2 p c b s p c b 1 7 2 /1 9 2 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 8 p c b s p c b 1 9 7 o ct ac h lo o ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 0 7 p c b s p c b 1 8 0 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 8 7 p c b s p c b 1 9 3 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .1 3 2 p c b s p c b 1 9 1 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 < 0 .0 0 1 t ab le s 2 co n ti n u ed . . . . 8 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d p c b s p c b 1 7 0 /1 9 0 h ep ta ch lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 3 p c b s p c b -2 0 2 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 n a p c b s p c b 1 9 9 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 0 8 p c b s p c b 1 9 6 /2 0 3 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 2 2 p c b s p c b 1 9 5 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 5 p c b s p c b 1 9 4 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 0 5 p c b s p c b 2 0 5 o ct ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 < 0 .0 0 1 p c b s p c b 2 0 8 n o n ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 0 2 p c b s p c b 2 0 7 n o n ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 < 0 .0 0 1 p c b s p c b 2 0 6 n o n ac h lo ro b ip h en y l g c -l r m s , g c -e c d < 0 .0 0 2 0 .0 1 3 p c b s p c b 2 0 9 d ec ac h lo ro b ip h en y l g c -l r m s , g c -e c d 0 .0 8 0 .0 0 4 s u m o p an d p p 'd d t re la te d σ d d t s u m o x y ch lo rd an e. c & tch lo rd an e, c & tn o n ac h lo r, h ep ta ch lo r, h ep ta ch lo r ep o x id e σ c h l s u m α, β, γh c h σ h c h s u m te tr a, p ec b z, h c b σ c b z s u m α, β + su lf at e σ en d o su lf an s u m p c b s σ p c b s u m m o n o -d i c b s σ m o n o -d i s u m tr i c b s σ tr i s u m te tr a cb s σ te tr a s u m p en ta c b s σ p en ta s u m h ex a c b s σ h ex a t ab le s 2 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 9 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d s u m h ep ta c b s σ h ep ta s u m o ct a c b s σ o ct a s u m n o n ad ec a c b s σ n o n ad ec a e n d o su lf an αe n d o su lf an < 0 .0 0 2 e n d o su lf an βe n d o su lf an < 0 .0 0 2 e n d o su lf an e n d o su lf an su lf at e < 0 .0 0 2 p er fl u o ro su lf o n at es p f b s p er fl u o ro b u ta n e su lf o n at e l c -m s /m s < 0 .0 0 1 p er fl u o ro su lf o n at es p f h x s p er fl u o ro h ex an e su lf o n at e l c -m s /m s < 0 .0 0 1 p er fl u o ro su lf o n at es p f h p s p er fl u o ro h ep ta n e su lf o n at e l c -m s /m s < 0 .0 0 1 p er fl u o ro su lf o n at es p f o s p er fl u o ro o ct an e su lf o n at e l c -m s /m s 0 .2 9 9 p er fl u o ro su lf o n at es p f d s p er fl u o ro d ec an e su lf o n at e l c -m s /m s < 0 .0 0 1 p er fl u o ro su lf o n at es p f o s a p er fl u o ro o ct an es u lf o n am id e l c -m s /m s < 0 .0 0 1 p er fl u o ro su lf o n at es σ p f s a s s u m p er fl u o ro su lf o n at es l c -m s /m s p er fl u o ro ca rb o x y la te s p f h x a p er fl u o ro h ex an o at e l c -m s /m s 0 .1 7 9 p er fl u o ro ca rb o x y la te s p f h p a p er fl u o ro h ep ta n o at e l c -m s /m s 0 .0 3 6 p er fl u o ro ca rb o x y la te s p f o a p er fl u o ro o ct an o at e l c -m s /m s 0 .2 1 3 p er fl u o ro ca rb o x y la te s p f n a p er fl u o ro n o n an o at e l c -m s /m s 0 .0 4 6 p er fl u o ro ca rb o x y la te s p f d a p er fl u o ro d ec an o at e l c -m s /m s 0 .2 5 6 p er fl u o ro ca rb o x y la te s p f u n a p er fl u o ro u n d ec an o at e l c -m s /m s 0 .0 8 5 p er fl u o ro ca rb o x y la te s p f d o a p er fl u o ro u n d o d ec an o at e l c -m s /m s 0 .1 5 9 p er fl u o ro ca rb o x y la te s p f t a p er fl u o ro tr id ec an o at e l c -m s /m s 0 .0 8 9 p er fl u o ro ca rb o x y la te s p f t ra p er fl u o ro te tr ad ec an o at e l c -m s /m s 0 .1 3 1 p er fl u o ro ca rb o x y la te s σ p f c a s s u m p er fl u o ro ca rb o x y la te s l c -m s /m s t ab le s 2 co n ti n u ed . . . . 10 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d t o x ap h en e t o x ap h en e t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 t o x ap h en e h o m o lo g s h ex a t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 t o x ap h en e h o m o lo g s h ep ta t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 t o x ap h en e h o m o lo g s o ct a t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 t o x ap h en e h o m o lo g s n o n a t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 t o x ap h en e h o m o lo g s d ec a t ec h n ic al st an d ar d g c -e c n im s < 0 .0 5 c h lo ro b o rn an es 2 p ar la r 11 -1 2 g c -e c n im s < 0 .0 5 c h lo ro b o rn an es p ar la r 1 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es h ex -s ed g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 2 1 b 7 -4 9 9 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es h ep -s ed g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 2 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 3 2 b 7 -5 1 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 2 6 b 8 -1 4 1 3 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 3 1 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 3 8 b 8 -7 8 9 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 3 9 b 8 -5 3 1 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 4 0 -4 1 b 8 -1 4 1 4 /1 9 4 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 4 2 b 8 -8 0 6 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 4 4 b 8 -2 2 2 9 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 5 0 b 9 -1 6 7 9 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 5 1 b 8 -7 8 6 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 5 6 b 9 -1 0 4 6 g c -e c n im s < 0 .0 2 t ab le s 2 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 11 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d c h lo ro b o rn an es p ar la r 5 8 b 9 -7 1 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 5 9 b 9 -1 0 4 9 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 6 2 b 9 -1 0 2 5 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 6 3 b 9 -2 2 0 6 g c -e c n im s < 0 .0 2 c h lo ro b o rn an es p ar la r 6 9 b 1 0 -1 11 0 g c -e c n im s < 0 .0 2 p b d e s b d e 1 7 d ib ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 2 8 /3 3 tr ib ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 4 9 te tr ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 4 7 te tr ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 3 5 p b d e s b d e 6 6 te tr ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 0 0 p en ta b ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 9 9 p en ta b ro m o d ip h en y l et h er g c -e c n im s < 0 .0 2 2 p b d e s b d e 8 5 p en ta b ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 5 4 h ex ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 5 3 h ex ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 3 8 h ex ab ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 8 3 h ep ta b ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 1 9 0 h ep ta b ro m o d ip h en y l et h er g c -e c n im s < 0 .0 0 2 p b d e s b d e 2 0 9 d ec ab ro m o d ip h en y l et h er g c -e c n im s < 0 .5 5 8 s u m p b d e s σ p b d e s o th er b f r s t b p -a e a ll y l 2 ,4 ,6 -t ri b ro m o p h en y l et h er g c -e c n im s < 0 .0 0 2 o th er b f r s p t b x te tr ab ro m o x y le n e g c -e c n im s < 0 .0 0 2 o th er b f r s t b p -d b p e 2 -b ro m o al ly l 2 ,4 ,6 -t ri b ro m o p h en y l et h er g c -e c n im s < 0 .0 0 2 t ab le s 2 co n ti n u ed . . . . 12 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 2 co n ti n u ed m d l 3 m d l c la ss 1 c o m m o n n am e c h em ic al o r o th er n am e in st ru m en ta l an al y si s g c -m s g c -e c d o th er b f r s p b b e p en ta b ro m o b en ze n e g c -e c n im s < 0 .0 0 2 o th er b f r s t b c t te tr ab ro m o -o -c h lo ro to lu en e g c -e c n im s < 0 .0 0 2 o th er b f r s p b t o p en ta b ro m o to lu en e g c -e c n im s < 0 .0 0 2 o th er b f r s p b e b p en ta b ro m o et h y lb en ze n e g c -e c n im s < 0 .0 0 2 o th er b f r s d p t e /t b p -d b p e 2 ,3 -d ib ro m o p ro p y l 2 ,4 ,6 tr ib ro m o p h en y l et h er g c -e c n im s < 0 .0 0 2 o th er b f r s h b b h ex ab ro m o b en ze n e g c -e c n im s < 0 .0 0 2 o th er b f r s b b -1 0 1 p en et ab ro m o b ip h en y l g c -e c n im s < 0 .0 0 2 o th er b f r s p b b a p en ta b ro m o b en zy l ac ry la te g c -e c n im s < 0 .0 0 2 o th er b f r s e h t eb b 2 -e th y l1 -h ex y l 2 ,3 ,4 ,5 te tr ab ro m o b en zo at e g c -e c n im s < 0 .0 0 2 o th er b f r s h b c d d h ex ab ro m o cy cl o d o d ec an e g c -e c n im s < 0 .0 0 2 o th er b f r s b t b p e b is (t ri b ro m o p h en o x y ) et h an e g c -e c n im s < 0 .0 0 2 o th er b f r s b e h t b p b is (2 -e th y l1 -h ex y l) te tr ab ro m o p h th al at e g c -e c n im s < 0 .0 0 2 o c o sy n -d p s y n -d ec h lo ra n e g c -e c n im s < 0 .0 0 2 o c o an ti -d p a n ti -d ec h lo ra n e g c -e c n im s < 0 .0 0 2 o th er b f r s o b in d o ct ab ro m o tr im et h y lp h en y li n d an e g c -e c n im s < 0 .0 0 2 1 o c o = o th er ch lo ri n at ed o rg an ic , o c p = o rg an o ch lo ri n e p es ti ci d e re la te d , p c b = p o ly ch lo ri n at ed b ip h en y l, p b d e = p o ly b ro m in at ed d ip h en y l et h er , b f r = b ro m in at ed fl am e re ta rd an t. 2 n o m en cl at u re o f ch lo ro b o rn an es b as ed o n a n d re w s an d v et te r 1 9 9 5 . 3 m d l = m et h o d d et ec ti o n li m it = 3 * s d o f b la n k v al u es .“ n a” = n o t an al y se d .w h er e n o n d et ec t v al u es w er e p re se n t in b la n k s a“ < ” v al u es w er e u se d .t h es e ar e in st ru m en t d et ec ti o n li m it s b as ed o n s /n o f ap p ro x im at el y 1 0 :1 . alces vol. 53, 2017 larter et al. – pops in moose livers 13 table s3. comparison of gc-ms and gc-ecd analysis of 3 moose liver samples. see table s2 for full list analytes represented by each group. moose moose moose moose moose moose gc-ms ecd gc-ms ecd gc-ms ecd gen-14 gen-14 gen-15 gen-15r gen-16 gen-16 target analytes ng/g ww ng/g ww ng/g ww ng/g ww ng/g ww ng/g ww σddt <0.002 <0.002 <0.002 <0.002 <0.002 0.01 σchl 0.09 0.02 0.02 0.05 0.02 0.02 σhch 0.11 0.13 0.07 0.11 0.07 0.12 hcb 0.19 0.17 0.24 0.12 0.20 0.15 σpcb 0.62 0.30 1.07 0.74 1.03 0.89 smono-di 0.22 0.06 0.44 0.11 0.21 0.13 σ-tri 0.26 0.13 0.11 0.15 0.11 0.31 σ-tetra 0.08 0.08 0.33 0.19 0.42 0.27 σ-penta 0.05 <0.002 0.18 0.17 0.06 0.02 σ-hexa 0.02 0.01 0.02 0.03 0.02 0.07 σ-endosulfan 0.01 0.07 0.03 0.04 0.03 0.12 table s4. recoveries of internal standards during or prior to sample extraction. % recovery % recovery compound gc-ecd analysis (n = 7) sd gc-ms analysis (n = 10) sd 1,3-dbb 84.7 3.5 85.6 23.9 1,3,5-tbb 80.2 3.8 112 29.2 1,2,4,5-ttbb 87.7 4.3 138 35.3 δ-hch 79.0 5.2 110 39.3 endrin ketone 91.8 5.1 59.4 37.2 pcb 30 108 3.3 94.0 5.1 pcb 204 120 4.4 98.6 3.0 d16-ghbcdd1 84.1 17.7 13c-bde-2091 67.4 15.8 13c12-pcb133 82.3 8.0 1hbcdd and bde-209 were determined in n=7 samples by gc-ncims following gc-ecd analysis. 14 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 5 . r ec o v er ie s o f a o c p s an d p b d e st an d ar d sp ik e (n = 1 ) d u ri n g an al y si s o f th e m o o se li v er sa m p le s. a n al y te % a n al y te % a n al y te % p b d e 1 7 9 5 h ex ac h lo ro b u ta d ie n e 5 5 αen d o su lf an 9 7 p b d e 2 8 /3 3 1 0 1 1 ,2 ,4 ,5 -t t c b 7 0 ci sc h lo rd an e 7 2 p b d e 4 9 8 6 1 ,2 ,3 ,4 -t t c b 6 8 t ra n sn o n ac h lo r 6 7 p b d e 7 1 9 6 p e c b 6 6 d ie ld ri n 7 9 p b d e 4 7 1 3 8 3 ,4 ,5 -t ri ch lo ro v er at ro le 8 7 p ,p -d d e 6 7 p b d e 6 6 1 0 3 αh c h 6 3 o p -d d d 8 6 p b d e 1 0 0 11 6 βh c h 7 8 e n d ri n 6 9 p b d e 9 9 1 3 3 h c b 7 1 b -e n d o su lf an 7 7 p b d e 8 5 1 0 7 3 ,4 ,5 ,6 -t et ra ch lo ro v er at ro le 7 6 p ,p ’d d d 7 2 p b d e 1 5 4 1 0 0 p en ta ch lo ro an is o le 6 8 ci sn o n ac h lo r 6 0 p b d e 1 5 3 1 0 4 γh c h (l in d an e) 6 7 o ,p ’d d t 7 2 p b d e 1 3 8 8 5 h ep ta ch lo r 5 9 p ,p ’d d t 6 8 p b d e 1 8 3 9 2 a ld ri n 6 4 m et h o x y ch lo r 1 0 1 p b d e 1 9 0 9 2 h ep ta ch lo r e p o x id e 7 9 m ir ex 7 5 p b d e 2 0 9 7 1 o x y ch lo rd an e 8 1 h b c d d 7 3 o ct ac h lo ro st y re n e 6 2 b t b p e 7 8 tr an sch lo rd an e 6 2 d b d p e 11 7 o ,p ’d d e 7 2 alces vol. 53, 2017 larter et al. – pops in moose livers 15 t ab le s 6 . a ri th m et ic an d g eo m et ri c m ea n co n ce n tr at io n s o f o c p /o c o s an d p c b s in m o o se li v er fr o m th e d eh ch o an d s o u th s la v e re g io n s o f th e n o rt h w es t t er ri to ri es (n g /g w et w ei g h t an d li p id w ei g h t; n = 7 fo r ea ch re g io n ). “< ” v al u es ar e in st ru m en t d et ec ti o n li m it s w h er e ar it h m et ic m ea n s w er e < 0 .0 0 1 -0 .0 0 2 n g / g w w . d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw % li p id 6 .3 6 .2 5 .3 7 .2 5 .7 5 .7 5 .0 6 .5 h ex ac h lo ro b u ta d ie n e 0 .0 11 0 .0 1 0 0 .0 0 7 0 .0 1 4 0 .1 7 0 .1 7 0 .1 2 0 .2 4 0 .0 0 5 0 .0 0 4 0 .0 0 3 0 .0 11 0 .0 8 0 0 .0 7 4 0 .0 5 0 0 .1 6 1 ,2 ,4 ,5 -t et ra ch lo ro b en ze n e 0 .2 5 0 .2 4 0 .1 7 0 .3 0 4 .3 5 4 .2 2 2 .9 8 5 .5 0 0 .1 6 0 .1 4 0 .0 8 0 .3 5 2 .6 8 2 .5 2 1 .5 5 5 .3 8 1 ,2 ,3 ,4 -t et ra ch lo ro b en ze n e 0 .2 8 0 .2 7 0 .1 8 0 .3 5 4 .8 5 4 .6 8 3 .1 6 5 .8 7 0 .1 6 0 .1 5 0 .0 9 0 .3 6 2 .7 9 2 .6 4 1 .8 6 5 .5 4 p en ta ch lo ro b en ze n e 0 .1 1 0 .0 3 7 0 .0 0 8 0 .3 2 1 .8 5 0 .5 9 0 .1 1 5 .9 1 0 .1 7 0 .1 6 0 .1 0 0 .4 5 3 .0 0 2 .7 4 1 .7 5 6 .8 5 h ex ac h lo ro b en ze n e 0 .1 7 0 .1 7 0 .1 0 0 .2 4 2 .8 5 2 .6 7 1 .4 5 4 .3 3 0 .3 4 0 .2 6 0 .1 1 0 .9 3 6 .0 2 4 .6 4 1 .6 2 1 6 .6 3 ,4 ,5 ,6 -t et ra ch lo ro v er at ro le < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p en ta ch lo ro an is o le 0 .0 4 0 0 .0 1 7 0 .0 0 5 0 .1 5 0 .7 0 0 .2 8 0 .0 7 2 .6 6 0 .0 2 9 0 .0 2 3 0 .0 0 4 0 .0 7 0 .5 2 0 .4 1 0 .0 8 1 .2 3 αh c h 0 .0 4 1 0 .0 2 4 0 .0 0 6 0 .1 1 0 .7 0 0 .3 8 0 .0 9 1 .9 5 0 .0 5 6 0 .0 5 2 0 .0 3 1 0 .1 0 0 .9 9 0 .9 1 0 .4 9 1 .5 8 βh c h 0 .1 0 0 .0 9 7 0 .0 5 9 0 .1 7 1 .6 6 1 .5 6 1 .1 1 2 .9 3 0 .0 7 9 0 .0 7 2 0 .0 4 0 0 .1 7 1 .3 8 1 .2 7 0 .7 1 2 .5 4 γh c h 0 .0 5 6 0 .0 3 5 0 .0 0 5 0 .1 4 0 .9 5 0 .5 7 0 .0 7 2 .4 8 0 .0 6 3 0 .0 5 8 0 .0 3 2 0 .1 3 1 .1 1 1 .0 2 0 .5 6 1 .9 2 p en ta ch lo ro n it ro b en ze n e < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 h ep ta ch lo r < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 a ld ri n < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 d ac th al < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 o ct ac h lo ro st y re n e < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 h ep ta ch lo r e p o x id e 0 .0 4 4 0 .0 3 3 0 .0 1 0 0 .0 8 0 0 .7 1 6 0 .5 3 0 .1 6 1 .4 7 0 .0 4 6 0 .0 3 7 0 .0 1 7 0 .1 5 0 .7 8 0 .6 5 0 .2 7 2 .2 3 t ab le s 6 co n ti n u ed . . . . 16 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw o x y ch lo rd an e 0 .0 3 3 0 .0 2 0 0 .0 0 4 0 .1 0 0 .5 1 0 0 .3 2 0 .0 6 4 1 .4 8 0 .0 3 9 0 .0 2 5 0 .0 0 9 0 .1 4 0 .6 6 0 .4 4 0 .1 5 2 .0 8 tr an sch lo rd an e < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 ci sch lo rd an e 0 .0 1 7 0 .0 11 < 0 .0 0 2 0 .0 3 8 0 .2 9 8 0 .1 7 4 0 .0 1 4 0 .6 9 7 0 .0 2 6 0 .0 2 2 0 .0 1 0 0 .0 6 5 0 .4 5 0 .3 8 3 0 .1 4 8 1 .0 0 0 tr an sn o n ac h lo r 0 .0 3 1 0 .0 1 6 0 .0 0 4 0 .0 5 8 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 d ie ld ri n < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 ci sn o n ac h lo r < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 e n d ri n < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 αe n d o su lf an 0 .0 1 3 0 .0 1 0 0 .0 0 4 0 .0 2 4 0 .2 0 4 0 .1 6 0 .0 5 3 0 .3 5 0 .0 0 6 0 .0 0 4 < 0 .0 0 2 0 .0 2 5 0 .1 1 0 .0 6 6 0 .0 2 1 0 .5 0 βe n d o su lf an 0 .0 0 7 0 .0 0 6 0 .0 0 3 0 .0 1 5 0 .1 0 9 0 .0 9 2 0 .0 3 6 0 .2 2 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 e n d o su lf an su lf at e 0 .0 2 6 0 .0 1 5 0 .0 0 3 0 .0 9 2 0 .4 3 4 0 .2 4 0 .0 5 9 1 .7 3 0 .0 0 5 0 .0 0 4 0 .0 0 2 0 .0 0 8 0 .0 8 6 0 .0 7 9 0 .0 3 4 0 .1 3 o ,p ’d d e < 0 .0 0 2 < 0 .0 2 < 0 .0 2 < 0 .2 p ,p ’d d e 0 .0 1 3 0 .0 1 3 0 .0 1 3 0 .0 1 4 < 0 .0 2 < 0 .0 2 < 0 .2 o ,p ’d d d < 0 .0 0 2 < 0 .0 2 < 0 .0 2 < 0 .2 p ,p ’d d d 0 .0 1 7 0 .0 1 7 0 .0 1 7 0 .0 1 7 < 0 .0 2 < 0 .0 2 < 0 .2 o ,p ’d d t < 0 .0 0 2 < 0 .0 2 < 0 .0 2 < 0 .2 p ,p ’d d t < 0 .0 0 2 < 0 .0 2 < 0 .0 2 < 0 .2 m et h o x y ch lo r < 0 .0 0 2 < 0 .0 2 < 0 .0 2 < 0 .2 m ir ex 0 .0 0 4 0 .0 0 2 < 0 .0 0 2 0 .0 1 4 0 .0 7 1 0 .0 4 0 0 .0 1 4 0 .2 6 0 .0 0 4 0 .0 0 3 < 0 .0 0 2 0 .0 1 3 0 .0 7 3 0 .0 4 9 0 .0 1 2 0 .2 6 t o x ap h en e 0 .8 4 0 0 .6 4 2 0 .2 7 8 2 .1 7 1 3 .6 1 0 .3 4 .3 7 3 0 .7 1 .0 6 0 .8 5 0 0 .1 6 7 1 .8 3 1 8 .8 1 5 .0 2 .5 7 3 2 .7 h ex ac h lo ro b o rn an es 0 .0 2 6 0 .0 0 8 < 0 .0 0 2 0 .1 0 3 0 .4 1 9 0 .1 3 1 0 .0 1 4 1 .4 6 0 .0 1 2 0 .0 0 7 0 .0 0 5 0 .0 6 2 0 .2 3 5 0 .1 2 6 0 .0 7 7 1 .2 3 7 t ab le s 6 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 17 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw h ep ta ch lo ro b o rn an es 0 .3 6 6 0 .0 4 7 0 .0 0 3 1 .5 0 5 .8 3 0 .7 5 3 0 .0 3 9 2 1 .3 0 .1 2 3 0 .0 3 9 0 .0 0 5 0 .4 9 4 2 .0 7 7 0 .6 8 2 0 .0 7 7 7 .7 2 o ct ac h lo ro b o rn an es 0 .0 8 6 0 .0 3 9 < 0 .0 0 2 0 .2 5 1 1 .5 0 0 .6 2 9 0 .0 1 4 4 .7 2 0 .2 6 7 0 .1 7 4 0 .0 0 5 0 .4 3 1 4 .7 0 3 .0 7 0 .0 7 7 7 .7 0 n o n ac h lo ro b o rn an es 0 .3 1 8 0 .3 11 0 .2 2 5 0 .4 4 7 5 .1 0 4 .9 9 3 .5 4 6 .7 2 0 .4 8 1 0 .2 6 0 0 .0 0 5 1 .2 6 8 .5 0 4 .6 0 0 .0 7 7 2 2 .6 d ec ac h lo ro b o rn an es 0 .0 4 8 0 .0 4 3 0 .0 1 5 0 .0 8 5 0 .7 7 0 0 .6 8 3 0 .2 11 1 .1 8 3 0 .1 8 6 0 .1 0 4 0 .0 1 0 0 .5 2 5 3 .4 3 1 .8 3 0 .2 0 8 1 0 .5 p 2 6 < 0 .0 2 < 0 .0 2 < 0 .2 p 5 0 < 0 .0 2 < 0 .0 2 < 0 .2 p 6 2 < 0 .0 2 < 0 .0 2 < 0 .2 p c b -1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 p c b -3 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 p c b 4 /1 0 0 .0 0 5 0 .0 0 5 0 .0 0 4 0 .0 0 9 0 .0 8 6 0 .0 8 3 0 .0 5 1 0 .1 3 0 .0 2 8 0 .0 0 8 0 .0 0 5 0 .1 9 0 .5 0 2 0 .1 4 0 .0 7 7 3 .3 9 p c b 7 /9 0 .0 3 6 0 .0 2 0 0 .0 0 5 0 .0 6 4 0 .5 5 3 0 .3 2 8 0 .0 7 9 1 .0 7 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 6 0 .0 0 6 0 .0 0 6 0 .0 0 4 0 .0 0 8 0 .0 9 2 0 .0 9 0 0 .0 6 0 0 .1 3 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 8 /5 0 .0 1 6 0 .0 1 3 0 .0 0 5 0 .0 2 6 0 .2 1 3 0 .1 5 0 0 .0 2 0 0 .3 7 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 1 2 /1 3 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 5 0 .1 1 7 0 .0 11 < 0 .0 0 2 0 .4 2 0 2 .0 5 0 .1 7 2 0 .0 1 4 7 .7 1 0 .1 0 6 0 .0 4 8 0 .0 0 5 0 .3 3 0 1 .9 0 7 0 .8 5 5 0 .0 8 9 5 .8 9 p c b 1 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 8 0 .0 4 5 0 .0 4 5 0 .0 3 9 0 .0 5 3 0 .6 8 0 .6 7 6 0 .5 6 6 0 .7 6 < 0 .0 0 2 < 0 .0 2 p c b 1 7 0 .0 3 8 0 .0 2 5 0 .0 0 5 0 .0 9 0 0 .6 3 0 .3 9 6 0 .0 8 9 1 .6 5 0 .0 2 1 0 .0 1 5 0 .0 0 5 0 .0 6 0 0 .3 6 2 0 .2 5 8 0 .0 7 7 0 .9 2 p c b 2 7 /2 4 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 4 0 .0 4 4 0 .0 3 7 0 .0 1 6 0 .0 7 1 < 0 .0 0 2 < 0 .0 2 p c b 1 6 /3 2 0 .0 0 9 0 .0 0 4 0 .0 0 0 0 .0 1 8 0 .1 3 8 0 .0 6 0 0 .0 0 4 0 .2 6 < 0 .0 0 2 < 0 .0 2 t ab le s 6 co n ti n u ed . . . . 18 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw p c b 2 6 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 2 5 0 .0 0 6 0 .0 0 5 0 .0 0 2 0 .0 1 3 0 .1 1 3 0 .0 7 6 0 .0 3 3 0 .2 5 2 < 0 .0 0 2 < 0 .0 2 p c b 3 1 /2 8 0 .0 2 9 0 .0 1 5 0 .0 0 5 0 .0 8 5 0 .4 7 0 .2 3 5 0 .0 7 9 1 .6 0 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 2 0 /3 3 /2 1 0 .0 4 3 0 .0 1 4 0 .0 0 4 0 .2 4 0 .7 4 0 .2 1 7 0 .0 5 7 4 .2 6 0 .0 6 8 0 .0 1 5 0 .0 0 5 0 .4 3 5 1 .0 6 1 0 .2 5 9 0 .0 7 8 6 .6 9 p c b 2 2 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 0 3 0 .0 2 7 0 .0 2 0 0 .0 1 0 0 .0 4 4 < 0 .0 0 2 < 0 .0 2 p c b 3 7 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 4 6 0 .0 3 2 0 .0 1 4 0 .0 9 2 0 .0 1 8 0 .0 0 7 0 .0 0 5 0 .1 1 0 0 .3 5 1 0 .1 3 0 0 .0 7 7 2 .2 0 p c b 5 3 0 .0 0 4 0 .0 0 4 0 .0 0 4 0 .0 0 4 0 .0 8 3 0 .0 8 3 0 .0 8 3 0 .0 8 3 < 0 .0 0 2 < 0 .0 2 p c b 4 5 0 .0 11 0 .0 1 0 0 .0 0 8 0 .0 1 3 0 .1 8 0 .1 7 0 0 .1 2 0 0 .2 4 < 0 .0 0 2 < 0 .0 2 p c b 4 6 0 .0 0 6 0 .0 0 5 0 .0 0 3 0 .0 0 9 0 .1 0 0 .0 8 4 0 .0 4 4 0 .1 6 < 0 .0 0 2 < 0 .0 2 p c b 7 3 /5 2 0 .0 7 9 0 .0 2 0 < 0 .0 0 2 0 .2 7 1 .3 3 0 .3 3 0 .0 1 4 4 .2 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 4 3 /4 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 4 8 /4 7 /7 5 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 4 4 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 5 9 /4 2 0 .0 1 9 0 .0 0 7 < 0 .0 0 2 0 .0 3 6 0 .2 9 8 0 .1 1 8 0 .0 1 6 0 .6 7 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 7 1 /4 1 /6 8 /6 4 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 6 3 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 5 6 0 .0 5 6 0 .0 5 6 0 .0 6 < 0 .0 0 2 < 0 .0 2 p c b 7 4 /6 1 0 .0 1 0 0 .0 0 2 0 .0 0 0 0 .0 6 0 0 .1 6 8 0 .0 4 0 0 .0 0 7 0 .9 4 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 7 0 /7 6 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 8 0 /6 6 0 .0 6 8 0 .0 1 5 < 0 .0 0 2 0 .1 3 1 .0 5 4 0 .2 4 0 .0 1 6 2 .4 9 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 5 6 /6 0 0 .0 4 1 0 .0 0 8 < 0 .0 0 2 0 .1 2 0 .7 0 6 0 .1 2 0 .0 1 4 2 .2 0 0 .0 4 6 0 .0 3 1 0 .0 0 5 0 .1 4 0 .8 1 0 .5 4 6 0 .0 7 7 2 .5 0 t ab le s 6 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 19 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw p c b 8 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 9 5 /9 3 0 .0 0 2 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 3 7 0 .0 2 6 0 .0 1 4 0 .0 9 2 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 9 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 9 2 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 8 4 /9 0 /1 0 1 /8 9 0 .0 1 8 0 .0 0 3 < 0 .0 0 2 0 .1 0 0 0 .2 8 6 0 .0 4 0 0 .0 1 4 1 .8 3 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 9 0 .0 8 8 0 .0 7 7 0 .1 0 0 p c b 9 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 8 3 /1 0 8 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 9 7 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 8 6 /1 11 /1 2 5 /1 1 7 /8 7 / 11 6 /1 1 5 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 2 0 /8 5 0 .0 2 6 0 .0 0 8 < 0 .0 0 2 0 .0 7 5 0 .4 5 2 0 .1 2 2 0 .0 1 4 1 .3 8 0 .0 2 8 0 .0 2 5 0 .0 1 5 0 .0 4 0 0 .4 8 0 .4 5 0 .2 3 0 .7 1 p c b 11 0 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 8 2 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 0 7 /1 0 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 2 3 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 11 8 /1 0 6 0 .0 0 6 0 .0 0 6 0 .0 0 6 0 .0 0 6 0 .0 8 7 0 .0 8 7 0 .0 8 7 0 .0 8 7 < 0 .0 0 2 < 0 .0 2 p c b 11 4 0 .0 0 0 0 .0 0 0 0 .0 0 0 0 .0 0 0 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 0 3 < 0 .0 0 2 < 0 .0 2 p c b 1 0 5 /1 2 7 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 0 6 0 .0 3 1 0 .0 2 1 0 .0 1 4 0 .1 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 2 6 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 5 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 t ab le s 6 co n ti n u ed . . . . 20 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw p c b 1 3 5 /1 4 4 0 .0 0 8 0 .0 0 2 < 0 .0 0 2 0 .0 5 1 0 .1 4 9 0 .0 2 8 0 .0 1 4 0 .9 5 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 3 9 /1 4 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 3 1 /1 6 5 /1 4 2 /1 4 6 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 5 1 0 .0 3 7 0 .0 1 4 0 .0 9 2 0 .0 0 6 0 .0 0 5 0 .0 0 5 0 .0 1 0 0 .1 0 1 0 .0 9 6 0 .0 7 7 0 .2 0 0 p c b 1 5 3 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 5 5 0 .0 4 0 0 .0 1 4 0 .0 9 2 0 .0 1 4 0 .0 0 7 0 .0 0 5 0 .0 8 0 0 .2 6 0 .1 2 5 0 .0 7 7 1 .4 2 9 p c b 1 3 2 /1 6 8 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 4 1 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 4 8 0 .0 4 8 0 .0 4 8 0 .0 4 8 < 0 .0 0 2 < 0 .0 2 p c b 1 3 7 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 5 5 0 .0 4 0 0 .0 1 4 0 .0 9 2 0 .0 0 6 0 .0 0 5 0 .0 0 5 0 .0 1 0 0 .1 0 0 .0 9 6 0 .0 7 7 0 .1 7 9 p c b 1 6 3 /1 6 4 /1 3 8 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 5 8 /1 6 0 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 6 0 .0 1 6 0 .0 1 6 0 .0 1 6 < 0 .0 0 2 < 0 .0 2 p c b 1 2 9 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 0 3 0 .0 5 8 0 .0 5 8 0 .0 5 8 0 .0 5 8 < 0 .0 0 2 < 0 .0 2 p c b 1 5 9 0 .0 0 5 0 .0 0 3 < 0 .0 0 2 0 .0 1 2 0 .0 7 6 0 .0 4 9 0 .0 1 6 0 .2 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 2 8 /1 6 7 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 0 5 0 .0 8 5 0 .0 8 5 0 .0 8 5 0 .0 8 5 < 0 .0 0 2 < 0 .0 2 p c b 1 5 6 0 .0 0 2 0 .0 0 2 < 0 .0 0 2 0 .0 0 7 0 .0 3 8 0 .0 2 5 0 .0 0 8 0 .1 4 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 5 7 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 6 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 8 2 /1 8 7 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 11 0 .0 4 5 0 .0 2 5 0 .0 1 4 0 .2 0 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 8 3 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 1 4 0 .0 5 9 0 .0 2 9 0 .0 1 4 0 .2 6 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 7 4 /1 8 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 7 7 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 7 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 t ab le s 6 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 21 t ab le s 6 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw p c b 1 7 2 /1 9 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 6 0 .0 1 6 0 .0 1 4 0 .0 2 0 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 8 0 0 .0 0 3 0 .0 0 2 < 0 .0 0 2 0 .0 0 7 0 .0 4 1 0 .0 2 7 0 .0 1 4 0 .1 1 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 9 3 0 .0 0 9 0 .0 0 2 < 0 .0 0 2 0 .0 5 8 0 .1 7 0 .0 3 4 0 .0 1 4 1 .0 8 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 9 1 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 7 0 /1 9 0 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 11 0 .0 11 0 .0 11 0 .0 11 < 0 .0 0 2 < 0 .0 2 p c b -2 0 2 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 9 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 9 6 /2 0 3 0 .0 0 2 0 .0 0 2 < 0 .0 0 2 0 .0 0 5 0 .0 3 6 0 .0 3 0 0 .0 1 6 0 .0 9 0 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 1 9 5 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 1 9 4 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 0 2 0 .0 2 2 0 .0 2 1 0 .0 1 6 0 .0 3 3 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 2 0 5 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 2 0 8 0 .0 0 5 0 .0 0 2 < 0 .0 0 2 0 .0 3 2 0 .0 9 9 0 .0 2 7 0 .0 0 9 0 .5 9 3 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 0 2 0 .0 1 8 0 .0 1 8 0 .0 1 5 0 .0 2 0 p c b 2 0 7 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 2 0 6 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 p c b 2 0 9 < 0 .0 0 2 < 0 .0 2 < 0 .0 0 2 < 0 .0 0 2 < 0 .0 2 22 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 7 . a ri th m et ic an d g eo m et ri c m ea n co n ce n tr at io n s o f b ro m in at ed an d ch lo ri n at ed fl am e re ta rd an ts (n g /g w et w t an d n g /g li p id w t, n = 7 ). “< ” v al u es ar e in st ru m en t d et ec ti o n li m it s w h er e ar it h m et ic m ea n s w er e < 0 .0 0 1 -0 .0 0 2 n g /g w w . “n a” in d ic at es n o t av ai la b le . d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g / g lw n g / g lw n g / g lw n g / g lw b d e 1 7 < 0 .0 0 2 0 .0 0 1 < 0 .0 0 2 0 .0 0 2 0 .0 11 0 .0 11 < 0 .0 0 7 0 .0 2 2 < 0 .0 0 2 0 .0 0 1 < 0 .0 0 2 0 .0 0 2 0 .0 0 9 0 .0 0 9 < 0 .0 0 8 0 .0 1 0 b d e 2 8 /3 3 0 .0 0 3 0 .0 0 1 0 .0 0 1 0 .0 1 0 0 .0 4 6 0 .0 2 3 0 .0 0 7 0 .1 4 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 9 0 .0 0 9 0 .0 0 8 0 .0 1 0 b d e 4 7 0 .1 2 7 0 .0 3 7 0 .0 0 5 0 .5 1 3 2 .1 7 0 .5 8 5 0 .0 6 9 9 .4 0 0 .0 5 7 0 .0 4 8 0 .0 2 3 0 .1 3 3 0 .9 9 8 0 .8 4 5 0 .3 5 3 2 .3 7 0 b d e 4 9 0 .0 0 4 0 .0 0 2 0 .0 0 1 0 .0 1 2 0 .0 7 0 .0 3 0 .0 0 7 0 .2 2 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 9 0 .0 0 9 0 .0 0 8 0 .0 1 0 b d e 6 6 0 .0 0 2 0 .0 0 1 0 .0 0 0 0 .0 0 5 0 .0 3 2 0 .0 1 8 0 .0 0 4 0 .0 7 6 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 1 0 .0 0 9 0 .0 0 9 0 .0 0 8 0 .0 1 0 b d e 8 5 0 .0 0 6 0 .0 0 1 0 .0 0 1 0 .0 3 8 0 .1 11 0 .0 2 1 0 .0 0 7 0 .6 8 9 0 .0 0 4 0 .0 0 1 0 .0 0 1 0 .0 1 5 0 .0 6 3 0 .0 2 0 0 .0 0 8 0 .2 3 1 b d e 9 9 0 .1 4 8 0 .0 3 6 0 .0 0 5 0 .6 6 8 2 .5 6 0 .5 8 1 0 .0 6 9 1 2 .2 6 0 .0 8 1 0 .0 5 4 0 .0 1 9 0 .2 6 6 1 .4 1 0 .9 5 1 0 .3 0 0 4 .7 5 b d e 1 0 0 0 .0 2 6 0 .0 0 8 0 .0 0 1 0 .1 3 4 0 .4 6 6 0 .1 2 9 0 .0 0 7 2 .4 7 0 .0 1 8 0 .0 1 2 0 .0 0 4 0 .0 5 3 0 .3 0 4 0 .2 1 8 0 .0 6 3 0 .9 4 5 b d e 1 3 8 < 0 .0 0 2 0 .1 11 0 .0 2 1 0 .0 0 7 0 .6 9 < 0 .0 0 2 0 .0 6 3 0 .0 2 0 0 .0 0 8 0 .2 3 b d e 1 5 3 0 .0 1 4 0 .0 0 7 0 .0 0 1 0 .0 6 1 0 .2 5 0 0 .1 0 6 0 .0 0 8 1 .1 2 0 .0 0 8 0 .0 0 3 0 .0 0 1 0 .0 2 5 0 .1 3 5 0 .0 5 5 0 .0 0 8 0 .3 8 9 b d e 1 5 4 0 .0 1 0 0 .0 0 4 0 .0 0 1 0 .0 4 8 0 .1 8 1 0 .0 6 3 0 .0 0 8 0 .8 8 3 0 .0 0 5 0 .0 0 1 0 .0 0 1 0 .0 2 0 0 .0 8 6 0 .0 2 2 0 .0 0 8 0 .3 2 0 b d e 1 8 3 < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 b d e 1 9 0 < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 b d e 2 0 9 0 .1 5 0 .0 7 7 0 .0 0 4 0 .5 4 3 2 .2 6 1 .2 4 0 .0 7 7 7 .6 9 n a n a t b p -a e < 0 .0 0 2 < 0 .0 1 0 .0 2 7 0 .0 0 3 < 0 .0 0 2 0 .2 0 0 0 .4 9 1 0 .0 4 5 0 .0 0 8 3 .5 7 t b p d b p e < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 b e h t b p < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 t ab le s 7 co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 23 t ab le s 7 co n ti n u ed d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a ri th m ea n g m m in m ax a n al y te n g / g w w n g / g w w n g / g w w n g / g w w n g /g lw n g /g lw n g /g lw n g /g lw n g / g w w n g / g w w n g / g w w n g / g w w n g / g lw n g / g lw n g / g lw n g / g lw b t b p e < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 d p t e < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 e h t eb b < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 h b b < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 o b in d < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 p b b a < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 p b b e < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 p b e b < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 p b t o < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 p t b x < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 sy n -d p < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 an ti -d p < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 t b c t < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 h b c d < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 b b -1 0 1 < 0 .0 0 2 < 0 .0 1 < 0 .0 0 2 < 0 .0 1 24 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 8. a ri th m et ic an d ge om et ri c m ea n co nc en tr at io ns of in di vi du al p fa s s in m oo se li ve r sa m pl es (n g/ g w w ). d eh ch o d eh ch o d eh ch o d eh ch o s o u th s la v e s o u th s la v e s o u th s la v e s o u th s la v e n g /g w w n g /g w w n g /g w w n g /g w w n g /g w w n g /g w w n g /g w w n g /g w w a n al y te a m g m m in m ax a m g m m in m ax p f b a < 0 .1 6 < 0 .1 6 p f p ea < 1 .2 < 1 .2 p f h x a < 0 .1 5 < 0 .1 5 p f h p a < 0 .0 0 5 < 0 .0 0 5 p f o a 0 .0 6 9 0 .0 6 7 0 .0 5 6 0 .1 1 3 0 .0 9 0 0 .0 8 5 0 .0 5 0 0 .1 3 8 p f n a 0 .2 5 5 0 .2 4 3 0 .1 7 4 0 .4 4 8 0 .1 8 8 0 .1 7 1 0 .0 8 0 0 .3 7 0 p f d a 0 .2 4 0 0 .2 0 4 0 .0 7 8 0 .6 0 1 0 .1 5 9 0 .1 4 9 0 .0 6 9 0 .2 4 1 p f u n a 0 .1 6 7 0 .1 5 2 0 .0 9 6 0 .2 7 5 0 .0 9 4 0 .0 9 0 0 .0 5 8 0 .1 3 8 p f d o a 0 .0 0 8 0 .0 0 5 0 .0 0 3 0 .0 2 0 0 .0 1 7 0 .0 1 6 0 .0 0 8 0 .0 2 3 p f t ri a 0 .0 11 0 .0 0 6 0 .0 0 3 0 .0 3 2 0 .0 2 3 0 .0 2 1 0 .0 1 0 0 .0 4 1 p f t a 0 .0 0 4 0 .0 0 3 < 0 .0 0 3 0 .0 1 5 0 .0 1 2 0 .0 0 8 0 .0 0 3 0 .0 4 5 p f h x d a 0 .0 0 4 0 .0 0 3 0 .0 0 3 0 .0 1 0 0 .0 4 9 0 .0 3 2 0 .0 1 0 0 .1 7 3 p f b s 0 .0 5 5 0 .0 5 2 0 .0 2 6 0 .0 8 7 0 .0 2 9 0 .0 1 8 0 .0 0 3 0 .0 5 1 p f h x s 0 .0 6 3 0 .0 5 7 0 .0 2 2 0 .1 0 6 0 .0 1 2 0 .0 0 7 0 .0 0 3 0 .0 5 4 p f h p s < 0 .0 0 5 < 0 .0 0 5 p f o s 0 .4 3 7 0 .3 7 7 0 .2 1 0 0 .9 9 3 0 .2 5 2 0 .2 4 4 0 .1 7 4 0 .3 3 5 p f o s _ l in ea r 0 .1 8 9 0 .1 8 9 0 .1 7 3 0 .2 0 2 0 .1 8 3 0 .1 7 2 0 .0 8 8 0 .2 6 2 p f d s 0 .0 0 9 0 .0 0 5 0 .0 0 3 0 .0 3 4 < 0 .0 0 5 p f o s a 0 .0 0 8 0 .0 0 4 0 .0 0 3 0 .0 3 0 0 .0 1 2 0 .0 0 8 0 .0 0 3 0 .0 3 2 alces vol. 53, 2017 larter et al. – pops in moose livers 25 t ab le s 9 a . p ea rs o n co rr el at io n m at ri x fo r m aj o r h al o g en at ed o rg an ic co n ta m in an ts in m o o se li v er . c o rr el at io n co ef fi ci en ts in b o ld ar e si g n if ic an t at p < 0 .0 5 . a g e % li p id l o g σ p f c a l o g σ p f s a l o g σ 1 3 p b d e l o g σ p c b l o g σ e n d o su lf an l o g t o x ap h en e l o g σ d d t l o g σ c h l l o g σ h c h l o g σ c b z l o g σ m o n o _ d i c b l o g σ t ri -c b l o g σ t et ra c b l o g σ p en ta c b a g e 1 % li p id 0 .2 6 6 1 l o g σ p f c a 0 .2 4 8 0 .4 0 6 1 l o g σ p f s a 0 .0 7 6 0 .5 9 0 0 .6 9 9 1 l o g σ 1 3 p b d e 0 .1 8 7 0 .3 4 1 0 .0 1 8 0 .1 3 8 1 l o g σ p c b 0 .2 8 5 0 .0 4 2 � 0 .0 0 1 � 0 .0 3 3 0 .4 6 0 1 l o g σ e n d o su lf an 0 .0 0 6 0 .3 4 1 0 .3 2 8 0 .6 1 8 0 .0 0 5 0 .3 0 9 1 l o g t o x ap h en e � 0 .2 8 8 � 0 .1 5 2 � 0 .4 9 8 � 0 .5 0 5 � 0 .2 0 1 � 0 .0 4 6 � 0 .0 5 4 1 l o g σ d d t � 0 .3 6 8 0 .4 2 � 0 .0 3 8 0 .4 1 8 � 0 .0 8 6 0 .1 0 3 0 .6 2 3 0 .2 6 8 1 l o g σ c h l 0 .1 0 7 0 .1 0 3 0 .0 5 � 0 .0 9 6 0 .6 0 2 0 .3 5 8 � 0 .4 2 3 � 0 .3 1 4 � 0 .5 0 5 1 l o g σ h c h 0 .1 4 9 � 0 .2 4 0 � 0 .2 5 1 � 0 .4 0 5 � 0 .0 11 � 0 .2 9 4 � 0 .7 9 � 0 .1 9 5 � 0 .6 9 8 0 .5 1 4 1 l o g σ c b z 0 .3 8 � 0 .4 9 6 � 0 .1 6 3 � 0 .5 5 5 0 .0 3 2 0 .0 5 2 � 0 .6 2 2 � 0 .1 5 4 � 0 .8 2 1 0 .4 5 7 0 .7 5 1 l o g σ m o n o _ d ic b 0 .2 2 7 � 0 .5 5 6 � 0 .3 7 1 � 0 .8 2 9 � 0 .3 2 4 0 .1 4 7 � 0 .5 1 4 0 .2 1 2 � 0 .5 5 9 0 .0 8 9 0 .4 8 8 0 .6 7 1 1 l o g σ t ri -c b � 0 .0 2 8 � 0 .4 8 1 � 0 .1 1 8 � 0 .5 8 3 � 0 .3 2 5 0 .0 3 6 � 0 .6 4 3 � 0 .0 3 8 � 0 .5 0 2 0 .3 2 1 0 .5 9 4 0 .5 9 2 0 .7 8 1 1 l o g σ t et ra -c b 0 .4 8 5 � 0 .3 1 9 � 0 .4 2 6 � 0 .6 0 5 0 .1 5 4 0 .6 1 3 � 0 .2 8 5 0 .0 9 1 � 0 .2 6 7 0 .2 0 1 0 .2 3 3 0 .5 4 6 0 .7 0 9 0 .4 5 9 1 l o g σ p en ta -c b 0 .5 1 8 � 0 .3 8 4 � 0 .0 1 8 � 0 .3 3 9 � 0 .3 0 3 0 .2 2 3 � 0 .2 0 2 � 0 .2 2 9 � 0 .4 2 6 0 .0 5 9 0 .3 4 4 0 .6 2 8 0 .6 8 1 0 .6 6 4 0 .6 6 7 1 l o g σ h ex ac b � 0 .2 9 4 0 .4 2 1 0 .3 6 5 0 .5 7 6 � 0 .1 9 6 � 0 .0 8 1 0 .5 5 5 0 .2 7 7 0 .6 8 6 � 0 .3 0 9 � 0 .6 5 9 � 0 .7 1 7 � 0 .6 6 2 � 0 .4 5 3 � 0 .5 8 8 � 0 .4 1 6 26 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 9 b . p ea rs o n co rr el at io n m at ri x fo r se le ct ed in d iv id u al h al o g en at ed o rg an ic co n ta m in an ts in m o o se li v er . c o rr el at io n co ef fi ci en ts in b o ld ar e si g n if ic an t at p < 0 .0 5 . a g e % li p id l o g h c b d l o g 1 2 4 5 t ec b z l o g 1 2 3 4 t ec b z l o g p ec b z l o g h c b l o g p c a l o g αh c h l o g βh c h l o g γh c h l o g h ep ta -c h lo r ep o x id e l o g o x y ch lo rd an e a g e 1 .0 0 0 % li p id 0 .2 6 6 1 .0 0 0 l o g h c b d 0 .0 3 5 0 .5 6 9 1 .0 0 0 l o g 1 2 4 5 _ t ec b z 0 .3 4 1 � 0 .5 1 3 � 0 .6 0 8 1 .0 0 0 l o g 1 2 3 4 _ t ec b z 0 .3 4 7 � 0 .5 1 3 � 0 .5 9 9 1 .0 0 0 1 .0 0 0 l o g p ec b z 0 .3 4 2 � 0 .4 5 7 � 0 .5 5 2 0 .9 7 2 0 .9 7 2 1 .0 0 0 l o g h c b 0 .1 0 3 � 0 .4 2 6 � 0 .5 1 0 0 .4 6 7 0 .4 5 8 0 .4 7 6 1 .0 0 0 l o g p c a 0 .4 3 6 � 0 .2 9 8 � 0 .5 2 9 0 .7 7 8 0 .7 7 0 0 .8 2 8 0 .5 4 3 1 .0 0 0 l o g αh c h 0 .2 4 0 � 0 .5 8 7 � 0 .5 2 0 0 .9 2 0 0 .9 1 9 0 .8 7 2 0 .5 9 1 0 .6 6 4 1 .0 0 0 l o g βh c h � 0 .1 7 8 0 .4 0 8 0 .4 1 6 � 0 .1 3 0 � 0 .1 2 8 � 0 .1 1 7 � 0 .3 4 1 � 0 .2 1 7 � 0 .1 7 0 1 .0 0 0 l o g γh c h 0 .1 3 9 � 0 .3 5 1 � 0 .5 5 1 0 .6 4 1 0 .6 4 1 0 .6 9 7 0 .4 1 4 0 .5 3 7 0 .5 1 2 0 .1 5 5 1 .0 0 0 l o g h ep ta ch lo r ep o x id e 0 .3 2 9 0 .0 2 1 � 0 .0 5 9 0 .5 3 1 0 .5 3 4 0 .5 4 2 0 .1 1 6 0 .5 5 6 0 .3 2 7 0 .1 7 2 0 .3 0 2 1 .0 0 0 l o g o x y ch lo rd an e � 0 .0 3 8 0 .2 6 5 � 0 .0 2 7 0 .1 1 4 0 .1 1 2 0 .2 1 2 0 .3 3 5 0 .3 0 8 0 .1 4 6 0 .2 8 2 0 .3 6 4 0 .3 0 6 1 .0 0 0 l o g ci sch lo rd an e 0 .2 0 1 � 0 .5 9 4 � 0 .3 7 6 0 .7 2 1 0 .7 2 3 0 .6 7 9 0 .2 8 3 0 .4 6 2 0 .6 3 3 � 0 .2 8 2 0 .3 8 3 0 .5 8 6 � 0 .2 3 8 l o g αe n d o su lf an 0 .1 2 4 0 .3 6 9 0 .3 2 7 � 0 .3 2 0 � 0 .3 1 4 � 0 .2 8 3 � 0 .4 5 7 0 .0 2 3 � 0 .4 9 4 � 0 .0 2 5 � 0 .4 0 6 0 .3 8 0 � 0 .1 4 0 l o g en d o su lf an su lf at e 0 .0 7 2 0 .4 3 2 0 .6 3 5 � 0 .7 2 9 � 0 .7 2 6 � 0 .7 0 0 � 0 .4 9 9 � 0 .5 2 8 � 0 .6 5 0 � 0 .1 8 0 � 0 .6 9 3 � 0 .4 3 6 � 0 .4 4 4 l o g m ir ex � 0 .1 0 2 � 0 .4 6 4 0 .0 3 6 0 .1 8 6 0 .1 9 7 0 .1 2 7 � 0 .0 4 7 � 0 .2 7 4 0 .3 2 1 0 .1 3 4 0 .1 8 2 � 0 .1 5 4 � 0 .3 3 7 l o g b d e 4 7 0 .4 6 2 � 0 .2 2 5 � 0 .3 11 0 .6 3 1 0 .6 2 8 0 .6 3 2 0 .3 3 3 0 .7 3 6 0 .4 4 8 � 0 .0 6 8 0 .4 2 5 0 .8 3 0 0 .2 5 8 l o g p f n a 0 .1 4 8 0 .4 5 2 0 .5 4 8 � 0 .3 11 � 0 .3 0 6 � 0 .2 8 5 � 0 .4 9 8 � 0 .3 3 6 � 0 .4 3 7 0 .4 1 5 � 0 .0 7 1 0 .3 3 3 � 0 .0 4 0 l o g p f d a 0 .3 9 1 0 .3 3 9 0 .2 9 5 � 0 .2 11 � 0 .2 0 8 � 0 .2 6 4 � 0 .0 1 0 � 0 .2 3 4 � 0 .2 4 6 0 .1 5 2 � 0 .1 0 0 0 .2 9 6 � 0 .0 5 9 l o g p f u n a 0 .0 2 5 0 .5 0 2 0 .6 2 5 � 0 .6 7 7 � 0 .6 7 1 � 0 .6 9 4 � 0 .4 3 3 � 0 .6 6 5 � 0 .7 0 5 0 .0 8 5 � 0 .6 1 7 � 0 .0 1 0 � 0 .2 8 2 l o g p f b s � 0 .0 0 5 0 .2 4 5 0 .5 8 9 � 0 .1 8 5 � 0 .1 7 0 � 0 .1 7 3 � 0 .4 7 5 � 0 .2 4 6 � 0 .2 0 1 0 .7 4 3 � 0 .1 0 1 0 .2 1 3 0 .0 0 1 l o g p f h x s 0 .1 9 3 0 .3 5 9 0 .6 7 8 � 0 .4 0 8 � 0 .3 9 7 � 0 .3 3 6 � 0 .1 6 4 � 0 .1 4 9 � 0 .2 8 1 � 0 .1 6 0 � 0 .6 2 7 0 .0 5 3 0 .1 5 1 l o g p f o s � 0 .0 0 8 0 .5 5 5 0 .6 11 � 0 .7 0 5 � 0 .7 0 4 � 0 .7 1 5 � 0 .2 3 6 � 0 .5 9 0 � 0 .7 0 2 0 .3 6 8 � 0 .3 8 1 � 0 .0 2 3 � 0 .0 8 7 alces vol. 53, 2017 larter et al. – pops in moose livers 27 l o g o x y ch lo rd an e l o g ci sch lo rd an e l o g αe n d o sl fa n l o g en d o su lf an su lf at e l o g b d e 4 7 l o g p f n a l o g p f d a l o g p f u n a l o g p f b s l o g p f h x s a g e % li p id l o g h c b d l o g 1 2 4 5 _ t c b z l o g 1 2 3 4 _ t c b z l o g p ec b z l o g h c b l o g p c a l o g αh c h l o g βh c h l o g γh c h l o g h ep ta ch lo r ep o x id e l o g o x y ch lo rd an e l o g ci sch lo rd an e 1 l o g αe n d o su lf an � 0 .0 7 6 1 l o g en d o su lf an su lf at e � 0 .3 5 3 0 .4 5 9 1 l o g m ir ex 0 .4 4 4 � 0 .4 7 1 � 0 .0 4 4 1 l o g b d e 4 7 0 .6 3 5 0 .1 1 9 � 0 .5 3 1 � 0 .0 8 7 1 l o g p f n a 0 .1 5 1 0 .3 0 2 0 .3 4 6 0 .2 3 8 0 .1 8 1 l o g p f d a 0 .1 8 5 0 .1 2 8 0 .2 1 2 0 .2 2 8 0 .2 9 3 0 .7 3 1 l o g p f u n a � 0 .1 5 2 0 .3 8 2 0 .6 0 7 0 .0 5 3 � 0 .1 6 1 0 .6 9 5 0 .7 1 7 1 l o g p f b s � 0 .1 6 9 0 .3 2 3 0 .0 8 5 0 .2 7 6 � 0 .0 4 4 0 .3 8 0 .1 8 2 0 .2 3 7 1 l o g p f h x s � 0 .2 6 7 0 .4 8 2 0 .5 5 2 � 0 .2 0 9 � 0 .0 8 5 0 .1 2 6 0 .1 1 9 0 .4 1 9 0 .2 7 2 1 l o g p f o s � 0 .3 1 7 0 .3 1 3 0 .4 8 7 � 0 .0 1 5 � 0 .1 8 5 0 .6 5 8 0 .7 5 9 0 .8 4 2 0 .3 6 9 0 .2 5 7 28 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 9 c . p ea rs o n co rr el at io n m at ri x fo r se le ct ed p c b co n g en er s (c o -e lu te rs re p o rt ed as si n g le g c p ea k s) in m o o se li v er . c o rr el at io n co ef fi ci en ts in b o ld ar e si g n if ic an t at p < 0 .0 5 . a g e % l ip id l o g c b 1 5 l o g c b 3 1 _ 2 8 l o g c b 7 3 _ 5 2 l o g c b 5 6 _ 6 0 l o g c b 1 0 5 _ 1 2 7 l o g c b 1 5 3 l o g c b 1 5 6 l o g c b 1 8 0 a g e 1 % l ip id 0 .2 6 6 1 l o g c b 1 5 � 0 .4 5 5 0 .0 6 6 1 l o g c b 3 1 _ 2 8 0 .3 8 8 � 0 .4 2 4 � 0 .3 0 9 1 l o g c b 7 3 _ 5 2 � 0 .3 7 2 0 .3 5 1 0 .1 5 5 � 0 .7 2 1 1 l o g c b 5 6 _ 6 0 0 .3 9 5 � 0 .0 5 4 � 0 .2 8 5 0 .2 4 9 0 .2 1 3 1 l o g c b 1 0 5 _ 1 2 7 0 .4 2 4 � 0 .5 3 2 � 0 .1 8 7 0 .9 2 2 � 0 .7 7 0 0 .2 4 3 1 l o g c b 1 5 3 � 0 .2 3 9 � 0 .2 2 3 � 0 .0 0 3 � 0 .3 6 7 0 .6 4 6 0 .2 2 0 � 0 .3 11 1 l o g c b 1 5 6 0 .1 2 9 � 0 .5 7 2 0 .2 5 0 0 .3 2 6 � 0 .4 6 6 0 .0 4 3 0 .6 2 5 0 .0 1 4 1 l o g c b 1 8 0 � 0 .1 1 3 0 .0 2 1 � 0 .0 2 5 � 0 .5 0 0 0 .4 7 4 � 0 .1 2 3 � 0 .4 9 2 0 .7 9 6 � 0 .1 9 4 1 l o g c b 1 9 6 _ 2 0 3 � 0 .0 6 3 0 .1 3 9 � 0 .2 1 2 � 0 .5 3 8 0 .4 3 8 � 0 .1 8 1 � 0 .5 5 8 0 .6 1 2 � 0 .2 9 9 0 .8 8 6 alces vol. 53, 2017 larter et al. – pops in moose livers 29 t ab le s 1 0 a . c o m p ar is o n o f m ea n co n ce n tr at io n s (l o g n g /g w et w t) o f m aj o r h al o g en at ed o rg an ic s in m o o se li v er fr o m th e d eh ch o an d s o u th s la v e (s s r ) re g io n s u si n g th e s tu d en ts tte st as su m in g se p ar at e v ar ia n ce ; si g n if ic an t v al u es ar e b o ld ed (p < 0 .0 5 ). v ar ia b le r eg io n n m ea n s ta n d d ev m ea n d if fe re n ce l o w er 9 5 % l im it u p p er 9 5 % l im it t d f p -v al u e l o g σ p f c a d eh ch o 7 � 0 .1 5 2 0 .1 6 2 0 .0 5 7 � 0 .1 0 5 0 .2 1 8 0 .7 8 3 1 0 .1 2 9 0 .4 5 2 s s r 7 � 0 .2 0 9 0 .1 0 2 l o g σ p f s a d eh ch o 7 � 0 .2 8 4 0 .1 8 9 0 .2 4 5 0 .0 6 1 0 .4 2 9 3 .0 9 .3 8 2 0 .0 1 4 s s r 7 � 0 .5 2 9 0 .1 0 5 l o g σ p b d e d eh ch o 7 � 0 .5 8 9 0 .5 1 6 0 .3 5 4 � 0 .1 7 5 0 .8 8 4 1 .4 7 4 1 0 .9 2 3 0 .1 6 9 s s r 7 � 0 .9 4 3 0 .3 7 3 l o g σ p c b d eh ch o 7 � 0 .2 1 7 0 .1 7 7 0 .2 4 0 .0 1 6 0 .4 6 5 2 .3 4 1 11 .7 4 2 0 .0 3 8 s s r 7 � 0 .4 5 8 0 .2 0 6 l o g σ e n d o su lf an d eh ch o 7 � 1 .4 7 8 0 .3 7 3 0 .3 6 1 � 0 .0 0 2 0 .7 2 4 2 .2 3 8 9 .3 9 6 0 .0 5 1 s s r 7 � 1 .8 3 9 0 .2 0 8 l o g t o x ap h en e d eh ch o 7 � 0 .1 9 2 0 .3 2 9 � 0 .1 9 0 � 0 .5 3 3 0 .1 5 3 � 1 .2 1 5 11 .1 9 7 0 .2 4 9 s s r 7 � 0 .0 0 2 0 .2 5 0 l o g σ c h l d eh ch o 7 � 1 .0 5 2 0 .2 9 6 0 .0 3 3 � 0 .2 8 1 0 .3 4 6 0 .2 3 11 0 .8 2 3 s s r 7 � 1 .0 8 5 0 .2 3 7 l o g σ h c h d eh ch o 7 � 0 .8 5 9 0 .1 4 3 � 0 .1 2 3 � 0 .2 7 7 0 .0 3 � 1 .7 6 11 .5 5 3 0 .1 0 5 s s r 7 � 0 .7 3 6 0 .1 1 8 l o g σ c b z d eh ch o 7 � 0 .4 7 9 0 .4 4 3 � 0 .3 5 1 � 0 .7 6 7 0 .0 6 6 � 1 .9 5 4 7 .7 2 2 0 .0 8 8 s s r 7 � 0 .1 2 8 0 .1 7 l o g σ m o n o _ d ic b d eh ch o 7 � 1 .5 0 3 1 .0 8 6 � 1 .0 8 1 � 2 .1 1 8 � 0 .0 4 � 2 .3 7 5 8 .6 1 6 0 .0 4 3 s s r 7 � 0 .4 2 2 0 .5 2 0 l o g σ t ri -c b d eh ch o 7 � 1 .8 4 4 0 .8 3 0 � 0 .6 0 8 � 1 .4 0 2 0 .1 8 7 � 1 .7 3 9 8 .7 3 5 0 .1 1 7 s s r 7 � 1 .2 3 6 0 .4 0 8 l o g σ t et ra -c b d eh ch o 7 � 1 .3 0 1 0 .6 9 0 0 .0 0 9 � 0 .6 3 9 0 .6 5 8 0 .0 3 3 7 .7 1 2 0 .9 7 5 s s r 7 � 1 .3 1 0 .2 6 4 l o g σ h ex ac b d eh ch o 7 � 1 .3 4 9 0 .6 2 1 0 .3 4 2 � 0 .2 5 1 0 .9 3 6 1 .3 1 4 8 .6 3 5 0 .2 2 3 s s r 7 � 1 .6 9 2 0 .2 9 9 30 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 1 0 b . c o m p ar is o n o f m ea n co n ce n tr at io n s (l o g n g /g w et w t) o f m aj o r in d iv id u al h al o g en at ed o rg an ic co n ta m in an ts in m o o se li v er fr o m th e d eh ch o an d s o u th s la v e (s s r ) re g io n s u si n g th e s tu d en ts tte st as su m in g se p ar at e v ar ia n ce ; si g n if ic an t v al u es ar e b o ld ed (p < 0 .0 5 ). v ar ia b le r eg io n n m ea n s ta n d d ev m ea n d if fe re n ce l o w er 9 5 % c i u p p er 9 5 % c i t d f p -v al u e l o g h c b d d eh ch o 7 � 1 .9 8 2 0 .1 3 0 .4 0 6 0 .2 5 4 0 .5 5 8 5 .8 1 1 2 0 .0 0 1 s s r 7 � 2 .3 8 8 0 .1 3 1 l o g 1 2 4 5 -t ec b z jm r 7 � 1 .8 0 4 1 .1 1 9 � 0 .9 3 2 � 1 .9 6 8 0 .1 0 3 � 2 .1 8 4 6 .2 2 4 0 .0 7 s s r 7 � 0 .8 7 2 0 .1 5 3 l o g 1 2 3 4 -t ec b z jm r 7 � 1 .7 8 5 1 .1 4 4 � 0 .9 3 2 � 1 .9 9 0 .1 2 6 � 2 .1 3 9 6 .1 8 1 0 .0 7 5 s s r 7 � 0 .8 5 3 0 .1 4 l o g p ec b z d eh ch o 7 � 1 .4 3 6 0 .7 2 5 � 0 .5 9 9 � 1 .2 7 1 0 .0 7 3 � 2 .1 4 7 0 .0 7 2 s s r 7 � 0 .8 3 7 0 .1 6 l o g h c b d eh ch o 7 � 0 .7 7 9 0 .1 3 5 � 0 .2 2 9 � 0 .5 0 5 0 .0 4 6 � 1 .9 0 8 0 .0 9 2 s s r 7 � 0 .5 4 9 0 .2 9 l o g p c a d eh ch o 7 � 2 .0 3 0 .6 7 � 0 .3 9 7 � 1 .0 5 1 0 .2 5 6 � 1 .3 6 1 0 0 .2 0 4 s s r 7 � 1 .6 3 3 0 .3 8 1 l o g αh c h d eh ch o 7 � 1 .6 2 3 0 .5 1 3 � 0 .3 4 1 � 0 .8 1 8 0 .1 3 7 � 1 .6 8 3 7 .0 7 0 .1 1 8 s s r 7 � 1 .2 8 2 0 .1 5 4 l o g βh c h d eh ch o 7 � 1 .1 0 9 0 .0 8 1 0 .0 4 4 � 0 .1 0 8 0 .1 9 7 0 .6 5 7 8 .9 2 0 .5 2 3 s s r 7 � 1 .1 5 3 0 .1 5 9 l o g γh c h d eh ch o 7 � 1 .6 8 4 0 .3 1 3 � 0 .4 4 � 0 .7 4 1 � 0 .1 3 8 � 3 .3 0 7 8 .9 3 0 .0 0 6 s s r 7 � 1 .2 4 4 0 .1 6 0 .1 1 8 l o g h ep ta ch lo rep o x id e d eh ch o 7 � 1 .4 8 0 0 .3 8 7 � 0 .0 0 2 � 0 .3 8 5 0 .3 8 � 0 .0 1 4 9 .9 2 3 0 .9 8 9 s s r 7 � 1 .4 7 7 0 .2 3 6 l o g o x y ch lo rd an e d eh ch o 7 � 1 .7 0 .5 0 8 � 0 .0 4 0 � 0 .5 7 7 0 .4 9 7 � 0 .1 6 11 0 .8 7 3 s s r 7 � 1 .6 6 0 .4 0 1 l o g ci sch lo rd an e d eh ch o 7 � 1 .9 6 5 0 .5 5 2 � 0 .2 8 0 � 0 .8 0 3 0 .2 4 3 � 1 .2 3 8 0 .2 5 4 s s r 7 � 1 .6 8 6 0 .2 4 1 t ab le s 1 0 b co n ti n u ed . . . . alces vol. 53, 2017 larter et al. – pops in moose livers 31 t ab le s 1 0 b co n ti n u ed v ar ia b le r eg io n n m ea n s ta n d d ev m ea n d if fe re n ce l o w er 9 5 % c i u p p er 9 5 % c i t d f p -v al u e l o g αen d o su lf an d eh ch o 7 � 2 .0 9 1 0 .4 2 4 0 .3 5 3 � 0 .1 4 7 0 .8 5 4 1 .5 4 1 2 0 .1 5 0 s s r 7 � 2 .4 4 5 0 .4 3 5 l o g en d o su lf an su lf at e d eh ch o 7 � 1 .8 2 5 0 .4 7 9 0 .5 2 1 0 .0 6 2 0 .9 8 0 2 .5 8 9 0 .0 3 0 s s r 7 � 2 .3 4 6 0 .2 3 6 l o g m ir ex d eh ch o 7 � 2 .6 0 4 0 .4 3 7 � 0 .0 3 1 � 0 .5 3 6 0 .4 7 4 � 0 .1 3 5 1 2 0 .8 9 5 s s r 7 � 2 .5 7 3 0 .4 3 0 l o g b d e 4 7 d eh ch o 7 � 1 .4 3 7 0 .8 5 1 � 0 .0 9 7 � 0 .8 9 0 0 .6 9 7 � 0 .2 9 7 0 .7 8 3 s s r 7 � 1 .3 4 1 0 .2 7 7 l o g p f n a d eh ch o 7 � 0 .6 1 5 0 .1 4 2 0 .1 5 3 � 0 .0 5 7 0 .3 6 3 1 .6 1 11 0 .1 3 7 s s r 7 � 0 .7 6 8 0 .2 0 7 l o g p f d a d eh ch o 7 � 0 .6 9 0 .2 6 0 .1 3 8 � 0 .1 2 6 0 .4 0 2 1 .1 5 11 0 .2 7 4 s s r 7 � 0 .8 2 7 0 .1 7 9 l o g p f u n a d eh ch o 7 � 0 .8 1 7 0 .2 0 .2 2 6 0 .0 2 8 0 .4 2 5 2 .5 4 1 0 0 .0 2 9 s s r 7 � 1 .0 4 4 0 .1 2 6 l o g p f o s d eh ch o 7 � 0 .4 2 3 0 .2 4 3 0 .1 8 9 � 0 .0 4 3 0 .4 2 1 1 .8 6 9 0 .0 9 8 s s r 7 � 0 .6 1 2 0 .1 1 7 l o g p f b s d eh ch o 7 � 1 .2 8 0 .1 5 4 0 .4 6 7 � 0 .0 7 8 1 .0 1 2 2 .0 3 5 6 .8 2 7 0 .0 8 2 s s r 7 � 1 .7 4 7 0 .5 8 7 l o g p f h x s d eh ch o 7 � 1 .2 4 6 0 .2 2 6 0 .9 4 0 .4 8 1 .4 4 .6 6 3 8 .5 0 4 0 .0 0 1 s s r 7 � 2 .1 8 6 0 .4 8 3 l o g p c b 1 5 d eh ch o 7 � 1 .9 6 9 1 .2 9 � 0 .6 4 5 � 1 .9 0 3 0 .6 1 4 � 1 .1 5 1 0 0 .2 7 8 s s r 7 � 1 .3 2 5 0 .7 3 3 l o g p c b 1 7 d eh ch o 7 � 1 .6 0 7 0 .4 5 4 0 .1 9 9 � 0 .2 5 9 0 .6 5 8 0 .9 6 11 0 .3 5 8 s s r 7 � 1 .8 0 6 0 .3 0 7 t ab le s 1 0 b co n ti n u ed . . . . 32 pops in moose livers – larter et al. alces vol. 53, 2017 t ab le s 1 0 b co n ti n u ed v ar ia b le r eg io n n m ea n s ta n d d ev m ea n d if fe re n ce l o w er 9 5 % c i u p p er 9 5 % c i t d f p -v al u e l o g c b 2 0 _ 3 3 _ 2 1 d eh ch o 7 � 1 .8 6 9 0 .6 1 6 0 .0 6 4 � 0 .6 8 5 0 .8 1 2 0 .1 9 1 2 0 .8 5 5 s s r 7 � 1 .9 3 2 0 .6 6 8 l o g p c b 5 6 _ 6 0 d eh ch o 7 � 2 .1 0 9 1 .0 1 � 0 .6 7 5 � 1 .6 1 9 0 .2 6 8 � 1 .6 8 7 0 .1 3 5 s s r 7 � 1 .4 3 3 0 .3 4 2 l o g p c b 1 5 3 d eh ch o 7 � 2 .6 0 7 0 .3 6 8 � 0 .4 7 8 � 0 .9 6 2 0 .0 0 6 � 2 .1 6 11 0 .0 5 3 s s r 7 � 2 .1 2 9 0 .4 5 5 alces vol. 53, 2017 larter et al. – pops in moose livers 33 persistent organic pollutants in the livers of moose harvested in the southern northwest territories, canada analytical methods pcbs, ocos, and bfrs pfass quality assurance title_bkm_6 3904.p65 alces vol. 39, 2003 heikkilä et al. – moose and tree species composition 203 the impact of moose browsing on tree species composition in finland risto heikkilä1, pertti hokkanen1, mira kooiman2, nilay ayguney3, and cyril bassoulet4 1finnish forest research institute, vantaa research centre, p.o. box 18, 01301 vantaa, finland; 2stationsweg 46, 2991 rn badenrecht, the netherlands; 3konutkent 1-2 b-5 blok, c-girts no 4, 06530 ankara, turkey; 4enitab, apt. 110 rés montesquieu, 33170 gradignan, france abstract: the attitude usually adopted in finnish forestry regarding the moose (alces alces) has traditionally been that it influences scots pine in young mixed stands and therefore intensive treatments have been recommended to favor monocultures. the need to maintain diversity across the landscape is, however, changing attitudes. we tested the hypothesis that selective browsing can influence the composition of tree species in young stands, both in managed and natural forests. moose browsing effect on sapling heights was compared in exclosures and adjacent open areas in the southand mid-boreal forest zones of central and north finland at the end of the 1990s. moose appeared to impact young trees by reducing height growth, thereby reducing the possibility of selected broadleaved species to reach maturity. the number of aspen trees can obviously be expected to greatly decrease as a result of regenerating suckers being browsed by moose. rowan considerably declined under browsing pressure. on the other hand, the results also suggest that moose browsing may be beneficial by releasing conifers from competition among tree species in managed forests. in this sense, the relationship between browsed birches and the condition of conifers is crucial. browsing obviously reduces tree species diversity in areas of high moose density. however, some trees sheltered by neighboring ones are not browsed, and more information is needed about optimal treatment of young stands. in finland’s relatively small nature conservation areas, repeated browsing can quickly retard the height of slowly-regenerating broadleaf species. this browsing impact may lead to ecosystem changes without significantly impacting moose populations, the management of which by hunting is restricted in the set-aside natural forests and conservation areas. alces vol. 39: 203-213 (2003) key words: browsing, cleaning, damage, diversity, forest management, silviculture, tree species composition the ecology of moose (alces alces) in the boreal forests of finland is related to forest management. the moose population in the country was recently estimated to have reached its all time record level, which was 113,000-125,000 individuals in winter 2002 (ruusila et al. 2002). the current situation is similar to that in the late 1970s, when demands to reduce moose damage in forests and vehicle collisions resulted in a high degree of culling. since the adoption of intensive forest regeneration in the 1960s, the role of moose in relation to the characteristics of young stands has received research interest; specifically, how to mitigate the impact of moose browsing on young forest stands. since the late 1950s there has been discussion of the impact of moose browsing on forest succession; that is, how to minimize the impact of browsing on non-commercially treated young stands. especially young scots pine (pinus sylvestris) stands, which are often intensively cleaned, both mechanically and chemically in order to improve the condition of commercial stands. moose and tree species composition – heikkilä et al. alces vol. 39, 2003 204 the removal of broadleaf trees 5-10 years after regeneration has been commonly practiced. until the late 1980s it was argued that the presence of broadleaf trees in forest stands, aspen (populus tremula), rowan (sorbus aucuparia), and willows (salix spp.), is undesirable because they attract moose and augment browsing activity, thereby impacting growth of commercial species (yli-vakkuri 1956, löyttyniemi and lääperi 1988). the argument that the harmful impact of moose is likely related to the presence of broadleaf trees justified total removal of tree species favoured by moose. for instance, a total removal of these species has been recommended as a means to reduce the expected moose damage as moose move to their winter ranges (lääperi 1995). also, it was found that a relatively high broadleaf density could expose young pine to browsing risk (heikkilä and härkönen 1993). removing the most suitable moose forage means a considerable reduction in forage availability without it having a damage-reducing effect (härkönen 1998). the extreme opinions regarding the overall harmfulness of broadleaf trees provided the grounds for their large-scale total removal, and this policy is upheld by some people even today. after the termination of chemical treatment, a lot of mechanical work is now needed to keep stands clean manually. however, favouring of pine monocultures has been considered to be beneficial both from the viewpoint of growing pine as well as that of minimizing moose damage (kärkkäinen 1998). on the other hand, managing for an admixture of birch has been found to provide opportunities for alternatives in maintaining profitable forestry (mielikäinen 1980). nevertheless, total cleaning of deciduous mixtures is still relatively common in practice, although other methods are also being developed. new ecological and economical aspects are widely discussed to promote both forest yield and maintain the multiple goals of social interest groups. demands for diversity, forest protection, and the use of forests for recreation, call for new knowledge regarding both moose and forest management. recent demands for conserving biological diversity, as well as concerns over the status of nature conservation areas, raises new moose-related management issues. from the silvicultural viewpoint, there is the opinion that moose browsing only increases costs without any appreciable benefits. as for conserving biodiversity, the question is whether or not moose population densities can be kept high and not retard the maturation of certain species, which might then lead to long-term “unnatural” ecosystem changes in forest stands. in finland, the forest areas are mainly managed and nature conservation areas are relatively small. however, they are subjected to continuous browsing by moose populations, the density of which often is high in relation to food availability. therefore, it is important to define the ecological role of moose as a component of conserved natural ecosystems. we hypothesized that moose browsing can influence the development of young stands for several years after regeneration has started. the aim of this study is to compare the early succession of young stands with and without the impact of browsing. results of recent investigations in managed and set-aside forests are presented with reference to tree species composition. study areas the experimental areas were situated in 3 different managed forest areas in central finland. the forests represent typical small-scale management, where compartment size is 2-5 ha. the main commercial alces vol. 39, 2003 heikkilä et al. – moose and tree species composition 205 tree species are scots pine and norway spruce (picea abies), but birch species (betula pendula and b. pubescens) occur in patches within the conifer stands. all the experimental stands had been managed normally, which means planting 1 year after clearcutting and a light soil preparation. in kuru (n 61º55’, e 23º47’) 2 of the stands were planted with pine and 1 stand with spruce in 1990. the stand in keuruu (n 62º26’, e 24º14’) was planted with spruce in 1990 and in viitasaari (n 63º14’, e 25º27’) 6 stands were planted with pine in 1984. in viitasaari, the cleaning treatments were done in the same way, with the use of a brush saw to cut stumps to 10-20 cm in height, both within and outside the exclosures in 1988 (cf. härkönen et al. 1998). in other areas, cleaning was not needed during the study period. in the nature conservation areas, studies were conducted in 2 nature parks. one exclosure was established in vesijako, central finland (n 61º21’, e 25º06’) and the other in pisavaara, northern finland (n 66º16’, e 25º06’). in these parks natural regeneration of old spruce-dominated forests is slowly proceeding in openings made by disturbance. t h e e s t i m a t e d d e n s i t y o f m o o s e populations was 0.7-1.0 individuals per km2, according to information received from the game management districts (keskisuomen riistanhoitopiiri, yearly reports). methods in managed forests, exclosures were built in viitasaari and kuru in 1989 (25 x 50 m in size), and keuruu in 1994 (20 x 30 m in size). in the nature park of vesijako, the exclosure was built in 1996 (20 x 30 m in size), and in pisavaara in 1997 (30 x 40 m in size). the exclosures were placed randomly in young stands following the criterion that the stand characteristics should be similar in the exclosures and adjacent open control areas. comparisons were made between exclosures and adjacent open areas 5-10 m away. inventories were made by applying a network of 9 plots, each 20 sq. metres in area, situated systematically at similar distances inside and outside the exclosures. fieldwork was carried out in 2000 in kuru, keuruu, and vesijako, and in 2001 in viitasaari and pisavaara. the heights of trees > 50 cm were measured on all sample plots. moose browsing was determined by counting all browsed shoots per tree, which made it possible to estimate the browsing pressure for comparisons. the dominant height for each species, indicating the trees most likely to reach maturity in future stands, was calculated from the heights of the 100 tallest trees of each species. statistical computations were made using 1-way anova in comparisons among groups of trees and t-tests for individual species comparisons between exclosures and adjacent open areas (spss advanced statistics, spss incorporated, chicago, illinois, usa). results in the managed forest experimental stands, the mean heights of tree species differed significantly overall between the exclosures and adjacent open areas (f = 143.7, p < 0.001). all species were on average taller within than outside the exclosures, other than juniper (juniperus communis) and grey alder (alnus incana), which were not significantly different, and spruce, which was significantly taller outside than inside the exclosure (fig. 1). in managed forests of viitasaari and keuruu, moose, on average, had a significant impact only on willows and rowan, the mean heights of which remained at approximately 1 metre (table 1). in kuru, scots pine and broadleaf trees were significantly shorter in the open areas. norway spruce, on the other hand, was taller outside the moose and tree species composition – heikkilä et al. alces vol. 39, 2003 206 t ab le 1 . d en si ty a nd h ei gh t ( cm ) o f t re e sp ec ie s in e xc lo su re s (e ) a nd o pe n ar ea s (o ) o f y ou ng s ta nd s in m an ag ed fo re st s. m ea ns a re g iv en w it h th ei r st an da rd e rr or s. st ud y a re as k ur u k eu ru u v iit as aa ri t re es /h a h ei gh t t re es /h a h ei gh t t re es /h a h ei gh t t re e sp ec ie s e o e o e o e o e o e o sc ot s 2, 66 4 ± 86 5 2, 84 9 ± 1, 07 1 23 5 ± 9 19 5 ± 8* ** 25 9 ± 8 29 6 ± 9 68 ± 5 84 ± 8 2, 25 6 ± 31 2, 30 3 ± 74 42 3 ± 11 36 3 ± 41 pi ne n or w ay 2, 71 9 ± 1, 01 3 2, 25 7 ± 1, 12 2 11 3 ± 7 15 9 ± 8* ** 1, 29 5 ± 67 4 1, 20 3 ± 59 8 13 9 ± 8 14 6 ± 8 1, 50 4 ± 10 4 1, 66 4 ± 88 19 1 ± 24 13 5 ± 9 sp ru ce si lv er 2, 64 5 ± 91 8 3, 01 5 ± 97 6 21 2 ± 8 14 8 ± 5* ** 77 7 ± 11 78 1, 01 7 ± 37 6 16 0 ± 17 15 4 ± 12 81 8 ± 58 85 5 ± 65 20 2 ± 23 19 9 ± 29 b irc h d ow ny 8, 14 0 ± 1, 57 2 8, 65 8 ± 1, 89 9 16 9 ± 4 14 0 ± 2* ** 64 8 ± 1, 19 6 1, 05 5 ± 48 8 15 2 ± 11 17 9 ± 15 10 ,2 46 ± 2 26 12 ,4 36 ± 6 37 19 4 ± 5 17 8 ± 16 b irc h a sp en 92 1 ± 84 46 1 ± 35 12 7 ± 23 10 9 ± 13 w ill ow s 1, 36 9 ± 1, 01 1 1, 31 3 ± 1, 06 2 12 0 ± 7 73 ± 3 ** * 55 6 ± 10 2 40 7 ± 17 5 22 1 ± 25 79 ± 5 ** * 6, 16 6 ± 31 8 3, 93 8 ± 22 6 16 7 ± 8 12 7 ± 5* * r ow an 1, 44 3 ± 1, 11 4 96 2 ± 68 5 12 8 ± 7 67 ± 2 ** * 10 ,2 56 ± 2 ,0 11 9, 91 6 ± 1, 87 9 17 1 ± 3 10 8 ± 2* ** 1, 95 5 ± 88 78 0 ± 77 17 4 ± 7 93 ± 4 ** * a ld er 18 ± 1 1 23 5 ± 19 0 14 0 ± 46 36 4 ± 0 ju ni pe r 16 8 ± 95 28 0 ± 15 1 88 ± 1 2 82 ± 1 3 ** p < 0 .0 1, ** * p < 0 .0 01 . alces vol. 39, 2003 heikkilä et al. – moose and tree species composition 207 exclosure, where broadleaf trees had been kept low in height by browsing. the sprucedominated young stand in keuruu had an exceptionally closely-spaced stock of rowan, the average height of which was relatively low outside the exclosures. browsing by moose greatly reduced dominant tree heights, especially aspen, willows, and rowan (fig. 2). these species were transformed to low-growing vegetation, except for willows in viitasaari and rowan in keuruu. the relatively tall-growing willow species in viitasaari was sallow (salix caprea), which had not been severely browsed in some plots where they were protected by other neighbouring trees. the rowans in keuruu were abundant and because of stem density several individual trees of this species likely were not browsed (t = 2.256, p = 0.054). the dominant 100 individuals of tree species such as silver birch, aspen, willows, and rowan were, on average, significantly shorter under selective browsing outside than inside the exclosures in managed forests (f = 17.1, p = 0.001) (fig. 2). downy birch and dominant conifers were not affected, however. the dominant heights of scots pine and norway spruce were greater outside than within the exclosures in viitasaari (fig. 3), which was the oldest experimental stand. there the dominant height of all broadleaf species was significantly affected by browsing. in both managed and natural forests, moose had only occasionally impacted norway spruce and grey alder. however, regeneration of the latter species was not abundant but instead formed sparsely-distributed groups. the two naturally-regenerated young stands were different. the older stormcreated opening in pisavaara was rich in tree species, which were more abundant and taller, compared to the younger and more closed forest in vesijako (table 2). fig. 1. mean heights of tree species in exclosures and open areas in experimental young stands of managed forests (kuru, keuruu, and viitasaari). means are given with their standard errors. significance: **p < 0.01, ***p < 0.001. *** *** *** ** ** ** ** p in e s p ru c e s ilv e r b ir c h o w n y b ir c h a sp e n w ill o w s r o w a n j u n ip e r g re y a ld e r0 100 200 300 400 cm exclosure open height fig. 2. dominant heights of tree species in experimental young stands of managed forests (kuru, keuruu, and viitasaari). means are given with their standard errors. significance: ***p < 0.001. p in e s p ru c e s ilv e r b ir c h d o w n y b ir c h a s p e n w ill o w s r o w a n j u n ip e r0 200 400 600 800 *** *** *** *** cm height exclosure open moose and tree species composition – heikkilä et al. alces vol. 39, 2003 208 the impact of browsing in these two locations was considerable soon after regeneration had started. in the case of vesijako, the slowly-regenerating aspen and rowan vegetation was soon subject to browsing. in pisavaara, moose had also browsed intensively on both of the birch species, which was an indication of high browsing pressure. broadleaf species were losing their dominance in the dense norway spruce vegetation (fig. 4). in vesijako, the average differences were not as clear, despite the retarded growth of aspen. dominant height there was also low, due to browsing, when compared to dominant height within the exclosure (fig. 5). browsing intensities between the study areas could not be directly compared due to differences in age and tree species. the average number of browsed twigs did not significantly differ among the 3 managed forest areas (f = 2.6, p = 0.11). the highest values were obtained in kuru, 5.1 (± 1.3 se), and keuruu, 3.1 (± 1.5 se), while viitasaari provided the lowest, 1.6 twigs/ tree (± 0.4 se). as the stand in the latter area was several years older, the annual browsing pressure had evidently been relatively low. comparisons of browsing pressure in natural forests could not be made due to considerable differences in age and tree species composition. hares (lepus timidus) and black shoot blight (venturia populina) also caused some damage. the former occurred in some of the plots in viitasaari, but not frequently. the latter was relatively common among young aspen in the vesijako natural forest and occurred both within and outside the exclosures. discussion the average tree heights in young stands of managed finnish forests indicate that both scots pine and norway spruce are dominant, in keeping with silvicultural goals. the slight, but still significant, difference in mean heights of trees between exclosures and open areas was to be expected due to the impact of moose browsing. experimental forests can be regarded as being risky areas, where economically significant damage can occur. however, the estimated local moose population densities indicate that the expected intensity of damage does not usually jeopardize successful regeneration (cf. heikkilä and härkönen 1993). the study forests therefore represent relatively large areas where moose populations normally concentrate in winter in above-average densities. in these areas, considerable damage can occur in more restricted, smaller high-risk areas that cannot be exactly predicted because of yearly changes in moose movements. the average heights of broadleaved tree species consistently declined under browsing pressure. this occurs in young stands due to both silvicultural cleaning and fig. 3. dominant heights of tree species in experimental young stands of the viitasaari managed forest 11 years after fencing. means are given with their standard errors. significance: ***p < 0.001. p in e s p ru c e s ilv e r b ir c h d o w n y b ir c h a sp e n w ill o w s r o w a n j u n ip e r0 200 400 600 800 *** *** *** *** *** *** *** cm *** height exclosure open alces vol. 39, 2003 heikkilä et al. – moose and tree species composition 209 * p < 0 .0 5, ** p < 0 .0 1, ** * p < 0 .0 01 . t ab le 2 . d en si ty a nd h ei gh t ( cm ) o f t re e sp ec ie s in e xc lo su re s (e ) a nd o pe n ar ea s (o ) o f y ou ng s ta nd s in n at ur al fo re st s. m ea ns a re g iv en w it h th ei r st an da rd e rr or s. s tu dy a re as v es ija ko pi sa va ar a t re es /h a h ei gh t t re es /h a h ei gh t t re e sp ec ie s e o e o e o e o n or w ay s pr uc e 1, 61 0 ± 40 5 1, 39 0 ± 38 0* * 37 ± 6 25 ± 3 * 3, 88 8 ± 93 4 10 ,0 50 ± 2, 55 4* * 12 7 ± 10 10 5 ± 5* si lv er b irc h 2, 11 2 ± 49 1 2, 94 3 ± 1, 00 8* 15 5 ± 11 78 ± 5 ** * d ow ny b ir ch 1, 05 6 ± 31 5 3, 61 2 ± 1, 00 9 13 6 ± 16 86 ± 7 ** * a sp en 3, 22 0 ± 1, 43 5 3, 61 0 ± 1, 26 8 78 ± 5 56 ± 3 ** * 83 0 ± 25 0 1, 38 9 ± 65 4 10 7 ± 10 79 ± 6 ** w ill ow s 61 1 ± 10 2 11 1 ± 11 1 14 7 ± 23 60 ± 3 ** r ow an 88 9 ± 30 2 1, 11 1 ± 33 1* 69 ± 1 1 52 ± 4 ** * 7, 33 2 ± 1, 24 5 2, 11 6 ± 44 8* * 87 ± 3 70 ± 3 ** * a ld er 2, 44 2 ± 1, 17 3 72 1 ± 60 2 21 8 ± 24 21 9 ± 47 ju ni pe r 56 ± 0 11 2 ± 0 10 0 ± 0 60 ± 0 moose and tree species composition – heikkilä et al. alces vol. 39, 2003 210 the impact of moose browsing. cleaning greatly reduced the numbers of stems of both downy birch and aspen in viitasaari (härkönen et al. 1998), whereas the relatively wide spacing of rowan was obviously due to browsing. it is a matter of opinion how intensively early non-commercial treatment should be done, although excess broadleaved vegetation has to be removed to release conifer growth. moose browsing may be beneficial if the density of competing tree species is reduced, since the proportion of dominant pine is economically important in the future stands. the pines were taller in height outside the exclosures in viitasaari, indicating a release effect likely caused by moose browsing in open areas. thus the release may have compensated for the effect of moose stem breakages that can reduce pine height (heikkilä and löyttyniemi 1992). the per-tree effect of browsing declines with increasing food availability in terms of stand density (vivås and saether 1987, heikkilä and mikkonen 1992). according to studies by thompson and curran (1993) in newfoundland, moose browsing can be beneficial in thinning young dense stands. tree species selected by moose, such as aspen, willows, and rowan (bergström and hjeljord 1987), were generally of shorter height under browsing, and only in keuruu was the closely-spaced rowan vegetation still present in the dominant tree layer. mature trees of these species are likely to be reduced in numbers over large areas due to browsing. this depends also on the intensity of early cleaning as well as on stand treatments applied in first thinning cuttings. the sustainability of young trees varies between species (saether 1990), the resistance of birches to browsing being relatively high due to numerous twigs. these species react to browsing by producing numerous new shoots which are palatable to moose (bergström 1984). the response of sparsely-branched rowan easily leads to loss in height development, following which this tree species will most likely fail to reach *** *** *** ** ** s p ru ce s ilv e r b ir ch d o w n y b ir ch a sp e n w ill o w s r o w a n g re y a ld e r0 100 200 300 400 500 *** * cm exclosure open height fig. 4. dominant heights of tree species in experimental young stand of the pisavaara nature conservation area. means are given with their standard errors. significance: *p < 0.05, **p < 0.01, ***p < 0.001. fig. 5. dominant heights of tree species in experimental young stand of the vesijako nature conservation area. means are given with their standard errors. significance: **p < 0.01, ***p < 0.001. *** ** ** s p ru ce a s p e n r o w a n 0 40 80 120 160 200 cm exclosure open height alces vol. 39, 2003 heikkilä et al. – moose and tree species composition 211 maturity as it is a weak competitor to other species. conserving tree species diversity necessitates space to be created by silvicultural treatments favouring species selected by moose. the survival strategy of aspen is based on dense suckering (zackrisson 1985), whereas the spreading of abundantly-produced rowan seeds enables this species to reproduce far from mature trees. both of these species are known to favour relatively fertile sites where norway spruce is the main timber species. rowan has been found to regenerate abundantly in pinedominated young stands, and since it is a weak competitor, can be favoured in early stand treatments. conflicts may arise between economy and diversity-oriented ecology, if, for example, it is decided to favour fast-growing scots pine on fertile sites instead of ecologically-suitable norway spruce. because moose “take the remains”, browsing the reduced number of stems after stand treatments, a threat to aspen survival may arise, reducing this species over large areas in the future as a result of both forestry and browsing (heikkilä and härkönen 2001). in finland, however, the species selected by moose (e.g., aspen, rowan, and sallow) can become threatened by browsing mainly in relatively restricted, highest-density moose winter range areas. this conclusion can be drawn also from the results of national forest inventories conducted since 1950, in which the frequency of these species has been constant since 1995 (reinikainen et al. 2001). on the other hand, grey alder, not commonly browsed by moose, is ranked lower in numbers per hectare than moose-selected trees (finnish forest research institute 2001). in addition to moose, the occasional feeding by hares probably indicates low-density populations. in natural forests of vesijako and pisavaara, moose browsing significantly reduced regeneration of broadleaf species. the height differences indicate that it is hard for young broadleaf trees to reach heights above the snow level. birches were available in pisavaara, where they also were severely affected and remained low in height under browsing. the conditions in generally small nature conservation areas are exceptional, because the size of moose populations depends mainly on the surrounding forests. the number of moose utilizing broadleaf species in natural forests can be considerable from year to year. the carrying capacity of slowly regenerating nature conservation areas is relatively low. however, the situation is different from isle royale (michigan, usa), for example, where moose are controlled by wolves, and in island conditions do not likely move away as a result of reduced food resources. there the tree species diversity increased as a result of browsing, but the palatable tree species remained low-growing (risenhoover and maass 1987). in finland, on the other hand, the small natural forest areas, utilized by moose populations, are controlled only by hunting outside the conserved forests. consequently, the continuous consumption by moose in natural forests is not a natural element of the ecosystem because it depends on the decisions made for managing moose under the conditions of commercial forests. conclusions the broadleaved tree species selected by moose occur in varying densities in young stands. in managed forests, moose often “ t a k e t h e r e m a i n s ” f o l l o w i n g e a r l y silvicultural treatments in stands managed for conifer release. browsing causes damage to young pines and can also retard the height development of tree species such as aspen and rowan, reducing their competitive advantage and retarding maturation. in finnish forests, these species are still commoose and tree species composition – heikkilä et al. alces vol. 39, 2003 212 monly available, indicating a threat to tree species diversity to be found in restricted high-density moose areas only. on the other hand, the impact of moose browsing can be beneficial in releasing conifers from competition with broadleaf species. more knowledge is still needed for successful modelling of the structure of young stands and optimal stand treatments. silvicultural goals should also include diversity aspects important to moose in terms of food resources of good quality. the relationships between tree species abundance, consideration for diversity, and impact of moose browsing in managed forests constitute a complicated ecological-economic problem. regeneration of forest stands in small nature conservation areas can be negatively affected by continuous browsing: this does not result in a corresponding reduction in moose populations because of benefits afforded them by the surrounding managed forests. the fact that hunting is not allowed in conserved forests likely increases the browsing effect. consequently, ecosystem changes can be expected to occur in natural forests without negatively impacting moose populations, which will continue to challenge integration of moose and forest management objectives. acknowledgements we wish to thank sisko salminen and maija raahenmaa for preparing the data as well as two anonymous referees for valuable comments. references bergström, r. 1984. rebrowsing on birch (betula pendula and b. pubescens) stems by moose. alces 19:3-13. , and o. hjeljord. 1987. moose and vegetation interactions in northwestern europe and poland. swedish wildlife research supplement 1:213228. finnish forest research institute. 2001. the finnish statistical yearbook of foresty 2001. helsinki, finland. härkönen, s. 1998. effects of silvicultural cleaning in mixed pine-deciduous stands on moose damage to scots pine (pinus sylvestris). scandinavian journal of forest research 13:429-436. , r. heikkilä, w. e. faber, and å. pe h r s o n. 1998. the influence of silvicultural cleaning on moose browsing in young scots pine stands in finland. alces 34:409-422. heikkilä, r., and s. härkönen. 1993. moose (alces alces l.) browsing in young scots pine stands in relation to the characteristics of their winter habitats. silva fennica 27:127-143. , and . 2001. large mammals and forest ecosystem management. international symposium: ecosystem management in boreal forest landscapes. koli, finland. may 27-30, 2001. , and k. löyttyniemi. 1992. growth response of young scots pines to artificial stem breakage simulating moose damage. silva fennica 26:19-26. , and t. mikkonen. 1992. effects of density of young scots pine (pinus sylvestris) stand on moose (alces alces) browsing. acta forestalia fennica 231. k ä r k k ä i n e n , m . 1 9 9 8 . p u h t a a n männyntaimikon kasvatus kannattavinta. maaseudun tulevaisuus 17: 2. (in finnish). lääperi, a. 1995. hirvi. metsävahinkojen vähentäminen. 28 p. (in finnish). löyttyniemi, k., and a. lääperi. 1988. hirvi ja metsätalous. summary: moose in finnish forestry. university of helsinki, department of agricultural and forest zoology. report 13. m i e l i k ä i n e n , k . 1 9 8 0 . m ä n t y k o i v u s e k a m e t s i k ö i d e n r a k e n n e j a kehitys. summary: structure and develalces vol. 39, 2003 heikkilä et al. – moose and tree species composition 213 opment of mixed pine and birch stands. communicationes instituti forestalis fenniae 99.3.82s. reinikainen, a., r. mäkipää, i. vanhamajamaa, and j-p. hotanen, editors. 2 0 0 1 . k a s v i t m u u t t u v a s s a metsäluonnossa. summary: changes in the frequency and abundance of forest a n d m i r e p l a n t s s i n c e 1 9 5 0 . kustannusosakeyhtiö tammi, helsinki, finland. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17:357-364. ruusila, v., m. pesonen, r. tykkyläinen, and m. wallén 2002. hirvikanta lähes e n n a l l a a n s u u r i s t a k a a t o m ä ä r i s t ä huolimatta. riistantutkimuksen tiedote 180:1-12. (in finnish). saether, b-e. 1990. the impact of different growth patterns on the utilization of tree species by a generalist herbivore, the moose alces alces: implications for optimal foraging theory. behavioural mechanisms of food selection. nato asi series, volume g 20:323-340. thompson, i. d., and w. j. curran. 1993. a reexamination of moose damage to balsam fir white birch forest in central newfoundland: 27 years later. canad i a n j o u r n a l o f f o r e s t r e s e a r c h 23:1388-1395. vivås, h. j., and b-e. saether. 1987. interactions between a generalist herbivore, the moose alces alces, and its food resources: an experimental study of winter foraging behaviour in relation to browse availability. journal of animal ecology 56:509-520. y l i v a k k u r i , p . 1 9 5 6 . m ä n n y n kylvötaimistojen hirvivahingoista pohjanmaalla. summary: moose damage in seedling stands of pine in ostrobotnia. silva fennica 88:1-17. zackrisson, o. 1985. some evolutionary aspects of the life history characteristics of broadleaved tree species found in the boreal forest pages 17-36 in b. hägglund and g. peterson, editors. broadleaves in boreal silviculture – an obstacle or an asset? swedish university of agricultural sciences, department of silviculture. report 14. 203 previous meeting sites of the north american moose conference and workshop 1963 st. paul, minnesota 1964 st. paul, minnesota 1966 winnipeg, manitoba 1967 edmonton, alberta 1968 kenai, alaska 1970 kamloops, british columbia 1971 saskatoon, saskatchewan 8th 1972 thunder bay, ontario 9th 1973 québec city, québec 10th 1974 duluth, minnesota 11th 1975 winnipeg, manitoba 12th 1976 st. john’s, newfoundland 13th 1977 jasper, alberta 14th 1978 halifax, nova scotia 15th 1979 soldotna kenai, alaska 16th 1980 prince albert, saskatchewan 17th 1981 thunder bay, ontario 18th 1982 whitehorse, yukon territory 19th 1983 prince george, british columbia 20th 1984 québec city, québec 21st 1985 jackson hole, wyoming 22nd 1986 fredericton, new brunswick 23rd 1987 duluth, minnesota 24th 1988 winnipeg, manitoba 25th 1989 st. john’s, newfoundland 26th 1990 regina and ft. qu’apelle, saskatchewan 27th 1991 anchorage and denali national park, alaska 28th 1992 algonquin park, ontario 29th 1993 bretton woods, new hampshire 30th 1994 idaho falls, idaho 31st 1995 fundy national park, new brunswick 32nd 1996 banff national park, alberta 33rd 1997 fairbanks, alaska in conjunction with the 4th international moose symposium 34th 1998 québec city, québec 35th 1999 grand portage, minnesota 36th 2000 whitehorse, yukon territory 37th 2001 carrabassett valley, maine 38th 2002 hafjell, norway in conjunction with the 5th international moose symposium 39th 2003 jackson hole, wyoming 40th 2004 corner brook, newfoundland and labrador 41st 2005 whitefish, montana 42nd 2006 baddeck, nova scotia 43rd 2007 prince george, british columbia 44th 2008 6th international moose symposium, yakutsk, russia 45th 2009 pocatello, idaho future meetings 46th 2010 international falls, minnesota 47th 2011 jackson hole, wyoming f:\alces\vol_38\pagema~1\3803.pdf alces vol. 38, 2002 gosse et al. aircraft types and moose surveys 47 comparison of fixed-wing and helicopter searches for moose in a mid-winter habitat-based survey john gosse1, brian mclaren2 , and ewen eberhardt1 1terra nova national park, glovertown, nf, canada a0g 2l0; 2government of newfoundland & labrador, department of forest resources & agrifoods, p.o. box 2222, gander, nf, canada a1v 5t4 abstract: we conducted a mid-winter habitat-based survey in terra nova national park and an adjacent hunted area (moose management area 27) to compare the reliability and accuracy of using fixed-wing and helicopter aircraft for counting moose. forest inventory mapping was the primary consideration in defining block boundaries because this readily available information could be easily interpreted by observers during aircraft navigation, and because map classes could be chosen in a way expected to reduce variability in moose distribution. blocks were also classified from forest inventory mapping as being either open (mean crown closure of all stands < 50%), or dense (mean crown closure of all stands > 50%). we tested the precision of fixed-wing and helicopter aircraft for counting moose in blocks with open and dense crown cover by increasing the time spent during second searches with each aircraft type. more moose were seen in open blocks during second searches with increased flying time in both fixed-wing aircraft (100%) and helicopters (160%) than in dense forest cover blocks (12% and 43%, respectively). we also compared the accuracy of the two aircraft types in each crown cover class by recounting the same blocks at a similar intensity. verifying the accuracy of fixed-wing counts with helicopter searches of the same 8 blocks (the same crew flew approximately the same time), we found that the helicopter counts were on average 78% higher. we conclude that for highest accuracy and best classification of animals during a moose survey, helicopter counting is superior to fixed wing counting. alces vol. 38: 47-53 (2002) key words: fixed-wing, habitat-based survey, helicopter, moose, newfoundland, terra nova national park the stratified random block design (gasaway et al. 1986) is a widely used and recommended technique for surveying moose (alces alces) populations (timmermann and buss 1998). a consistent problem, especially for aerial census of forested habitats, is determining the number of animals missed due to poor visibility (samuel and pollock 1981, timmermann 1993). corrections for bias are often incorporated into a final population estimate, but visibility bias depends on many untested factors such as the study area, snow and weather conditions, observer experience, dominant vegetation, and the type of aircraft used in the survey (peterson and page 1993, rivest et al. 1995, anderson and lindzey 1996). another problem is that topographic maps are often used to navigate within blocks but may lead to imprecise survey results where there is insufficient detail for identifying block boundaries. creating survey blocks containing uniform habitat types can reduce variability in moose distribution within blocks (gasaway et al. 1986), and can provide an opportunity to assess moose visibility by recounting in uniform canopy cover situations. in this study, we used geographic information system (gis) technology and forest inventory databases to generate detailed field maps and to imaircraft types and moose surveys gosse et al. alces vol. 38, 2002 48 prove navigational and survey accuracy. wildlife managers in other regions have recently incorporated global positioning systems (gps) and on-board computer mapping to enhance navigation and overall survey efficiency (lynch and shumaker 1995, poole et al. 1999). gps was not used in this moose survey since block boundaries were designed to follow landscape features that were easily recognizable from the air. our objectives were to: (1) test the precision of using fixed-wing aircraft and helicopters to count moose in blocks with open and dense crown cover by increasing the time spent during second searches with each aircraft type; and (2) compare the accuracy of the 2 aircraft types in each crown cover class by recounting moose in the same blocks at a similar intensity. study area terra nova national park (tnnp; 48° 34' 00" n, 54° 00' 00" w) is a large protected area (410 km2) located in the north central boreal forest subregion (meades and moores 1994) of eastern newfoundland (fig. 1). the maritime climate in this area is characterized by brief, cool summers and relatively moderate winters. mean seasonal temperatures range from -5.8°c in february to 16.4°c in july, and mean annual precipitation is approximately 1,200 mm (deichmann and bradshaw 1984). topography is hilly (elevation < 200 m) and forests comprise about 70% of the area (gauthier et al. 1977). forest communities in tnnp are largely dominated by latesuccessional black spruce (picea mariana), although mixed balsam fir (abies balsamea) stands are prevalent along coastal areas. most forest stands are interspersed with fens, barrens, and small water bodies resulting in a naturally fragmented landscape. moose management area 27 (mma 27) is located to the immediate west of tnnp and is 3,620 km2 in area (fig. 1). forest types, climate, and topography are similar to tnnp. commercial timber harvesting has changed forest age distribution in the northern and eastern areas to younger age classes, with about half of stands < 40 years old and only 13% of stands > 80 years old. moose density was last assessed in 1989 at 1.7 / km2 (mercer 1995). the annual harvest in mma 27 is 200-300 moose per year (150-200 either-sex and 150-200 fig. 1. location of the 2 study areas, terra nova national park (tnnp) and moose management area 27 (mma 27) in eastern newfoundland. alces vol. 38, 2002 gosse et al. aircraft types and moose surveys 49 male-only licences have been allocated annually to this management area since 1990). methods a modification of the stratified random block survey (gasaway et al. 1986) was used to count moose in tnnp and mma 27 from mid-january to late march 2001. though the survey was conducted over a relatively wide time period, snow depth and condition did not change significantly and their potential effect on moose distribution was considered minimal. prior to stratification of the 2 units into regions of suspected uniform moose density, potential survey block boundaries were digitized using arcview gis 3.2. boundaries encompassed areas of similar habitat. exact block areas were calculated using the xtools extension for arcview; these ranged from 4-6 km2. forest inventory mapping was used to define all boundaries because observers could easily interpret this readily available information during aircraft navigation. map classes could be chosen in a way to reduce variability in moose distribution. easily discernible map classes chosen to aid in navigation included lake shorelines, streams, roadsides, forest edges along bogs, barrens and clearcuts, and abrupt transitions between stands of different species composition. continuous patches of productive and insect-damaged balsam fir were incorporated within single blocks wherever possible, moose winter distribution being primarily associated with these forest types. blocks were also classified from forest inventory mapping as being either open (mean crown closure of all cover types < 50%), or dense (mean crown closure of all cover types > 50%). stratification over both units was carried out in a cessna-185 aircraft flying parallel strips spaced approximately 200 m apart at an altitude of 50-150 m. a front-seat navigator and 2 rear-seat observers were instructed to find expected areas of uniform moose density for the purpose of identifying 2 strata (high and low moose density) in tnnp and 4 strata in mma 27 (areas were expected to be very high, high, low, and very low moose density). only 30% of mma 27 was flown during stratification because of the large area and finite resources for the project. the remaining area was stratified based on past survey results and expected moose densities in different habitat types. observers also made checks at this time of the assessment by gis of suitable census block boundaries and the designation of open and dense cover blocks. blocks were then randomly assigned to be censused according to stratum: sampling effort in the very highdensity stratum (mma 27 only) equaled 100% because there were only 4 blocks in this stratum, in the high-density stratum approximately 20%, in the low-density stratum approximately 5-10%, and in the very low-density stratum (mma 27 only) approximately 5%. whenever possible, census counts followed fresh snowfalls by < 48 hours and were always conducted when snow depth was > 60 cm. observer seat assignment was the same as during stratification and a similar flight pattern was followed. between 10 and 60 minutes were allotted for each census block. population estimates followed gasaway et al. (1986) and were derived from helicopter counting only, since most blocks were surveyed with this aircraft type. to test the reliability of fixed-wing and helicopter aircraft in blocks with open and dense crown cover we approximately doubled the amount of time spent during second searches and used new crew members in the same aircraft. doubling of effort was not always achieved (particularly for helicopter surveys) because of the high amount of time spent during initial searches. for this test, 6 blocks were recounted from fixed-wing aircraft (3 in open and 3 in aircraft types and moose surveys gosse et al. alces vol. 38, 2002 50 dense cover), and 8 blocks were recounted using helicopters (4 in open and 4 in dense cover). data collected from western newfoundland during january 2000 were used to supplement helicopter recounts in blocks with open crown cover since poor snow conditions prevented us from completing the desired replicate in this crown cover class during our survey. to compare the accuracy of fixed-wing and helicopter aircraft for sighting moose, the same crew spent approximately the same amount of time in each aircraft during recounts in 8 blocks (3 in open and 5 in dense cover). all recounts were done within 1-2 hours following the initial survey to reduce the probability of moose moving into adjacent blocks. moose were classified as male and female adults and unclassified yearlings and calves. sex determination was based on presence of the vulva patch and/or the p r e s e n c e o r a b s e n c e o f a n t l e r s (timmermann 1993). results survey results indicated that there were 308 ± 114 moose (90% ci) in tnnp and 2,140 ± 380 moose in mma 27. mean moose density was 0.75 and 0.59/ km2 in tnnp and mma 27, respectively. during the survey, a total of 72 moose were observed in tnnp and 301 moose in mma 27. of the total count in mma 27, 203 (67%) were observed in the very highdensity stratum, largely in recently cutover areas. more intensive second searches (100177% longer) from fixed-wing aircraft produced 36% more moose in the total counts for 6 blocks (table 1). second searches of blocks from helicopter with a more modest increase of flying time (31-144%), resulted in 74% more moose in 8 blocks (table 2). we found that the fixed-wing counts averaged 56% of the helicopter counts (42% s.d.) when counting the same 8 blocks (fig. 2). moose classification was possible from a helicopter but difficult from a fixedwing aircraft. about 80% of moose observed were classified into sex and ageclass from a helicopter, while only 8% of moose observed from the fixed-wing aircraft could be classified into age and sex categories. overall, more moose were seen in second searches of open blocks, 100% more (fixed-wing) and 160% more (helicopter), compared to dense forest cover blocks, 12% more (fixed-wing) and 43% more (helicopter; tables 1 and 2). in the fixed-wing and helicopter comparisons (fig. 2), count replicates 1-5 were in dense cover blocks and replicates 6-8 in open cover blocks. correction factors representing the increase in moose seen from a helicopter compared to a fixed-wing aircraft were 67% (58% s.d.) and 49% (36% s.d.) in open and dense blocks, respectively. discussion focusing block boundaries on habitat characteristics is an effective means to help delineate blocks, stratify, and assess accuracy of moose population surveys. observfig. 2. total number of moose counted in the same blocks by helicopter and fixed-wing aircraft. the aircraft correction factor, acf, is the ratio of counts in fixed-wing to the counts in helicopter and is reported with its standard deviation, sd. note that 28% less time was spent using helicopter for replicates 1, 7, and 8. alces vol. 38, 2002 gosse et al. aircraft types and moose surveys 51 ers felt that aircraft navigation and block boundary identification was facilitated by use of forest cover type maps. these maps also allowed us to use forest cover in our assessment of visibility bias. gasaway et al. (1986) recommended separate correction factors for each density stratum but variability in correction factors does not contribute much to the confidence interval in final population estimates in some areas (crête et al. 1986). we thus recommend generally that counting accuracy be assessed as much as possible under different flying conditions rather than by stratum. in our example, if at least half of the moose were missed in open blocks, then more than half of the moose present were missed because of poorer visibility through dense forest cover. this observation allows us to conclude that a correction factor of > 2 is table 1. flying time (min) and number of moose seen during first and second searches of open and dense blocks using fixed-wing (cessna-185) aircraft. the second search covered the same area as the first search, but was carried out by different observers in the same aircraft. first search second search cover replicate flying time moose flying time moose (min) count (min) count open 1 17 0 34 1 2 22 2 46 3 3 13 1 36 2 dense 1 30 4 60 4 2 38 4 76 5 3 15 0 30 0 table 2. flying time (min) and number of moose seen during first and second searches of open and dense blocks using a bell 206-b or 206-l helicopter. the second search covered the same area as the first search, but was carried out by different observers in the same aircraft. first search second search cover replicate flying time moose flying time moose (min) count (min) count open 1 25 3 49 3 2 7 0 13 6 3 13 2 19 2 4 13 0 17 2 dense 1 45 4 71 6 2 34 2 51 1 3 59 5 90 6 4 27 3 66 7 aircraft types and moose surveys gosse et al. alces vol. 38, 2002 52 warranted for largely forested survey units and is consistent with the findings of oosenbrug and ferguson (1992) who reported a mean sightability correction factor of 2.6 for 4 heavily forested blocks in eastcentral newfoundland. moose sightability decreases with an increase in forest cover (drummer and aho 1998, quayle et al. 2001) and this decline may be near zero visibility in the most dense cover classes (anderson and lindzey 1996). bergerud and manuel (1969) report considerable variation in the number of moose seen during recounts in open and dense cover blocks in central newfoundland. we see no support in current newfoundland survey procedures for the statement by gasaway et al. (1986) that > 95% of moose present in open canopied forests would be found had our search intensity been 4 min/km2; in fact, we exceeded this intensity in most of our sample blocks (including blocks in open areas). the best estimate of visibility bias during aerial survey may result from testing for the visibility of radio-collared animals (crête et al. 1986, peterson and page 1993, drummer and aho 1998). attempts have also been made to compare stratified random block survey to a mark-recapture population estimate with collared animals (oosenbrug and ferguson 1992), but random mixing of collared and uncollared moose in large survey areas may not be a sound assumption. for highest accuracy and best classification of animals during a moose survey, our study shows that helicopter counting is superior to fixed-wing counting. to save costs where classification results are not crucial to a survey, initial fixed-wing counts with limited helicopter recounting may achieve lower variability along with acceptable accuracy in population estimates (crête et al. 1986). our overall fixed-wing accuracy comparing helicopter counts (aircraft correction factor, acf = 0.56) was relatively higher than the 0.29 acf achieved by crête et al. (1986). this was probably the result of a combination of more open blocks and longer flying time. flying with fixed-wing produced lower variability in counting, with increased flying time (higher overall effort correction factor, ecf = 0.73; table 1) than helicopter (ecf = 0.58; table 2) in both open and dense cover. acknowledgements we thank all participating field staff from terra nova national park (tnnp) and the provincial department of forest resources and agrifoods (dfra). special thanks to truman porter, shawn avery, christine doucet (dfra), and janet feltham (tnnp) for extra assistance with data management. we are grateful to pilots baxter slade, craig moss (newfoundland helicopters), gene ploughman (thorburn aviation), rick adams, chris adams (springdale aviation), paul garrett, and ron whiffen (universal helicopters) for assistance during the survey. funding for this project was provided by tnnp and dfra. references anderson, c. r., jr., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24:247-259. bergerud, a. t., and f. manuel. 1969. aerial census of moose in central newfoundland. journal of wildlife management 33:910-916. crête, m., l. p. rivest, h. jolicoeur, j. m. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose with observations made from fixed-wing aircraft in southern quebec. journal of applied ecology 23:751-761. deichmann, k. h., and d. b. bradshaw. 1984. terra nova national park realces vol. 38, 2002 gosse et al. aircraft types and moose surveys 53 source description and evaluation. environment canada. terra nova national park, glovertown, newfoundland and labrador, canada. drummer, t. d., and r. w. aho. 1998. a sightability model for moose in upper michigan. alces 34:15-19. gasaway, w. c., d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. fairbanks, alaska, usa. gauthier, poulin, theriault, ltd. 1977. a biophysical classification of terra nova national park. a report for parks canada, atlantic region. historic properties, halifax, nova scotia, canada. lynch, g. m., and g. e. shumaker. 1995. gps and gis assisted moose surveys. alces 31:145-151. meades, w. j., and l. moores. 1994. forest site classification manual: a field guide to the forest types of newfoundland. forestry canada and newfoundland department of forestry and agriculture, frda report number 003. st. john’s, newfoundland, canada. mercer, w. e. 1995. moose management plan for newfoundland. report on file with wildlife division, newfoundland and labrador, st. john’s, newfoundland, canada. oosenbrug, s. m., and s. h. ferguson. 1992. moose mark-recapture survey in newfoundland. alces 28:21-29. peterson, r. o., and r. e. page. 1993. detection of moose in midwinter from fixed-wing aircraft over dense forest cover. wildlife society bulletin 21:8086. poole, k. g., g. mowat, and d. pritchard. 1999. using gps and gis for navigation and mark-recapture for sightability correction in moose inventories. alces 35:1-10. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37:43-54. rivest, l. p., f. potvin, h. crepeau, and g. daigle. 1995. statistical methods for aerial surveys using the double count technique to correct visibility bias. biometrics 51:461-470. samuel, m. d., and k. h. pollock. 1981. correction of visibility bias in aerial surveys where animals occur in groups. j o u r n a l o f w i l d l i f e m a n a g e m e n t 45:993-997. timmermann, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29:35-46. , and m. e. buss. 1998. population and harvest management. pages 559615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. 13 using snow urine samples to assess the impact of winter ticks on moose calf condition and survival daniel ellingwood1, peter j. pekins1, and henry jones2 1department of natural resources and the environment, university of new hampshire, durham, nh, 03824, usa; 2maine department of inland fisheries and wildlife, bangor, me, 04401, usa. abstract: snow urine samples collected in northern new hampshire, usa were used to measure urea nitrogen (un) and creatinine (c) content to develop ratios for tracking the nutritional restriction of individual moose (alces alces) through winter (2014–2017), inclusive of the adult winter tick (dermacentor albipictus) engorgement period. samples (n = 215) were collected from 55 moose (38 calves, 17 cows) on a twice monthly schedule from late january through snowmelt or calf mortality (march – early april). early winter un:c ratios from cows, surviving calves, and calves that ultimately died from infestation of winter ticks were similar and reflected a normal winter diet low in protein. a heightened un:c ratio (> 3.5 mg/dl) was measured in march which aligned with peak feeding by adult winter ticks, and presumably reflected accelerated protein deficit associated with blood loss. this increase was not observed population-wide despite shared habitat, occurring only in calves with mortal weight loss and anemia associated with heavy winter tick infestation. measurement of un:c ratios from snow urine samples proved an effective method to measure the temporal impact of winter tick infestation, and march samples can support other metrics used to estimate calf mortality. alces vol. 55: 13–21 (2019) key words: creatinine, epizootic, new hampshire, moose, nutritional restriction, snow urine, urea nitrogen, winter ticks new hampshire has experienced at least 5 winter tick (dermacentor albipictus) epizootics (> 50% calf mortality) in the last decade, an unprecedented rate of occurrence (musante et al. 2010, bergeron et al. 2013, jones et al. 2019). concurrent with this heightened frequency of epizootics, new hampshire’s moose (alces alces) population has declined ~45% over the past 15 years (nhfg 2015). the cause of this decline in northern new hampshire is the negative influence of winter ticks on calf survival which can be < 30% annually, and associated lower productivity of yearling and adult cows (musante et al. 2010, jones et al. 2017, 2019). the fate of moose calves (8–12 months old) during an epizootic is determined by their relative condition (i.e., body weight, fat stores) and the severity of tick infestation as it relates to metabolic impacts, including blood loss, protein deficiency, and subsequent weight loss (samuel 2004, musante et al. 2007, ellingwood 2018). as such, monitoring the condition of calves through winter and spring should provide insight into the relative and temporal influence of tick loads on metabolic imbalance, survival, and productivity. measurement of urea nitrogen from urine samples in snow serves as an informative measure of individual condition, reflecting muscle tissue catabolism in animals with an otherwise low-protein diet (seal et al. snow urine sampling – ellingwood et al. alces vol. 55, 2019 14 1972, kirkpatrick et al. 1975, bahnak et al. 1979). radio-marked calves provide for a unique opportunity to monitor and assess specific physiological parameters from individuals with known fate through the peak mortality period associated with winter tick parasitism. study area the study area (berlin) is located within coos county and includes sections of wildlife management units (wmu) b, c1, and c2 in the towns of berlin, milan, dummer, success, cambridge, millsfield, stark, and second college grant (fig. 1). fig.1. location of the study area in wildlife management units a2, b, c1, and c2 in coos county, new hampshire, usa. alces vol. 55, 2019 snow urine sampling – ellingwood et al. 15 the landscape is bisected by the androscoggin river and is relatively mountainous, bordered to the west by the kilkenny range and the south by the mahoosuc range. landcover is predominately commercial forest in which deciduous areas are dominated by yellow (betula alleghaniensis) and paper birch (b. papyrifera), american beech (fagus grandifolia), and sugar maple (acer saccharum), with softwood stands characterized by black spruce (picea mariana), red spruce (p. rubens), balsam fir (abies balsamea), and white cedar (thuja occidentalis) (degraaf et al. 1992). logging operations remove 1–3% of timber annually, and optimal moose habitat (4–16 year-old growth) increased 2.5x between 2001 and 2015 to equal > 17% of forest cover (dunfey-ball 2017). habitat quality is considered good and not a limiting factor to the local moose population (bergeron et al. 2011, dunfey-ball 2017). the average date of first snowfall is 14 november, with permanent snow cover typically in december (dunfey-ball 2017). this site was the location of a comprehensive study of moose population dynamics in 2001–2005 when density was estimated as ~0.8 moose/km2 (musante et al. 2010). the most recent population estimate is ~0.6 moose/km2, and from 2014–2018, > 200 moose were fit with radio-collars as part of a larger study. winter tick-related calf mortality was 62%, 74%, 77%, and 30% in 2014, 2015, 2016, and 2017, respectively (jones et al. 2019, p. j. pekins, unpublished data). methods capture and monitoring calves (~8 months old) were captured in early january 2014 – 2017 via net-gunning and aerial darting techniques (aero tech inc., clovis, new mexico, usa in 2014 and 2015; native range capture services, elko, nevada, usa in 2016 and 2017). moose were fitted with either vhf (n = 76; m2610b, advanced telemetry systems, isanti, minnesota, usa; mod-600, telonics, mesa, arizona, usa) or gps radio-collars (n = 104; gps plus vertex survey collar, vectronic aerospace gmbh, berlin, germany). the tick load of each moose was measured by summing the number of ticks on 4, 10 cm transects on both the shoulder and rump (sine et al. 2009, bergeron and pekins 2014). the sum of those 8 transects is used as an index to compare relative tick loads between individuals and across years. the vhf radio-collars transmitted at 55 pulses/min (ppm) while active, and switched to 110 ppm after 4 h without movement, signifying a mortality event. the gps radio-collars transmitted 2 location fixes/day (00:00 and 12:00 hr est) which were viewed and downloaded via vectronics aerospace software (gps plus x v10.4), and also had a vhf beacon transmitting at the same pulse rates as the vhf radio-collars. these radio-collars switched to mortality mode and sent a mortality alert via email after 5 h of inactivity. all calves were necropsied in the field following standard procedures (mason and madden 2007, munson 2015), including an external examination, body weight measurement, and collection of tissue samples for subsequent examination at the new hampshire veterinary diagnostics lab (durham, new hampshire) to determine cause of death (jones et al. 2019). urea nitrogen:creatinine sampling snow urine samples were used to track the nutritional restriction of individual animals through the winter, inclusive of the adult tick engorgement period. samples were collected from radio-marked calves and accompanying unmarked adult cows presumed to be the mothers. an effort was made to collect a sample from each individual snow urine sampling – ellingwood et al. alces vol. 55, 2019 16 every 2 weeks beginning in late january and extending through snowmelt or mortality (march – early april); the goal was to collect 3–6 temporal samples per individual. samples were collected within 24 h of urination by locating the bedding site and/or tracks in the snow at coordinates transmitted by the gps radio-collar at 00:00 hr that day; tracks were followed until a sample was identified. moose with vhf radio-collars were located using ground telemetry techniques (mech 1983) and back-tracked to collect urine samples. in cases where an adult accompanied the calf, samples were distinguished between the pair by the size of tracks and bed nearest to the sample. consistent with the methods used by delgiudice et al. (1988), the most concentrated portion was collected in plastic bags using rubber gloves to avoid contamination. samples were subsequently thawed at room temperature and aliquoted into 2 ml cryovials. these aliquots were stored frozen until submission to biovetusa (burnsville, minnesota) for measurement of urea nitrogen (un) and urinary creatinine (c) content (mg/dl). these data were expressed as a ratio (un:c) to correct for the dilution of each sample by snow (delgiudice et al. 1988); c is proportional to muscle mass and remains near constant in individuals over a given day (delgiudice and seal 1988). analysis included those individuals that were sampled most consistently, while attempting to achieve near equal representation of surviving and dead calves. the un:c ratio of urine samples were log e transformed to stabilize the variance in the dataset prior to analysis (delgiudice et al. 1989). samples were initially grouped by fate (mortality or survivor), and samples within each group were pooled at 2-week intervals (each individual was sampled every other week). a two-way anova was used to examine differences in the mean un:c ratio of samples collected from cows, surviving calves, and dead calves across collection intervals. tukey’s range (hsd) test was used to make post-hoc comparisons between multiple collection periods. results winter mortality of radio-marked calves in 2014 – 2016 averaged 71%, dropping to 30% in 2017; the leading cause of death (> 90%) was winter tick parasitism (ellingwood 2018, jones et al. 2019). a total of 158 snow urine samples were collected from 38 radio-marked calves (23 winter mortalities and 15 survivors) in the winters of 2014–2017. samples (n = 57) were also collected from 17 unmarked adult cows accompanying a portion of these marked calves; all adult cows were presumed to survive through winter. maximum annual snow depth during the collection periods ranged from 17.8 – 71.1 cm. from ~15 january – 14 february, there were no statistical differences between un:c ratios of unmarked cows, surviving calves, and calves that died (p > 0.05). from ~ 15 feb – 15 march, the mean un:c ratio of calves that died and those that survived diverged significantly (p < 0.05), whereas the un:c ratio of cows and surviving calves remained similar. the un:c ratios of unmarked adult cows stayed stable throughout the sampling season (x̅ = 2.35 ± 0.85), with individual samples ranging from 0.8–4.54 with no difference detected between collection intervals (p > 0.05; fig. 2). the average un:c ratio of calves that died increased through the second week of march, peaking at 4.68 ± 2.93; the ratio of survivors was lower and relatively stable during this same time period (x̅ = 2.43 ± 0.74; table 1). in the weeks following this spike in un:c ratios, the average ratio of calves that died never returned to levels < 3.5, and mortality occurred 1–5 weeks after their un:c ratio peaked (x̅ = 3 weeks). alces vol. 55, 2019 snow urine sampling – ellingwood et al. 17 spring snowmelt prevented sample collection from 14 calves within 2 weeks of death. discussion moose are on a negative nutritional plane from winter through early spring, although winter mortality from malnutrition is considered rare in the northeastern united states. adult survival rates in late winter average 97% and starvation has not been identified as a source of mortality regionally (musante et al. 2010, jones et al. 2019). a low level of undernutrition during winter is typical and reversible, as fat stores are used to supplement the reduced nutritional quality of winter browse (delgiudice and seal 1988, delgiudice and moen 1997, schwartz and renecker 2007). winter diets have limited digestible protein which is reflected by low un:c ratios in moose urine, and as fat stores deplete, moose increase catabolism of muscle protein (rich in nitrogen) fig. 2. mean (± se) un:c ratios in snow urine samples collected from radio-marked calves and unmarked adult cow moose in northern new hampshire (january-april, 2014–2017). dotted line at un:c ratio = 3.5 represents the survival threshold described by delgiudice (1995). table 1. summary of un:c ratios of snow-urine samples (n) from moose in 2014–2017, new hampshire, usa. original un:c data are presented here with statistical comparisons made after data were log e transformed. dead calves had significantly higher (p < 0.05; *) un:c ratios from late february onward. collection interval surviving calves dead calves unmarked adult cows n x̅ sd n x̅ sd n x̅ sd 15–31 jan. 12 2.17 0.85 13 2.57 1.04 7 1.96 0.60 1–14 feb. 11 2.50 0.80 19 2.88 0.97 12 2.37 0.87 15–28 feb 15 2.99 0.75 18 3.58* 1.19 11 2.56 0.77 1–14 mar. 14 2.43 0.74 15 4.68* 2.93 9 2.42 0.93 15–31 mar. 10 2.47 0.53 12 3.66* 0.88 13 2.50 1.01 1–14 apr. 10 2.18 0.87 9 4.27* 2.37 5 1.96 0.77 snow urine sampling – ellingwood et al. alces vol. 55, 2019 18 to meet protein and energy requirements, causing an abrupt spike in un:c ratios (delgiudice et al. 1987, delgiudice and seal 1988). ratios > 3.5 represent serious deterioration in body condition with estimated losses of 0.5–0.8 kg/day of lean body (muscle) mass for a 400 kg moose (delguidice et al. 1997). for an average (174 kg) calf expressing similar un:c ratios, this would equate to 0.2–0.3 kg/day or ~5% weight loss in 30 days. low un:c levels (< 3.5) were measured in this study from january through late february each year, suggesting that all cows and calves were in an early phase of undernutrition. the spike in un:c ratios measured in calves that died is symptomatic of animals experiencing prolonged undernutrition and catabolizing muscle. similar effects have been measured in white-tailed deer (odocoileus virginianus), elk (cervus canadensis), and other moose populations due to limited forage diversity, availability, and quality associated with marginal habitat and deep snow (delgiudice et al. 1989, 1991a, 1991b). all dead calves expressed signs of anemia and emaciation including severe weight loss (23 ± 7%) in 3.5 months (ellingwood 2018). this degree of weight loss is comparable to that measured in whitetailed deer experiencing prolonged and irreversible levels of malnutrition (delgiudice and seal 1988). in contrast to the abovementioned examples, the impact of undernutrition in this study was not evident population-wide or associated with habitat or winter severity. calves that died experienced a un:c spike in march, whereas surviving calves and adult cows did not. surviving calves and adults shared the same habitat as the dead calves, and forage availability in the study area is not considered limiting (bergeron et al. 2011, dunfey-ball 2017). limited mobility and increased mortality due to snow depth for moose occurs at >90 cm (coady 1974); however, in the winters of 2014–2017, average snow depth was ≤ 71 cm in northern new hampshire and the highest rate of calf mortality occurred in the year of lowest maximum snow depth (jones et al. 2019). the timing of the spike in dead calves aligns with the peak feeding period of adult winter ticks in midto late march (samuel 2004). arguably, calves were experiencing an increasing protein and energy deficit due to the physiological requirement to replenish blood (protein) loss associated with feeding by winter ticks. tick load estimates derived from half-hide counts of dead calves averaged 47,000 (jones et al. 2019), and compensation for blood loss associated with this infestation level requires 50 to >100% of the daily protein requirement of a calf during the peak 2-week feeding period of adult winter ticks (musante et al. 2007). metabolically, this depletion of muscle mass (protein) mimics that of a starving animal, producing a similar response in the un:c ratios. for example, delgiudice and seal (1988) identified un:c ratios ranging from 4 >23 mg/dl in urine samples collected from white-tailed deer experiencing prolonged undernutrition. the average un:c ratio peaked at 4.68 for dead calves; however, given the average period (20 ± 13 days) between the date of the last sample collection and mortality, our data do not reflect the higher ratios that occur closer to death. the tick load measured at january capture was the primary determinant of calf fate, as tick load and probability of survival were inversely related. for calves with low and moderate tick counts, body weight has some counter-balancing effect on survival, with heavier calves (> 174 kg) expressing heightened resistance to mortality (ellingwood 2018). this influence was also demonstrated in higher survival of heavier calves in alces vol. 55, 2019 snow urine sampling – ellingwood et al. 19 northern maine with similar tick loads. the body weight of 73% of calves captured in new hampshire was less than the average weight (190 kg) in northern maine where winter tick epizootics remain uncommon (ellingwood 2018; l. kantar, maine department of inland fisheries and wildlife, unpublished data). given the high proportion (17.5%) of optimal habitat (4–16 year old regenerating growth) in the region (dunfey-ball 2017), these low weights presumably reflect the carry-over effects of compromised reproductive cows impacted by high annual tick loads (musante et al. 2010, jones et al. 2017). temporal monitoring of un:c ratios identified the occurrence and timing of nutritional restriction of calves that ultimately died from blood loss to winter ticks. while all moose exhibited some degree of undernutrition, the spike observed in the un:c ratio of dead calves was related directly to the feeding period of adult female winter ticks. in more southern moose populations where winter ticks are of most concern, un:c ratios from snow urine samples could be used to assess calf condition, identify a potential epizootic (delgiudice et al. 1997), and predict mortality rate in the population. collection and analysis of snow urine samples from calves in the second and third week in march should be adequate to identify the proportion with un:c ratio > 3.5 and provide a reasonable estimate of the seasonal calf mortality rate. the ability to measure this metric provides critical recruitment information and is a useful proxy for productivity, both necessary to develop effective management strategies. acknowledgements funding for this project was provided through wildlife restoration program grant no. f13af01123 (nh w-104-r-1) to n.h. fish and game department from the u.s. fish and wildlife service, division of wildlife and sport fish restoration, and safari club international foundation. this research was made possible through the access granted by property owners including american forest management, the conservation fund, plum creek, t. r. dillon, and wagner forest management ltd. the efforts of field technicians j. debow and t. soucy were critical for the collection of these data. references bahnak, b. r., j. c. holland, l. j. verme, and j. j. ozoga. 1979. seasonal and nutritional effects on serum nitrogen constituents in white-tailed deer. journal of wildlife management 43: 454–460. bergeron, d. h., and p. j. pekins. 2014. evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire. alces 50: 1–15. _______, _______, h.f. jones, and w.b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. _____, _____, and k. rines. 2013. temporal assessment of physical characteristics and reproductive status of moose in new hampshire. alces 49: 39–48. coady, j. w. 1974. influence of snow on behavior of moose. le naturaliste canadian 101: 417–436. degraff, r. m., m. yamasaki, w. b. leak, and j. w. lanier. 1992. new england wildlife: management of forested habitats. general technical report ne-144. usda forest service, northeast forest experiment station, radnor, pennsylvania, usa. delguidice g.d. 1995. assessing winter nutritional restriction of northern deer with urine in snow: considerations, potential, and limitations. wildlife society bulletin 23: 687–693. snow urine sampling – ellingwood et al. alces vol. 55, 2019 20 _____, l. d. mech, and u. s. seal. 1988. chemical analyses of deer bladder urine and urine collected from snow. wildlife society bulletin 16: 324–326. _______, _______, and _______. 1989. physiological assessment of deer populations by analysis of urine in snow. journal of wildlife management 53: 284–291. _______, _______, _______, and p. d. karns. 1987. winter fasting and refeeding effects on urine characteristics in white-tailed deer. journal of wildlife management 51: 860–864. _______, and r. moen. 1997. simulating nitrogen metabolism and urea nitrogen: creatinine ratios in ruminants. journal of wildlife management 61: 881–894. ______, r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restriction, ticks, and numbers of moose on isle royale. journal of wildlife management 61: 895–903. _______, _____, and u. s. seal. 1991b. differences in urinary chemistry profiles of moose on isle royale during winter. journal of wildlife diseases 27: 407–416. _____, and u. s. seal. 1988. classifying winter undernutrition in deer via serum and urinary urea nitrogen. wildlife society bulletin 16: 27–32. _____, f. j. singer, and u. s. seal. 1991a. physiological assessment of winter nutritional deprivation in elk of yellowstone national park. journal of wildlife management 55: 653–664. dunfey-ball, k. 2017. moose density, habitat and winter tick epizootics in a changing climate. m. s. thesis. university of new hampshire, durham, new hampshire, usa. ellingwood, d. 2018. assessing the impact of winter tick epizootics on moose condition and population dynamics in northern new hampshire. m. s. thesis. university of new hampshire, durham, new hampshire, usa. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of inter tick epizootics. alces 53: 85–98. _______, _______, _______, d. ellingwood, i. sidor, a. lichtenwalner, and m. o’neil. 2019. mortality assessment of calf moose during successive years of winter tick epizootics in new hampshire and maine. canadian journal of zoology: 97: 22–30. kirkpatrick, r.l., d.e. buckland, w.a. abler, p.f. scanlon, j.b. whalen, and h.e. buckhart. 1975. energy and protein influences on blood urea nitrogen of white-tailed fawns. journal of wildlife management 39: 692–698. mason, g. l., and d. j. madden. 2007. performing the field necropsy examination. veterinary clinics food animal practice 23: 503–526. mech, d. l. 1983. handbook of animal radio tracking. university of minnesota press, minneapolis, minnesota, usa. munson, l. 2015. necropsy of wild animals. wildlife health center, school of veterinary medicine, university of california davis, davis, california, usa. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. _____, _____, and _____. 2010. char acteristics and dynamics of a regional moose (alces alces) population in the northeastern united states. wildlife biology 16: 185–204. new hampshire fish and game department (nhfg). 2015. new hampshire game management plan: 2016–2025. new hampshire fish and game department, concord, new hampshire, usa. samuel, w. m. 2004. white as a ghost: winter ticks and moose. natural history alces vol. 55, 2019 snow urine sampling – ellingwood et al. 21 series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. schwartz, c. c., and l. a. renecker. 2007. nutrition and energetics. pages 441–478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. seal, u. s., l. j. verme, j. j. ozoga, and a. w. erickson. 1972. nutritional effects on thyroid activity and blood of white-tailed deer. journal of wildlife management 36: 1041–1051. sine, m. e., k. morris, and d. knupp. 2009. assessment of a line transect method to determine winter tick abundance on moose. alces 45: 143–146. 15 metrics of harvest for ungulate populations: misconceptions, lurking variables, and prudent management r. terry bowyer1, kelley m. stewart2 , vernon c. bleich2, jericho c. whiting3, kevin l. monteith4, marcus e. blum2, and tayler n. lasharr4 1institute of arctic biology, university of alaska fairbanks, fairbanks, alaska 99775, usa; 2department of natural resources and environmental sciences, university of nevada reno, mail stop 186, 1000 valley road, reno, nevada 89512, usa; 3department of biology, brigham young university-idaho, rexburg, idaho, usa; 4haub school of environment and natural resources, and wyoming cooperative fish and wildlife research unit, department of zoology and physiology, university of wyoming, 804 east fremont st., laramie, wyoming 82072, usa abstract: biologists often must use incomplete information to make recommendations concerning harvest of large mammals. consequently, those recommendations must draw on a firm understanding of the ecology of the species in question, along with selection of the most applicable population characteristics on which to base harvest—both essential components for prudent management. density-dependent processes, which are ubiquitous among populations of large mammals, may be counterintuitive because of unexpected patterns in recruitment coincident with changes in population size. misconceptions concerning population dynamics of ungulates also can occur when demographics are based solely on correlations with environmental factors. further, the concept of a harvestable surplus can be misleading for managing ungulate populations, because of the parabolic relationship between population size and number of recruits—harvest determines the surplus rather than vice versa. understanding consequences of mortality, especially relative components of compensatory or additive mortality, also is necessary. knowledge of the proximity of an ungulate population to ecological carrying capacity (k) is required to fully assess whether most mortality is compensatory or additive. we describe selected life-history traits and population characteristics of ungulates useful in parametrizing where populations are in relation to k, thereby allowing for a reasonable harvest despite some uncertainty in population size. we advocate an adaptive-management approach while monitoring those life-history traits to evaluate the suitability of a particular harvest strategy. alces vol. 56: 15–38 (2020) key words: adaptive management, additive mortality, compensatory mortality, density dependence, harvest metrics, harvestable surplus, harvesting females, life-history characteristics, modeling, prudent management humans have engaged in organized hunting for millennia (hull 1964). in the past 2 centuries, considerable effort has been focused on restoring and conserving populations of wildlife in north america (leopold 1933, allen 1954, trefethen 1975, bowyer et al. 2019), and conducting research to understand effects of harvest on those populations (mccullough 1979, 2001, connelly et al. 2012, monteith et al. 2013). herein, we define harvest to be the legal and regulated killing of game species by licensed (i.e., authorized) hunters, as typically practiced throughout most of north america, europe, and parts of africa. hunting also incorporates ethical considerations such as “fair chase,” whereby hunters must not take an unfair advantage over animals they pursue (posewitz metrics of harvest – bowyer et al. alces vol. 56, 2020 16 1994). culling—another term for hunting that may not meet those preceding criteria— typically occurs when removals are designed to reduce population size to diminish conflicts with agriculture or to regulate predators (quirós-fernández et al. 2017), to combat the spread of diseases (mysterud et al. 2019), or for the removal of individuals with undesirable phenotypic traits (torres-porras et al. 2007); these are not subjects of this essay. regulation of hunting effort, means of take, seasons, and harvest quotas reflect a biological and ethical approach to harvest management that has been aspired to since leopold (1933). indeed, a sound and scientifically based harvest is the benchmark that exemplifies prudent management of wildlife, and is one of seven pillars of the north american model of wildlife conservation (organ et al. 2010, mahoney and geist 2019). biologists, however, frequently must make harvest recommendations with a limited understanding of population characteristics, often because fiscal constraints hinder intensive monitoring or surveys (dinsmore and johnson 2012). in such instances, management decisions must rely firmly on an understanding of the ecology of any species, in addition to data available for a particular area. managing harvest based on imperfect information might be necessary, but selecting the correct population characteristics on which to base prudent management is essential. our purpose is to examine populations metrics for large mammals in general, and ungulates in particular, within the context of density-dependent processes. in addition, we clarify how such measurements relate to the reliable and prudent management of ungulate populations. life-history characteristics of large mammal populations many demographic characteristics exist with which to assess populations of large mammals (caughley 1977, williams et al. 2001, skalski et al. 2005 for reviews). population dynamics of those species have been widely studied (schmidt and gilbert 1978, mccullough 1979, sauer and boyce 1983, skogland 1985, boyce 1989, 2018, grøtan et al. 2009, and others). moreover, life-history traits of large mammals differ markedly from those of small mammals (caughley and krebs 1983), and those differences help form the basis for understanding and assessing the population dynamics of large mammals (mccullough 1979, bowyer et al. 2014). large mammals possess numerous attributes consistent with a slow-paced life history. these species exhibit a type i survivorship curve, wherein survival of young initially declines, sometimes markedly (gaillard et al. 1998), but then quickly approaches an asymptote with survivorship remaining high throughout mid-life, followed by high mortality late in life that is reflected in a precipitous decay of the curve (deevey 1947). concomitantly, large mammals exhibit slow development, a delay in age at first reproduction, are iteroparous, possess small litters with large-bodied progeny, are long-lived, provide high maternal investment in young, and exhibit a low intrinsic rate of increase (r) (stubbs 1977, gaillard et al. 2000). this suite of characteristics leads to strong density dependence among large mammals not only in their demographics, but also in their population dynamics (bowyer et al. 2014). large body size buffers them against environmental extremes, and the slow life-history characteristics result in strong competitive abilities of ungulates compared with small mammals. further, individuals of species exhibiting density dependence and having long lives may forgo or restrict allocation to reproduction to increase probability of their survival (monteith et al. 2014a), or tradeoff current against future reproduction (morano alces vol. 56, 2020 metrics of harvest – bowyer et al. 17 et al. 2013). these traits should not be considered a dichotomy (i.e., large vs. small mammals); rather, they should be viewed as a continuum across the range of life histories. moreover, not every large mammal will subscribe perfectly to these life-history traits (stearns 1977, kleiman 1981, mccullough 1999). nonetheless, density dependence resulting from these attributes is the critical factor in understanding and managing populations of large mammals (mccullough 1979, fowler 1981, bowyer et al. 2014). density dependence and population characteristics unimpeded population growth of density-dependent species typically follows an s-shaped (or logistic) curve of number of individuals over time (verhulst 1838). this curve, which often is depicted as symmetrical, shows exponential growth up to an inflection point, and then moves toward an asymptote at ecological carrying capacity, k (mccullough 1979). the curve need not be perfectly symmetrical and can include an overshoot of or oscillations around k (mccullough 1999). nonetheless, this growth curve provides a heuristic framework for understanding the population dynamics of large mammals, so long as it is recognized that some departures from this basic pattern can occur (fowler 1981, mccullough 1999). the principal reason the curve of population size over time is s shaped in ungulates relates primarily to nutrition (fowler 1981, fowler and smith 1981). as the population increases, per capita availability of food declines, eventually causing negative effects on reproduction and survival; those effects become especially prominent once abundance surpasses the inflection point of the curve (mccullough 1979, 1999, monteith et al. 2014a). indeed, several of those traits that characterize large mammals as having a slow-paced life history change with the size of the population relative to k, including survival of young, age at first reproduction, litter size, and weight of neonates (albon et al. 1983, gaillard et al. 2000, eberhardt 2002, bowyer et al. 2014). this relationship between population size and k, rather than density per se, has critically important implications for population dynamics, because k differs among environments and can change over time. hence, dynamics of populations at differing densities is dependent on their relation to k (kie et al. 2003). the ultimate cause of population regulation—food—can vary in complex ways and interact with other factors, such as weather (mitchell et al. 2015), human disturbance (lendrum et al. 2012), disease (eve and kellogg 1977, sams et al. 1996), or immune function (downs et al. 2015). therefore, care must be taken when interpreting factors other than density dependence to avoid errors in managing populations of ungulates. there is a risk of making such errors by accepting the fit of observations to an explanation as evidence for the correctness of that premise. the danger lies in that those observations may be consistent with other hypotheses; correlation may not reflect causation (mccullough 1979). with respect to ungulates, such correlations may be spurious, because nutritional condition may result in other factors being correlated with but not the actual cause of the observed population dynamics. convincing empirical and experimental evidence has documented the widespread occurrence of strong density dependence among ungulates (mccullough 1979, 2001, sauer and boyce 1983, kie and white 1985, fowler 1987, boyce 1989, stewart et al. 2005, bonenfant et al. 2009, monteith et al. 2014a), especially for sexually dimorphic artiodactyls. a useful approach for understanding how density dependence underpins population dynamics of ungulates involves metrics of harvest – bowyer et al. alces vol. 56, 2020 18 plotting the number of recruits (the number of animals successfully added to the population in a reproductive effort) as a function of population size (mccullough 1979, fowler 1981; fig. 1). this parabolic curve illustrates that number of young added to the population is low at extremely low numbers because few adults exist to produce young, and low at high numbers (near k) because few young survive—most young are successfully added to the population at intermediate numbers, a point termed maximum sustained yield (msy). msy is the product of population size and recruitment rate, and lies at the peak of the parabola, which also indicates the maximum annual harvest that a population can sustain under a given set of ecological conditions without causing a decline in numbers (fig. 1). this model is not fully age structured and considers only adults and young. the assumption is that recruited young compensate for adults removed in the harvest. consequently, the age structure of the population becomes younger with an increasing harvest that reduces population size relative to k (bowyer et al. 1999). the concept of ecological carrying capacity (k) is central to understanding density dependence (mccullough 1979). traditionally, k has been defined by the number of animals that a particular environment can support at equilibrium (caughley 1977, mccullough 1979), a concept that is reflected in the conceptual models we present. long-term changes in k (either increases or decreases) can be brought about by perturbations to habitats, including via mechanical manipulation, fire, climate change, grazing, or population overshoots of k (klein 1968, bleich and hall 1982, kie et al. 2003, holl and bleich 2010, holl et al. 2012, berger et al. 2018). short-term changes in k also occur. for example, temporary variation in productivity or winter severity can alter the extent of density-dependent processes in any particular year (loison and langvatn 1998, bowyer et al. 2014, monteith et al. 2014a). although average population size may decrease with a fluctuating k (boyce and daley 1980), the effect of those interannual shifts in k is comparatively modest relative to those brought about by large perturbations to the environment. harvest of ungulates is sustainable, in part, because of their density-dependent attributes (kokko 2001, bowyer et al. 2014, 2019). harvest reduces population size, resulting in increased availability of food on a per capita basis. in situations when populations become increasingly food limited as they exceed msy and approach k, heightened nutrition that follows reductions in population size can increase survival, fig. 1. the parabolic relationship between recruitment number (i.e., the number of young successfully added to the population) and population size for an ungulate population. msy is maximum sustained yield, which is the maximum harvest (or other mortality) that can be sustained by the population, fry is a fixed removal yield, and k is the number of individuals that the environment can support under equilibrium conditions (adapted from bowyer et al. 2014). the yield curve is derived from the near-linear inverse relationship between recruitment rate (young/adult) and population size (mccullough 1979:88, 93; boyce 1989:84). alces vol. 56, 2020 metrics of harvest – bowyer et al. 19 fecundity, or both, and thereby compensate for animals harvested (owen-smith 2006). density-dependent responses to reduced population size include increased survival of young (eberhardt 2002, bonenfant et al. 2009, monteith et al. 2014a), large body mass of neonates (keech et al. 2000), and enhanced rates of growth (mccullough 1979, schmidt et al. 2007, monteith et al. 2018). in addition, large litter sizes (keech et al. 2000, hanson et al. 2009), rapid growth to large body size (monteith et al. 2009), high pregnancy rates (stewart et al. 2005), and early age at first reproduction (monteith et al. 2014a, jensen et al. 2018) are associated with a high nutritional plane typical of a population well below k, whether from harvest or other causes (gasaway et al. 1992, hayes et al. 2003). those density-dependent responses facilitate resilience to harvest and promote persistence of hunted populations of ungulates (bowyer et al. 2019). the harvestable surplus and density dependence leopold (1933) proposed the concept of the harvestable surplus—populations of most animals produce more young than are necessary to ensure persistence. accordingly, that excess could be harvested without adversely affecting the population. this idea may be particularly relevant for species with comparatively fast-paces life histories. those species exhibit j-shaped growth curves, are especially sensitive to annual variation in weather, and exhibit no evident relationship between population density and mortality. hence, for those density-independent species, the number of surplus animals in a particular year determines the harvest (leopold 1933, errington 1945). those animals constitute leopold’s “doomed surplus”—death might occur from a variety of causes, including harvest, but remaining animals allow the population to rebound the following year with few adverse effects from harvesting the surplus individuals (leopold 1933). populations of many upland game birds are limited by primarily by weather (perkins et al. 1997, flanders-wanner et al. 2004, terhune et al. 2019), and management often follows this harvest paradigm. many organisms can have density-dependent components to their life histories (sibly et al. 2005), even for species with fast-paced characteristics. the essential question is whether population density or weather-related events primarily regulate their populations. mccullough (1979) contended that the concept of a harvestable surplus was of limited value in understanding the harvest of highly density-dependent species, which includes most ungulates. the principal conceptual difference between harvest paradigms is that for density-independent species the surplus determines the harvest, whereas for density-dependent species the harvest determines the surplus. for animals exhibiting a strong influence of density dependence (i.e., characterized by marked changes in vital rates under changing densities relative to k), the harvest, through its effects on abundance, becomes a determinant of the surplus in subsequent years (fig. 1). progressively increasing the harvest (thereby reducing the population) along the x-axis of population size from k toward msy in figure 1, theoretically results in an increase in the number of young recruited into the population until population size falls below msy—harvest is regulating the surplus. convincing empirical data support this premise (mccullough 1979). leopold’s (1933) concept of a harvestable surplus is inadequate for managing density-dependent species, because recruitment of young when the population is near k would be low, and there would be little surplus (fig. 1). decreasing harvest to account for that poor recruitment ultimately would metrics of harvest – bowyer et al. alces vol. 56, 2020 20 be counter-productive, with the result that the population remains near k and again exhibits poor recruitment the following year, even though harvest might have been reduced with the expectation that it would compensate for poor recruitment. the best outcome from such management is a loss of hunting opportunity; the worst is a population in poor physical condition, with small-bodied individuals that are more likely to succumb to adverse weather or other maladies such as predation or disease (bowyer et al. 2014). harvests that reduce the population well below msy also will diminish recruitment (fig. 1). managing populations below msy is unlikely to be sustainable. unless the intent is a large reduction in population size, such management can be risky because of the vagaries of dealing with small populations (lande and barrowclough 1987). judging where the population is with respect to k will be addressed later. compensatory versus additive mortality another attribute of populations of large mammals that can muddle interpretation of data needed for prudent management is the shifting pattern of compensatory and additive morality (errington 1946). compensatory mortality occurs when one source of mortality compensates for another (e.g., animals killed during hunting season would have died anyway from harsh winter conditions or predation—bartmann et al. 1992, boyce et al. 1999). with additive mortality, sources of death are summed (e.g., animals killed by hunters would be added to those dying from other causes, but in the absence of harvest would not have otherwise died). additive mortality varies with the dynamics of ungulate populations and the proximity of those populations to k (mccullough 1979, bowyer et al. 2014, monteith et al. 2014a). females that exist in populations at low to moderate numbers with respect to k tend to be in excellent nutritional condition (mccullough 1979, monteith et al. 2014a); attempts to recruit young into the population often are successful, because they have the necessary resources to complete gestation and provision young—food is not limiting (mccullough 1979; fig. 2). as the population increases toward k, however, intraspecific competition intensifies and per capita availability of food diminishes. with increased competition and fewer resources, nutritional condition of females declines, resulting in lower recruitment of young into the breeding population (mccullough 1979, bishop et al. 2009, bowyer et al. 2014; fig. 2). at those higher numbers, females attempt to rear more offspring than the environment can support; these are young that might perish from a variety of sources, but irrespective of the cause of mortality, they are destined to die—mortality is compensatory (fig. 2). in some situations, compensatory mortality also can occur as a result of seasonally determined processes of density dependence (boyce et al. 1999). harvest may not influence spring breeding or pre-harvest numbers of animals. with “seasonality,” density dependence following harvest can increase seasonal abundance or annual survival, resulting in compensatory mortality. this outcome occurs via the interaction between reduction in size and the density-dependent response of the population (boyce et al. 1999). a critically important consideration is that ungulate populations undergoing additive effects of mortality tend to be those at low density with respect to k, with mortality becoming increasingly compensatory as population size and, thus, nutritional limitation increases (monteith et al. 2014a). for populations near k, predators killing young is less of a concern, and there is no need to reduce the alces vol. 56, 2020 metrics of harvest – bowyer et al. 21 size of the planned harvest to compensate for other sources of mortality (fig. 2). additive and compensatory mortality, however, do not represent a discrete dichotomy. notice how prescribing a 30% harvest, for example, induces mortality that varies in its distribution between successful and attempted recruits as the population changes in size (fig. 2). these differing patterns of mortality also mean that predator control near k is unlikely to enhance survival of young, whereas such control may be justified at lower population sizes where mortality is additive (bowyer et al. 2013, 2014). density dependence, weather, and population modeling populations of wildlife have been modeled using a variety of approaches (lack 1954, starfield and bleloch 1986, royama 1992, caswell 2001, mackenzie et al. 2006, and others). kokko (2007) recommended the inclusion of density dependence in population models; failing to do so results in a lessgeneral model, which reduces the usefulness of that approach, especially for large mammals. indeed, a sustainable harvest is not possible without incorporating the concept of density dependence (mendelssohn 1976). numerous reasons exist why density-dependence can be difficult to measure (mccullough 1990, bowyer et al. 2014), and outcomes from densitydependent processes may be counterintuitive (e.g., harvest potentially increasing the number of recruits; fig. 1). nonetheless, incorporating density dependence is essential to building harvest models, no matter how tempting it might be to construct simpler models based on variables that are easier to measure. in highly stochastic environments, food resources available during any particular year may vary widely (mackie et al. 1990, marshal et al. 2005, heffelfinger et al. 2017). hence, capacity of the habitat to support large herbivores can vary annually, creating what could appear to be an absence of density dependence when density itself displays no obvious relationship with nutritional or demographic variables (mccullough 1999). density dependence, however, remains firmly in operation because resources available per capita are a function of both population abundance (i.e., density) and the availability of food within a particular year (mccullough 1990, monteith et al. 2014a). consequently, a seemingly absent short-term relationship with density does not necessarily imply an absence of density dependence (mccullough 1990, kie et al. 2003). fig. 2. changes in number of successful recruits, as well as unsuccessful attempts to recruit, in relation to increasing size of an ungulate population. females attempt to add more young to the population than can be sustained by the environment as a function of its carrying capacity (k). note that mortality becomes increasingly more compensatory (one source of mortality substitutes for another) as the population approaches k. in contrast, a similar level of mortality becomes increasingly additive (one source of mortality is added to another) as populations size backs further away from k, because the number of young that females attempt to recruit approaches the number they can recruit given improvements in nutrition (adapted from monteith et al. 2014a). prescribing a 30% harvest, for example, results in mortality that varies in its distribution between successful and attempted recruits as the population changes in size. metrics of harvest – bowyer et al. alces vol. 56, 2020 22 one shortcoming of basing models for ungulates solely on density-independent variables, such as rain, snowfall or temperature, is that those variables can interact with population density of animals in relation to k (monteith et al. 2014a). ungulate populations close to k will be in poor nutritional condition, and weather is likely to help or hinder those individuals disproportionally compared with animals existing at lower numbers where their physical condition is good (bowyer et al. 2014 for review). at sufficiently high numbers relative to k, density dependence may outweigh even density independent events that might be beneficial (stewart et al. 2005). conversely, severe weather, such as extreme drought, may be overridden by effects of increased nutrition related to reduced population numbers (thalmann et al. 2015). furthermore, body mass of reindeer (rangifer tarandus) was more important than spring phenology in determining production of young in a severe arctic environment, largely because of carryover effects from reserves accumulated in previous seasons (veiberg et al. 2016). spurious correlations between weather and fitness components such as pregnancy, young recruited, and survival can exist for density-dependent species and may lead erroneously to the conclusion that weather, rather than population size, is regulating a population, especially for populations near k. incorporating density-independent variables for ungulates based on such misconceptions can result in models that will not cope with variation in population size relative to k. in this example, weather metrics, including winter severity, are lurking variables (i.e., those that are correlated with the variable of interest but are not its primary cause), and their misinterpretation can lead to the mismanagement of populations. as in our previous example, reducing harvest because of a perceived effect of weather—such as from high snowfall or severe drought for a population near k—would result in the loss of potential hunting opportunity, and poor recruitment again the following year because the population remained near k. we concede that there are rare weather events that kill animals without regard to their physical condition (bleich and pierce 2001, o’gara 2004, bleich 2018), but such events cannot be common or few animals would persist in those environments. we argue that the starting point for models and tests of hypotheses concerning the population dynamics of ungulates should begin with the key assumption of density dependence. we further propose that results from correlational studies implicating weather as a cause of changes in demographic traits be viewed with care and skepticism, in the absence of ascertaining the position of the population in relation to k. this is a critical point in selecting which metrics to use in evaluating effects of harvest. role of the sexes and harvest of females sexual segregation—the differential use of space, forage, or habitat by the sexes outside the mating season—occurs widely among polygynous ungulates (bowyer 2004). ungulates exhibit primarily polygynous mating systems, wherein relatively few large males are necessary to inseminate females within a population (darwin 1872, geist 1974). increased polygyny intensifies malemale competition for mates, which in turn has led to the evolution of increased sexual dimorphism in body size and weapons (geist 1966, weckerly 1998, loison et al. 1999, emlen 2015). avian models, which ascribe sexual dimorphism to intersexual competition, do not suffice for mammals (ralls 1977). spatial segregation of the sexes tends to be most pronounced near and following parturition, when provisioning of young is alces vol. 56, 2020 metrics of harvest – bowyer et al. 23 critically important, and neonates are most vulnerable to predation (bowyer 2004). indeed, differences in body size between sexes of ungulates foster intersexual disparities in susceptibility to predators, with females and young being more vulnerable than adult males (berger 1991, bleich et al. 1997, bowyer et al. 2001). avoidance of predators by females and neonates can result in marked differences in behavior (bleich 1999) and their resultant use of space compared with males and nonparturient females (barten et al. 2001). in addition, sexual dissimilarities in digestive morphology and function occur, and also can explain sexual segregation in ruminants based on the allometry of metabolic requirements, minimal food quality, and retention of digesta (barboza and bowyer 2000, 2001). adult males eat abundant forages high in fiber, because ruminal capacity prolongs retention, and consequently allows greater use of fiber for energy than in females (fig. 3). females, which typically are smaller-bodied than males, are better in postruminal digestion of forage, especially when energy and protein requirements needed for reproduction increase (monteith et al. 2014b). lactating females also increase rumen size, as well as the length and width of rumen papillae beyond that of nonreproductive females (zimmerman et al. 2006). increased nutrient requirements of pregnant females, including the costs of remodeling their digestive tracts to facilitate lactation, underlie differential use of habitats and forages and can lead to sexual segregation (barboza and bowyer 2000, 2001). although a number of hypotheses have been forwarded to explain sexual segregation, predation (bleich et al. 1997) and the gastrocentric model (barboza and bowyer 2000, 2001) are the prevailing views concerning how this phenomenon relates to the spatial ecology of ungulates (stewart et al. 2011 for review). males and females may partition space at fine scales in species such as white-tailed deer (odocoileus virginianus), or large scales as in moose (alces alces) (mccullough 1979, stewart et al. 2003, oehlers et al. 2011). increasing population density (i.e., the number of individuals relative to k) results in greater overlap in the distribution of the sexes and a reduction in degree of sexual segregation (stewart et al. 2015), but with that overlap comes a divergence in diets of males and females (kie and bowyer 1999, schroeder et al. 2010). the upshot is that males and females avoid competing for resources, and arguably should be managed as if they were separate species (bowyer 2004), including developing separate management plans for the sexes. consequently, the harvest of males does little to promote population productivity when compared with the harvest of females (mccullough 1979). for sexually dimorphic ungulates, harvesting males to such low numbers that females might not be fertilized occurs infrequently (schwartz et al. 1992, laurian et al. 2001). a particularly heavy harvest of males will reduce their age structure, and thereby reduce the size of males and their horn-like structures (jenks et al. 2002, monteith et al. 2013). throughout most of north america, harvest of female ungulates is either rare or limited but varies by species and area (monteith et al. 2013, 2018). management paradigms typically have focused on the harvest of males, ostensibly because of the low-risk and conservative approach that male harvest offers for maintaining population size. mysterud et al. (2002) provides insights into how the harvest of males might affect population dynamics. nonetheless, a male-biased harvest is expected to have a negligible influence on population dynamics (freeman et al. 2014), because abundance of males has little effect on nutrition of females, and thereby recruitment of young (mccullough metrics of harvest – bowyer et al. alces vol. 56, 2020 24 1979, 2001, monteith et al. 2018). indeed, harvest that only targets males can have little controlling influence on population sizes except under exceptionally high rates (mccullough 1979, milner-gulland et al. 2003), and primarily only influences the age structure of males in the population. harvest of females may provide an important, yet undervalued and underused, management tool for regulating density-dependent processes for many ungulate populations by holding population numbers below k (monteith et al. 2018). females play the predominant role in the dynamics of most ungulate populations. consequently, harvest of adult females can allow managers to manipulate population sizes to decrease nutritional limitations and competition for resources (mccullough 1979, 2001, solberg et al. 2002, doak et al. 2016). fig. 3. model of intake and digestive function in nonreproductive females (middle) compared with large males (above) and lactating females (below). width of arrows reflects amount of food intake, length of arrows indicates rate of digesta passage, and shading indicates density of nutrients in food. diagrams of the digestive tract are stippled to reflect potential changes in fibrosity of food for males and increases in postruminal size and function of lactating females (modified from barboza and bowyer 2000, zimmerman et al. 2006). alces vol. 56, 2020 metrics of harvest – bowyer et al. 25 further, when maintaining or increasing size of secondary sexual characteristics (i.e., horns, antlers, or pronghorns) in males is a management objective, harvest of females may be an effective management option (monteith et al. 2018). as per capita resources decline with rises in population size, females are limited in the resources they can allocate to offspring (festa-bianchet and jorgenson 1998, monteith et al. 2009), and females that are in poor condition will produce sons that have smaller bodies and smaller horn-like structures that can persist into adulthood (festa-bianchet et al. 2004, solberg et al. 2004, monteith et al. 2009, 2013, 2018). another strategy for increasing the size of horn-like organs is allowing males to reach the age of asymptotic body growth prior to harvest (stewart et al. 2000, jenks et al. 2002, monteith et al. 2009). although maintaining populations at or near k may be viewed as an ideal management outcome when considering only animal abundance, if management goals are aligned with increased nutritional condition or large horn-like structures, maintaining populations at a moderate population size via female harvest can result in a productive and more stable population with a greater yield of large males (mccullough 1978, monteith et al. 2018). predation also has the potential to hold ungulate populations at densities well below k (gasaway et al. 1992, hayes et al. 2003). the role of k in understanding predator-prey dynamics and its relevance to management has been discussed elsewhere (person et al. 2001, bowyer et al. 2005). metrics for harvest many methods exist to estimate abundance and trends of animal populations (krebs 1998, maier et al. 2005, ryan 2011, pierce et al. 2012a for reviews), most of which entail considerable expense, time, and sometimes risk (boyce 2012). our approach has been constrained chiefly to understanding harvest metrics; thus, we will focus on harvest-based characteristics that provide information concerning populations. among methods for estimating population size from harvest are population reconstructions (i.e., cohort analysis) (mccullough 1979, bowyer et al. 1999, ueno et al. 2009), and catch-per-unit-effort (cpue), which in addition to other shortcomings, requires the size of a population to be reduced sufficiently to observe changes over time (bishir and lancia 1996, bowyer et al. 1999, schmidt et al. 2005). those methods, however, may require several years to parametrize equations necessary to estimate population size. we contend that specific attention to understanding the position of population abundance in relation to k offers another potential and more immediate approach for managing populations of ungulates. sophisticated metrics, such as a reliable estimate of population size, may not be necessary for management purposes, although most managers would welcome such detailed information. we maintain that prudent management of ungulate populations relies on one critical point—knowing where the population is in relation to k. the importance of that knowledge is illustrated by considering a harvest strategy termed the fixed removal yield (fry) by mccullough (1979), which results in a population being on the righthand “hump” of the recruitment parabola (fig. 1). a harvest at this level is the near-maximum harvest that a population can sustain without causing a decline in numbers, but is less than the maximum harvest (msy) to avoid an inadvertent overkill. moreover, such a harvest would be buffered by compensatory mortality (fig. 2). an identical amount of sustained harvest applied to metrics of harvest – bowyer et al. alces vol. 56, 2020 26 a population on the left-hand side of the parabola would drive a population toward extirpation (fig. 1). if the objective is a high yield, managing for fry is prudent. nevertheless, without some benchmark to identify k, applying such a harvest strategy can be precarious. estimating k can be a daunting proposition. regression methods of plotting recruitment rate over population size can be used to estimate k, but that technique tends to overestimate k, and may entail many years of data collection (mccullough 1979, bowyer et al. 1999). forage-based models (hobbs and swift 1985, beck et al. 2006) for estimating k have been developed but are labor intensive; forage measurements also may lag declines of large herbivores. other methods (boyce 1989, forsyth and caley 2006) exist to determine k, but those approaches require large data sets and can be costly; again, years may be needed to obtain the information necessary to parameterize those models (bowyer et al. 2005, 2013, monteith et al. 2014a). issues related to the conservation and management of ungulates likely would have been decided before many of the aforementioned models could be developed (bowyer et al. 2013). in addition, habitat or environmental changes might have occurred, potentially invalidating conclusions from the resulting models. monteith et al. (2014a) proposed the use of “animal-indicated nutritional carrying capacity” (ncc), and identified methods necessary to parametrize that metric. ncc is based on the nutritional condition of individuals comprising a population when r = 0 (i.e., no population change). animals in poor nutritional condition infer a population near or above ncc—where resources sufficient to sustain good body condition are not available. a population consisting of individuals in relatively good nutritional condition typifies a population below ncc, where resources exist to support population growth (monteith et al. 2014a). this approach offers a workable means for assessing ncc and tracking the status of populations over time. although, this approach was developed for mule deer (o. hemionus), ncc should be a concept useful for managing numerous species of ungulates, in part, because it allows for interannual variation in k. the life-history characteristics and population parameters in table 1 provide a conceptual framework with which to assess the nutritional status of ungulate populations and calibrate their relationship to msy and k. these variables change with the size of a population relative to msy and k, and can be used to help assess effects of harvest. knowing the size of a population may not be essential for management purposes—what is needed is an understanding of where the population is in relation to k. moreover, ungulate populations exhibit a sequence of changes in life-history traits that tend to be modified as the population approaches k. the most sensitive of those characteristics is declining recruitment of young, followed by increasing age at first reproduction, declining litter size, lower rates of pregnancy, and finally diminishing adult survival (gaillard et al. 2000, eberhardt 2002). care should be taken to collect appropriate data concerning multiple metrics to ensure a reliable assessment of the status of a population. for example, young to adult ratios commonly are used to index population productivity. interpretation of such ratios are inherently chancy, however, because of their double-variable nature (caughley 1974, theberge 1990, person et al. 2001, bowyer et al. 2013). nevertheless, ratios are tempting to use because they are readily obtainable (bowyer et al. 2013), and because mortality in adult females should be relatively constant given the type i survivorship curve of ungulates (deevey 1947). nonetheless, alces vol. 56, 2020 metrics of harvest – bowyer et al. 27 sufficiently strong density-dependent mechanisms can adversely affect adult survival (pierce et al. 2012b). a situation in which both adult females and young were decreasing would result in a ratio indicative of a population that was unchanged, when the population was in decline. if survival of adult females were known, however, the ratio could be interpreted (monteith et al. 2014a), but those data would require additional monitoring of survival of adult females instead of merely sampling ratios. caughley (1974) warned that the use of such ratios could be problematic without some indication of population size and trajectory, and if r were known, the ratio would be superfluous; others have echoed similar concerns for ratio data (theberge 1990, person et al. 2001, bowyer et al. 2013). we contend that measures of animal condition and related metrics (table 1) offer a stronger basis for prudent management so long as several variables (table 1) are considered in concert. we make no specific recommendations on which metrics in table 1 to employ to accomplish that goal—those would be a function of the species under consideration and which variables might be collected most effectively and economically. likewise, we make no recommendations as to where to manage an ungulate population in relation to k. such management decisions are both socioeconomic and biological in nature, and likely vary with the objectives or responsibilities of the management agency. we do, however, provide the background to interpret the likely biological outcomes from such management decisions. adaptive management several studies of density dependence in ungulates have been based on experimental manipulations of density over large areas with free-ranging animals (mccullough 1979, 2001, stewart et al. 2005); such research allows for much stronger inference concerning cause and effect than from observational or correlative studies. moreover, the use of a hypothetico-deductive approach provides a rigorous framework for testing predictions, including those concerning density dependence, because that approach allows for falsification (popper 1959). illogic ensuing from observational or correlational studies that rely solely on inductive reasoning can fall victim to popper’s white-swan fallacy—no number of sightings of white swans can verify the hypothesis that all swans are white, because the observation of a single black swan falsifies that premise (popper 1959). when more in-depth research is possible, critical tests of properly framed hypotheses will answer questions about why particular phenomena occur, and hold the potential to be generalizable, thereby yielding an overall understanding and greater certainty in processes underpinning population regulation. less-general tests, focused on questions about what happened, inevitably will be narrow in nature and scope, and necessarily will require repeated testing to judge the importance of a phenomenon. these patterns are particularly germane to understanding population dynamics of ungulates. hypotheses that consider harvest within the broad framework of density-dependent processes are likely to provide a sound understanding of population dynamics and yield prudent management. studies that assess only a few environmental factors related to productivity are too restrictive and too narrow to provide a reliable understanding of population dynamics, and offer, at best, a chancy tactic for management. yet, hypothetico-deductive tests stemming from in-depth research are not always possible. adaptive management has been advocated as a method for dealing with the presence of uncertainty in biological systems metrics of harvest – bowyer et al. alces vol. 56, 2020 28 (walters and hilborn 1978, westgate et al. 2013). the harvesting of ungulate populations clearly holds a degree of ambiguity given that biologists often must set management objectives with limited information concerning the status of populations (dinsmore and johnson 2012). where more structured and reliable data related to population dynamics are unavailable because of time or expense, we suggest that varying the harvest of females and monitoring selected life-history characteristics and related metrics (table 1) can resolve the status of a population relative to whether the harvest is sustainable, and thereby result in prudent management decisions. also, the sequence in which those life-history traits are observed can provide insights into where the population is relative to k. a similar approach has been used for adaptive management of stock-recruitment curves for fisheries (smith and walters 1981). there are several caveats to this approach. not all density-dependent variables increase linearly with population size, including population growth and recruitment of young into the population. further, the parabolic relationship between recruitment number and population size may not be symmetrical (mccullough 1999, sibly et al. 2005). nonlinearities can confound interpretation needed for adaptive management particularly when linear responses were expected. for instance, a harvest of females that initially resulted in increased recruitment of young, might eventually cause a decline in recruitment if that harvest exceeded msy (fig. 1). maternal effects are widespread in cervids (freeman et al. 2013) and might further confuse interpretations from employing an adaptive-management approach. body size and antlers of young white-tailed deer took several generations to respond to enhanced nutritional condition of their small mothers (monteith et al. 2009). moose that were born small relative to larger offspring failed to compensate in size over time (keech et al. 1999), whereas young caribou (r. tarandus) did compensate (dale et al. 2008). table 1. variation in life-history and population characteristics of ungulates in relation to the proximity of the population to msy (maximum sustained yield) and k (ecological carrying capacity) (modified from bowyer et al. 2014). life-history and populations characteristics ≤ msy near k physical condition of adult females better poorer pregnancy rate of adult females higher lower pause in annual reproduction by adult females less likely more likely yearlings pregnanta usually seldom corpora lutea counts of adult femalesa higher lower litter sizea higher lower age at first reproduction for females younger older weight of neonates heavier lighter mortality of youngb additive compensatory diet quality higher lower population age structure younger older age at extensive tooth wear older younger asome species of ungulates may exhibit limited variability in particular characteristics. badditive and compensatory mortality would be inferred from other variables in this table. alces vol. 56, 2020 metrics of harvest – bowyer et al. 29 further uncertainty in interpreting population dynamics may be caused by delayed density dependence, wherein recruitment is lagged further than expected following harvest (fryxell et al. 1991, lande et al. 2006). deteriorated rangelands may take time to recover from overgrazing or other perturbations (heady 1975), and ungulate populations inhabiting such ranges may not respond immediately to harvest in the expected manner. in addition, interspecific competition holds potential to lower k for sympatric ungulates (stewart et al. 2002), and some large carnivores can regulate ungulate populations at densities below k (gasaway et al. 1992, tatman et al. 2018). diseases and parasites likewise can affect populations of large mammals (cassirer and sinclair 2007, jones et al. 2017). provided that managers are mindful of these caveats, the variables in table 1 offer a useful method for judging where the population is with respect to k, and thereby determining harvests without requiring estimates of population size. clearly, an adaptive management approach offers many strengths for prudently managing ungulate populations, but being knowledgeable, observant, and patient is necessary. context and conclusions the conservation of mammals worldwide is a pressing concern (ceballos and ehrlich 2002). unregulated or illegal hunting continues to be a threat to the conservation of some mammals, especially in underdeveloped countries (van vliet et al. 2015). terrestrial families within the certartiodactyla (i.e., the even-toed ungulates) are especially vulnerable to threats to their existence (bowyer et al. 2019). ungulates possess life-history traits that make them more susceptible than other mammals to extinction, including large body size and slow-paced life histories (cardillo et al. 2005). based on the iucn red list, threats to mammals from hunting still exist (bowyer et al. 2019). many of those threats, however, concerned the historical depletion of a species, or illegal killing; nowhere is legal and regulated recreational hunting of mammals recognized as a threat (bowyer et al. 2019). in north america, europe, and parts of africa, hunting has been the foundation for successful programs to ensure conservation of critically important habitats or to restore wild populations (geist 1995, organ et al. 2010, krausman and bleich 2013). selecting the best metrics for managing the harvest of ungulates is a refinement of existing management practices and a critically important step in their scientific stewardship, but this procedure needs to be viewed in the proper context. today, in north america and europe, populations of wild mammals are thriving, and legal hunting remains not only a cornerstone to financial return in support of their persistence, but also a useful tool to regulate their abundance at either ecologically or socially acceptable levels (organ et al. 2010). acknowledgements we thank j. a. jenks, c. d. mitchell, and m. s. boyce for their helpful and insightful comments on our manuscript. this is professional paper 129 from the eastern sierra center for applied population ecology. references albon, s. d., b. mitchel, and b. w. stains. 1983. fertility and body weight in female red deer. journal of animal ecology 52: 969–980. doi: 10.2307/4467 allen, d. l. 1954. our wildlife legacy. funk and wagnalls, new york, new york, usa. barboza, p. s., and r. t. bowyer. 2000. sexual segregation in dimorphic deer: a new gastrocentric hypothesis. journal of mammalogy 81: 473–489. doi: 10.1093/ jmammal/81.2.473 metrics of harvest – bowyer et al. alces vol. 56, 2020 30 _____, and ____. 2001. seasonality of sexual segregation in dimorphic deer: extending the gastrocentric model. alces 37: 275–292. barten, n. l., r. t. bowyer, and k. j. jenkins. 2001. habitat use by female caribou: tradeoffs associated with parturition. journal of wildlife management 65: 77–92. doi: 10.2307/ 3803279 bartmann, r. m., g. c. white, and l. h. carpenter. 1992. compensatory mortality in a colorado mule deer population. wildlife monographs 121: 1–39. beck, j. l., j. m. peek, and e. k. strand. 2006. estimates of elk summer range nutritional carrying capacity constrained by probabilities of habitat selection. journal of wildlife management 70: 283–294. doi: 10.2193/002 2-541x (2006) 70[283:eoesrn] 2.0.co;2 berger, j. 1991. pregnancy incentives, predation constraints, and habitat shifts: experimental and field evidence for wild bighorn sheep. animal behaviour 41: 61–77. doi: 10.1016/s0003-3472(05) 80503-2 _____, c. hartway, a. gruzdev, and m. johnson. 2018. climate degradation and extreme icing events constrain life in cold-adapted mammals. scientific reports 8: 1156. doi: 10.1038/s41598 018-19416-9 bishir, j. w., and r. a. lancia. 1996. on catch-effort methods of estimating animal abundance. biometrics 52: 1457– 1466. doi: 10.2307/2532859 bishop, c. j., g. c. white, d. j. freddy, b. e. watkins, and t. r. stephenson. 2009. effect of enhanced nutrition on mule deer population rate of change. wildlife monographs 172: 1–28. doi: 10.2193/ 2008-107 bleich, v. c. 1999. mountain sheep and coyotes: patterns of predator evasion in a mountain ungulate. journal of mammalogy 80: 283–289. doi: 10.2307/ 1383228 _____. 2018. mass mortalities of migratory mule deer (odocoileus hemionus): implications for ecosystem function, conservation, and management? western wildlife 5: 7–12. _____, r. t. bowyer, and j. d. wehausen. 1997. sexual segregation in mountain sheep: resources or predation? wildlife monographs 134: 1–50. _____, and s. a. holl. 1982. management of chaparral habitat for mule deer and mountain sheep in southern california. pages 247–254 in c. e. conrad and w. c. oechel, technical coordinators. proceedings of the symposium on the dynamics and management of mediterranean-type ecosystems. usda forest service, general technical report psw-58. june 22–26, 1981, san diego, california, usa. _____, and b. m. pierce. 2001. accidental mass mortality of migrating mule deer. western north american naturalist 61: 124–125. bonenfant, c., j.-m. gaillard, t. coulson, m. festa-bianchet, a. loison, m. g. leif, e. loe, p. blanchard, n. pettorelli, n. owen-smith, j. du toit, and p. duncan. 2009. empirical evidence of density-dependence in populations of large herbivores. advances in ecological research 41: 313–357. doi: 10.1016/s0065-2504(09)00405-x bowyer, r. t. 2004. sexual segregation in ruminants: definitions, hypotheses, and implications for conservation and management. journal of mammalogy 85: 1039–1052. doi: 10.1644/bbl-002.1 _____, v. c. bleich, k. m. stewart, j. c. whiting, and k. l. monteith. 2014. density dependence in ungulates: a review of causes and consequences, with some clarifications. california fish and game 100: 550–572. _____, m. s. boyce, j. r. goheen, and j. l. rachlow. 2019. conservation of the world’s mammals: status, protected areas, community efforts, and hunting. alces vol. 56, 2020 metrics of harvest – bowyer et al. 31 journal of mammalogy 100: 923–941. doi: 10.1093/jmammal/gyy180 _____, j. g. kie, d. k. person, and k. l. monteith. 2013. metrics of predation: perils of predator-prey ratios. acta theriologica 58: 329–340. doi: 10.1007/ s13364-013-0133-1 _____, d. r. mccullough, and g. e. belovsky. 2001. causes and consequences of sociality in mule deer. alces 37: 371–402. _____, m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–89. _____, d. k. person, and b. m. pierce. 2005. detecting top-down versus bottom-up regulation of ungulates by large carnivores: implications for conservation of biodiversity. pages 342– 361 in j. c. ray, k. h. redford, r. s. steneck, and j. berger, editors. large carnivores and the conservation of biodiversity. island press, covelo, california, usa. boyce, m. s. 1989. the jackson elk herd: intensive wildlife management in north america. cambridge university press, cambridge, united kingdom. _____. 2012. managing moose by the seat of your pants. theoretical population biology 82: 340–347. doi: 10.1016/j. tpb.2012.03.002 _____. 2018. wolves for yellowstone: dynamics in time and space. journal of mammalogy 99: 1021–1031. doi: 10.1093/jmammal/gyy115 _____, and d. j. daley. 1980. population tracking of fluctuating environments and natural selection for tracking ability. american naturalist 115: 480–491. doi: 10.1086/283575 _____, a. r. e. sinclair, and g. c. white. 1999. seasonal compensation of predation and harvesting. oikos 87: 419–426. doi: 10.2307/3546808 cardillo, m., k. e. jones, j. bielby, o. r. bininda-emonds, w. sechrest, c. d. orme, and a. purvis. 2005. multiple causes of high extinction risk in large mammal species. science 309: 1239– 1241. doi: 10.1126/science.1116030 cassirer, f. e., and a. r. e. sinclair. 2007. dynamics of pneumonia in a bighorn sheep metapopulation. journal of wildlife management 71: 1080–1088. doi: 10.2193/2006-002 caswell, h. 2001. matrix population models: construction, analysis, and interpretation. 2nd edition. sinauer associates, cumberland, massachusetts, usa. caughley, g. 1974. interpretation of age ratios. journal of wildlife management 38: 557–562. doi: 10.2307/3800890 _____. 1977. analysis of vertebrate populations. john wiley and sons, new york, new york, usa. _____, and c. j. krebs. 1983. are big mammals simply little mammals writ large? oecologia 59: 7–17. doi: 10.1007/ bf00388066 ceballos, g., and p. r. ehrlich. 2002. mammal population losses and the extinction crisis. science 296: 904–907. doi: 10.1126/science.1069349 connelly, j. w., j. h. gammonley, and t. w. keegan. 2012. harvest management. pages 202–231 in n. j. silvy, editor. the wildlife techniques manual: management. 7th edition, volume 2. john hopkins university press, baltimore, maryland, usa. dale, b. w., l. g. adams, w. b. collins, k. jolly, p. valkenburg, and r. tobey. 2008. stochastic and compensatory effects limit persistence of variation in body mass of young caribou. journal of mammalogy 89: 1130–1135. doi: 10.1644/07-mamm-a-137.1 darwin, c. 1872. the expression of emotions in man and animals. john murray, london, united kingdom. deevey, e. s., jr. 1947. life-tables for natural populations of animals. quarterly review of biology 22: 283–314. doi: 10.1086/395888 metrics of harvest – bowyer et al. alces vol. 56, 2020 32 dinsmore, s. j., and d. h. johnson. 2012. population analysis in wildlife biology. pages 349–380 in n. j. silvy, editor. the wildlife techniques manual: research. 7th edition, volume 1. john hopkins university press, baltimore, maryland, usa. doak, p., c. j. carroll, and k. kielland. 2016. harvest of female moose at high density: modelling the impacts of harvest on population size and biomass yield. wildlife biology 22: 153–159. doi: 10.2981/wlb.00163 downs, c. j., k. m. stewart, and b. l. dick. 2015. investment in constitutive immune function by north american elk experimentally maintained at two different population densities. plos one 10: e0125586. doi: 10.1371/journal.pone.0125586 emlen, d. j. 2015. animal weapons: the evolution of battle. henry holt and co., new york, new york, usa. eberhardt, l. l. 2002. a paradigm for population analysis of long-lived vertebrates. ecology 83: 2841–2854. doi: 1 0 . 1 8 9 0 / 0 0 1 2 9 6 5 8 ( 2 0 0 2 ) 083[2841:apfpao]2.0.co;2 errington, p. l. 1945. some contributions of a 15-year local study of the northern bobwhite to a knowledge of population phenomena. ecological monographs 15: 1–34. doi: 10.2307/1943293 _____. 1946. predation and vertebrate populations. quarterly review of biology 21: 144–177. doi: 10.1086/395220 eve, j. h., and f. e. kellogg. 1977. management implications of abomasal parasites in southeastern white-tailed deer. journal of wildlife management 41: 169–177. doi: 10.2307/3800590 festa-bianchet, m., d. w. coltman, l. turelli, and j. t. jorgenson. 2004. relative allocation to horn and body growth in bighorn rams varies with resource availability. behavioral ecology 15: 305–312. doi: 10.1093/beheco/arh014 _____, and j. t. jorgenson. 1998. selfish mothers: reproductive expenditure and resource availability in bighorn ewes. behavioral ecology 9: 144–150. doi: 10.1093/beheco/9.2.144 flanders-wanner, b. l., g. c. white, and l. l. mcdaniel. 2004. weather and prairie grouse: dealing with effects beyond our control. wildlife society bulletin 32: 22–34. doi: 10.2193/ 00917648(20 04)32[22:wapgdw]2.0.co;2 forsyth, d. m., and p. caley. 2006. testing the irruptive paradigm of large herbivore dynamics. ecology 87: 297–303. doi: 10.1890/05-0709 fowler, c. w. 1981. density dependence as related to life history strategy. ecology 62: 602–610. doi: 10.2307/1937727 _____. 1987. a review of density dependence in large mammal populations. current mammalogy 1: 401–441. doi: 10.1007/978-1-4757-9909-5_10 _____, and r. d. smith, editors. 1981. dynamics of large mammal populations. john wiley and sons, new york, new york, usa. freeman, e. d., r. t. larsen, k. clegg, and b. r. mcmillian. 2013. long-lasting effects of maternal condition in free ranging cervids. plos one 8: e5873. _____, _____, m. e. peterson, c. r. anderson, k. r. hersey, and b. r. mcmillian. 2014. effects of male-biased harvest on mule deer: implications for rates of pregnancy, synchrony, and timing of parturition. wildlife society bulletin 38: 806–811. doi: 10.1002/ wsb.450 fryxell, j. m., d. j. t. hussell, a. b. lambert, and p. c. smith. 1991. time lags and population fluctuations in white-tailed deer. journal of wildlife management 55: 377–385. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low-densities in alaska and yukon and implications for conservation. wildlife monographs 120: 1–59. alces vol. 56, 2020 metrics of harvest – bowyer et al. 33 gaillard, j.-m., m. festa-bianchet, and n. g. yoccoz. 1998. population dynamics of large herbivores: constant adult survival and variable recruitment. trends in ecology and evolution 13: 58–63. doi: 10.1016/s0169-5347(97)01237-8 _____, _____, _____, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual review of ecology and systematics 31: 367–393. geist, v. 1966. the evolution of horn-like organs. behaviour 27: 175–214. doi: 10.1163/156853966x00155 _____. 1974. on the relationship of social evolution and ecology in ungulates. annual review of ecology and systematics 8: 193–207. _____. 1995. north american policies of wildlife conservation. pages 75–129 in v. geist and i. mct. cowan, editors. wildlife conservation policy. detselig enterprises, limited, calgary, alberta, canada. grøtan, v., b.-e. sæther, m. lillegård, e. j. so l b e r g, and s. en g e n. 2009. geographical variation in the influence of density dependence and climate on the recruitment of norwegian moose. oecologia 161: 685–695. hanson, l. b., m. s. mitchell, j. b. grand, d. b. jolley, b. d. sparklin, and s. s. ditchkoff. 2009. effect of experimental manipulation on survival and recruitment of feral pigs. wildlife research 36: 185–191. doi: 10.1071/wr08077 hayes, r. d., r. f. frnell, r. m. p. ward, j. carey, m. dehn, g. w. kuzyk, a. m. baer, c. l. gardner, and m. o’donoghue. 2003. experimental reduction of wolves in the yukon: ungulate responses and management implications. wildlife monographs 152: 1–35. heady, h. f. 1975. rangeland management. mcgraw-hill, new york, new york, usa. heffelfinger, l. j., k. m. stewart, a. p. bush, j. s. sedinger, n. w. darby, and v. c. bleich. 2017. timing of precipitation in an arid environment: effects on population performance of a large herbivore. ecology and evolution 2017: 1–12. hobbs, n. t., and d. m. swift. 1985. estimates of habitat carrying-capacity incorporating explicit nutritional constraints. journal of wildlife management 49: 814–822. holl, s. a., and v. c. bleich. 2010. responses of large mammals to fire and rain in the san gabriel mountains, california. northern wild sheep and goat council proceedings 17: 139–156. _____, _____, b. w. callenberger, and b. bahro. 2012. simulated effects of two fire regimes on bighorn sheep: the san gabriel mountains, california, usa. fire ecology 8: 88–103. doi: 10.4996/ fireecology.0803088 hull, d. b. 1964. hounds and hunting in ancient greece. university of chicago press, chicago, illinois, usa. hutchinson, g. e. 1978. an introduction to population ecology. yale university press, new haven, connecticut, usa. jenks, j. a., w. p. smith, and c. s. deperno. 2002. maximum sustained yield harvest versus trophy management. journal of wildlife management 66: 528–535. doi: 10.2307/3803186 jensen, w. f., j. j. maskey, jr., j. r. smith, and e. s. michel. 2018. reproductive parameters of moose during population expansion in north dakota. alces 54: 27–36. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. keech, m. a., r. d. boertje, r. t. bowyer, and b. w. dale. 1999. effects of birth weight on growth of young moose: do low-weight neonates compensate? alces 35: 51–57. _____, r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and metrics of harvest – bowyer et al. alces vol. 56, 2020 34 t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450–462. doi: 10.2307/3803243 kie, j. g., and r. t. bowyer. 1999. sexual segregation in white-tailed deer: density-dependent changes in use of space, habitat selection, and dietary niche. journal of mammalogy 80: 1004–1020. _____, _____, and k. m. stewart. 2003. ungulates in western forests: habitat requirements, population dynamics, and ecosystem processes. pages 296–340 in c. j. zabel and r. g. anthony, editors. mammal community dynamics: management and conservation in the coniferous forests of western north america. cambridge university press, new york, new york, usa. _____, and m. white. 1985. population dynamics of white-tailed deer (odocoileus virginianus) on the welder wildlife refuge, texas. southwestern naturalist 30: 105–118. kleiman, d. g. 1981. correlations among life history characteristics of mammalian species exhibiting two extreme forms of monogamy. pages 332–344 in r. d. alexander, and d. w. tinkle, editors. natural selection and social behavior. chiron press, asheville, north carolina, usa. klein, d. r. 1968. the introduction, increase, and crash of reindeer on st. matthew island. journal of wildlife management 32: 350–367. kokko, h. 2001. optimal and suboptimal use of compensatory responses to harvesting: timing of hunting as an example. wildlife biology 7: 141–150. _____. 2007. modelling for field biologists and other interesting people. cambridge university press, cambridge, united kingdom. krausman, p. r., and v. c. bleich. 2013. conservation and management of ungulates in north america. international journal of environmental studies 70: 372–382. krebs, c. j. 1989. ecological methodology. harper collins, new york, new york, usa. lack, d. 1954. the natural regulation of animal populations. oxford university press, oxford, united kingdom. lande, r., and g. f. barrowclough. 1987. effective population size, genetic variation, and their use in population management. pages 87–123 in m. soulé, editor. viable populations for conservation. cambridge university press, new york, new york, usa. _____, s. engen, b.-e. sæther, and t. coulson. 2006. estimating density dependence from time series of population age structure. american naturalist 168: 76–87. doi: 10.1086/504851 laurian, c., j.-p. quellet, r. courtois, l. breton, and s. st-onge. 2001. effects of intensive harvesting on moose reproduction. journal of applied ecology 37: 515–531. lendrum, p. e., c. r. anderson, jr., r. a. long, j. g. kie, and r. t. bowyer. 2012. habitat selection by mule deer during migration: effects of landscape structure and natural-gas development. ecosphere 3:art 82. doi: 10.1890/ es12-00165.1 leopold, a. 1933. game management. charles scribner’s sons, new york, new york, usa. loison, a., j.-m. gaillard, c. pelabon, and n. g. yoccoz. 1999. what factors shape sexual size dimorphism in ungulates? evolutionary ecology research 1: 611–633. _____, and r. langvatn. 1998. short-and long-term effects of winter and spring weather on growth and survival of red deer in norway. oecologia 116: 489– 500. doi: 10.1007/s004420050614 mackenzie, d. i., j. d. nichols, j. a. royle, k. h. pollock, l. l. bailey, and alces vol. 56, 2020 metrics of harvest – bowyer et al. 35 j. e. hines. 2006. occupancy estimation and modeling: inferring patterns and dynamics of species occurrence. elsevier, new york, new york, usa. mackie, r. j., k. l. hamlin, d. f. pac, g. l dusek, and a. k. wood. 1990. compensation in free-ranging deer populations. transactions of the north american wildlife and natural resources conference 55: 518–526. mahoney, s. p., and v. geist, editors. 2019. the north american model of wildlife conservation (wildlife management and conservation). johns hopkins university press, baltimore, maryland, usa. maier, j. a. k., j. ver hoef, a. d. mcguire, r. t. bowyer, l. saperstein, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics: effects of scale. canadian journal of forest research 35: 2233–2243. marshal, j. p., p. r. krausman, and v. c. bleich. 2005. dynamics of mule deer forage in the sonoran desert. journal of arid environments 60: 593–609. doi: 10.1016/j.jaridenv.2004.07.002 mccullough, d. r. 1978. essential data required on population structure in field studies of threatened herbivores. pages 302–317 in threatened deer. iucn, morges, switzerland. _____. 1979. the george reserve deer herd: ecology of a k-selected species. university of michigan press, ann arbor, michigan, usa. ______. 1990. detecting density dependence: filtering the baby from the bathwater. transactions of the north american wildlife and natural resources conference 55: 534–543. _____. 1999. density dependence and life-history strategies of ungulates. journal of mammalogy 80: 1130–1146. _____. 2001. male harvest in relation to female removals in a black-tailed deer population. journal of wildlife management 65: 46–58. doi: 10.2307/ 3803276 mendelssohn, r. 1976. optimization problems with a leslie matrix. american naturalist 110: 339–349. milner-gulland, e. j., o. m. bukreeva, t. coulson, a. a. lushchekina, m. v. kholodova, a. b. bekenov, and i. a. grachev. 2003. conservation: reproductive collapse in saiga antelope harems. nature 422: 135–135. doi: 10.1038/ 422135a mitchell, c. d., r. channey, k. aho, j. g. kie, and r. t. bowyer. 2015. density of dall’s sheep in alaska: effects of predator harvest? mammal research 60: 21–28. monteith, k. b., k. l. monteith, r. t. bowyer, d. m. leslie, jr., and j. a. jenks. 2014b. reproductive effects on fecal nitrogen as an index of diet quality: an experimental assessment. journal of mammalogy 95: 301–310. doi: 10.1644/ 12-mamm-a-306.1 monteith, k. l., v. c. bleich, t. r. stephenson, b. m. pierce, m. m. conner, j. g. kie, and r. t. bowyer. 2014a. life-history characteristics of mule deer: effects of nutrition in a variable environment. wildlife monographs 186: 1–56. _____, r. a. long, v. c. bleich, j. r. heffelfinger, p. r. krausman, and r. t. bowyer. 2013. effects of harvest, culture, and climate on trends in size of horn-like structures in trophy ungulates. wildlife monographs 183: 1–26. doi: 10.1002/wmon.1007 _____, _____, t. r. stephenson, v. c. bleich, r. t. bowyer, and t. n. lasharr. 2018. horn size and nutrition in mountain sheep: can ewe handle the truth? journal of wildlife management 82: 67–84. _____, l. e. schmitz, j. a. jenks, j. a. delger, and r. t. bowyer. 2009. growth of male white-tailed deer: consequences of maternal effects. journal of mammalogy 90: 651–660. doi: 10.1644/08-mamm-a-191r1.1 morano, s., k. m. stewart, j. s. sedinger, c. a. nicolai, and m. vavra. 2013. metrics of harvest – bowyer et al. alces vol. 56, 2020 36 life-history strategies of north american elk: trade-offs associated with reproduction and survival. journal of mammalogy 94: 162–172. mysterud, a., t. coulson, and n. c. stenseth. 2002. the role of males in the dynamics of ungulate populations. journal of animal ecology 71: 907–915. _____, o. strand, and c. m. rolandsen. 2019. efficacy of recreational hunters and marksmen for host culling to combat chronic wasting disease. wildlife society bulletin 43: 683–692. oehlers, s. a., r. t. bowyer, f. huettmann, d. k. person, and w. b. kessler. 2011. sex and scale: implications for habitat selection by alaskan moose alces alces gigas. wildlife biology 17: 67–84. doi: 10.2981/10-039 o’gara, b. w. 2004. mortality. pages 379– 408 in b. w. o’gara and j. d. yoakum, editors. pronghorn: ecology and management. university press of colorado, boulder, colorado, usa. organ, j. f., s. p. mahoney, and v. geist. 2010. born in the hands of hunters: the north american model of wildlife conservation. the wildlife professional 4: 22–27. owen-smith, n. 2006. demographic determination of the shape of density dependence for three african ungulate populations. ecological monographs 76: 93–109. perkins, a. l., w. r. clark, t. z. riley, and p. a. vohs. 1997. effects of landscape and weather on winter survival of ringnecked pheasant hens. journal of wildlife management 61: 634–641. doi: 10.2307/3802171 person, d. k., r. t. bowyer, and v. van ballenberghe. 2001. density dependence of ungulates and functional responses of wolves: effects on predator-prey ratios. alces 37: 253–273. pierce, b. l., r. r. lopez and n. j. silvy. 2012a. estimating animal abundance. pages 284–310 in n. j. silvy, editor. the wildlife techniques manual: management. 7th edition, volume 1. john hopkins university press, baltimore, maryland, usa. pierce, b. m., v. c. bleich, k. l. monteith, and r. t. bowyer. 2012b. top-down versus bottom-up forcing: evidence from mountain lions and mule deer. journal of mammalogy 93: 977–988. popper, k. r. 1959. the logic of scientific discovery. martino publishing, mansfield center, connecticut, usa. posewitz, j. 1994. beyond fair chase: the ethic and tradition of hunting. globe pequot press, guilford, connecticut, usa. quirós-fernández, f., j. marcos, p. acevedo, and c. gortázar. 2017. hunters serving the ecosystem: the contribution of recreational hunting to wild boar population control. european journal of wildlife research 63: 57. ralls, k. 1977. sexual dimorphism in mammals: avian models and unanswered questions. american naturalist 111: 917–938. doi: 10.1086/283223 royama, t. 1992. analytical population dynamics. chapman and hall, new york, new york, usa. ryan, j. m. 2011. mammalogy techniques manual, 2nd edition. lulu, raleigh, north carolina, usa. sauer, j. r., and m. s. boyce. 1983. density dependence and survival of elk in northwestern wyoming. journal of wildlife management 47: 31–37. doi: 10.2307/ 3808049 sams, m. g., r. l. lochmiller, c. w. qualls, d. m. leslie, jr., and m. e. payton. 1996. physiological correlates of neonatal mortality in an overpopulated herd of white-tailed deer. journal of mammalogy 77: 179–190. schmidt, j. i., j. m. ver hoef, and r. t. bowyer. 2007. antler size of alaskan moose alces alces gigas: effects of population density, hunter harvest and use of guides. wildlife biology 13: 53–65. doi: 10.2981/0909-6396(2007 )13[53: asoama]2.0.co;2 alces vol. 56, 2020 metrics of harvest – bowyer et al. 37 _____, _____, j. a. k. maier, and r. t. bowyer. 2005. catch per unit effort for moose: a new approach using weibull regression. journal of wildlife management 69: 1112–1124. schmidt, j.l. and d.l. gilbert, editors. 1998. big game of north america, ecology and management. stackpole books, harrisburg, pennsylvania, usa. schroeder, c. a., r. t. bowyer, v. c. bleich, and t. r. stephenson. 2010. sexual segregation in sierra nevada bighorn sheep, ovis canadensis sierrae: ramifications for conservation. arctic, antarctic, and alpine research 42: 476–489. doi: 10.1657/1938-4246-42. 4.476 schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula, alaska. alces 28: 1–13. sibly, r. m., d. barker, m. c. denham, j. hone, and m. pagel. 2005. on the regulation of populations of mammals, birds, fish, and insects. science 309: 607–610. skalski, j. r., k. e. ryding, and j. j. millslpaugh. 2005. wildlife demography: analysis of sex, age, and count data. elsevier academic press, new york, new york, usa. skogland, t. 1985. the effects of density-dependent resource limitations on the demography of wild reindeer. journal of animal ecology 54: 359–374. doi: 10.2307/4484 smith, a. d. m., and c. j. walters. 1981. adaptive management of stock-recruitment systems. canadian journal of fisheries and aquatic sciences 38: 690–703. solberg, e. j., b.-e. sæther, o. strand, and a. loison. 2002. dynamics of a harvested moose population in a variable environment. journal of animal ecology 68: 186–204. _____, a. loison, j.-m. gaillard, and m. heim. 2004. lasting effects of conditions at birth on moose body mass. ecography 27: 677–687. doi: 10.1111/j.0906-7590.2004.03864.x starfield, a. m., and a. l. bleloch. 1986. building models for conservation and wildlife management. macmillan publishing company, new york, new york, usa. stearns, s. c. 1977. the evolution of life history traits: a critique of the theory and a review of the data. annual review of ecology and systematics 8: 145–171. doi: 10.1146/annurev.es.08.110177. 001045 stewart, k. m., r. t. bowyer, b. l. dick, b. k. johnson, and j. g. kie. 2005. density-dependent effects on physical condition and reproduction in north american elk: an experimental test. oecologica 143: 85–93. _____, _____, j. g. kie, n. j. cimon, and b. k. johnson. 2002. temprospatial distributions of elk, mule deer, and cattle: resource partitioning and competitive displacement. journal of mammalogy 83: 229–244. doi: 10.1644/1545 -1542(2002 )083<0229:tdoemd> 2.0.co;2 _____, _____, _____, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36: 77–83. _____, _____, and p. j. weisberg. 2011. spatial use of landscapes. pages 181– 217 in d. g. hewett, editor. biology and management of white-tailed deer. crc press, boca raton, florida, usa. _____, t. e. fulbright, d. l. drawe, and r. t. bowyer. 2003. sexual segregation in white-tailed deer: responses to habitat manipulations. wildlife society bulletin 31: 1210–1217. _____, d. r. walsh, j. g. kie, b. l. dick, and r. t. bowyer. 2015. sexual segregation in north american elk: the role of density dependence. ecology and evolution 5: 709–721. stubbs, m. 1977. density dependence in life-cycles of animals and its importance in k-and r-selected strategies. journal of metrics of harvest – bowyer et al. alces vol. 56, 2020 38 animal ecology 56: 677–688. doi: 10.2307/3837 tatman, n. m., s. g. liley, j. w. cain iii, and j. w. pitman. 2018. effects of nutrition and calf predation on elk vital rates in new mexico. journal of wildlife management 82: 1417–1428. terhune, t. m., w. e. palmer, and s. d. wellendorf. 2019. northern bobwhite chick survival and effects of weather. journal of wildlife management 83: 963–974. thalmann, j. c., r. t. bowyer, k. a. aho, f. w. weckerly, and d. r. mccullough. 2015. antler and body size in blacktailed deer: an analysis of cohort effects. advances in ecology 2015: article id 156041. doi: 10.1155/2015/156041 theberge, j. b. 1990. potentials for misinterpreting impacts of wolf predation through prey: predator ratios. wildlife society bulletin 18: 188–192. torres-porras, j., j. carranza, and j. pérez-gonzález. 2009. selective culling of iberian red stags (cervus elaphus hispanicus) by selective montería in spain. european journal of wildlife reserch 55: 117–123. trefethen, j. b. 1975. an american crusade for wildlife. boone and crockett club, alexandria, virgina, usa. ueno, m., t. matshi, e. j. solberg, and t. takashi. 2009. application of cohort analysis to large terrestrial mammal harvest data. mammal study 34: 65–76. doi: 10.3106/041.034.0202 van vliet, n., j. fa, and r. nasi. 2015. managing hunting under uncertainty: from one-off ecological indicators to resilience approaches in assessing the sustainability of bushmeat hunting. ecology and society 20: 7. veiberg, v., l. e. loe, s. d. albon, r. j. irvine, t tverra, e. ropstad, and a. stein. 2016. maternal winter body mass and not spring phenology determine annual calf production in an arctic herbivore. oikos 126: 980–987. doi: 10.1111/ oik.03815 verhulst, p. f. 1838. notice sur al loi que al populations suit dans so accroissement. correspondence mathematiqe et physique 10:113–121. original not seen; citation from hutchinson (1978). walters, c. j., and r. hilborn. 1978. ecological optimization and adaptive management. annual review of ecology and systematics 9: 157–188. doi: 10.1146/annurev.es.09.110178. 001105 williams, b. k., j. d. nichols, and m. j. conroy. 2001. analysis and management of animal populations. academy press, san diego, california, usa. weckerly, f. w. 1998. sexual-size dimorphism: influence of mass and mating systems in the most dimorphic mammals. journal of mammalogy 79: 33–52. westgate, m. j., g. e. likens, and d. b. lindenmayer. 2013. adaptive management of biological systems: a review. biological conservation 158: 128–139. doi: 10.1016/j.biocon.2012.08.016 zimmerman, t. j., j. a. jenks, and d. m. leslie, jr. 2006. gastrointestinal morphology of female white-tailed and mule deer: effects of fire, reproduction, and feeding type. journal of mammalogy 87: 598–605. doi: 10.1644/ 05-mamm a-356r1.1 179 distinguished moose biologist past recipients thunder bay, ontario. 1992 not presented 1991 charles c. schwartz, alaska dept. of fish and game, soldotna, alaska. 1990 rolf peterson, michigan technological university, houghton, michigan. 1989 warren b. ballard, alaska dept. of fish and game, nome, alaska. 1988 vince f. j. crichton, manitoba dept. of natural resources, winnipeg manitoba. and michel crête, ministère du loisir, de la chasse et de la péche, service de la faune terrestre, québec, pq. 1987 w. c. (bill) gasaway, alaska dept. of fish and game, fairbanks, alaska. 1986 h. r. (tim) timmermann, ontario ministry of natural resources, thunder bay, ontario. 1985 ralph ritcey, fish and wildlife branch, kamloops, british columbia. 1984 edmund telfer, canadian wildlife service, edmonton, alberta. 1983 albert w. franzmann, alaska division of fish and game, soldotna, alaska. 1982 a. (tony) bubenik, ontario ministry of natural resources, maple, ontario. 1981 patrick d. karns, minnesota division of fish and wildlife, grand rapids, minnesota. and al elsey, ontario ministry of natural resources, thunder bay, ontario. in 1974, prior to the establishment of the distinguished moose biologist award, the group recognized the pioneering moose research of the late laurits (larry) krefting, u.s. fish and wildlife service, with an individual award. 2010 michael w. schrage fond du lac resource management division, cloquet, minnesota. 2009 kenneth n. child, prince george, british columbia. 2007 kris j. hundertmark, university of alaska fairbanks, fairbanks, alaska. 2006 kristine m. rines, new hampshire fish and game department, new hampton, new hampshire. 2005 w. m. (bill) samuel, university of alberta, edmonton, alberta. 2004 w. eugene mercer, wildlife division, st. john's, newfoundland. 2003 arthur r. rodgers, ontario ministry of natural resources, thunder bay, ontario. 2002 bernt-erik sæther, norwegian university of science and technology, trondheim, norway. 2001 r. terry bowyer, university of alaska, fairbanks, alaska. 2000 gerry m. lynch, alberta environmental protection, edmonton, alberta. 1999 william j. peterson, minnesota department of natural resources, grand marais, minnesota. 1998 peter a. jordan, university of minnesota, st. paul, minnesota. 1997 margareta stéen, swedish university of agricultural sciences, uppsala, sweden. 1996 vic van ballenberghe, u.s. forest service, anchorage, alaska. 1995 not presented 1994 james m. peek, university of idaho, moscow, idaho. 1993 murray w. lankester, lakehead university, f:\alces\vol_38\pagemaker\3815. alces vol. 38, 2002 bowyer et al. morphology of moose antlers 155 geographical variation in antler morphology of alaskan moose: putative effects of habitat and genetics r. terry bowyer1, kelley m. stewart1, becky m. pierce1,2, kris j. hundertmark1,3, and william c. gasaway4 1institute of arctic biology, and department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775-7000, usa, e-mail ffrtb@uaf.edu; 2california department of fish and game, 407 west line street, bishop, ca 93514, usa; 3alaska department of fish and game, 43961 kalifornsky beach road, soldotna, ak 99669, usa; 4alaska department of fish and game (deceased) abstract: we assessed antler size of alaskan moose (alces alces gigas) with respect to the geographic region and dominant vegetation community (taiga or tundra) from which they were harvested from 1968 to 1983. our retrospective analysis indicated that moose from the copper river delta and alaska peninsula possessed the largest antlers, whereas those from southeast alaska, usa, had the smallest antlers. delta flood plains of the copper river offer a rich food supply for moose, and browse on the alaska peninsula also is plentiful; both areas have mild maritime climates and longer growing seasons than tundra and taiga habitats in interior alaska—large antlers in those moose populations likely were the result of superior nutrition. after controlling for age, antlers of moose from tundra communities were significantly larger than those inhabiting taiga. willows (salix spp.), which are an important food for moose, dominate braided rivers and associated riparian areas in tundra habitat, and provide a high-quality and stable food supply over time. fire and subsequent successional changes dominate taiga landscapes, which results in a variable food supply that is sometimes low in quality and quantity. again, forage abundance and quality likely play important roles in determining antler size for populations of alaskan moose inhabiting those plant communities. nonetheless, antlers of a. a. gigas from taiga regions in alaska, usa, were larger than those of a. a. andersoni from similar habitat in northeastern minnesota, usa, and saskatchewan, canada. in addition, moose from tundra habitat on the seward peninsula, alaska, which have colonized that area within the last ~30 years from the boreal forest, possessed antlers intermediate in size between moose inhabiting taiga and tundra. moreover, moose from forested areas of southeast alaska, which have a unique mitochrondial dna haplotype from other subspecies of moose, also had comparatively smaller antlers than other moose in alaska. those outcomes indicated that differences in antler size likely have a genetic in addition to a nutritional basis. we hypothesize that differences in antler size of alaskan moose in relation to habitat may have genetic as well as nutritional underpinnings related to openness of habitat, but more research is needed. finally, our results on antler morphology, in concert with information on pelage coloration and recent data on genetics, do not support hypotheses concerning a double migration, or eastern and western races of moose, forwarded to explain morphological variation in moose inhabiting the new world. likewise, we reject the hypothesis that ecotypical differences are primarily responsible for morphological variation in subspecies of moose inhabiting north america. alces vol. 38: 155-165 (2002) key words: age, alaskan moose, alces alces, antlers, ecotypes, genetics, habitat, morphology, nutrition, size, taiga, tundra morphology of moose antlers bowyer et al. alces vol. 38, 2002 156 size and conformation of cervid antlers are influenced by genetics, age, and nutrition (goss 1983). how those factors interact to determine antler size and shape, and whether antler morphology characterizes populations or subspecies, however, continues to be debated (bubenik 1983, gasaway et al. 1987). for instance, geist (1998) proposed that that there were eastern and western races of moose (alces alces). other geographical variation in morphology for subspecies of this large cervid was thought to be nutritional, and such differences were best regarded as ecotypes (geist 1998). bubenik (1998), however, hypothesized that smaller-antlered moose inhabiting forested regions (taiga moose) were genetically distinct from larger-antlered moose living in more open areas (tundra moose), and that such distinctions were worldwide. he proposed a double-migration hypothesis for moose entering the new world to explain extant morphological variation (bubenik 1998). peterson (1955) delineated 4 subspecies of moose in north america based on skull morphology, and bowyer et al. (1991) described pelage and behavioral differences among subspecies. hundertmark et al. (2003) confirmed those subspecific differences using mitochondrial dna (mtdna). moreover, gasaway et al. (1987) documented clear differences in antler size among subspecies of moose, allowing the possibility of genetic underpinnings of that variability. whether such differences are the result of genetics, nutrition, or both factors, however, remain unresolved. moose are a useful species to evaluate effects of nutrition and genetics on antler morphology because antlers of this large cervid have been well studied (bubenik et al. 1978, bubenik 1998), including relations with body mass (sæther and haagenrud 1985; solberg and sæther 1993, 1994; stewart et al. 2000), age (sæther and haagenrud 1985, stewart et al. 2000, bowyer et al. 2001), mineral composition (moen and pastor 1998), and theoretical interactions of population density, harvest, and genetics (hundertmark et al. 1998). in addition, differences in morphology among subspecies have been confirmed with genetic analyses (hundertmark et al. 2002a, 2002b, 2003). finally, data on age and antler size are available from alaska, for moose (a. a. gigas) inhabiting taiga and tundra habitats, and from northeastern minnesota, usa, and saskatchewan, canada, for another subspecies (a. a. andersoni) living in taiga (gasaway et al. 1987). we hypothesized that if nutrition was influential in determining antler size, we would find the largest-antlered alaskan moose (a. a. gigas) living in areas where forage was abundant, as well as differences in moose living in tundra compared with those from taiga habitat. conversely, if genetics were the overriding determinant of antler size, we postulated that the largest difference would be between a. a. gigas from taiga and a. a. andersoni from that same habitat type. we recognize that these hypotheses are not mutually exclusive, but contend that, in concert with other data on morphology and genetics, we could test ideas concerning the evolution and morphology of subspecies of moose in north america. study area we subset our data by game management units (gmus) established by the alaska department of fish and game, and assigned a habitat type based on the predominant vegetation community in each unit (fig. 1). taiga extended from the eastern boarder with canada westward across the interior; moose harvested in that habitat were from gmus 12-15, 20-22, and 24. moose from southeast alaska inhabited coastal coniferous forests and were haralces vol. 38, 2002 bowyer et al. morphology of moose antlers 157 vested from gmu 1. moose killed in the remaining gmus, which were classified as tundra, included 5, 7, 11, 16-17, 19, 23, and 25-26. we further subdivided our data regionally because habitat in the copper river delta (gmu 6) and alaska peninsula (gmu 9) differed markedly from other areas. likewise, we separated data from the seward peninsula (gmu 22) because moose had recently colonized that tundra area from nearby taiga. methods antler measurements used in our retrospective analysis were collected originally from hunter-harvested moose during 196883 across game management units in alaska (gasaway et al. 1987). those measurements were made mostly by employees of the alaska department of fish and game, who were experienced in gathering such data. data on antler spread, palm length and width, beam circumference, and number of antler tines were obtained from a subset of data that included associated information on age (n = 1,501). methods used to measure antlers were provided by gasaway et al. (1987), stewart et al. (2000), and bowyer et al. (2001). age of moose was determined by counts of tooth cementum annuli (sergeant and pimlott 1959, gasaway et al. 1978). we used principal component analysis (mcgarigal et al. 2000) to obtain an overall index to antler size. principal component 1 (pc1) explained 73% of the variation in m e a s u r e m e n t s o f m o o s e a n t l e r s ; eigenvectors associated with pc1 had similar loadings (0.30-0.35) for the various antler characteristics (stewart et al. 2000, bowyer et al. 2001). pc1 exhibited patterns of rapid increase with age from 1 to 6 fig. 1. map of game management units (gmus) and vegetation types in alaska, usa, used in our analysis of antler size of moose (adapted from albert et al. 2001). morphology of moose antlers bowyer et al. alces vol. 38, 2002 160 (peek 1974, weixelman et al. 1998). moose populations throughout much of alaska are held at low density by large mammalian carnivores (van ballenberghe 1987, gasaway et al. 1992, van ballenberghe a n d b a l l a r d 1 9 9 4 , b a l l a r d a n d v a n ballenberghe 1998, bowyer et al. 1998). consequently, biases from density-dependent effects on physical condition of cervids (sensu mccullough 1979, bowyer et al. 1 9 9 9 ) a n d , u l t i m a t e l y , a n t l e r s i z e (mccullough 1982, stewart et al. 2000) would not be expected. declines in body and antler size, which are positively correlated in cervids (clutton-brock 1982, mccullough 1982, bowyer 1986, stewart et al. 2000), would be expected with increasing population density relative to carrying capacity (k). those well-documented relationships offer strong evidence against the ecotype hypothesis of geist (1998). if morphological differences among subspecies were mostly the result of nutrition, then the magnitude of morphological variation observed among subspecies should be present in a population undergoing a rapid change in population size. although nutritional stunting can occur among cervids, sufficiently large changes in morphology, including differences in pelage markings, within a population undergoing even enormous changes in numbers have not been described (klein 1968, mccullough 1979). indeed, we are unaware of a nutritional mechanism that would cause the marked differences in pelage color and behavior described for subspecies of moose in north america (bowyer et al. 1991). the presence of a white morph in moose that is not an albino, and white females giving birth to reddish-brown young (franzmann 1981, armstrong and brown 1986), strongly support the hypothesis of a genetic component to differences in pelage coloration that cannot be attributed to ecotypes. that same interpretation likely holds for subspecific differences in antlers of moose. over the past ~30 years, moose have colonized the seward peninsula, which is mostly tundra, from nearby taiga habitat. those moose possess antlers that are intermediate in size between moose inhabiting tundra and taiga habitats (fig. 2). although we believe that the intermediate antler size of moose on the seward peninsula likely has nutritional underpinnings, we cannot completely rule out genetics as a cause for that difference because the response in size was neither immediate nor as large as those in other tundra regions of the state (fig. 2). subspecies of cervids inhabiting more open terrain tend to be more social, and have larger body and antler sizes, and more conspicuously marked pelage than those from densely vegetated forests (cowan 1936, peek et al. 1974, hirth 1977, geist 1987, bowyer et al. 1991, molvar and bowyer 1994). more research is needed to determine if antler size was under selection related to more open habitat for moose living on the seward peninsula, as well as for moose inhabiting other open landscapes. antler conformation for alaskan moose (a. a. gigas) differs from other subspecies in north america in their tendency to exhibit a “butterfly” configuration of main and brow palms (gasaway et al. 1987, bowyer et al. 2001). bubenik (1983) further proposed that in forest-dwelling subspecies (i.e., a. a. andersoni, a. a. shirasi, and a. a. americana) the orientation of palms curved upward to form a dish, whereas in moose from the tundra (e.g., a. a. gigas) the palms were comparatively flat. gasaway et al. (1987), however, rejected that hypothesis by comparing ratios of antler characteristics from various subspecies of moose— few differences existed in the overall shape of antlers. moose from wooded habitats also were postulated to have a narrower antler spread than those living in tundra (bubenik et al. 1978, bubenik 1983). alces vol. 38, 2002 bowyer et al. morphology of moose antlers 161 gasaway et al. (1987) concluded that if forest-dwelling moose have evolved antlers that are adapted to dense woodlands, they have done so by altering size rather than shape of antlers—a supposition supported by our results (figs. 2 and 3). moreover, we hypothesize that differences in antler size between alaskan moose inhabiting taiga and moose from forested areas of northeastern minnesota are mostly genetic. moose from both of those areas are subjected to predation (peek et al. 1976, gasaway et al. 1992); consequently, nutrition is not likely the cause of that disparate difference in antler spread (fig. 3). we further hypothesize that differences between the size of antlers of a. a. andersoni from northeastern minnesota and saskatchewan may be nutritional (fig. 3). such differences would be expected because of more intense predation in the minnesota population (peek et al. 1976), and the concomitant increase in physical condition of those moose from being farther away from k than moose from saskatchewan (sensu mccullough 1979, bowyer et al. 1999). the hypothesis of bubenik (1998) that taiga and tundra moose are genetically distinct also can be rejected, as can the hypothesis of geist (1998) for the existence of eastern and western races of moose. moose subspecies inhabiting tundra in the russian far east (a. a. buturilini) possess a different chromosome number (2n = 68) and are not closely related to a. a. gigas (2n = 70) from tundra in alaska (hundertmark et al. 2002b). moreover, alaskan moose, which geist (1998) places with moose from the russian far east, have the same fundamental chromosome number as other subspecies of moose in north america, and are more closely related to other subspecies in the new world than subspecies from eurasia (hundertmark et al. 2002b). similar morphology of a. a. gigas and a. a. buturilini likely is a result of convergent evolution resulting from living in more open habitats than other subspecies of moose (hundertmark et al. 2002b). differences in antler size between a. a gigas from taiga in alaska, and a. a. andersoni from that same habitat in minnesota and saskatchewan (fig. 3), implicate genetics as the cause of such geographic variation. moreover, moose from forested areas of southeast alaska recently have been identified as possessing a unique haplotype of mtdna lacking in other subspecies of moose (hundertmark et al. 2003). although we caution that our sample size was small (fig. 2), males from southeast alaska also had much smaller antlers than moose from taiga habitat in other regions of alaska, further indicating that selection operating in isolated populations affects antler size. consequently, differences in antler size among populations of moose are not completely a result of the type of habitat they occupy. in addition, moose from southeast alaska may be a unique subspecies. we believe, however, that morphological data presented herein, and genetic data from hundertmark et al. (2002a, 2002b, 2003) are not yet sufficient to draw that conclusion—more research is needed. neither the hypothesis of geist (1998) nor bubenik (1998) was supported by our study of antler size in moose. clearly, both nutrition and genetics are involved in the size and shape of antlers (williams et al. 1994), but not in the manner proposed by either geist (1998) or bubenik (1998). more research is required to understand precisely how nutritional and genetic factors interact, and how they might be related to founder effects during dispersal into new habitat, and how natural selection operates on the size of moose inhabiting more open habitat compared with those living in closed boreal forest. we proposed that studies of dna microsatellites, which would allow greater resolution of genetic differences among morphology of moose antlers bowyer et al. alces vol. 38, 2002 162 populations (broders et al. 1999), would be a fruitful next step in resolving this important question. acknowledgements we are indebted to those individuals who originally made measurements of moose antlers, and who are properly acknowledged in gasaway et al. (1987). the alaska department of fish and game, and the institute of arctic biology at the university of alaska fairbanks funded this research. data on vegetation communities of alaska were obtained from http://climchange.cr.usgs.gov/info/lite/alaska/fig1.gif. references albert, d. m., r. t. bowyer, and s. d. miller. 2001. effort and success of brown bear hunters in alaska. wildlife society bulletin 29:501-508. armstrong, e. r., and g. brown. 1986. white moose, alces alces, sightings in northern ontario. canadian field-naturalist 100:262-263. ballard, w. b., and v. van ballenberghe. 1998. moose-predator relationships: research and management needs. alces 34:91-105. bowyer, r. t. 1986. antler characteristics as related to social status of male southern mule deer. southwestern naturalist 31:289-298. , m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35:73-89. , j . l . r a c h l o w , v . v a n ballenberghe, and r. d. guthrie. 1991. evolution of a rump patch in alaskan moose: an hypothesis. alces 27:12-23. , k. m. stewart, j. g. kie, and w. c. gasaway. 2001. fluctuating asymmetry in antlers of alaskan moose: size matters. journal of mammalogy 82:814824. , v. van ballenberghe, and j. g. kie. 1997. the role of moose in landscape processes: effects of biogeography, population dynamics, and predation. pages 265-287 in j. a. bissonette, editor. wildlife and landscape ecology: effects of pattern and scale. springer, new york, new york, usa. , , and . 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79:1332-1344. , , and . 2003. moose (alces alces). pages 931-964 in g.a. feldhamer, b. c. thompson, and j.a. chapman, editors. wild mammals of north america: biology, management, and economics. second edition. the johns hopkins university press, baltimore, maryland, usa. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. c. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8:1309-1315. bubenik, a. b. 1983. the behavioural aspects of antlerogensis. pages 389449 in r. d. brown, editor. antler development in cervidae. casear kleberg wildlife research institute, kingsville, texas, usa. . 1998. evolution, taxonomy, and morphology. pages 77-123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , o . w i l l i a m s , a n d h . r . timmermann. 1978. some characteristics of antlerogensis in moose. a preliminary report. proceedings of the north american moose conference and alces vol. 38, 2002 bowyer et al. morphology of moose antlers 163 workshop 14:157-177. clutton-brock, t. h. 1982. the function of antlers. behaviour 79:108-125. coady, j. w. 1974. influence of snow on behavior of moose. naturaliste canadien 101:417-436. cowan, i. mct. 1936. distribution and variation in deer (genus odocoileus) of the pacific coastal region of north america. california fish and game 22:155-246. , w. s. hoar, and j. hatter. 1950. the effect of forest succession upon the quantity and upon the nutritive values of woody plants used as food by moose. canadian journal of research d 28:249-271. franzmann, a. w. 1981. alces alces. mammalian species 154:1-7. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. , d. b. harkness, and r. a. rausch. 1 9 7 8 . a c c u r a c y o f m o o s e a g e determinations from incisor cementum layers. journal of wildlife management 42:558-563. , d. j. preston, d. j. reed, and d. d. roby. 1987. comparative antler morphology and size of north american moose. swedish wildlife research supplement 1:311-325. geist, v. 1987. on speciation in ice age mammals, with special reference to cervids and caprids. canadian journal of zoology 65:1067-1084. . 1998. deer of the world: their evolution, behaviour, and ecology. stackpole books, mechanicsburg, pennsylvania, usa. goss, r. j. 1983. deer antlers: regeneration, function, and evolution. academic press, new york, new york, usa. hirth, d. h. 1977. social behavior of white-tailed deer in relation to habitat. wildlife monographs 53. hundertmark, k. j., r. t. bowyer, g. f. shields, and c. c. schwartz. 2003. mitochondrial phylogeography of moose (alces alces) in north america. journal of mammalogy 84:718 728. , g. f. shields, r. t. bowyer, and c. c. schwartz. 2002a. genetic relationships deduced from cytochrome-b sequences among moose. alces 38:113122. , , i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. schwartz. 2002b. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. molecular phylogenetics and evolution 22:375-387. , t. h. thelen, and r. t. bowyer. 1998. effects of population density and selective harvest on antler phenotypes in simulated moose populations. alces 34:375-383. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64:450-462. klein, d. r. 1968. the introduction, increase, and crash of reindeer on st. matthew island. journal of wildlife management 32:350-367. loranger, a. j., t. n. bailey, and w. w. larned. 1991. effects of forest succession after fire in moose wintering habitats on the kenai peninsula, alaska. alces 27:100-109. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. wildlife monographs 136. morphology of moose antlers bowyer et al. alces vol. 38, 2002 164 mccullough, d. r. 1979. the george reserve deer herd: population ecology of a k-selected species. university of michigan press, ann arbor, michigan, usa. . 1982. antler characteristics of george reserve white-tailed deer. journal of wildlife management 46:821826. mcgarigal, k., s. cushman, and s. stafford. 2000. multivariate statistics for wildlife and ecology research. springer, new york, new york, usa. moen, r., and j. pastor. 1998. a model to predict nutritional requirements for antler growth in moose. alces 34:59-74. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75:621630. neter, j., w. wasserman, and j. h. kutner. 1985. applied linear statistical models: regression, analysis of variance, and experimental designs. second edition. irwin, homewood, illinois, usa. peek, j. m. 1974. a review of moose food h a b i t s t u d i e s i n n o r t h a m e r i c a . naturaliste canadien 101:195-215. . 1998. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , r. e. leresche, and d. r. stevens. 1974. dynamics of moose aggregations in alaska, minnesota, and montana. journal of mammalogy 55:126-136. , d. l. ulrich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. sæther, b.-e., and h. haagenrud. 1985. geographical variation in the antlers of norwegian moose in relation to age and size. journal of wildlife management 49:983-986. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23:315-321. solberg, e. j., and b.-e. sæther. 1993. fluctuating asymmetry in the antlers of moose (alces alces): does it signal male quality? proceedings of the royal society of london b 245:251-255. , and . 1994. male traits as life-history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammalogy 75:1069-1079. stewart, k. m., r. t. bowyer, j. g. kie, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36:77-83. telfer, e. s. 1978. cervid distribution, browse and snow cover in alberta. journal of wildlife management 42:352361. , and j. p. kelsall. 1984. adaptations of some large north american mammals for survival in snow. ecology 65:1828-1834. van ballenberghe, v. 1987. effects of predation on moose numbers: a review of recent north american studies. swedish wildlife research supplement 1:431-460. , and w. b. ballard. 1994. limit a t i o n a n d r e g u l a t i o n o f m o o s e populations: the role of predation. canadian journal of zoology 72:1345-1354. weixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces vol. 38, 2002 bowyer et al. morphology of moose antlers 165 alces 34:213-238. williams, j. d., w. f. krueger, and d. h. hamel. 1994. heritabilities for antler characteristics and body weight in yearling white-tailed deer. heredity 73:7883. 175 45th north american moose conference and workshop rainy river community college / voyageurs national park, international falls, minnesota 23-26 june 2010 conference attendees, more than 40 members of the public attended the free event. invited speakers included lee frelich, art rodgers, mark lenarz, erika butler, steve windels, and rolf peterson. all speakers touched on the potential effects of climate change on moose, sparking some heated exchanges at times with members of the audience. the event was co-sponsored by the voyageurs national park association and the minnesota deer hunters association. contributed paper sessions continued on friday morning with seven presentations about moose health and disease issues. after lunch, delegates had the option to attend a moose necropsy workshop or a boat tour of voyageurs national park; both events were well attended. the workshop consisted of participants rotating through four stations covering a range of topics related to necropsies of moose and other mammals. the stations were hosted by kevin keel, murray lankester, rolf peterson, and erika butler/mike schrage. many of those who attended said it was the highlight of the conference. the boat tour was nearly canceled because of severe thunderstorms but the weather cleared to a beautiful sunny june afternoon at the last-minute. complimentary beverages were flowing and the leisurely guided cruise along the shores of rainy lake provided some wildlife viewing highlights, including a swimming deer (right on schedule), nesting bald eagles, and many other waterbirds. the banquet was held friday night at the thunderbird resort on rainy lake. the evening was kicked off with a stirring slideshow tribute by vince crichton to the founders of the north american moose conference, pat karns and al elsey. pat and al organized the first conference in st. paul, mn in 1963, with the theme “moose in a warming world”, the 45th north american moose conference and workshop was held at rainy river community college in international falls, minnesota from june 23-26, 2010. the conference was hosted by voyageurs national park, with substantial support from the university of minnesota-duluth/natural resources research institute, minnesota department of natural resources, fond du lac band of lake superior chippewa, 1854 treaty authority, central lakes college, and lakehead university. there were 114 paid delegates from throughout north america (15 states and 8 provinces), sweden, and norway. this nearrecord total included a great turnout by the minnesota delegation (47) and 17 students, which bodes well for future conferences. many delegates arrived on wednesday and attended the welcome reception that evening to reconnect with old friends and colleagues. dennis simon, wildlife section chief with the minnesota department of natural resources, gave welcoming remarks on behalf of the minnesota delegation to open the conference thursday. a half-day of conventional papers followed including sessions on thermal ecology of moose and moose management (moose love broccoli, apparently). the morning session ended with ken childs’ capstone presentation as the 2009 recipient of the distinguished moose biologist award. eight contributed posters and vendors from four telemetry companies were present for the entire conference. the conference reconvened after lunch at the historic backus auditorium in downtown international falls for the plenary session, continuing on the theme of “moose in a warming world.” in addition to the regular 176 and both men passed away in late 2009, only 25 days apart. catherine haas and david wattles received the newcomer awards and rolf peterson received the senior’s award. the members of the organizing committee received the order of alces, and volunteers from voyageurs national park were recognized for their efforts on behalf of the conference. minnesota’s own mike schrage received the 2010 distinguished moose biologist award for his contributions to moose management and research in the state. the annual banquet auction followed, with martha minchak taking the lead in showing items to attendees. rolf peterson was so determined that the auction make a profit that while he didn’t figuratively give the shirt off his back, he literally donated the moose tie off his front! a full day of contributed papers followed on saturday with sessions addressing advancements in techniques for moose research and management, moose-habitat relationships, effects of moose on forest health, and humanmoose interactions. a traditional walleye lunch, including fry bread and hand-harvested wild rice, was provided and prepared by the red lake band of chippewa indians. the alces business meeting followed lunch, including a presentation of the new online submission process for the journal. ten delegates attended a pre-conference field trip to isle royale hosted by rolf peterson and staff from isle royale national park. highlights included tours of banksund cabin and passage island and a reenactment of the climactic scene from nevada barr’s crime novel winter study. a number of conference delegates also joined in a post-conference field trip along the north shore of lake superior organized by gord eason. the conference was an overwhelming success with 36 contributed papers, 6 plenary presentations, the capstone lecture, 8 posters, 4 vendors, a necropsy workshop, and a boat tour. many thanks are owed to the organizing committee, volunteers, and invited speakers who made the event so successful. we especially thank our sponsors, including (in alphabetical order) 1854 treaty authority, bureau of indian affairs, central lakes college, fond du lac band of lake superior chippewa, isle royale national park, lakehead university chapter of the wildlife society, minnesota deer hunters association, minnesota department of natural resources, natural resources research institute, red lake band of chippewa indians, vectronic aerospace, voyageurs national park, voyageurs national park association, and the minnesota chapter of the wildlife society. additional thanks to those who donated items to the auction. chairs: ron moen and steve windels host: voyageurs national park location: international falls, minnesota date: 23-26 june, 2010 number of delegates / participants: 114 44_table_of_contents v2.pdf alces 44 (2008) contents mountain caribou interactions with wolves and moose in central british columbia..........................dale r. seip 1 the importance of individual variation in defining habitat selection by moose in northern british columbia.............................michael p. gillingham and katherine l. parker 7 effects of plant compensation across sites on regression estimates of shoot biomass and length..... ...................................................................roy v. rea and michael p. gillingham 21 does moose browsing threaten european aspen regeneration in koli national park, finland?..................... ......sauli härkönen, kalle eerikäinen, riikka lähteenmäki, and risto heikkilä 31 differential habitat selection by moose and elk in the besa-prophet area of northern british columbia ....................................................michael p. gillingham and katherine l. parker 41 recovery of low bull:cow ratios of moose in interior alaska...................donald d. young jr. and rodney d. boertje 65 grain overload and secondary effects as potential mortality factors of moose in north dakota................... ......erika a. butler, william f. jensen, roger e. johnson, and jason m. scott 73 an examination of the absence of established moose (alces alces) populations in southeastern cape breton island, nova scotia, canada............................................ ........................................................karen beazley, helen kwan, and tony nette 81 what do we know about nocturnal activity of moose?.......................................................nicole a. klassen and roy v. rea 101 using gis to modify a stratified random block survey design for moose...................douglas c. heard, andrew b. d. walker, jeremy b. ayotte, and glen s. watts 111 assessing re-colonization of moose in new york with hsi models.......................................................................lisa hickey 117 (continued on inside back cover) management and challenges of the mountain pine beetle infestation in british columbia............................................chris ritchie 127 43nd north american moose conference and workshop.................. 137 order of alces 2007 ..................................................................................... 137 previous meeting sites of the north american moose conference and workshop ........................................................................ 138 kris j. hundertmark distinguished moose biologist 2007 recipient ................................. 139 distinguished moose biologist past recipients .............................. 140 distinguished moose biologist award criteria ............................... 141 editorial review committee ....................................................................... 142 additional copies available from: lakehead university bookstore, thunder bay, ontario, canada p7b 5e1 alces 39-44 price $40.00 canadian or u.s. each (including supplementary issues) alces 24-38 price $38.00 canadian or u.s. each (including supplementary issues) alces 17-23 price $20.00 canadian or u.s. each proceedings of the north american moose conference and workshop 8-16 price $20.00 canadian or u.s. each make cheques, money orders or purchase orders payable to lakehead university bookstore. all prices include 5% g.s.t., mailing and handling costs. prices are subject to change. acknowledgements brooke pilley and susan rodgers worked long hours formatting and typesetting manuscripts. alces home page further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our website: http://bolt.lakeheadu.ca/~alceswww/alces.html alces vol. 44, 2008 gillingham and parker variation in habitat selection 7 the importance of individual variation in defining habitat selection by moose in northern british columbia michael p. gillingham and katherine l. parker natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia, canada v2n 4z9, email: michael@unbc.ca abstract: understanding resource use and selection has been central to many studies of ungulate ecology. global positioning satellite (gps) collars, remote sensing, and geographic information systems (gis) now make it easier to examine variation in use and selection by individuals. resource selection functions, however, are commonly developed for global (all animals pooled) models and important information on individual variability may be lost. using data from 14 female moose (alces alces) collared in the muskwa-kechika management area of northern british columbia, we examined differences among global and individual resource selection models for 5 seasons (winter, late winter, calving, summer, and fall). the global models indicated that moose selected for mid-elevations, and for deciduous burns and carex sedge areas in all seasons. resource selection models for individuals, however, indicated that no individuals selected the same attributes as the global models. we also examined selection ratios among seasons with individual moose as replicates, and within individuals with bootstrapping techniques. we discuss the importance of considering individual variation in defining resource selection and habitat use by moose and contrast the results of selection ratios and resource selection models. we also use these data to illustrate some of the pitfalls that can be encountered using the 2 methodologies. alces vol. 44: 7-20 (2008) key words: alces alces, habitat selection, home range, individual variation, resource selection, selection ratio to better inform management strategies, wildlife research has long focused on understanding use of habitats and, when combined with the availability of resources, what animals select and avoid on the landscape. early studies using radio telemetry examined populationlevel habitat selection (e.g., neu et al. 1974, reviewed by thomas and taylor 1990) usually concentrating specifically on habitat types (reviewed by alldredge and griswold 2006). the availability of global positioning satellite (gps) collars and advances in remote sensing and geographic information systems (gis) now enable researchers to more easily examine variation in selection among individuals (thomas and taylor 1990, 2006). numerous recent studies on selection by ungulates including moose (e.g., osko et al. 2004, dussault et al. 2005, poole and stuartsmith 2005, 2006, poole et al. 2007) have used resource selection functions (rsf, sensu boyce and mcdonald 1999, boyce et al. 2002, manly et al. 2002), although other multivariate approaches have also been employed (nikula et al. 2004). the rsf models provide a broadscale perspective of general selection patterns on the landscape (boyce and mcdonald 1999, manly et al. 2002). they also accommodate any type of habitat variables (categorical and continuous) and easily incorporate spatial data acquired from gis or remote sensing (boyce and mcdonald 1999). our best understanding of variation in resource selection by ungulates comes from study designs in which use and availability of resources are measured for individual animals variation in habitat selection – gillingham and parker alces vol. 44, 2008 8 (design iii in thomas and taylor 1990, 2006). gps locations can provide relatively accurate estimates of use by ungulates (but see d’eon et al. 2002, frair et al. 2004). multiple assumptions, however, are inherent in estimating resource availability for individual animals. various studies have assumed that random points drawn from the landscape or on an overlay of all home ranges (e.g., poole and stuart-smith 2006), from individual home ranges (e.g., nielson et al. 2002, gillies et al. 2006), from a buffer of potential movement radius (e.g., arthur et al. 1996), or from a radius of movement between consecutive fixes (i.e., matched-case; e.g., johnson et al. 2002) constitute a sample of individual resource availability. when availability is estimated for all animals as a group (design ii in thomas and taylor 2006), or when individuals are combined (pooled) in the analysis of data, an estimate of variation in individual selection potentially is lost. despite wide use of rsfs, there continues to be debate about their appropriateness (i.e., keating and cherry 2004, but see johnson et al. 2006). thomas and taylor (2006) identified several problems with researchers not meeting the statistical assumptions of selection modeling, but the main concerns include pooling of unequal sample sizes when individual animals are combined into global models, and a variety of issues around the estimation of unused points. some of the statistical concerns regarding selection models are difficult for researchers to accommodate. for example, the inherent attributes of telemetry (mortality, premature collar failure, fix-acquisition bias) are problematic because researchers should exclude animals or fixes in analyses (i.e., throw away data) in order to balance a sample design before building global models. until new statistical techniques emerge in the literature, however, rsfs will continue to be used by wildlife biologists. understanding selection of habitat attributes, in addition to measures of habitat use only, allows for a better understanding of the relative values of specific habitats in different landscapes. while studying resource selection by moose in northern british columbia, we observed that many individual collared animals completely avoided specific habitats. this avoidance may have occurred because of low animal densities or juxtaposition of particular habitats within seasonal ranges, but also because of avoidance of specific attributes such as low forage availability or predation risk (gillingham and parker 2008). the avoided habitats differed among individual moose such that when taken as a group, all available habitats were used by some collared animals in all seasons. in statistical modeling of resource selection, habitats that are completely (or nearly always) avoided must be dropped from individual resource selection models because of issues of complete, or near-complete separation (i.e., no or very low use of some levels of categorical variables). simpler analytical techniques such as selection ratios (manly et al. 2002) are not subject to the same constraints, but they cannot deal with continuous variables such as elevation and distance to specific features. to highlight the potential importance of individual variation in moose behaviour and selection, we constructed both individual and global rsf models and contrasted the results. as a baseline for habitat selection, we also examined individual and pooled selection ratios for collared moose in different seasons. we suggest that some of the biological information that may be lost in pooling animals warrants equal consideration with some of the statistical arguments that apply to the study of resource selection. study area the study area was located between 57°11’ and 57°15’ n, and 121°51’ and 124°31’ w, south of the prophet river and including the besa river, within the muskwa ranges and rocky mountain foothills. it covered a total alces vol. 44, 2008 gillingham and parker variation in habitat selection 9 area of ~740,887 ha within the muskwa-kechika management area (mkma) in northern british columbia. the besa-prophet study area is characterized by numerous east-west drainages and south-facing slopes. the underlying sedimentary rock formations are folded and faulted, and as is common along the eastern slopes of the rockies, potentially contain significant oil and gas reserves. at this time there is relatively little access into the besaprophet region other than several permanent outfitter camps and 1 government-designated, all-terrain vehicle trail. the majority of human activity occurs during the summer and fall with the start of hunting seasons; some snowmobile activity occurs during winter. there are primarily 3 biogeoclimatic zones in the besa-prophet study area: boreal white and black spruce (picea glauca and p. mariana) at lower elevations, spruce-willowbirch (salix spp., betula glandulosa) at midelevations (~1300-1600 m), and alpine tundra above ~1600 m (meidinger and pojar 1991). valleys at ~800-1300 m are lined with white spruce, some lodgepole pine (pinus contorta) and trembling aspen (populus tremuloides) on dry sites, and black spruce, willow-birch communities on poorly drained sites. there also are slopes that have been burned by the british columbia ministry of environment and local guide outfitters to enhance ungulate populations, primarily stone’s sheep (ovis dalli stonei). the spruce-willow-birch zone of the subalpine area is characterized by an abundance of willow and scrub birch, as well as some balsam fir (abies lasiocarpa) and white spruce often in krummholz form, and various grasses, sedges, and fescues (festuca spp.). alpine areas consist of permanent snowfields, rock, mat vegetation, and grasslands (demarchi 1996). methods fifteen adult female moose were fitted with gps collars (gtx, advanced telemetry systems, isanti, mn) in march 2003. collars were programmed to record locations every 6 hours for a 1-year sampling period. we defined 5 seasons that were distinguished by biological and ecological characteristics for our analyses of habitat selection by moose: winter (1 november–28 february) that corresponded to the formation of sex-specific groups following rut; late winter (1 march–15 may) when movement rates were lowest (gillingham and parker, unpublished); calving (16 may–15 june) when parturient females became solitary and the onset of plant greening occurred; summer (16 june–15 august) from plant green-up through peak vegetation biomass to the start of plant senescence; and fall (16 august–31 october) when senescence of vegetation occurred, males and females formed mixed sex groups, and females came into estrus. rsf model inputs vegetation classification – the vegetation classification system for the besa-prophet study area was developed using remote-sensing imagery and 227 field training sites (lay 2005). fifteen general vegetation associations were classified with a 2001 landsat enhanced thematic mapper (tm) image with 25-m resolution. we amalgamated several of these associations into 10 habitat classes to ensure that we had sufficient samples sizes for our analyses and an overall classification accuracy of >80%. classes were lumped according to similarity of vegetation and elevation, and associations relevant to moose (table 1). the 2 burn classes may also include some other small disturbed areas such as avalanche chutes, which could not be distinguished separately with remote-sensing imagery. we used the normalized difference vegetation index (ndvi) derived from landsat tm and enhanced thematic mapper (etm) to describe seasonal changes in vegetation (model described in gustine et al. 2006a). the tm (4 june and 22 july 2001) and etm (15 august 2001) images were used as a measure of vegetation biomass (june, july, and august) variation in habitat selection – gillingham and parker alces vol. 44, 2008 10 and vegetation quality (change in ndvi from june to july and from july to august). we assumed that the vegetation classes and relative differences in biomass and quality among classes were comparable among years in our study area. other gis inputs – we obtained elevation, slope, and aspect layers from a digital elevation model (dem) in the 1:20,000 british columbia terrain and resource inventory management program (british columbia ministry of crown lands 1990). to minimize issues of perfect separation between used and available points, we modeled aspect as 2 continuous variables (i.e., northness and eastness; roberts 1986); we did not assign an aspect to pixels with a slope ≤1°. northness (the cosine of aspect) values of 1.00 and -1.00 suggest selection for north and south aspects, respectively, whereas values near 0.00 suggest selection for east and west aspects. eastness (the sine of aspect) values show selection for east (i.e., 1.00) and west (i.e., -1.00) aspects; values of 0.00 show selection for northern/ southern exposures. we also defined potential risk of predation to moose using resource selection functions with logistic regression models by season from gps-collared wolves (canis lupus) and grizzly bears (ursus arctos) in the besaprophet area (details of predator models are in gustine et al. 2006a, b, walker et al. 2007). grizzly bears and wolves are assumed to be the most significant, large mammalian predators in the muskwa-kechika management area (bergerud and elliott 1998). the predationrisk models included slope, aspect, elevation, habitat class, fragmentation (an index of vegetation diversity), and distance to linear features (e.g., seismic lines). we generated a risk surface to define which areas have the highest selection values for grizzly bears or wolves in each season by applying the coefficients from models to each 25 x 25-m pixel in the besa-prophet, based on its topographic and vegetation features. we scaled values from 0 to 1 to standardize selection surfaces, and then assumed that the risk of predation to moose from grizzly bears and wolves was directly related to selection values from the rsfs of those species. habitat class description non-vegetated rock and rock habitats; permanent snowfields or glaciers and water bodies. elymus burn recently burned and open disturbed sites dominated by elymus innovatus. deciduous burn older burned and disturbed areas containing populus tremuloides and populus balsamifera shrubs (<2 m) and trees (≥2 m). subalpine deciduous shrubs ≥1600 m in elevation; and spruce-shrub transition zone at middle to upper elevations with white and hybrid spruce (picea glauca and p. glauca x engelmanni), and dominated by birch and willow. stunted spruce low productivity sites often on north-facing slopes with picea glauca of limited tree height and percent cover. pine-spruce white and hybrid spruce-dominated communities; and lodgepole pinedominated communities. riparian low-elevation, wet areas with black (picea mariana) and hybrid spruce; often with standing water in spring and summer; exposed gravel bars adjacent to rivers and creeks. alpine dry alpine tundra habitat ≥1600 m characterized by dryas spp.; and wet alpine tundra habitat ≥1600 m dominated by cassiope spp. and sedge (carex spp.) meadows. low shrub deciduous shrubs <1600 m dominated by birch and willow. carex wetland meadows dominated by sedges (carex spp.) at elevations <1600 m, with intermittent salix shrubs. table 1. description of the 10 habitat classes used to describe habitat selection by moose in the besaprophet area of northern british columbia. alces vol. 44, 2008 gillingham and parker variation in habitat selection 11 determining use and availability gps locations from telemetered moose were screened for fix quality (points with positional dilution of precision >25 were dropped) and for improbable fixes (spatial viewer, unpublished visual basic program; m. p. gillingham). to determine availability of resources for individual animals, we used all movement rates from consecutive 6-hour gps fixes for each animal in a season, and determined the 95th percentile distance traveled during 6 hours. our reasoning was that 95% of the time an animal typically moves within this movement potential (arthur et al. 1996). the movement potential, therefore, generally represents how far an animal could have moved and the movements shorter than the potential represent choices that the animal made. the remaining 5% of movements included longer distances traveled during a 6-hour period within a season and were likely evoked by less common conditions (e.g., migratory movements and transitional movements prior to calving; gustine et al. 2006a, 2006b). a circular buffer with the corresponding 95th percentile radius was then placed around each gps location (used point) and we randomly selected 5 points from that area to represent availability. the circular buffer was defined by distance only and did not exclude physical constraints or barriers to movements such as cliffs. nonetheless, we believe this is a better representation of what was available in the vicinity of the moose versus selecting points from a very large minimum convex polygon (mcp) of home-range size or a kernel based on density of use. to examine possible issues of lack of independence among animals (e.g., 2 or more collared animals spending large amount of time in close proximity), we calculated the minimum distance between every collared animal whenever a location was obtained. finally, we checked to ensure that no 2 points were used twice and that there was no overlap between used and available points (manly et al. 2002). we then used a raster gis (imageworks xpace; pci geomatics corp. 2001) to query attributes in all gis layers for used and random points. we did not consider the used and available points to be matched (i.e., casecontrolled) because buffers for random points were selected based on seasonal movement potential and not the distance moved from the last fix (i.e., we used a design iii rather than a design iv; thomas and taylor 2006). resource selection modeling we developed 11 a priori, ecologically plausible models (table 2) to describe resource selection in an information theoretic framework using akaike’s information criterion (aic; burnham and anderson 2002), and evaluated the relative importance of each of the variables in the models using selection coefficients (β) from logistic regression. we used tolerance scores (threshold of <0.20) to assess all model variables for collinearity, which can inflate selection coefficients and error terms (menard 2002). the same suite of models was used for individual moose and all moose (pooled data), but not all models were run in all seasons (i.e., no risk of predation by bears during hibernation; some variables were dropped because of collinearity in models; table 2). we ranked the suite of models using aic values corrected for small sample size (aicc; burnham and anderson 2002) and then validated all top models for individuals and pooled animals using k-fold, cross-validation (boyce et al. 2002) and an averaged spearman’s rank correlation coefficient. within each model set (i.e., by individual and season), we calculated akaike weights (wi), which are an estimate of the relative weight of evidence that the top model is the best within a model set. in cases for which there was not a single model with wi ≥0.95, we considered competing models until the sum of wi was ≥0.95 (burnham and anderson 2002). for each model set, the selection coefficients (β) in competing models variation in habitat selection – gillingham and parker alces vol. 44, 2008 12 were averaged based on their relative wi. all statistical analyses were run in stata (version 9.2; statacorp 2006) and we used the add-in desmat (hendrickx 1999) for deviation coding of categorical variables. logistic regression models do not provide reliable estimates if there is either complete or near-complete separation (few cases of presence or absence) in levels of categorical variables (menard 2002). in our study, this occurred whenever individual moose completely avoided an available habitat or used it very infrequently in a season. to avoid issues of separation, for each individual we dropped both used and available points in habitats for which either use or available points were rare (i.e., <5 points). therefore, strong avoidance of an available habitat by individual moose is not reflected in the final individual resource selection functions if that avoidance was complete or near-complete. estimates of variation around selection for both individual and global models were obtained directly from fitting logistic regressions. selection ratios because we dropped several habitat classes for each moose in our rsf modeling, we also calculated selection ratios (manly et al. 2002) so that rarely used habitats were not ignored in our analyses. we took the ratio of used (gps locations) to available (random) points; available habitat types were divided by 5 (because we chose 5 random points per fix) before calculating each ratio. each individual moose was treated as a replicate and selection ratios by habitat class were averaged across individual animals within each season. to compare selection ratios to the β coefficients from the individual rsf models, we estimated variation in selection ratios for each individual by bootstrapping the used and available points. for each animal and season, we randomly selected 100 replicates; in each repmodel calving summer fall winter late winter elevation1+aspect+habitat2 yes yes yes yes yes elevation+slope+aspect+habitat yes yes yes yes yes wolf 3+habitat yes yes yes yes yes elevation+slope+aspect+wolf+bear3+biomass+habitat yes yes yes elevation+slope+aspect+wolf+bear+quality+habitat yes yes aspect+wolf+bear+biomass+habitat yes yes yes aspect+wolf+bear+quality+habitat yes yes elevation+slope+aspect+wolf+biomass+habitat yes yes yes elevation+slope+aspect+wolf+quality+habitat yes yes aspect+wolf+biomass+habitat yes yes yes aspect+wolf+quality+habitat yes yes table 2. candidate models for all animals pooled (global) and individuals, developed a priori to describe resource selection by moose by season in the besa-prophet area of northern british columbia. vegetation biomass for calving, summer, and fall were based on ndvi values from june, july, and august, respectively. vegetation quality, assessed by the change in ndvi between summer months, was used only in calving and summer models. no risk of predation by grizzly bears was included during hibernation (winter and late winter seasons). 1 elevation was modeled as a quadratic with both a linear and squared term. 2 habitat classes are described in table 1. 3 wolf and bear represent risk of predation by wolves and grizzly bears, respectively; see text for details. alces vol. 44, 2008 gillingham and parker variation in habitat selection 13 licate 80% of the available used points (along with their corresponding random available points) were sampled. these data were then used to estimate a within-animal and season variance for each selection ratio. results we obtained 14,534 gps locations from 14 of the collared moose. the fix rates (i.e., the number of gps fixes recorded as a percentage of the number of attempted gps fixes) averaged 76.7 ± 0.03% (x ± se, range among individuals = 56-90%). animals were assumed to be independent. average distances between individuals at any one point in time ranged from 8.6 km in late winter to 12.2 km in summer, with a maximum value of 54.5 km between individuals during fall. closest locations among collared animals also occurred in fall, but fewer than 1.3% of locations were within 250 m and <1.9% were within 500 m of another collared individual. there was relatively good agreement in the signs of significant selection coefficients for continuous variables when individual and global models were compared (table 3). in no seasons, however, did the sign of the coefficients that were significant in the individual models correspond completely with significant attributes in the global model. for example, moose appeared to always select for mid elevations (positive linear term and negative quadratic term) based on global models, but this was reversed for at least 1 individual animal during both the calving and winter seasons. given that there were 14 individual models for each season (except in winter when n = 12), many of the significant global coefficients corresponded to similar selection in less than half of the individual models (table 3). individual variation in selection associated with the continuous variables also indicated that there were several parameters that were significant in some individual models that were not supported by the global models. for example, slope was important in 9 of the 14 individual summer models, but there was no significant selection for slope in the global summer model. pooling of animals to build the global seasonal models resulted in rsf models in which all habitat classes were included (fig. parameter calving summer fall winter1 late winter + – all + – all + – all + – all + – all elev (km) 7 1 7.44 10 0 30.15 7 0 24.07 6 2 4.12 7 0 8.75 elev (km2) 1 7 -2.39 0 10 -10.7 0 7 -8.23 1 7 -1.52 0 7 -2.85 slope 1 6 -0.03 1 8 0 8 -0.03 2 4 2 4 -0.01 northness 0 1 0 1 1 0 0 1 eastness 1 0 0 1 0 1 wolf risk 1 2 2 0 1.06 4 0 2 0 1 2 bear risk 1 1 -1.18 2 2 biomass 0 1 1 3 -1.7 1 5 -0.76 quality 2 1 1.99 3 0 0.45 table 3: comparison of significant selection coefficients by season for continuous variables from individual and global (all animals) resource selection models for 14 female moose in the besa-prophet area of northern british columbia. for each season, the number under the + indicates the number of individual final models that showed significant selection for that parameter; the number under the – indicates the number of individuals that significantly avoided that attribute. the significant β coefficients in each seasonal global model for all animals are shown under ‘all’. 1 only 12 animals were used in the winter models because of collar failure. variation in habitat selection – gillingham and parker alces vol. 44, 2008 14 1a). even though all habitat classes were available to all moose, individual animals frequently avoided habitat classes such that those classes could not be included in individual models. in winter, 2 moose used all 10 habitat classes, but in all other seasons each of the collared moose completely avoided at least 1 habitat class. as many as 7 (calving and fall), 6 (late winter and summer), and 4 (winter) habitats were dropped from all analyses because of near-complete separation for individual animals. the global models were able to incorporate all habitat classes without issues of complete separation, but there were no ‘average’ individuals in our sample that exhibited this habitat selection. because many of the individuals completely avoided the same habitat classes in a given season, inference in the global seasonal models for some habitat classes was actually based on as few as 1 3 individuals. the global rsf models suggested selection for, and avoidance of, many more habitat classes than did the selection ratios (fig. 1a and 1b), in part because of the small sample sizes for the selection ratios (number of individuals). most habitat classes that were identified as important by the selection ratios, however, were also important in the rsf models, with the exception of subalpine and riparian habitats during calving, and stunted spruce, pine-spruce, and riparian habitats during fall (fig. 1). there were no instances in which a habitat that was significantly selected in the rsf models or selection ratios was significantly avoided in the other. although the bootstrapped estimates of individual selection ratios were only an approximation of selection, they were not nearly as affected by near-complete avoidance of specific habitat classes in particular seasons as rsf models were (i.e., selection ratios could be computed as long as there was at least some use of a habitat type; table 4). because of the large number of habitat classes that were dropped from the rsf modeling due to separation issues, the majority of individual rsf models for moose did not show strong selection or avoidance of any habitats (n = 14 models in all seasons except winter for which n = 12; table 4: rsf β+ and β–). in contrast, the bootstrapped estimates of selection ratios provided much stronger evidence of selection and avoidance of most habitat classes in most seasons (although these estimates were not being corrected for the continuous variables that also were incorporated in the rsf models). for example, selection ratios indicated that twice as many individual moose selected for subalpine in summer and against alpine in fall as when determined by rsf models. in the 0 1 2 3 4 5 -3 -2 -1 0 1 2 calving summer fall winter late winter ***** * * * * * * ** ** * * * *** * * * ** * * * * * ** * ** * * * *** ** * * * * * * * * * a b no n-v ege tate d ely mu s b urn de cid uou s b urn su bal pin e stu nte d s pru ce pin e-s pru ce rip aria n alp ine low sh rub ca rex no n-v ege tate d ely mu s b urn de cid uou s b urn su bal pin e stu nte d s pru ce pin e-s pru ce rip aria n alp ine low sh rub ca rex r s f s el ec tio n  c oe ffi ci en ts s el ec tio n r at io fig. 1. comparison of seasonal selection of habitat classes by female moose based on selection coefficients (β) of global resource selection models (a) and selection ratios (b) in the besa-prophet study area in northern british columbia. error bars represent 1 se; * indicate significant β coefficients (a; significantly different from 0) or selection ratios (b; significantly different from 1). alces vol. 44, 2008 gillingham and parker variation in habitat selection 15 extreme, 10 of 14 moose selected for carex associations during calving when analyzed by selection ratios, yet no significant selection by any individual occurred in the rsf models. discussion resource selection functions are a powerful tool for incorporating both continuous and categorical variables in studies of resource selection and they are widely used for many wildlife species. our results, however, suggest that care should be taken in interpreting global (cross-animal) rsf models even when all of the statistical assumptions of the analyses are met. in particular, although most models are developed to make populationlevel inferences, the variation in individual selection may be important to researchers as well as to outcomes of global models. the work of thomas and taylor (1990) focused considerable attention on examining individual variation in selection (design iii) and led to the development of new methods of modeling resource selection (e.g., manly et al. 2002), but many studies inherently avoid examining individual variation by building global models. recently, new analytical approaches allow for individual effects. thomas et al. (2006), for example, presented a bayesian random-effects model to assess resource selection. in addition, approaches for explicitly accounting for individual animals have been developed in discreet-choice models (e.g., buskirk and millspaugh 2006). these models provide simultaneous estimation of both individualand population-level selection. individual effects can also be included in more traditional rsf approaches (gillies et al. 2006). there is a difference, however, between accounting for individual effects in models and examining important individual habitat calving summer fall winter1 late winter rsf sr rsf sr rsf sr rsf sr rsf sr β+ β– r+ r– β+ β– r+ r– β+ β– r+ r– β+ β– r+ r– β+ β– r+ r– nonvegetated 0 1 0 3 0 4 0 6 1 3 1 6 0 4 0 8 0 2 0 7 elymus burn 1 1 7 2 2 1 4 7 4 0 3 7 1 1 3 8 2 0 4 5 deciduous burn 3 1 11 2 5 0 10 2 8 0 8 3 6 1 7 4 2 0 8 5 subalpine 3 3 5 1 10 3 11 0 12 0 9 0 10 2 2 0 5 7 stunted spruce 1 0 7 4 5 0 4 9 0 1 1 10 1 0 4 6 6 1 7 2 pine-spruce 2 2 5 9 5 0 9 5 1 3 2 10 2 6 3 8 2 1 5 8 riparian 1 1 3 9 3 1 3 7 0 2 3 7 1 0 3 3 1 0 3 9 alpine 1 0 0 7 0 5 0 10 0 4 0 10 0 1 1 5 low shrub 1 0 7 5 3 4 10 3 9 0 12 0 6 0 9 3 2 0 6 5 carex 10 3 1 1 6 2 6 1 1 0 3 3 2 4 table 4. comparison of significant categorical habitat classes using selection coefficients (β) from individual resource selection (rsf) models and individual bootstrapped selection ratios (sr; see text) for 14 female moose in the besa-prophet area of northern british columbia. for each season, the number under the β+ indicates the number of individual final rsf models that showed selection for that parameter; the number under the β– indicates the number of individuals that avoided that attribute in rsf models. similarly, r+ and r– correspond to the number of individuals that showed significant selection ratios for and against the habitat class, respectively. 1 only 12 animals were used in the winter models because of collar failure. variation in habitat selection – gillingham and parker alces vol. 44, 2008 16 differences in selection. we argue that the challenge (both statistical and methodological) is to understand what individual animals select and avoid. although techniques are available for analyzing presence-only data (see pearce and boyce 2006), we cannot examine the choices animals make without comparing used and available points. from a management perspective, there is a demand for global rsf models. in principle, global models derived from appropriately pooling data (see review of thomas and taylor 2006) from individuals should provide the average response of the population if the sample of individuals is representative of the selection strategies within the population. in addition, many studies of resource selection that use rsf models result in resource selection probability functions (rspf; manly et al. 2002) that are mapped as surfaces in a gis. despite some problems interpreting these surfaces (e.g., keating and cherry 2004), they can spatially and concisely depict the results of global models, but unfortunately without individual variation. in our study, there was generally good agreement between global and individual models, although there were some important differences. in order to meet statistical assumptions, data were dropped for categorical (habitat class) variables, and in some cases, these necessary statistical procedures may have affected the biological interpretation of selection by moose on the landscape. only the pine-spruce habitat was used by all individuals in all seasons. when several individuals almost completely avoided a particular habitat class in a given season, inference in the global seasonal models was based on a few individuals and the resultant selection coefficients were sometimes misleading. for example, during late winter, use of carex habitats by 11 of 14 female moose was so limited that carex could not be included in individual models, even though it was almost completely avoided by those individuals. when all animal locations and habitat availability were incorporated, however, the global model for this season indicated that moose selected for carex because of the behaviour of a few individuals. the same situation occurred in fall when only 4 moose used carex habitats enough to be included in individual models, but again the global model indicated selection for carex. when comparing all animal-use locations to all available locations without regard to sample sizes per individual, the global model appears to be biologically misleading. in all, we recorded 5 instances for which at least half the moose avoided a habitat class so extensively that it had to be dropped from individual models, but the global model indicated selection for that class. in addition, there were 5 instances in which habitats were dropped for more than half of the moose, but for which selection coefficients in the global model did not indicate avoidance. there were also 6 season-habitat models for which more than half of the animals completely avoided the habitat, and the ‘correct’ conclusion was drawn from the rsf models because the remaining animals showed significant avoidance of those habitats even though that inference was being drawn from as few as 1 individual. it is important to note that selection by individuals reflects the choices that they have to draw from, which may include different seasonal range configurations separating foraging from resting areas, for example, or individual demands related to physiological condition. these differences among animals are real and, therefore, it is important to understand the variability within the population. selection ratios can be calculated from a single use point in a habitat class, unlike rsf selection coefficients, and seemingly may provide a better measure of habitat selection. they are also much simpler to calculate. selection ratios and similar selection indices, however, can only accommodate use and avoidance of categorical variables (alldredge and griswold 2006), and the influence of continuous alces vol. 44, 2008 gillingham and parker variation in habitat selection 17 variables is not incorporated. theoretically, continuous variables such as elevation or slope could be partitioned into categories and then combinations of multiple categorical variables using selection ratios or similar indices could be assessed, but the interpretation of those results would be difficult. therefore, in most cases researchers are probably dependent on analytical techniques that result in the exclusion of little-used habitat classes when examining the influence of both continuous and categorical variables on resource selection. in those instances when continuous variables such as slope, elevation, and aspect do not vary much on the landscape, selection ratios could be effective in quantifying selection of rarely used habitat classes. we are not advocating the substitution of selection ratios for rsf modeling. we have used these ratios to demonstrate what information was lost by rsf models alone in this study. if managers and biologists are interpreting rsf coefficients as the average response of animals in a population, they should also examine use and availability to ensure that the responses of all individuals are reflected in the coefficients and their measures of variation. categorical variables (e.g., habitat class) that are never or rarely used by individuals should be reported as measures of avoidance if those resources are available. the purpose of our analyses was to show the extent and influence of individual variation in defining habitat selection by moose. these findings are from relatively few individuals (n = 14) from a relatively short period of time (1 year). they provide initial insights into habitat selection by moose in northern british columbia, but more importantly, they appear to show a large range in variation among individuals. it is possible that with a gps sampling rate of >4 fixes per day and over multiple years, fewer habitat classes would have been dropped in seasonal analyses, although the landscape of the besa-prophet area is spatially heterogeneous and animals could easily have used and had access to all habitat classes within our 1 year of measurements. longer-term data sets would certainly lend themselves to more robust rsf analyses even though year effects might be introduced. within-individual variation probably also occurred in response to reproductive status (e.g., with or without a calf) and age, both of which can influence habitat selection in relation to nutritional demands and predation risk (e.g., bowyer et al. 1999). unfortunately, we did not have exact age and reproductive information for the gps-collared animals in this study. nonetheless, if the individual variation that we observed within these few animals was representative of many more animals at the population level, knowledge of the different selection strategies should be important to wildlife biologists and managers. our findings indicate that these strategies may be masked using global selection models. in this study, the acquisition rates of the collars were low and may reflect features of the terrain (d’eon et al. 2002, frair et al. 2004), vegetation (rempel et al. 1995, d’eon et al. 2002), or leaf cover (dussault et al. 1999, d’eon 2003). the habitat class that would likely have the poorest signal attenuation would be the pine-spruce habitat, but it was the only habitat in which use was recorded for all animals. we do not know which habitat classes may have been under-represented in our samples. by using stationary collars we potentially could have developed corrections for each habitat class (frair et al. 2004), but even the behaviour of individual animals has been shown to influence fix rates (d’eon 2003, graves and waller 2006). nonetheless, because we used the same data, our comparisons of individual and global models and between rsf models and selection ratios per se would not be affected by fix acquisition biases. unless habitats are categorized into very broad types and all habitats are used extensively by all individuals, issues of complete, or near-complete separation will continue variation in habitat selection – gillingham and parker alces vol. 44, 2008 18 to pose statistical problems when analyzing resource selection data. when using rsf models, researchers should be explicit about which habitat classes (or levels of any categorical variables) are dropped during analyses so that this information (i.e., total avoidance of specific attributes) is made known in addition to model results. if we had not set out to examine individual variation in selection by moose, we would not have realized that inference about habitat selection was based on very few individuals for several habitat classes in different seasons. therefore, we believe that caution must be taken when pooling individuals (in addition to stated statistical limitations) not only because of the potential loss of important individual variation, but also because all animals probably do not exhibit the average responses predicted by global models. researchers with access to long-term data sets with numerous individuals and a high frequency of sampling in heterogeneous environments should attempt to define whether incorporating more individuals in a global model can ever encompass the range of individual strategies for a given population, or whether knowledge of the different selection strategies within a population is more important to effective management of habitats. acknowledgements support for this project was provided by the muskwa-kechika trust fund, the british columbia ministry of environment and the university of northern british columbia. we extend particular thanks to g. williams who helped familiarize us with the besa-prophet landscape. r. b. woods of the british columbia ministry of environment captured and collared all of the animals monitored in this study. we are also grateful to those who assisted in data visualization and analyses (j. b. ayotte, s. g. emmons, d. d. gustine, r. j. lay, b. milakovic, a. b. d. walker), and to external reviewers who provided very constructive comments on the manuscript. references alldredge, j. r., and j. griswold. 2006. design and analysis of resource selection studies for categorical resource variables. journal of wildlife management 70:337346. arthur, s. m., b. j. manly, l. l. mcdonald, and g. w. garner. 1996. assessing habitat selection when availability changes. ecology 77:215-227. bergerud, a. t., and j. p. elliott. 1998. wolf predation in a multiple-ungulate system in northern british columbia. canadian journal of zoology. 76:1551-1569. bowyer, r. t., v. van ballenberghe, j. g. kie, and j. a. k. maier. 1999. birthsite selection by alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80:1070-1083. boyce, m. s., and l. l. mcdonald. 1999. relating populations to habitats using resource selection functions. trends in ecology and evolution 14:268-272. _____, p. r. vernier, s. e. nielsen, and f. k. schmiegelow. 2002. evaluating resource selection functions. ecological modelling 157:281-300. british columbia ministry of crown lands. 1990. terrain and resource inventory management in british columbia specifications and guidelines for geomatics: digital baseline mapping at 1:20,000. ministry of crown lands for the government of british columbia, victoria, british columbia, canada. burnham, k. p., and d. r. anderson. 2002. model selection and multi-model inference: a practical information–theoretic approach. second edition. springerverlag, new york, new york, u.s.a. buskirk, s. w., and j. j. millspaugh. 2006. metrics for studies of resource selection. journal of wildlife management 70:358-366. d’eon, r. g. 2003. effects of a stationary alces vol. 44, 2008 gillingham and parker variation in habitat selection 19 gps fix-rate bias on habitat selection analysis. journal of wildlife management 67:858-863. _____, r. serrouya, g. smith, and c. o. kochanny. 2002. gps radiotelemetry error and bias in mountainous terrain. wildlife society bulletin 30:430-439. demarchi, d. a. 1996. introduction to the ecoregions of british columbia. british columbia wildlife branch, ministry of environment, lands and parks, victoria, british columbia, canada. dussault, c. r., r. courtois, j. ouellet, and j. huot. 1999. evaluation of gps telemetry collar performance for habitat studies in the boreal forest. wildlife society bulletin 27:965-972. _____, j. ouellet, r. courtois, j. huot, l. breton, and h. jolicoeur. 2005. linking moose habitat selection to limiting factors. ecography 28:619-628. frair, j. l., s. e. nielsen, e. h. merrill, s. r. lele, m s. boyce, r. h. m. munro, g. b. stenhouse, and h. l. beyer. 2004. removing gps collar bias in habitat selection studies. journal of applied ecology 41:201-212. gillies, c. s., m. hebblewhite, s. e. nielsen, m. a. krawchuk, c. l. aldridge, j. l. frair, d. j. saher, c. e. stevens, and c. l. jerde. 2006. application of random effects to the study of resource selection by animals. journal of animal ecology 75:887-898. gillingham, m. p., and k. l. parker. 2008. differential habitat selection by moose and elk in the besa-prophet area of northern british columbia. alces 44: 41-63. graves t. a., and j. s. waller. 2006. understanding the causes of missed global positioning system telemetry fixes. journal of wildlife management 70:844-851. gustine, d. d., k. l. parker, r. j. lay, m. p. gillingham, and d. c. heard. 2006a. calf survival of woodland caribou in a multi-predator ecosystem. wildlife monograph 165. _____, _____, _____, _____, and _____. 2006b. interpreting resource selection at different scales for woodland caribou in winter. journal of wildlife management 70:1601-1614. hendrickx, j. 1999. using categorical variables in stata. stata technical bulletin 52:2-8. johnson, c. j., k. l. parker, d. c. heard, and m. p. gillingham. 2002. a multiscale behavioral approach to understanding the movements of woodland caribou. ecological applications 12:1840-1860. _____, s. e. nielsen, e. h. merrill, t. l. mcdonald, and m. s. boyce. 2006. resource selection functions based on use-availability data: theoretical motivation and evaluation methods. journal of wildlife management 70:347-357. keating, k. a., and s. cherry. 2004. use and interpretation of logistic regression in habitat selection studies. journal of wildlife management 68:774-789. lay, r. j. 2005. use of landsat tm and etm+ to describe intra-season change in vegetation with consideration for wildlife management. m.sc. thesis, university of northern british columbia, prince george, british columbia, canada. manly, b. f., l. l. mcdonald, and d. l. thomas. 2002. resource selection by animals: statistical design and analysis for field studies. second edition. chapman-hall, london, u.k. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. british columbia ministry of forests, victoria, british columbia, canada. menard, s. 2002. applied logistic regression analysis. second edition. sage, thousand oaks, california, u.s.a. neu, c. w., c. r. byers, and j. m. peek. 1974. a technique for analysis of utilizationavailability data. journal of wildlife management 38:541-545. variation in habitat selection – gillingham and parker alces vol. 44, 2008 20 nielsen, s. e., m. s. boyce, g. b. stenhouse, and r. h. m. munro. 2002. modeling grizzly bear habitats in the yellowhead ecosystem of alberta: taking autocorrelation seriously. ursus 13:45-56. nikula, a., s. heikkinen, and e. helle. 2004. habitat selection of adult moose alces alces at two spatial scales in central finland. wildlife biology 10:121-135. osko, t. j., m. n. hiltz, r. j. hudson, and s. m. wasel. 2004. moose habitat preferences in response to changing availability. journal of wildlife management 68:576-584. pci geomatics corporation. 2001. pci works version 7.0. richmond hill, ontario, canada. pearce, j. l., and m. s. boyce. 2006. modelling distribution and abundance with presence-only data. journal of applied ecology 43:405-412. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammalogy 88:139-150. _____, and k. stuart-smith. 2005. finescale winter habitat selection by moose in interior montane forests. alces 41:1-8. _____, and _____. 2006. winter habitat selection by female moose in western interior montane forests. canadian journal of zoology 84:1823-1832. rempel, r. s., a. r. rodgers, and k. f. abraham. 1995. performance of a gps animal location system under boreal forest canopy. journal of wildlife management 59:543-551. roberts, d. w. 1986. ordination on the basis of fuzzy set theory. vegetatio 66:123-131. statacorp. 2007. stata version 9.1. college station, texas, u.s.a. thomas, d. l., d. johnson, and b. griffith. 2006. a bayesian random effects discrete-choice model for resource selection: population-level selection inference. journal of wildlife management 70:404-412. _____, and e. j. taylor. 1990. study designs and tests for comparing resource use and availability. journal of wildlife management 54:322-330. _____, and _____. 2006. study designs and tests for comparing resource use and availability ii. journal of wildlife management 70:324-336. walker, a. b. d, k. l. parker, m. p. gillingham, d. d. gustine, and r. j. lay. 2007. habitat selection and movements of stone’s sheep in relation to vegetation, topography and risk of predation. ecoscience 14:55-70. f:\alces\vol_38\pagema~1\3806.pdf alces vol. 38, 2002 koitzsch — moose habitat suitability index model 8 9 application of a moose habitat suitability index model to vermont wildlife management units ky b. koitzsch1 wildlife and fisheries biology program, university of vermont, burlington, vt 05405, usa abstract: habitat suitability index (hsi) models translate existing knowledge of a species’ habitat requirements into quantitative measures of habitat quality. the hsi is a numerical index that represents the ability of a given habitat to provide life requisites for a species on a scale from 0 (unsuitable habitat) to 1 (optimal habitat). habitat suitability index models are useful in natural resource planning for predicting the impacts of resource management practices on wildlife habitat. many moose (alces alces) hsi models require the labor-intensive collection of ground-level browse density data, which limits their applications for analyzing large landscapes required by moose. some, however, have been developed utilizing remotely sensed data to analyze large study areas. i tested the usefulness of one of these models, created for the lake superior region, to 2 wildlife management units (wmus) in vermont. areas of study wmus, “e1” and “i”, were 680 km2 and 729 km2, respectively. the model quantified 4 landscape-scale habitat variables representing annual cover types required by moose: percent area of regenerating forest, non-forested wetland, spruce/ fir forest, and deciduous/mixed forest. model analyses were performed using a geographic information system (gis). the model was useful in estimating relative habitat suitability of both wmus, identifying within-wmu habitat variation, quantifying change in habitat suitability following a natural habitat-altering event, and predicting temporal change in moose habitat due to changes in forest management practices. the model revealed significant differences in habitat suitability of 0.64 for wmu e1 and 0.34 for wmu i. to determine within-wmu habitat variation, both wmus were divided into 25-km2 evaluation units, which approximated the annual home range of moose in new england, and a hsi was calculated for each unit. habitat suitability of 81 km2 of wmu i increased from 0.30 to 0.53 due to an increase in regenerating forest following heavy canopy damage from an ice storm in january 1998. a reduction in habitat suitability from 0.81 to 0.35 of silvio o. conte national fish and wildlife refuge lands within wmu e1 was observed following a simulation in which all timber harvesting as a forest management practice was eliminated. initial validation of this model for analyzing moose habitat at the wmu-scale is supported by correlation of hsi output to moose harvest data for wmu e1 25-km2 evaluation units and by comparison of hsi to estimated moose densities for both wmus. alces vol. 38: 89-107 (2002) key words: alces alces, geographic information system (gis), habitat suitability index (hsi) model, moose, vermont, wildlife management unit (wmu) in the last 40 years, vermont has seen a considerable increase in the population size and distribution of eastern moose (alces alces americana). the population, estimated at 20 animals in 1960, was thought to exceed 2,500 in 1998, and continues to grow at a predicted rate of 1.10 moose/year (alexander 1993, alexander et al. 1998). in the same period of time, moose distribution has expanded from vermont’s extreme northeast corner to the entire state. in 1993 the vermont department of fish and wild1present address: p.o. box 953, waitsfield, vt 05673, usa moose habitat suitability index model — koitzsch alces vol. 38, 2002 9 0 life (vtdfw) initiated the first moose hunt in almost a century by issuing 30 harvest permits for one of vermont’s 26 wildlife management units (wmus). by 1999, the moose hunt was expanded to 10 wmus, representing approximately 51% of the state, and for the 2000 season 215 permits were issued. with an expanding moose population, public interest in non-consumptive uses of moose, such as viewing and photography, also have risen. in new england, moose habitat has been described in numerous studies (cioffi 1981, monthey 1984, crossley 1985, leptich and gilbert 1989, pruss and pekins 1992, thompson et al. 1995, alexander et al. 1998, and k. morris, maine department of inland fish and wildlife 1999, unpublished data). moose were found to require large habitats providing copious amounts of regenerating hardwood as their primary source of annual browse; young balsam fir (abies balsamea) at higher elevations as a source of winter browse; mature spruce (picea spp.) and balsam fir forests to escape the stressful effects of heat in summer and severe weather in winter (renecker and hudson 1986, 1990); and macrophyte-rich wetlands as a source of sodium in late spring and summer. pletscher (1987), while studying nutrient budgets for white-tailed deer in north-central new hampshire, demonstrated low sodium concentrations in terrestrial vegetation and suggested that deer made up for this deficiency by utilizing natural salt licks, artificial salt licks along salted roads, and aquatic vegetation. assuming sodium concentration in terrestrial vegetation throughout new england is low, as has been demonstrated for the lake superior region in work from isle royale (jordan et al. 1973, belovsky and jordan 1981), moose probably make up for sodium deficiencies in the same manner. wetland areas are also important for calving areas and as refugia from black bear (ursus americana) predation, and insects. by occupying large home ranges, moose in new england are able to meet their seasonal life requisites, which are often spatially separated. because of the presumed abundance and good quality of vermont’s moose habitat, the vtdfw does not routinely inventory or monitor moose habitat. it does, however, collect physical measurements from legally harvested and incidentally killed moose as indicators of population health and habitat condition. while these measurements may indicate that a moose population is approaching carrying capacity or is in decline due to habitat deficiencies, they cannot specify which habitat component is deficient. the present physical condition of vermont moose suggests that the herd is healthy and the habitat is productive (alexander et al. 1998). however, with increasing demands on vermont’s forests for recreation, forest products, conservation, and development, their ability to support moose could deteriorate. a tool that can inventory statewide moose habitat and predict habitat change due to change in forest management practices and natural disturbances will aid in the conservation and management of valuable moose habitats. this tool should be compatible with timber stand classification systems developed for timber management since forestry practices have such a great impact on moose habitat throughout its eastern range (hurley 1986). the tool also should be simple, inexpensive to apply and capable of analyzing large habitats required by moose. an effective tool to meet all these demands is the moose habitat suitability index (hsi) model (allen et al. 1987). the concept of the hsi model began with the development of the u.s. fish and wildlife service (usfws) sponsored habitat evaluation procedures (hep) in 1976. the hep quantified wildlife habitat based alces vol. 38, 2002 koitzsch — moose habitat suitability index model 9 1 on the habitat suitability index (hsi) and total area of available habitat. they were created in response to the national environmental policy act (nepa) of 1969, which required that the environmental impacts on wildlife from any activity involving federal funding or a federal permit be described prior to implementation of the project. this act made it necessary for biologists to relate wildlife species to their habitat and to predict species response to habitat alterations (thomas 1982). between 1982 and 1989, the usfws sponsored the development of over 160 hsi models for mammal, bird, reptile, amphibian, fish, and invertebrate species, and communities. these models translated existing knowledge of a species’ habitat requirements into standard, quantitative measures of habitat quality on a scale from 0 (unsuitable) to 1 (optimal). they are used to compare the ability of two or more study areas to provide habitat for a given species or to document habitat change over time within an individual study area. habitat suitability index models also predict the consequences of proposed natural resource management on wildlife habitats and identify suitable areas for development so negative impacts on wildlife habitats can be minimized. moose hsi model ii two moose hsi models (model i and model ii) were created by allen et al. (1987) as part of the usfws series and have served as the standard for more recent models (allen et al. 1991, courtois 1993, palidwor et al. 1995, hepinstall et al. 1996, rempel et al. 1997, romito et al. 1998, k. morris, maine department of inland fish and wildlife 1999, unpublished data). model i and ii were created for the evaluation of moose habitat in the lake superior region and were a product of a modeling workshop in duluth, minnesota in 1987. model i was designed to evaluate the abundance and quality of growing season and dormant season food and cover in study areas that approximate the size of annual habitats required by moose (~600 ha). intensive browse data collection is required for model i. model ii was designed to rapidly evaluate and compare the ability of relatively large areas to provide annual habitat for moose using remotely sensed data. for this study, the usefulness of model ii for analyzing large tracts of habitat was tested by applying it to 2 of vermont’s 26 wmus, which is the geographic unit used by the vtdfw for moose management. model ii relates cover-type composition to moose habitat suitability and incorporates 4 cover-type variables that provide annual life requisites for moose. model variables are: percent area of regenerating forests < 20 years old, used as a source of annual browse (variable 1); non-forested wetlands, used as a source of summer aquatic vegetation (variable 2); spruce/fir forests > 20 years old, used as a source of summer and winter cover (suitable stands need > 50% spruce/fir canopy) (variable 3); and upland deciduous or mixed forests > 20 years old, used for both annual browse and cover (suitable stands must have > 25% canopy cover of trees, of which < 50% of the canopy must be spruce/fir) (variable 4) (table 1). model variables were based on research conducted by peek et al. (1976) that described optimal moose habitat for northeast minnesota. twenty years was used as the cutoff age for regenerating forests because older trees are assumed to have little value as moose browse. model ii assumes that ideal year-round moose habitat requires the presence of all 4 habitat components and that model variables are weighted equally. however, if any of the 4 habitat components is missing from the evaluation area, other than wetlands, suitability will equal zero regardmoose habitat suitability index model — koitzsch alces vol. 38, 2002 9 2 less of the percent area of the other cover types. it also assumes a positive correlation between species abundance and habitat quality (allen et al. 1987). the degree of interspersion between food and cover is not addressed in model ii, however, in vermont where logging operations are relatively small and scattered, and non-forested wetland and mature spruce/fir forests are distributed throughout the state, it is assumed that interspersion is adequate. descriptions of preferred browse species and habitats of moose from studies in northern new hampshire (miller 1989, pruss and pekins 1992) and northern maine (cioffi 1981, monthey 1984, crossley 1985, leptich and gilbert 1989, thompson et al. 1995) are similar to those described for the lake superior region (allen et al. 1987) for which model ii was designed. climate, which dictates annual habitat preference, is also similar between these two regions. a classification system developed by wladimir köppen shows that climate in the great lakes region and new england is similar based on vegetation types and annual monthly means of temperature and precipitation (eichenlaub 1979). the u.s. department of commerce (1968) shows similarities between the two regions in annual minimum, maximum, and average daily temperatures, annual snowfall, mean annual numbers of days the minimum temperature was below 0oc, mean date for the last 0oc temperature in spring, and mean date for the first 0oc temperature in fall. because moose habitat composition and climate in vermont are similar to that of both northern maine and northern new hampshire, i consider it appropriate to assess moose habitat in vermont using model ii. specific objectives of this project were to: (1) generate gis coverages converting vegetation data into 4 cover types upon which model ii is based; (2) apply model ii to 2 wmus in vermont to predict habitat suitability; (3) predict within-wmu habitat suitability variation; (4) demonstrate the table 1. description of model variables and the life requisites they provide, variable suitability indices (si), and percent area of variables for optimum habitat suitability (allen et al. 1987, peek et al. 1976). variable # variable description life requisites provided variable % area of (%area) by variable suitability variables for indices (si) optimum habitat suitability 1 regenerating forest forage si 1 40 – 50 < 20 years old 2 non-forested aquatic forage, escape si 2 5 – 10 wetlands from insects and predation, thermoregulation 3 spruce / fir forest winter and summer cover si 3 5 – 15 > 20 years old 4 upland deciduous / forage and cover si 4 35 – 55 mixed forest > 20 years old alces vol. 38, 2002 koitzsch — moose habitat suitability index model 9 3 – 1,000 m in elevation except to the south where the nulhegan basin lies. the basin is drained by the nulhegan river, which flows eastward into the connecticut river. the state’s most extensive bogs and softwood swamps are located in the basin, which averages 350 450 m in elevation. this wmu is characterized by a mosaic of young, intermediate, and mature stands of trees due to its logging history and diverse geography. lowland areas are dominated by balsam fir, red spruce (picea rubens), black spruce (p. mariana), poplar (populus spp.), alder (alnus spp.), and paper birch (betula papyrifera). intermediate elevations contain primarily northern hardwood beech (fagus grandifolia) / birch (betula spp.) / maple (acer spp.) forest. sites above 800 m are predominately in red spruce and balsam fir stands. deciduous, mixed, and coniferous forests cover approximately 54%, 25%, and 15% of the area, respectively (d. williams, spatial analysis laboratory, university of vermont, unpublished data). this wmu has the greatest density of moose and the largest annual moose harvest of all wmus in the state. present estimate of moose density is 0.4 moose/km2 (c. alexander, vtdfw, personal communication). the second study area encompasses a 729 km2 portion of wmu i, and lies mostly within addison county. this portion will be referred to as wmu i for the study. wildlife management unit i is bordered by route 17 to the north, route 100 to the east, route 73 to the south, and route 116 to the west. it contains much of the northern half of the green mountain national forest and straddles the 1,000 – 1,200 m green mountain spine. to the east, the green mountains drop steeply into the mad and white river valleys, and to the west the mountains taper gradually into the champlain valley. wildlife management unit i is dominated by mature northern hardwood forests at mid change in hsi of wmu i after a 1996 ice storm destroyed over half of the forest canopy at upper elevations; (5) predict change in hsi of wmu e1 after ownership passed from a commercial wood products company to a federal entity; and (6) support model validation for vermont by correlating hsi values to population data from moose harvests. study area wildlife management unit e1 and a comparable-sized portion of wmu i were chosen as study areas (fig. 1) because they both are very important moose habitats in vermont, however, they vary greatly in vegetation composition, physiographic nature, and density of moose they support. wildlife management unit e1 (680 km2) is located within essex county in the northeast corner of vermont. e1 is bordered by canada to the north, the connecticut river to the east, route 105 to the south and route 114 to the west. it is roughly circular in shape and surrounded by mountains 600 fig. 1. location of study areas wmu e1 and wmu i in vermont. moose habitat suitability index model — koitzsch alces vol. 38, 2002 9 4 elevations, paper birch at high elevations, and pockets of spruce/fir on the highest peaks. unit i has fewer non-forested wetlands than e1. deciduous, mixed, and coniferous forests cover approximately 66%, 18%, and 10% of the area, respectively (d. williams, spatial analysis laboratory, university of vermont, unpublished data). wildlife management unit i is one of 3 additional units opened to hunting in 1999 and has an estimated moose density of 0.1 moose/km2 (c. alexander, vtdfw, personal communication). although wmu i is more densely populated by people than wmu e1, both areas represent 2 of the largest undeveloped tracts of land in the state. methods moose hsi model ii and hsi calculation model ii: hsi = (si 1 x si 2 x si 3 x si 4 )1/4 in the model, hsi is the habitat suitability index for the study area or evaluation unit and si 1 , si 2 , si 3 , and si 4 are suitability index (si) values for each of the 4 model variables. the hsi is the geometric mean of the 4 si values. percent areas of the 4 model variables are taken from variable cover type maps and plotted on suitability index graphs (figs. 2 and 3) to determine si values. suitability index graphs were created for the model following a description of optimal habitat from peek et al. (1976) (allen et al. 1987). percent area of variables falling within optimal ranges will produce a si = 1.0 and percent areas less than or greater than optimal ranges will produce a si < 1.0 (figs. 2 and 3) (allen et al. 1987). data for each variable were compiled into an arcview 3.1based gis (esri, redlands, california, usa). data were added to coverages by tablet digitizing with a calcomp drawing board ii and wintab digitizer software. variable 1 — percent area of regenerating forest. the area of regenerating forest within each wmu was determined from numerous data sources. the vermont forest resource advisory council (frac) quantified “heavy cuts” throughout vermont between 1977-1996 and reported and mapped their findings (vdfpr 1996). these “heavy-cut” maps were used as base maps for variable 1. “heavy cuts” were those visually determined from aerial flights or remotely sensed data to have been harvested below "c line". the c line represents the minimum amount of acceptable growing stock that makes a timber stand worth managing as defined by the fig. 2. suitability index curves showing relationship between percent area of regenerating forest and non-forested wetland variables and suitability index. optimal coverage of regenerating forest and non-forested wetland is 4050% and 5-10%, respectively (allen et al. 1987). alces vol. 38, 2002 koitzsch — moose habitat suitability index model 9 5 u.s. department of agriculture silvicultural stocking guides (long 1997). stand stocking level is a function of basal area per acre (ft2) and the number of trees per acre. in new england, stands harvested below the c line can be expected to have large quantities of early successional regeneration from species such as aspen (populus spp.), birch (betula spp.), cherry (prunus spp.), and maple (acer spp.). heavy cuts were easily discernible from maps because of their regular shape and obvious contrast from adjacent non-cut areas. ancillary data used to complete the variable 1 map through 1999 f o r w m u e 1 i n c l u d e d 1 9 9 9 d i g i t a l orthophotography quadrangles (doq) and 1997-1999 act 15 heavy-cut permits. forest service stand inventory data, which included all shelterwood, seed-tree, and clearcuts from 1977-march 1998, and 19971999 act 15 heavy-cut permits were used for wmu i. color-infrared aerial photographs, orthophotography, 1995 doqs (wmu i), and ground-truthing were used to verify data. the final map displayed areas of regenerating forest as closed polygons. from these maps (fig. 4), percent area of regenerating forests within the study areas was digitally queried. percent area in “heavy cuts” < 23 years old (1977–1999) was used as an estimator of regenerating forests < 20 years old because of data structure. i do not believe the addition of 2 years of data to variable 1 resulted in an overestimation of this variable, but rather made up for the small acreages of regenerating forests that were inadvertently missed while analyzing data. variable 2 — percent area of nonforested wetlands. non-forested wetlands included in this study followed suggested modifications to model ii by adair et al. (1991) and wetland classifications from cowardin et al. (1979). adair et al. (1991) fig. 3. suitability index curves showing relationship between percent area of spruce/fir and deciduous/mixed forest variables and suitability index. optimal coverage of spruce/fir forest and deciduous/mixed forest is 5-15% and 35-55%, respectively (allen et al. 1987). fig. 4. variable 1 regenerating forest. moose habitat suitability index model — koitzsch alces vol. 38, 2002 9 6 recommended that only wetlands with limnological conditions favoring macrophyte production should be included in the model. w e t l a n d s i n c l u d e d w e r e e m e r g e n t , unconsolidated bottom, rock bottom, and aquatic bed palustrine, scrub/shrub, dead forested, lower perennial riverine, littoral lacustrine, and beaver ponds. mylar and digital national wetland inventory (nwi) maps were used to determine total area of these non-forested wetland types in each wmu. national wetlands inventory data were used because they are readily available to the public, cover the entire united states, and are consistent with their wetland classifications. suitable wetlands were first identified on nwi mylar maps and then labeled as such on digital maps. where data were incomplete on digital maps, or had not yet been digitized, data were manually digitized from mylar maps. a coverage of suitable wetland polygons was then generated and percent area digitally queried (fig. 5). variable 3 — percent area of spruce/ fir forests. variable area was taken from a vegetation grid map of new hampshire and vermont created for the usfws gap analysis project (d. williams, spatial analysis laboratory, university of vermont, unpublished data). data used to create the map included 4 landsat thematic mapper (tm) satellite scenes acquired for spring, summer, and fall of 1992 / 1993, a small portion of the may 1995 scene, and ancillary data. resolution of the tm data was 30 m. the map was validated through interpretation of aerial videography linked to a global positioning system (gps). the map accurately classified 85% of 907 gps points examined (d. williams, spatial analysis laboratory, university of vermont, unpublished data). much of the error was associated with classification of mixed forest as either deciduous or coniferous forest. of the 7 land-cover types classified in this mapping project, only deciduous, coniferous, and mixed-forest classifications were pertinent to this study. the others were omitted from analysis. forests were classified as coniferous or deciduous if either contributed > 65% of stand species, or mixed if neither contributed >65%. the “coniferous forest” classification from this map represented spruce/fir forest for my analysis. since model ii requires that coniferous or deciduous forests contribute > 50% of stand species (allen et al. 1987), areas derived from this vegetation map may underestimate percent area of coniferous and hardwood forests and overestimate percent area of mixed forests. using arcview and arc/info spatial analysis gis software, the vegetation grid was clipped to the extent of the 2 study areas. polygon coverages of regenerating forest (variable 1) and non-forested wetland (variable 2) were then erased from the vegetation coverage leaving coniferous, mixed, and hardwood forest cover-types (fig. 6). percent area in spruce/fir forest fig. 5. variable 2 non-forested wetland. alces vol. 38, 2002 koitzsch — moose habitat suitability index model 9 7 was queried from the resulting coverage. variable 4 — percent area of deciduous/mixed forests. percent area in upland deciduous/mixed forest was derived in the same manner as variable 3 (fig. 7). within-wmu hsi variation to determine within-wmu hsi variation, each wmu was divided into 25-km2 hexagonal evaluation units, and a hsi was determined for each unit as previously described. twenty-five square kilometers was chosen as the evaluation unit size because it approximates the annual home range of moose in new england (crossley 1985, miller 1989, thompson et al. 1995, k. morris, maine department of inland fish and wildlife 1999, unpublished data). allen et al. (1987) and schultz and joyce (1992) recommended that size of the evaluation unit should approximate that of the animals’ home range for hsi analysis. a regular hexagonal pattern was chosen for evaluation unit shape because it is the best discontinuous sampling pattern for a spatial function (olea 1984). evaluation units were assigned to 3 habitat categories based on hsi values: least suitable habitat (hsi = 0.0–0.31), suitable habitat (hsi = 0.32– 0.66), and most suitable habitat (hsi = 0.67–1.0) (fig. 8). effects of habitat alteration on hsi of study area the effect of a rapid increase in regenerating forests on hsi was demonstrated by comparing hsi of 81 km2 of wmu i before and after heavy ice damaged much of the deciduous forest canopy above 1,000m in 1998 (fig. 9). the vermont department of forest parks and recreation (vdfpr) mapped statewide ice damage into 2 categories labeled “heavy” and “moderate”. forests were considered “heavily” damaged if > 50% of the forest canopy was damaged. this heavy damage to the forest canopy simulated the effect of harvest practices, which stimulate regeneration in the fig. 6. variable 3 – spruce/fir forest. fig. 7. variable 4 – deciduous/mixed forest. alces vol. 38, 2002 koitzsch — moose habitat suitability index model 9 9 type classification from the vegetation grid map (d. williams, spatial analysis laboratory, university of vermont, unpublished data). area of non-forested wetlands remained constant. model validation hsi model validation is a process that determines whether a model accurately predicts habitat quality from the animal’s perspective. initial validation of model ii was accomplished by correlating october moose harvest density (moose/km2) to hsi for 25km2 evaluation units within wmu e1. wildlife management unit i was omitted from this analysis because it had been hunted for just 1 year and only 6 moose were harvested. moose harvest locations recorded for wmu e1 from 1993-1999 were used for analysis. twenty-five kilometer square evaluation units with > 95% of their areas within wmu e1 were used in the analysis (n = 21). one hundred fifty-seven moose harvested were within the 21 evaluation units. the non-parametric spearman rank correlation coefficient was used for analysis because distribution of evaluation unit hsi values was not normal. for this analysis, it was assumed that moose, during the fall hunting season, are occupying habitats with the greatest hsi values, and therefore moose harvest should be highest in units with the greatest hsi. a positive correlation between hsi and moose harvest would support validation of the model. results moose hsi model ii revealed large differences in habitat suitability between wmu e1 (his = 0.64) and wmu i (his = 0.34) based on differences in variable composition (table 2). wildlife management unit e1 contained 17% regenerating forest (si = 0.42), 2% non-forested wetland (si = 0.48), 14% spruce/fir forest (si = 1.0), and 63% deciduous/mixed forest (s = 0.82). wildlife management unit i contained 4% regenerating forest (si = 0.11), 1% nonforested wetland (si = 0.32), 10% spruce/ fir forest (si = 1.0), and 84% deciduous/ mixed forest (si = 0.37). wildlife management unit e1 contained greater amounts of regenerating forest and non-forested wetland, and lesser amounts of deciduous/ mixed forest than wmu i. each wmu contained optimal amounts of spruce/fir forest, less than optimal amounts of regenerating forest and non-forested wetland, and more than optimal amounts of deciduous forests. within-wmu hsi variation was found in both study areas between 25-km2 evaluation units (fig. 8). evaluation units classified as “most suitable” (his = 0.67-1.0) had the greatest area of regenerating forest, non-forested wetland, and optimal area in spruce/fir forest, while the “least suitable” (his = 0.0-0.31) units had a lesser abundance of regenerating forest and nonforested wetland, and an overabundance of deciduous/mixed forest. wildlife management unit e1 had approximately 20% of its table 2. habitat suitability index values for wmu e1 and i (variable 1 = regenerating forest, variable 2 = non-forested wetland, variable 3 = spruce/fir forest, variable 4 = deciduous/mixed forest). percent area of model variables / suitability index (si) wmu area (km2) variable 1 variable 2 variable 3 variable 4 wmu hsi e1 680 16.90 / 0.42 1.72 / 0.48 14.30 / 1.00 63.40 / 0.82 0.64 i 729 4.46 / 0.11 0.73 / 0.32 9.76 / 1.00 83.76 / 0.37 0.34 moose habitat suitability index model — koitzsch alces vol. 38, 2002 100 within the 108 km2 of the silvio o. conte national fish and wildlife refuge parcel, habitat suitability was shown to decrease from 0.81 to 0.35 after 20 years of a simulated no-cut policy (table 4). over the 20-year simulation, a reduction of percent area of regenerating forest from 21% to 1% was offset by a corresponding 10% increase in both spruce/fir and deciduous/ mixed forests. within the parcel, si of regenerating forest decreased from 0.53 to 0.03 and si of spruce/fir and deciduous/ mixed forest also declined. within the entire wmu e1, hsi was predicted to decrease from 0.64 to 0.60. correlation of hsi to moose harvest density for 25-km2 evaluation units within wmu e1 revealed a spearman coefficient of r = 0.53 (p = 0.013) and an increasing trend in moose harvests with an increase in hsi of the evaluation unit (fig. 10). a hsi of 0.64 for wmu e1 and 0.32 for wmu i compared proportionately to estimated moose density of 0.40 moose/km2 and 0.10 moose/km2, respectively (c. alexander, vtdfw, personal communication). discussion moose hsi model ii predicted differences in habitat suitability between wmu table 3. change in hsi of 81 km2 of heavily damaged forest and wmu i following an ice storm in january 1998 (variable 1 = regenerating forest, variable 2 = non-forested wetland, variable 3 = spruce/ fir forest, variable 4 = deciduous/mixed forest). percent area of model variables / suitability index (si) study area variable 1 variable 2 variable 3 variable 4 hsi hid1 (b)2 81 2.95 / 0.07 0.00 / 0.20 26.60 / 0.86 70.10 / 0.67 0.30 hid1 (a)2 81 55.54 / 0.89 0.00 / 0.20 26.60 / 0.86 17.51 / 0.51 0.53 wmu i (b)2 729 4.46 / 0.11 0.73 / 0.32 9.48 / 1.00 78.92 / 0.47 0.36 wmu i (a)2 729 10.30 / 0.26 0.73 / 0.32 9.48 / 1.00 73.06 / 0.60 0.47 1hid = heavy ice damage. 2b = before ice damage, a = after ice damage. area in “most suitable” habitat and 80% in “suitable” (his = 0.32-0.67) habitat. “most suitable” habitats were located in the southcentral portion of wmu e1 and approximated the boundary of the nulhegan basin. one evaluation unit in the northeast corner also contained “most suitable” habitat. wildlife management unit i lacked any units in the “most suitable” category but had approximately 50% of its area in “suitable” habitat. “suitable” habitats were located in the western half of the study area and in the east-central portion. the model predicted that hsi increased from 0.30 to 0.53 in 81 km2 of wmu i, which was heavily damaged by an ice storm in 1998 (fig. 9) (table 3). an increase in percent area of regenerating forest from 3% to 56% caused an increase in si from 0.07 to 0.9. a corresponding decrease in deciduous/mixed forest from 70% to 18% caused a decrease in si from 0.67 to 0.51, but caused an increase in the deciduous/ mixed forest si for the entire wmu. due to ice storm damage, area in regenerating forest doubled in the entire wmu and caused an increase in si from 0.11 to 0.26. habitat suitability index of the entire wmu increased from 0.36 to 0.47 as a result of the storm. alces vol. 38, 2002 koitzsch — moose habitat suitability index model 101 e1 and wmu i that reflect differences in vtdfw moose densities of 0.4 moose/km2 and 0.1 moose/km2, respectively. calculated hsi values were 0.64 for wmu e1 and 0.34 for wmu i. this difference in hsi value was due to the greater percentage of regenerating forest and non-forested wetland, and the lesser percentage of mature deciduous/mixed forest in wmu e1 compared to wmu i. on a percentage basis, wmu e1 had 278% more regenerating forest, 136% more non-forested wetland, and 24% less deciduous/mixed forest than wmu i. these data support the theory of telfer (1978) and collins and helm (1997) who have indicated that the abundance of regenerating forest is often the most limiting factor to moose density. these data, which describe relative moose habitat suitability per wmu, will be useful to wildlife agencies for meeting their moose management objectives and for assisting members of the public in choosing their desired regional moose population levels in states where public opinion is considered for moose management. for instance, within vermont’s “moose investigation project statement”, the vtdfw identifies the need to determine relative habitat suitability per wmu using remotely sensed land-use databases, gis, and hsi models. the vtdfw also solicits public opinion when making moose management decisions. model ii also identified variation in habitat suitability within wmus. the abundance of “most suitable” habitat in wmu e1 was due to a concentration of regenerating forest and non-forested wetland habitats in the nulhegan basin. “suitable” habitat was found throughout the rest of the wmu where lesser amounts of these components exist, and “least suitable” habitat was found on the perimeter of the wmu associated with high concentrations of development and agriculture. wmu i contained no evaluation units with “most suitable” habitat because the wmu as a whole was deficient in regenerating forest and non-forested wetlands. in wmu i “suitable” evaluation units were found to the table 4. change in hsi of silvio o. conte national fish and wildlife refuge lands and entire wmu e1 following a simulated 20 year no-cut policy (variable 1 = regenerating forest, variable 2 = nonforested wetland, variable 3 = spruce/fir forest, variable 4 = deciduous/mixed forest). percent area of model variables / suitability index (si) study area area (km2) variable 1 variable 2 variable 3 variable 4 hsi silvio conte 108 21.20 / 0.53 4.09 / 0.85 19.00 / 0.95 51.63 / 1.00 0.81 silvio conte +20 108 1.00 / 0.03 4.09 / 0.85 29.60 / 0.82 62.24 / 0.84 0.35 wmu e1 680 16.90 / 0.42 1.72 / 0.48 11.32 / 1.00 63.40 / 0.82 0.64 wmu e1 +20 680 13.73 / 0.34 1.72 / 0.48 13.03 / 1.00 65.15 / 0.78 0.60 fig. 10. moose harvest density versus evaluation unit hsi for wmu e1. spearman correlation: r = 0.53, p = 0.013, n = 21. moose habitat suitability index model — koitzsch alces vol. 38, 2002 102 west of the green mountain spine in association with the highest concentrations of non-forested wetlands and regenerating forest (figs. 4 and 5). also, the east central portion of wmu i was heavily cut by the forest service in the 1980s and retains a “suitable” classification. because of its steep mountainous terrain, the remainder of wmu i contains less standing water, is less accessible to logging, and has a “least suitable” classification. differences in habitat suitability of adjacent evaluation units in both wmus can be attributed to differences in topography, which determines the ability of the land to develop wetlands and dictates accessibility for logging. in wmu i, suitable habitats are concentrated in flatter areas to the west of the green mountain ridge, and in wmu e1 habitat suitability of areas surrounding the nulhegan basin decreases as the basin rises up to the surrounding mountains. the potential effects of ice damage on habitat suitability were illustrated through hsi analysis before and after an ice storm in january 1998 damaged the deciduous canopy of 81 km2 of wmu i (fig. 9). increased hsi from 0.30 to 0.53 (an increase of 77%) was due to a significant increase of regenerating forest and a corresponding decrease in deciduous/mixed forest. this same increase in regenerating forest contributed to an increase in hsi of the entire wmu i from 0.36 to 0.47 (an increase of 31%). this effect of ice damage on hsi illustrates how natural events can rapidly affect habitat quality for moose. however, ice storm damage may not cause permanent canopy opening in the affected areas and the resulting increase in production of ground level browse may be shortlived. studies are presently under way to determine the long-term effects of the 1998 storm on the forest canopy, and these should reveal how long damaged areas will continue to provide regenerating browse. other environmental factors such as heavy defoliation by forest insects such as spruce budworm (choristoneura fumiferana), which defoliated 56% of all spruce and balsam fir in vermont in 1983 (r. kelley, vermont department of forests, parks and recreation, personal communication), tree disease, and wind-throw can have similar effects on moose habitat. the 108-km2 parcel purchased by the usfws in 1999 from champion international corporation contains much of the nulhegan basin. the nulhegan basin is arguably the most productive moose habitat in the state based on number of moose harvested, automobile-moose collisions, and sightings. of 187 moose harvests located in wmu e1 from 1993-1997, 30% were within the approximate boundary of the nulhegan basin. model ii predicted that after 20 years of a no-cutting policy, hsi of the parcel would decrease from 0.81 to 0.35 (a reduction of 57%), due to a significant reduction in regenerating forest and a corresponding increase in spruce/fir and deciduous/mixed forests. following 20 years of maturation, this forest will still provide valuable winter cover and non-forested wetland habitats, but would supply limited understory browse. if the suitability of the habitat is reduced by this amount, i project significant declines in the moose population and moose harvest in this area. habitat suitability index output also was used to support validation of model ii. the ideal method to validate hsi models is to compare model output to known population numbers of target species within the study area. since these data are usually unattainable, indicators of species abundance are utilized (clark and lewis 1983, laymon and barrett 1986, thomasma et al. 1991, robel et al. 1993). another common method used to validate models is to compare hsi output of known-use sites to random sites in order to test that the model can differentiate bealces vol. 38, 2002 koitzsch — moose habitat suitability index model 103 tween the two (allen et al. 1991, brennan 1991, apps and kinley 1998). for this study, moose harvest density was used as an indicator of species abundance for initial validation. a spearman correlation of hsi output to moose harvest density for 21wmu e1 evaluation units (r = 0.53, p = 0.013) indicated the tendency for moose harvest density to increase with an increase in hsi (fig. 10). a positive relationship between hsi and vtdfw estimated moose densities, for both wmus, also supports validation of model ii. additional research to further validate the use of model ii for vermont includes locating heavy use-sites using gps collars on moose and correlating these to hsi, and comparing hsi of heavy use-sites to that of random sites. a study to analyze the effects of road density on moose harvest, since hunter access to moose is most likely correlated to road access to hunting areas, also could be conducted. results of hsi evaluation only should be used to predict the potential of habitat to support moose and not as a predictor of population density. too many other factors exist, which can reduce moose abundance even when habitat is favorable, that are not addressed in the model. in vermont, these include traffic and road density that influence the number of moose/car collisions, deer density and infection rate of brain worm (parelaphostrongylus tenuis), winter severity, illegal harvest, black bear predation on moose calves, the number of hunting permits issued, and inter-specific competition for food. this study demonstrated that model ii is useful for analyzing large tracts of moose habitat. in view of the rate and scale at which humans alter the environment, it is important to look at habitat from a landscape perspective. too often we concern ourselves with ecological processes on a small scale, unaware of large changes occurring around us. with advances in higher resolution satellite imagery and the availability of a greater selection of satellite data scenes, the analysis of large habitats will become simpler. landsat 7 satellite data are currently available from the united states geological survey (usgs) at 30m resolution. landsat 7 records enhanced thematic mapper plus (etm+) data in 7 spectral bands plus an eighth panchromatic band, combines synoptic coverage, high spatial resolution (15 m from the panchromatic band), a high spectral band range (450-2350 nm), increased spatial resolution of the thermal ir band (band 6), and 5% radiometric calibration (usgs 2000, unpublished data). minimally processed data, known as level 0-r data, are available at 475 u.s. dollars/ scene and levels 1-r (radiometrically corrected) and 1-g (radiometrically and geometrically corrected) data are available at 600 u.s. dollars/scene. these costs are substantially lower than prices for current landsat 5 data (usgs, http:edc.usgs.gov/ buspartners/satellite satellite-program.html). because landsat 7 records image data of the entire world every 16 days, users can choose data from up to 22 different dates of the year for a particular study area. higher resolution imagery will enhance the ability to differentiate between regenerating and mature forest for hsi model applications for moose. suggested modifications to model ii renecker and hudson (1986, 1990) observed heat stress in moose, characterized by an increase in metabolism and respiration rate, when temperatures exceeded 14oc in summer and –5oc in winter. such stress can result in depressed foraging activity and weight loss. telfer (1984) observed that the southern limit of holarctic moose distribution corresponded closely to the 20oc july isotherm and that high temperatures that reduce reproductive performance might restrict the southern expansion moose habitat suitability index model — koitzsch alces vol. 38, 2002 104 of moose range into areas with adequate habitat otherwise. to reduce effects of heat stress, moose seek shade in dense cover and wet areas to bed, thereby reducing energy expenditure, respiration, and metabolism. from national oceanic and atmospheric administration (noaa) data (november 1998 – october 1999), i calculated the number of days that temperatures exceeded 20oc between may and september when moose are in summer pelage, and –5oc between october and april when moose are in winter pelage for both wmus. approximately 310 days in both island pond (wmu e1) and south lincoln (wmu i) exceeded these limits. temperatures at higher elevation beneath forest canopy where moose can escape heat would have been slightly lower, but these numbers still establish that moose are subjected to many days of heat stress. i believe there exists a threshold number of days above heat stress thresholds that moose simply cannot tolerate, and that habitat selection in vermont, and especially in southern new england, is temperature-dependent. the addition of a variable that quantifies the number of days, and the number of hours per day, temperatures exceed heat stress threshold likely would enhance accuracy of this model in new england and help predict the southern limit to moose range. also, if trends in global warming continue, the management of heat sensitive species such as moose and caribou (rangifer tarandus) will depend on determining the potential of traditional habitats to continue to provide for these animals. to validate this model with the additional variable for heat stress, moose habitat selection during times of heat stress should be monitored and correlated to hsi. the use of gps collars would be essential in acquiring these data. a study of moose activity and metabolism during these times also would provide data on behavioral and physiological changes associated with heat stress. collars fitted with an activity counter and temperature sensor could gather these data. management implications the results of this study indicate that in both wmus, percent area of deciduous/ mixed forest exceeds the amount for optimum habitat suitability, percent area of spruce/fir forest exists at optimal amounts, and the area of regenerating forests and non-forested wetlands exist in quantities well below that needed for optimum habitat suitability. non-forested wetland and regenerating forest are therefore most limiting to moose habitat suitability. with the number of beavers in the state increasing due to a decline in trapping, and legislation protecting wetland habitats, it appears that the present quantity and quality of nonforested wetland habitats will improve. however, a decreasing trend in heavy cutting throughout much of the state since the mid-1980s could reduce the occurrence of regenerating forests. to achieve the goals of moose management in vermont of maintaining moose populations at or above current densities, and to increase benefits associated with moose such as viewing and hunting, the continued use of forestry practices that create regenerating forests, and the continued protection of non-forested wetland habitat are desirable. resource managers also should strive to maintain habitat quality within the “suitable” and “most suitable” habitats as identified by model ii. since timber management has a great impact on moose habitat quality, (courtois 1993, palidwor et al. 1995, rempel et al. 1997, romito et al. 1998), forest and wildlife managers should strive to integrate the use of moose hsi models at the landscape scale into timber management practices. alces vol. 38, 2002 koitzsch — moose habitat suitability index model 105 acknowledgements this study was funded by remo pizzagalli and the pope & young conservation fund. thanks to cedric alexander for continued interest and guidance, and to the vermont department of fish and wildlife, vermont department of forest, parks & recreation, national wetland inventory, and vermont mapping program for providing data. thank you david hirth, jeffrey hughes, and ruth mickey for guidance as committee members and to the entire university of vermont spatial analysis lab staff for technical support. special thanks to lisa osborn who offered her endless support throughout the project. references adair, w., p. a. jordan, and j. tillma. 1991. aquatic forage ratings according to wetland type: modifications for the lake superior moose hsi. alces 27:140149. alexander, c. e. 1993. the status and management of moose in vermont. alces 29:187-195. , p. fink, l. e. garland, and f. hammond. 1998. moose management plan for the state of vermont 19982007. vermont department of fish and wildlife, waterbury, vermont, usa. allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. biological report 82(10.155), u.s. fish and wildlife service, fort collins, colorado, usa. , j. w. terrell, w. l. mangus, and e. l. lindquist. 1991. application and partial validation of a habitat model for moose in the lake superior region. alces 27:50-64. apps, c. d., and t. a. kinley. 1998. development of a preliminary habitat assessment and planning tool for mountain caribou in southeast british columbia. rangifer 10:61-72. belovsky, g.e., and p.a. jordan. 1981. sodium dynamics and adaptations of a m o o s e p o p u l a t i o n . j o u r n a l o f mammalogy 62:613-621. brennan, l. a. 1991. regional tests of a mountain quail habitat model. northwestern naturalist 72:100-108. cioffi, p. j. 1981. winter cover and browse selection by moose in maine. m.sc. thesis, university of maine, orono, usa. clark, j. d., and j. c. lewis. 1983. a validity test of a habitat suitability index model for clapper rail. proceedings of the annual conference, southeastern association of fish and wildlife agencies 37:95-102. collins, w. b., and d. j. helm. 1997. moose, alces alces, habitat relative to riparian succession in the boreal forest, susitna river, alaska. canadian fieldnaturalist 111:567-574. , and c. c. schwartz. 1998. logging in alaska’s boreal forest: creation of grasslands or enhancement of moose habitat. alces 34:355-374. courtois, r. 1993. description d’un indice de qualité d’habitat pour l’original (alces alces) au quebéc. gouvernment du quebéc, ministére du loisir, de la chasse et de la pêche, direction générale de la resource faunique, gestion intégrée des ressources, document technique 93/1. cowardin, l. m., v. carter, f. c. golet, and e. t. laroe. 1979. classification of wetlands and deep water habitats of the united states. u.s. fish and wildlife service fws/obs-79/31. crête, m. 1988. forestry practices in quebec and ontario in relation to moose population dynamics. forestry chronicle 64:246-250. crossley, a. 1985. summer pond use by moose in northern maine. m.sc. themoose habitat suitability index model — koitzsch alces vol. 38, 2002 106 sis, university of maine, orono, usa. eichenlaub, v. l. 1979. weather and climate of the great lakes region. university of notre dame press, notre dame, indiana, usa. hepinstall, j. a., l. p. queen, and p. a. jordan. 1996. application of a modified habitat suitability index model for moose. photogrammetric engineering and remote sensing 62:1281-1286. hurley, j. f. 1986. summary: development, testing and application of wildlifehabitat models – the managers viewpoint. pages 151-153 in j. verner, m. morrison, and c. j. ralph, editors. wildlife 2000: modeling habitat relationships of terrestrial vertebrates. university of wisconsin press, madison, wisconsin, usa. jordan, p.a., d. b. botkin, a. s. dominski, h. f. lowendorf, and g. e. belovsky. 1973. sodium as a critical nutrient for the moose of isle royale. proceedings of the north american moose conference and workshop 9:13-42. laymon, a. l., and r. h. barrett. 1986. developing and testing habitat-capability models: pitfalls and recommendations. pages 87-91 in j. verner, m. morrison, and c. j. ralph, editors. wildlife 2000: modeling habitat relationships of terrestrial vertebrates. university of wisconsin press, madison, wisconsin, usa. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53:880-885. long, s. 1997. drawing the line in the stand-a guide to understanding the c line. vermont woodlands, autumn issue:23-26. miller, b. k. 1989. seasonal movement patterns and habitat use of moose in northern new hampshire. m.sc. thesis, university of new hampshire, durham, new hampshire, usa. monthey, r. w. 1984. effects of timber harvesting on ungulates in northern maine. journal of wildlife management 48:279-285. olea, r. a. 1984. sample design optimization for spatial functions. mathematical geology 16:369-391. palidwor, k. l., d. w. schindler, and b. r. hagglund. 1995. habitat suitability index models within the manitoba model forest region: moose (alces alces). version 2. manitoba model forest incorporated, pine falls, manitoba, canada. peek, j. m., d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. pletscher, d. h. 1987. nutrient budgets for white-tailed deer in new england with special reference to sodium. journal of mammalogy 68:330-336. pruss, m. t., and p. j. pekins. 1992. effects of moose foraging on browse availability in new hampshire deer yards. alces 28:123-136. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517524. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditure and thermoregulatory response of moose. canadian journal of zoology 64:322-327. , and . 1990. behavioral and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspen-dominated forests. alces 26:66-72. robel, r .j., l. b. fox, and k. e. kemp. 1993. relationship between habitat suitability index values and ground counts alces vol. 38, 2002 koitzsch — moose habitat suitability index model 107 of beaver colonies in kansas. wildlife society bulletin 21:415-421. romito, t., k. smith, b. beck, m. todd, r. bonar, and r. quinlan. 1998. moose winter habitat: habitat suitability index model. version 5. weldwood of canada, hinton, alberta, canada. schultz, t. j., and l. a. joyce. 1992. a spatial application of a marten habitat model. wildlife society bulletin 20:7483. telfer, e. s. 1978. cervid distribution, browse and snow cover in alberta. journal of wildlife management 42:352361. . 1984. circumpolar distribution and habitat requirements of moose. pages 145-182 in r. olson, r. hasting, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. thomas, j. w. 1982. needs for and approaches to wildlife habitat assessment. transactions of the north american wildlife and natural resources conference 47:35-46. thomasma, l. e., t. d. drummer, and r. o. peterson. 1991. testing the habitat suitability index model for fisher. wildlife society bulletin 19:291-297. thompson, m. e., j. r. gilbert, g. j. matula, and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31:233-245. u.s. department of commerce. 1968. climatic atlas of the united states. u.s. government printing office, washington, d.c., usa. (vdfpr) vermont department of forests, parks and recreation. 1996. a report made by the vermont department of forests, parks and recreation for the assessment working group of the forest resource advisory council. waterbury, vermont, usa. f:\alces\vol_38\pagemaker\3813. alces vol. 38, 2002 morris aquatic vegetation 213 impact of moose on aquatic vegetation in northern maine karen i. morris maine department of inland fisheries and wildlife, 650 state st., bangor, me 04401, usa abstract: many ponds in northern maine have a low abundance of aquatic vegetation. five exclosures were built in 2 ponds with high moose use but little vegetation. all exclosures sustained ice damage each winter. one was damaged beyond repair after 3 years, 3 were lost during the fifth winter, and 1 lasted for 6 years. the number of plants rooted along a 20 m transect were counted in mid-august in the first, second, fourth, and fifth years of the study. all vegetation rooted in 24 1 m 2 plots (3 inside and 3 outside of each of the remaining exclosures) was pulled, dried, and weighed after the third growing season. ten plots (5 inside and 5 outside) from the 1 remaining exclosure were clipped and weighed after 6 growing seasons. plant biomass was greater in 3 of 4 protected than in unprotected areas after 3 years (p < 0.05) and in the 1 remaining exclosure after 6 years (p < 0.05). biomass increased within the exclosures from the third to the sixth year (p < 0.05) but there was no change in the unprotected area. alces vol. 38: 213-218 (2002) key words: alces, aquatic vegetation, exclosures, maine, moose in the late-1970s, what seemed to be a low abundance of aquatic plants was observed in many of the shallow ponds in northern maine. sightings of moose and moose tracks in these ponds indicated that these ponds were heavily used by moose. feeding by moose was suspected to be one possible cause for the near absence of aquatic plants in some ponds. however, no information is available on the condition of these ponds when moose were less abundant. moose commonly use aquatic plants as a source of sodium when sodium levels are low on nearby terrestrial browse (jordan et al. 1973; fraser 1979; fraser et al. 1982, 1984). several studies have documented that moose can reduce the abundance of aquatic plants (murie 1934, jordan et al. 1973, aho and jordan 1979, fraser and hristienko 1983). the objective of this study was to document the impact of moose browsing on the abundance of aquatic plants in 2 ponds in maine, usa. study area two ponds in north central maine that were included in crossley’s (1985) study of pond use by moose were selected for study. sixteen species of aquatic plants in these ponds ranged from 0.08% to 1.28% na (dry matter) compared to 0.0047-0.0098% in 10 species of commonly browsed terrestrial plants from the surrounding area (crossley 1985). aquatic plants seem to be the only source of concentrated sodium in the area. none of the radio collared moose in this area visited a recognizable mineral lick and none made the 35 km journey to the nearest salted highway (crossley 1985, leptich 1986, thompson 1987, maine department of inland fisheries and wildlife, unpublished data). leonard pond (15 ha) and bartlett pond (31 ha) are shallow (60-100 cm) bog lakes surrounded by floating mats of sphagnum spp. and associated plants. plant life in the ponds is sparse. the bottoms of both ponds are loose muck, allowing moose to swim and become completely submerged in waaquatic vegetation morris alces vol. 38, 2002 214 ter less than 1 m in depth. the bottom is visible throughout both ponds, although the water in leonard pond is noticeably dark stained. methods five moose exclosures, 3 in bartlett pond and 2 in leonard pond, were built in may 1983. an exclosure was placed in the center of each pond with the remainder spaced equally between the center of the pond and the shore farthest from the outlet. average august water depths in the exclosures were 24, 30, and 50 cm for bartlett pond, and 60 and 75 cm for leonard pond. each 4.9 m square (24 m2) exclosure was made of 4 hogwire panels and wooden posts. cedar fence posts (2 m) were pushed into the bottom at each corner and the middle of each side. when the muck bottom was exceptionally deep the corners of the exclosures were reinforced with posts up to 3.5 m long. the panels were 1.2 m high and set so that the top was near the water surface in may. the loose muck bottom made it impossible for moose to reach over the panels at water level or exert enough pressure to damage the exclosures. the exclosures were checked each may, and those that had sustained ice damage were repaired, if damage had not disturbed the bottom of the pond. the exclosure in 30 cm of water in bartlett pond was removed in the third year because the bottom was disturbed. during the third winter, the exclosure in 75 cm in leonard pond sank to the point that moose might have been able to reach plants near the edge of the exclosure. a second group of hogwire panels was placed on top of the original exclosure 2 weeks after ice-out. the abundance of plants inside and outside the exclosures was evaluated in mid-august for 6 years. in the third and sixth years of the study, biomass was measured by clipping and weighing plants from sample plots inside and outside the exclosures. in the other 4 years, relative abundance of plants inside and outside the exclosure was monitored by a line intercept technique. observations were made from a canoe using a glass bottom bucket to improve visibility. the canoe was placed inside the exclosure for each sampling session. we counted the number of plants that touched a 0.82 cm diameter 4 m long rod placed horizontally along the bottom of the pond. twenty meters of transect inside and outside of each exclosure were searched in each of 4 years. during the second year, it became apparent that the density of vegetation in some of the protected areas made the line intercept sampling method impractical. the rod could not be put in place, nor the plants counted, without disturbing the vegetation and therefore affecting the measurement. plant contacts with the rod over 100 were probably inaccurate so they were recorded as a minimum number. three growing seasons after construction of exclosures, all plants rooted on 3 plots (1 m2) inside and 3 plots outside of each exclosure, (2 in leonard pond and 2 in bartlett pond) were pulled, rinsed, oven dried to constant moisture content, and weighed to the nearest 0.1 g. the centers of 3 circular plots were located along a line 1 m inside a randomly selected panel, leaving the rest of the exclosure undisturbed. another 3 plots were located along a line 1 m outside of the same panel. the edges of the plots were 44 cm from the fence. five 1 m2 plots within the previously unclipped section of the remaining exclosure (bartlett pond) and 5 plots adjacent to the exclosure were clipped and weighed after 6 years of protection. the centers of these plots were located on lines 1 and 2.5 m from the panel farthest from the plots that were measured 3 years earlier. a t-test was used to compare biomass inside and outside the aquatic vegetation morris alces vol. 38, 2002 216 tite, decline in growth, loss of weight, and decreased productivity. severe restriction of salt intake may prevent reproduction (maynard and loosli 1969). belovsky (1981) suggested that the number of moose might be dependent on the availability of sodium. jordan (1987) noted that clinical deficiencies beyond a behavioral drive or adrenocortical hypertrophy are unlikely to be displayed by a free ranging population. however, he anticipated that there would be lower reproductive success or a population below what the terrestrial browse could support. there is some indication that maine moose have a higher than usual craving for salt, based on the length of time they spent using ponds. crossley (1985) found that moose in northern maine, especially cows with calves, were using ponds later into the summer than reported in several other studies (dunn 1975, best et al. 1977, fraser 1979, fraser et al. 1982). jordan et al. (1973) found moose on isle royale used aquatic plants into september during years when aquatic plants had been noticeably reduced by grazing, but that aquatic feeding declined in late july or early august in years when aquatic plants remained abundant. however, female moose are reported to use mineral licks into the fall in new hampshire (miller and litvaitis 1992) and quebec (couturier and barrette 1988). measuring the number of moose seen per hour in 1982, crossley (1985) recorded moose using 3 ponds, including the 2 in this study, later in the year than an earlier study in maine (dunn 1975). fire tower observers statewide recorded the number of moose they saw entering ponds in 1956 and from 1962 to 1967, and they noted the number of moose entering the water dropped off sharply after july (dunn 1975). crossley (1985) observed no decline in lake use by moose from late july through september. the number of moose entering a pond per hour (dunn 1975), and the number of moose seen per hour (crossley 1985), may not be directly comparable, because the length of time moose spend in a pond varies with time of year. furthermore, due to differences in the density of moose, the numbers of moose observed may not be comparable between the two studies, even if data were recorded in the same way. to make these data sets more comparable, the number of moose dunn (1975) reported entering the water/hr was multiplied by the average length of a visit (to the nearest hour) for that month. both sets were expressed as a percent of the maximum activity (fig. 2). the results indicate the pattern of moose use of ponds has either changed with time, or use in the study area differs from use statewide. ponds that appear to be heavily used by moose but have few aquatic plants, are common throughout northern maine (gentable 1. average biomass (g dry wt/m2 ± 1sd) inside and outside exclosures after 3 and 6 growing seasons. 3 years 6 years pond exclosure water depth inside outside inside outside (cm) bartlett a 25 41.6 ± 28.8 0.2 ± 0.2 108.3 ± 35.9 0.2 ± 0.2 bartlett c 50 6.1 ± 6.1 0.2 ± 0.1 leonard a 65 1.2 ± 0.3 0.9 ± 0.6 leonard b 50 4.5 ± 2.1 0.3 ± 0.4 august aquatic vegetation morris alces vol. 38, 2002 218 selection by moose in northern maine. m. s. thesis, university of maine, orono, maine, usa. maynard, l. a., and j. k. loosli. 1969. animal nutrition. mcgraw hill book company, new york, new york, usa. miller, b. k., and j. a. litvaitis. 1992. use of roadside saltlicks by moose, alces alces, in northern new hamps h i r e . c a n a d i a n f i e l d n a t u r a l i s t 106:112-117. murie, a. 1934. the moose of isle royale. university of michigan, museum of zoology, miscellaneous publications number 25. thompson, m. e. 1987. seasonal home range and habitat use by moose in northern maine. m. s. thesis, university of maine, orono, maine, usa. 4018.p65 alces vol. 40, 2004 mihajlovich and blake – herbicide and lichen 7 an evaluation of the potential of glyphosate herbicide for woodland caribou habitat management milo mihajlovich1 and peter blake2 1incremental forest technologies ltd., 7327-118a street, edmonton, ab, canada t6g 1v3; 2canadian forest products ltd., bag 100, grande prairie, ab, canada t8v 3a3 abstract: two studies evaluating the effects of glyphosate used for habitat management on caribou lichens and dwarf shrubs were undertaken. glyphosate substantially reduced blueberry cover at all rates tested in both cutover and uncut areas. glyphosate did not affect caribou lichen cover. alces vol. 40: 7-11 (2004) key words: caribou lichen, glyphosate, habitat management, tolerance woodland caribou (rangifer tarandus) are listed as a threatened species, in alberta and canada (by alberta sustainable resource development, fish and wildlife division and committee on the status of endangered wildlife in canada – anonymous 2002). as a threatened species, woodland caribou are on the blue list of species at risk of declining to non-viable levels and meriting special attention in forest management and forest management research. these efforts include developing an understanding of the factors hindering woodland caribou population recovery when management regimes are altered in their favour. wildlife managers suggest predation on woodland caribou young is, at present, the major factor limiting population response to more favourable management regimes. james (1999) and chowns (2003) suggest woodland caribou in alberta and ontario historically used habitats that limited contact with other ungulate species, thus reducing the frequency of encounter with predators (especially wolves, canis lupus). reduced frequency of encounter is suggested as a primary means of woodland caribou avoiding predation of their young by wolves. the boreal ecotype of woodland caribou achieves this through use of habitats unfavourable to other ungulate species (for example, jackpine, pinus banksiana lamb. – caribou lichen, cladina mitis (sandst.) hale & w. culb.) sites. the mountain ecotype of woodland caribou achieves this through spending a considerable portion (spring through late autumn) of the year in high elevation tundra site types, which are unfavourable to other high-elevation species. courtois et al. (2004) suggest joint management of caribou, moose, and wolves as a management scale (3,000 – 7,000 km2) strategy. understanding woodland caribou habitat use and the importance of predation in limiting woodland caribou recovery has led to an emphasis on managing habitat to reduce predator – prey interactions. one means of reducing predator – prey interaction is to provide woodland caribou with large, contiguous areas of habitat unattractive to other ungulates – thereby substantially reducing total ungulate population density and hence attractiveness to predators. to this end a number of very large cutblocks have been harvested in the boreal lower foothills ecoregion (beckingham et al. 1996) caribou management zone. the area of herbicide and lichen – mihajlovich and blake alces vol. 40, 2004 8 these large cutovers ranges from slightly larger than 100 ha to approximately 350 ha. it is presumed that the large size of these areas will reduce their appeal to large ungulates (moose, alces alces andersoni) which will, in turn, reduce wolf use of these areas (anonymous 2002). if these areas are managed for reduced appeal to browsing ungulates, management techniques to reduce browse development after harvesting may be of potential value. this would be especially appealing if major components of the diet of woodland caribou were less affected than moose browse. the authors identified the possible need for browse management in 1997 and assessed several options for longer-term browse reduction. broadcast herbicide application appeared to offer the most cost effective (biring et al. 1996) means, and least disruptive to caribou calving, of managing browse – due to treatment occurring several months after calving and occurring quickly with minimal human access on the ground during treatment. glyphosate is the most commonly used herbicide for forest management in canada – approximately 94% of f o r e s t m a n a g e m e n t h e r b i c i d e u s e i s glyphosate (ccfm 2004). approximately 38% of all forest stand tending in canada is with glyphosate herbicide. glyphosate is chosen for its effectiveness in controlling a broad spectrum of competing species (biring et al. 1996). glyphosate is classed as moderately persistent material in canadian soils (willis and macdowell 1983, tortenson 1985) with a half-life in soil of 20 – 100 days depending on soil conditions. health canada in the 1987 decision document on glyphosate registration for forestry deemed glyphosate to be a non-leaching (i.e., not soil mobile) based on an agriculture canada study of potential for soil mobility of pesticides (agriculture canada 1986). lautenshlager and sullivan (2002) recently reviewed the effects of herbicide treatments (primarily with glyphosate) on the biotic components of regenerating northern forests – they concluded that herbicides are a safe, effective tool for restoring conifers to previously conifer – dominated systems which have otherwise been replaced with hardwoods since europeans began harvesting those systems. oberg (2001) cites several articles that demonstrate terrestrial lichens (species not given) comprise the major component of the caribou winter diet. the case is made that both ecotypes of caribou in west-central alberta rely on caribou lichen as their primary winter food source. however, no data could be found on effects of broadcast glyphosate application, for browse management, on terrestrial lichens (caribou lichens) that form such a significant part of the woodland caribou’s diet. therefore the authors set up two experiments to examine the impact of broadcast glyphosate herbicide treatment on caribou lichen. g study area a lodgepole pine (pinus contorta var. latifolia) stand harvested and replanted to lodgepole pine 2 years prior to treatment was selected as a study site. located approximately 80 km southeast of grande prairie, alberta the site is located on a sandy soil. moisture regime is sub-mesic and soil nutrient levels appear to be low as key understory plants in the uncut stand were caribou lichens, bearberry (arctostaphylos uva-ursi), and scattered canada wild rye (elymus canadensis). methods two small-plot herbicide treatment trials were installed – one in the cutover and reforested area and a second in an adjacent uncut portion of the original lodgepole pine stand. treatments were glyphosate, applied as vision silviculture herbicide, at rates of alces vol. 40, 2004 mihajlovich and blake – herbicide and lichen 9 2, 4, and 6 l ha-1 (0, 712, 1424, and 2136 g (ae) glyphosate isopropylamine ha-1) with an untreated control (table 1). all treatments were applied using a carbon dioxide propelled small plot sprayer. application volume was 75 l ha-1, using 110015 flat fan nozzles. applications were made on september 9, 1997 during the typical season when herbicides might be applied for browse management – a time when browse may be controlled without negatively affecting conifer regeneration. weather conditions at treatment were: wind west at 2 km h-1, temperature 23°c, and relative humidity 58 percent. treatment plots were laid out prior to treatment. a single, fixed 0.25 m2 subplot for repeated assessment of subject plant cover was established and marked in each treatment plot. subplots were randomly located in each treatment plot and were used to overcome variability in cover and extent of subject plants. cover was assessed by ocular estimate. all estimates were by the same investigator and were made blind (without knowledge of the treatment being assessed). subject species were caribou lichen (a complex of cladina cenotea, c. rangiferana, and c. stellaris) and blueberry (vaccinium myrtilloides); bearberry was present but was not suffitable 1. browse control treatments evaluated. ciently uniform in distribution to provide an evaluation of treatment impact. caribou lichen cover was assessed 10 months after treatment (mat), blueberry cover was assessed 10 and 22 mat. a randomized, complete block experimental design was used. treatments were replicated 4 times. all comparisons were based on means of the 4 replicates. oneway analysis of variance was used to test for significant differences (p = 0.05) within a subject species, time period, and locale. if differences were found, tukey’s honestly significant difference separation of means test (university of missouri rolla 2002) was used to elucidate differences between treatments at that time period and locale. results results of cover assessments, at 10 and 22 months after treatment (mat), are given in table 2. broadcast glyphosate treatment, regardless of application rate did not significantly affect caribou lichen cover in either area (cutover and uncut). no visual symptoms of glyphosate activity were noted on the caribou lichen. sub-lethal glyphosate symptoms on susceptible species typically include: stunting (reduced elongation) of new stem or branch growth, dwarfing of new foliage, yellow or white color of new foliage, and clustering of foliage at branch tips. blueberry cover was reduced significantly, at both 10 and 22 mat by all glyphosate rates in both the cutover area and the uncut stand. cover reduction was directly related to glyphosate application rate with the two labelled application rates (4 and 6 l of product ha-1) resulting in greater reduction in cover than the lowest rate examined (2 l of product ha-1). reductions in blueberry cover were greater in the uncut area than in the cutover area (significance not tested). treatment number description application rate (l ha-1) active ingredient rate (g ha-1) 1 untreated control 0 0 2 ½ nominal brush rate 2 712 3 nominal brush rate 4 1424 4 maximum labeled rate 6 2136 herbicide and lichen – mihajlovich and blake alces vol. 40, 2004 10 discussion browse management to reduce use of recently reforested cutovers may be a means of limiting use of these areas by moose. it has been postulated that a reduction in moose utilization will, in turn, reduce the attractiveness of recent cutovers to predators (james 1999, courtois et al. 2004) thereby reducing predator impact on woodland caribou populations. herbicide use for browse reduction on large cutovers may be a feasible means of maintaining low browse densities (biring et al. 1996). results of this experiment suggest broadcast application of glyphosate herbicide for browse reduction would not negatively impact the supply of caribou lichen (and hence caribou forage) on treated areas. however, herbicide treatment would result in some reduction of blueberries thus reducing forage for species other than woodland caribou. conclusion broadcast application of glyphosate herbicide for browse management should not negatively impact caribou lichen extent on treated areas. table 2. changes in plant cover, 10 and 22 months after treatment. 1 lichen was a complex of cladina rangiferana, c. cenotea, and c. stellaris. 2 cover values are averages of ocular estimates of cover in 4 – 0.25 m2 permanent sub-plots established prior to treatment in each plot. thus each cover represents the average of a total of 16 cover assessments. 3 percent change values shown in bold font are “honestly significantly different” from pre-treatment cover values using tukey’s honestly significant difference (hsd) test. α = 0.05. acknowledgements the authors wish to thank dr. tom sullivan of the applied mammal research institute for his review and comments; canadian forest products limited for their support of this research; and two anonymous reviewers for their comments and suggestions. references agriculture canada. 1986. pesticide priority scheme for water monitoring program. unpublished report, pesticide directorate, food production and inspection branch, ottawa, ontario, canada. anonymous. 2002. threatened wildlife. alberta sustainable resource development, natural resource services, edmonton, alberta, canada. beckingham, j., i. g. w. corns, and j. h. archibald. 1996. field guide to ecosites of west-central alberta. natural resources canada, canadian forest service, northwest region, northern forestry centre, special report 9. edmonton, alberta, canada. biring, b. s., p. g. comeau, and j. o. locale glyphosate (g ha-1) at treatment 10 mat % change 10 mat3 at treatment 10 mat % change 10 mat3 22 mat % change 22 mat3 cutover 0 71 77 8.5 16.7 16.7 0 18.3 9.6 712 75 89 18.7 20 8.3 -58.5 15 -25 1424 80 83 3.8 11.7 3.3 -71.8 10 -14.5 2136 78 80 2.6 13.3 0 -100 3.3 -75.2 uncut 0 77 77 0 10 11.7 17 13.3 33 712 85 80 -5.9 11.7 4 -65.8 8.3 -29.1 1424 88 87 -1.1 16.7 2 -88 6.7 -59.9 2136 83 82 -1.2 13.3 0 -100 4 -69.9 lichen cover1, 2 blueberry cover2 alces vol. 40, 2004 mihajlovich and blake – herbicide and lichen 11 boateng. 1996. effectiveness of vegetation control methods in british columbia. canadian forestry service and british columbia ministry of forests. victoria, british columbia, canada. (ccfm) canadian council of forest ministers. 2004. pest control product use. http://nfdp.ccfm.org/compendium/pest/ index_e.php. chowns, t. j. 2003. state of the knowledge of woodland caribou in ontario. forestry research partnership. http:// w w w . f o r e s t r e s e a r c h . c a / product_catalogue/reports.htm. courtois, r., j.-p. ouellet, c. dussault, and a. gingras. 2004. forest management guidelines for forest-dwelling caribou in québec. forestry chronicle 80:598-607. james, a. r. c. 1999. effects of industrial development on the predator-prey relationships between wolves and caribou in northeastern alberta. ph.d. thesis, department of renewable resources, university of alberta, edmonton, alberta, canada. lautenschlager, r. a., and t. p.sullivan. 2002. effects of herbicide treatments on biotic components of regenerating northern forests. forestry chronicle 78:695-730. oberg, p. r. 2001. responses of mountain caribou to linear features in a westcentral alberta landscape. m.sc. thesis, department of renewable resources, university of alberta, edmonton, alberta, canada. tortensson, l. 1985. behaviour of glyphosate in soils and its degradation. page 137 in e. grossbard and d. atkinson, editors. the herbicide glyphosate. butterworths, london, u.k. university of missouri – rolla. 2002. virtual statistician website. accessed 2002. http://web.umr.edu/~psyworld/ tukeys4mean.htm. willis, g. h., and l. l. mcdowell. 1983. pesticides in agricultural runoff and their effects on downstream water quality. environmental toxicology and chemistry 1:267. alces37(1)_71.pdf 1 abundance of winter ticks (dermacentor albipictus) in two regenerating forest habitats in new hampshire, usa brent i. powers and peter j. pekins department of natural resources and the environment, university of new hampshire, durham, nh 03824, usa. abstract: recent decline in new hampshire’s moose (alces alces) population is attributed to sustained parasitism by winter ticks (dermacentor albipictus) causing high calf mortality and reduced productivity. location of larval winter ticks that infest moose is dictated by where adult female ticks drop from moose in april when moose preferentially forage in early regenerating forest in the northeastern united states. the primary objectives of this study were to: 1) measure and compare larval abundance in 2 types of regenerating forest (clear-cuts and partial harvest cuts), 2) measure and compare larval abundance on 2 transect types (random and high-use) within clear-cuts and partial harvests, and 3) identify the date and environmental characteristics associated with termination of larval questing. larvae were collected on 50.5% of 589 transects; 57.5% of transects in clear-cuts and 44.3% in partial cuts. the average abundance ranged from 0.11–0.36 ticks/m2 with abundance highest (p < 0.05) in partial cuts and on high-use transects in both cut types over a 9-week period; abundance was ~2 × higher during the principal 6-week questing period prior to the first snowfall. abundance (collection rate) was stable until the onset of < 0°c and initial snow cover (~15 cm) in late october, after which collection rose temporarily on high-use transects in partial harvests during a brief warm-up. the higher abundance of winter ticks on high-use transects indicates that random sampling underestimates tick abundance and relative risk of infestation of moose. calculating an annual index of infestation of winter ticks on moose is theoretically possible by integrating 3 factors: the infestation of harvested moose in october, the length of the questing period, and assuming a stable collection rate during the questing period. alces vol. 56: 1 – 13 (2020) key words: alces alces, dermacentor albipictus, forest habitat, infestation, moose, questing, tick abundance, winter ticks the influence of winter ticks (dermacentor albipictus) on population dynamics of moose (alces alces) in the northeastern united states (northeast) is well documented (musante et al. 2010, bergeron et al. 2013, jones et al. 2017, 2019, ellingwood et al. 2020). the physiological impact of blood loss on moose is directly associated with infestation level of winter ticks (musante et al. 2007), and recent research has addressed the physiology, ecology, and etiology of winter ticks (e.g., yoder et al. 2016, 2017a, 2017b, holmes et al. 2018). further, the presumed influence of climate change in the winter tick-moose relationship is that longer autumns and later onset of winter weather will extend the questing period of winter ticks (dunfey-ball 2017, jones et al. 2019). potential outcomes would include higher infestation levels, more frequent epizootics (>50% calf mortality), reduced productivity in yearling and adult cows, and sustained tick abundance on the landscape (musante et al. 2010, bergeron and pekins 2014, healy et al. 2018, 2020, jones et al. 2017, 2019). however, few studies winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 2 have attempted to measure field abundance of winter ticks (drew and samuel 1985, aalangdong 1994, addison et al. 2016), with only a single study in the northeast (bergeron and pekins 2014). as in typical host-parasite relationships, host density is directly related to parasite density with several studies indicating that increased moose density increases tick distribution and relative abundance (blyth 1995, pybus 1999, samuel 2007, bergeron and pekins 2014). field studies indicate that 85% of adult winter ticks are located within 60 cm of a moose carcass (drew and samuel 1985, 1986), and >95% of larvae are typically found within 1–2 m of the hatching location (drew and samuel 1985, 1986, addison et al. 2016) and ascend proximal vegetation the following autumn to quest for a host (drew and samuel 1985). likewise, in laboratory conditions yoder et al. (2016) found that larval ticks have limited mobility, crawling only ~1 m. recruitment of larval ticks is higher in open habitat than closed-canopy deciduous forest, except in hot and dry conditions (addison et al. 2016). therefore, distribution and questing location of winter ticks is where adult ticks drop from moose in marchapril, and the relative infestation risk is a function of environmental conditions and habitat use by moose. moose preferentially use young, regenerating forest habitat (4–16 years old) more than other cover types in spring and autumn (scarpitti et al. 2005, healy et al. 2018). further, the same animals demonstrate overlap in use of specific cuts during spring and autumn, suggesting a positive feedback loop of infestation (healy et al. 2018). in the single field study conducted in the northeast, larval abundance in clearcuts was generally related to moose density, but varied among and within clear-cuts (bergeron and pekins 2014). it is presumed that relative tick abundance is related to the previous years’ infestation level, and this earlier study was not preceded by or followed by an identified epizootic. this study was designed to measure larval abundance during autumnal questing in preferred cut habitat when tick abundance was presumably high following an epizootic in spring 2018 (61% calf mortality; powers 2019). study area the study area was in jericho state park located in the town of berlin entirely within wildlife management unit (wmu) c1 covering ~70 km2 in eastern coos county in northern new hampshire (utm 19 t 320970 e, 4926474 n; map in jones et al. 2017). moose density was estimated at 0.46– 0.87 moose/km2, down from 1.2 moose/km2 in 1998 (nhfg 2015). year-round access was through a network of former logging roads and off-highway recreational vehicle (ohrv)/snowmobile trails. the landscape was mostly lowland valleys with rolling hills and small water features (streams, rivers, ponds) scattered throughout. the predominant cover type was northern hardwood forest consisting of american beech (fagus grandifolia), sugar maple (acer saccharum), and paper and yellow birch (betula papyrifera and b. allegheniensis). conifer cover in low elevation areas consisted mostly of northern white cedar (thuja occidentalis), black spruce (picea mariana), red spruce (p. rubens), and balsam fir (abies balsamea); high elevation stands were red spruce and balsam fir (degraaf et al. 1992). the larger geographical area was the focus of a comprehensive moose habitat and survival study in 2002–2005 (scarpitti et al. 2005, musante et al. 2010), related studies of winter ticks and forest regeneration (bergeron et al. 2011, bergeron and pekins 2014), and since 2014, survival and productivity of moose alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 3 (jones et al. 2017, 2019, dunfey-ball 2017, healy et al. 2018, ellingwood et al. 2019). methods study plots were established in summer 2018 to measure larval abundance during the questing period in autumn 2018 (september– november). plots were established in two cut types: clear-cuts (n = 22) and partial harvests (e.g., geometric thinning) (n = 22) (fig. 1 and 2). each was within an age range associated with preferred foraging habitat (4–10 years), 4.04–4.85 ha in size, and with ample sign of moose use. moose use this area year-round and multiple radio-collared calves succumbed to infestation of winter ticks in springs 2014–2018. epizootic conditions occurred in the larger study area in spring 2018 (61% calf mortality) and 4 of the previous 5 years (jones et al. 2019). two treatments were defined in each plot: 1) random area within the plot ( similar to bergeron and pekins 2014), and 2) highuse areas that reflected concentrated moose activity. high-use areas were obvious foraging sites and movement corridors on trails and edges proximate to uncut forest that were readily identified from visual inspection and evidence of browsing (fig. 1 and 2). each plot was sampled at least 12 times 10 m fig. 1. schematic illustrating the sampling design in partial harvest plots in berlin, new hampshire, usa. partial harvests leave a mix of cut and uncut areas that create proximal foraging and bedding areas for moose. green clouds depict uncut portions of trees (canopy cover) and black lines depict cut area and skid trails that serve as pathways and foraging sites. orange lines depict typical location of high-use transects set within cut areas and skid trails. blue lines depict typical random transects avoiding high-use areas and spaced 10 m apart. winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 4 during the questing period (mid-september through mid-november); sampling continued until collection of larvae ceased. line transects were established weeks prior to sampling after visual inspections of each plot to identify random and high-use sampling locations within each plot. transects were spaced at least 10 m apart and no repeat sampling occurred of a transect either daily or during a subsequent visit. plots were sampled bi-weekly by flagging at least 4 transects (2 random, 2 high-use) per visit. flagging followed the basic technique used by others (drew and samuel 1985, piesman et al. 1986, ginsberg and ewing 1989, aalongdong 1994, bergeron and pekins 2014) in which a 1 m2 cotton cloth was dragged over vegetation to collect questing larvae. each transect flag was bagged (plastic ziplock) separately and frozen. subsequently, an entire count of larvae on each flag was performed to calculate abundance (ticks/m2; area = transect length (m) × 1 m2). a subset of plots (4 clear-cut and 4 partial harvest) were monitored continuously with remote data-loggers that measured hourly ambient temperature (±0.5ºc) from midaugust until late november at the typical questing height (125 cm) of larvae ( mcpherson et al. 2000). these data were analyzed relative to collection rate and tick abundance to investigate relationships between temperature, tick abundance, and relative questing activity. snow events were also monitored given the susceptibility of larvae to freezing/ desiccation (drew and samuel 1985). analysis the raw data exhibited the typical field-sampling problem of “zero-inflated” 10 m fig. 2. schematic illustrating the sampling design in clear-cut plots in berlin, new hampshire, usa. the clear-cut is white and set within a green cloud of unharvested forest; skid trails are depicted with black lines. orange lines depict high-use transects placed within foraging and movement pathways (e.g., edges and skid trails). blue lines depict random transects spaced 10 m apart through the uniform regenerating forest. alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 5 data, as ~50% of transects were tick-less (i.e., negative transects); therefore, the data were analyzed using a hurdle or “two-stage” linear model. the first stage was a logistic model that used the binary form of all transect data including negative transects; data were not log-transformed. in the second stage, the negative transects were removed and only positive transects were analyzed. after testing for normality, these data were subsequently log-transformed to fit a normal distribution. both were used to test if larval abundance was different between clear-cuts and partial harvests, and between random and high-use transects within cuts. a temporal analysis of larval abundance using all transect data, ambient temperature, and questing activity was performed with a linear mixed-effects model. fixed variables in the model included ambient temperature, transect type, date, snow depth, with plots as the random effect variable; analysis was performed in program r (ver. 3.4.4, austria). basic summary statistics were used to express and analyze ambient temperature measurements. the average daily temperature and abundance data were used to analyze temporal factors possibly influencing abundance within the model. a t-test was used to compare ambient temperature between plot types; analysis was performed in program r (ver. 3.4.4, austria). results a total of 589 transects were measured in the 44 plots from 15 september–20 november 2018. transect length ranged from 28–322 m (median = 177 m) in clear-cuts and 45–322 m (median = 177 m) in partial harvests (table 1). larval questing had initiated at the start of dragging on 15 september. the absolute number of larvae collected per transect ranged from 0–2,554 larvae. for all transects combined, the absolute average and maximum abundances were always higher on high-use than random transects in both cut types, with larger differences in partial cuts; a similar trend occurred on positive transects alone that had abundances ~1.5–2.5 × higher than the overall combined averages (table 1). the first stage model (all transects) indicated that abundance was 1.8 × times higher (p < 0.05) in partial harvests (0.24 ± 0.08 ticks/m2) than clear-cuts (0.13 ± 0.03 table 1. field abundance (ticks/m2) of larval ticks collected from 15 september–10 november 2018 in 22 clear-cut and 22 partial cut study plots, berlin, new hampshire, usa. positive transects were those where larvae were collected. random indicates transects that were distributed randomly within a plot. high-use indicates transects that were located in areas of concentrated moose activity (i.e., game trails and foraging areas). all transects clear-cut (random) clear-cut (high-use) partial harvest (random) partial harvest (high-use) # of transects 140 138 155 156 transect length (m) 74–321 28–322 70–322 45–322 abundance (se) 0.12 (0.02) 0.15 (0.04) 0.11 (0.03) 0.36 (0.13) max abundance 1.90 5.52 4.04 13.45 range (# ticks/transect) 0–459 0–975 0–527 0–2554 positive transects # of transects 74 86 66 72 abundance (se) 0.22 (0.04) 0.25 (0.06) 0.27 (0.23) 0.81 (0.29) range (# ticks/transect) 1–459 1–975 1–527 1–2554 winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 6 ticks/m2) (table 1). there was a strong trend (p = 0.13) toward higher abundance on highuse than random transects in both cut types. the second stage model indicated that abundance was 2.3 × higher (p = 0.05) in partial harvests (0.54 ± 0.35 ticks/m2) than clearcuts (0.24 ± 0.11 ticks/m2 (table 1). abundance in clear-cuts was similar (p = 0.47) on random (0.22 ± 0.04 ticks/m2) and high-use transects (0.25 ± 0.06 ticks/m2), whereas abundance was higher (p < 0.05) on highuse (0.81 ± 0.29 ticks/m2) than random transects (0.27 ± 0.23 ticks/m2) in partial harvests (table 1). two drags on high-use transects in partial harvests (13.2 and 13.4 ticks/m2) substantially elevated the mean abundance estimates in weeks 2 (1.25 ticks/m2) and 8 (0.98 ticks/m2) (table 2). these values reflected the collection of very large clusters of larvae (identified on the flags) and could be considered outliers relative to weekly estimates; their removal would more closely align the weekly estimates (0.10 and 0.05 ticks/m2). however, these data were retained in the weekly analyses because they represent important characteristics of local variation in larval abundance and ecology. weekly mean abundances were used to test for a temporal relationship because the infestation rate of moose is presumably correlated with the relative abundance of larvae. because snow and low ambient temperature at the end of week 6 measurably reduced the collection rate (abundance) in week 7 (table 2, fig. 4), linear regression was used to determine if abundance was constant (i.e., slope = 0) across the first 6 weeks assuming that collection rate was mostly unaffected by weather. further, because transect type was not related to abundance in clear-cuts, the means of random and high-use transects in clearcuts (table 2) were averaged to produce a weekly abundance; partial harvest data were not tested because abundance differed by transect type. the slope in clear-cuts was 0.0019 (90% ci = −0.004 to 0.456) and not different than 0 (p > 0.05), indicating that weekly abundance was stable in the first 6 weeks (fig. 3). the first stage model was rerun with the 6-week data and indicated that abundance was 1.7 × higher (p = 0.01) in partial harvests (0.25 ± 0.08 ticks/m2) than clear-cuts (0.15 ± 0.03 ticks/m2) (table 3). similarly, higher abundance (p = 0.02) occurred on high-use than random transects in partial harvests (0.36 ± 0.15 vs. 0.14± 0.05 ticks/m2) and clear-cuts (0.17 ± 0.03 vs. 0.14 table 2. weekly larval tick abundance (ticks/m2) from 15 september to 10 november 2018, berlin, new hampshire, usa. transect type indicated by “random” and “high-use” within both cut types. clear-cut random se clear-cut high-use se partial harvest random se partial harvest high-use se week 1 0.21 0.11 0.12 0.04 0.03 0.01 0.17 0.15 week 2 0.03 0.03 0.13 0.04 0.43 0.40 1.25 1.31 week 3 0.20 0.08 0.32 0.23 0.01 0.01 0.19 0.15 week 4 0.16 0.07 0.22 0.07 0.16 0.07 0.35 0.33 week 5 0.01 0.01 0.09 0.06 0.03 0.02 0.24 0.11 week 6 0.20 0.10 0.22 0.08 0.19 0.12 0.19 0.08 week 7 0.08 0.05 0.03 0.01 0.03 0.02 0.17 0.12 week 8 0.05 0.02 0.03 0.01 0.12 0.08 0.98 0.84 week 9 0.03 0.02 0.00 0.00 0.00 0.00 0.03 0.03 alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 7 ± 0.05 ticks/m2). abundance was similar on random transects in both cut types (0.14 ticks/m2) (table 3). interestingly, the second stage of the model indicated that abundance was 1.9 × higher (p = 0.03) in partial harvests (0.54 ± 0.16 ticks/m2) than clear-cuts (0.28 ± 0.05 ticks/m2), but transect type had no effect on abundance (p = 0.90) (table 3). absolute abundance on high-use transects was always higher than on random transects in both cut types (table 2). with the onset of cold temperatures and snow in late october (week 6), abundance declined in each plot and transect type in week 7 (fig. 4). however, a temporary increase in activity and collection occurred on fig. 3. the mean weekly abundance of winter ticks in clear-cuts from 15 september–26 october 2018, berlin, new hampshire, usa. each point is the weekly mean prior to the snow event on 26 october that induced substantial reduction in tick abundance. the vertical line at each point represent standard error. the dotted line represents the temporal linear relationship that indicated that abundance was stable during the 6 weeks. table 3. the 6-week field abundance (ticks/m2) of larval ticks collected in 15 september–27 october 2018 in 22 clear-cut and 22 partial cut study plots, berlin, nh. random indicates that transects were distributed randomly within a plot. high-use indicates transects that were located in areas of concentrated moose activity (i.e., game trails, and foraging areas). clear-cut (random) clear-cut (high-use) partial harvest (random) partial harvest (high-use) # of transects 105 105 106 107 transect length (m) 74–321 28–322 70–322 45–322 mean abundance (se) 0.14 (0.03) 0.17 (0.03) 0.14 (0.05) 0.36 (0.15) max abundance 1.90 5.52 4.04 13.45 range (# ticks/transect) 0–459 0–975 0–527 0–2554 positive transects # of transects 53 69 47 52 abundance (se) 0.27 (0.06) 0.30 (0.08) 0.30 (0.10) 0.74 (0.29) range (# ticks/transect) 1–459 1–975 1–527 1–2554 winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 8 5 november (week 8) in partial cuts when ambient temperature rose to 8.5°c; abundance in clear-cuts did not increase concurrently (table 2, fig. 4). by 10 november (week 9, table 2), abundance was functionally zero based on lack of collection and the obvious (observed) inability of the few collected larvae to crawl on the flag. the onset of sustained snow cover and temperatures <0°c coincided with a decline in larval abundance (p < 0.05). decline in abundance in both cut and transect type from 15 september to 20 november was correlated with date (p = 0.002). no individual effect was found with temperature or snow depth (p > 0.05); however, a significant interaction effect (p = 0.03) indicated their negative combined effect on abundance. the termination of questing was assumed as 10 november based on lack of collection and consistent ambient temperature <0ºc. the minimal length of the questing period was 56 days based on the sampling period (15 september– 10 november), but this is considered a conservative estimate because larvae were questing on 15 september. discussion winter tick epizootics are typically considered sporadic events (samuel 2004) and were undocumented in the northeast until the mid-2000s (musante et al. 2010); more recently, the frequency of epizootics is unprecedented in the northeast – 5 in 6 years (jones et al. 2019, powers 2019, ellingwood et al. 2020). not surprisingly, winter tick abundance on the landscape is poorly understood, in part, because epizootics were infrequent or unknown, and the fieldwork associated with measuring tick abundance is labor-intensive. similarly, little is known about the actual -15 -10 -5 0 5 10 15 20 25 wk 1 wk 2 wk 3 wk 4 wk 5 wk 6 wk 7 wk 8 wk 9 te m pe ra tu re (° c) week (15 september 16 november) clear-cut par�al harvest 1s t s no w fa ll brief warm-up fig. 4. average daily temperature in clear-cuts and partial harvests from 15 september–16 november 2018, berlin, new hampshire, usa. alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 9 distribution of larval ticks on the landscape relative to the dynamic nature of multiple variables including moose density, habitat/ forest diversity, habitat use and movement patterns of moose, and micro-environmental conditions that influence tick survival. this study provides novel information about tick abundance in 2 optimal foraging habitats of moose, length of the larval questing period, and conditions that terminate questing. although the average larval abundance on random transects in clear-cuts and partial harvests (0.12 and 0.11 ticks/m2; table 1) was similar to that measured previously in new hampshire (2-year average = 0.11; bergeron and pekins 2014), the average abundance on high-use transects was 1.4– 3.3 × higher (0.15 and 0.36 ticks/m2, respectively; table 1). further, the maximum abundance on random (1.9 ticks/m2) and high-use transects (5.52 ticks/m2) in clearcuts (table 1) was considerably higher than that (0.40–0.64 ticks/m2) measured a decade earlier, and ticks were collected in all clear-cuts whereas ~10% were without ticks in 2008–2009 (bergeron and pekins 2014). the average abundance was much higher in elk island national park in alberta, canada (1.36 ticks/m2) in the year preceding a moose die-off (aalangdong 1994, samuel 2007), except in week 2 and week 8 in partial harvests (table 2). it is not clear why the abundance in alberta was much higher than that measured after the spring 2018 epizootic, and why average abundance in clear-cuts in new hampshire was relatively stable since 2008–2009 despite multiple epizootics. the data reflect the difficulty and variability associated with measuring larval abundance, but also indicate that larval abundance likely increased over the past decade. furthermore, the abundance estimates provided here and in bergeron and pekins (2014) should be considered conservative for a number of reasons. most importantly, we have no ability to estimate the efficiency or detection probability of a single drag, but it is improbable that all larvae are collected with a single drag regardless of time of day or environmental conditions. we encourage multiple sampling of transects in future experiments to improve accuracy and abundance estimates. predictably, larvae were not distributed evenly within either cut type, as not all transects produced ticks and abundance was higher on high-use transects (table 1). both reflect non-random or preferred habitat use by moose, and maximum abundance always occurred on high-use transects in both cut types – 13.45 ticks/m2 in partial harvests and 5.52 ticks/m2 in clear-cuts. the similar abundances on random transects in this and the previous regional study (bergeron and pekins 2014) indicates that random sampling likely underestimates tick abundance, moose-tick encounter rates, and projected infestation rates. for example, the abundance estimates on positive transects was ~ 2 × higher than the overall average during the principal 6 weeks of questing (table 3). it is important to recognize that the earlier study reported a regional abundance, whereas this study was within a focal area of ~70 km2 with a moderate-high moose density experiencing winter tick-associated mortality (jones et al. 2019). the effect of winter conditions on questing was evident due to the combined influence of temperature and weather (drew and samuel 1985). specifically, overall abundance declined in both cut and transect types after the snowfall on 24 october (week 6; table 2, fig. 4). although the exposure time at <0°c lasted 3 days (25–27 october), the warm-up on 5 november (weeks 7 and 8) and associated increase in collection rate in partial harvests reflects the resilience of winter ticks at these conditions (holmes et al. 2018, addison et al. 2019), particularly on high-use transects in partial harvests winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 10 (table 2). the few larvae collected in week 9 were curled and immobile, characteristics consistent with thermally stressed larvae (holmes et. al. 2018), and were presumably collected due to their claw-like appendages. as addison et al. (2019), we found that a short-term warmup after an initial snowfall resulted in a temporary increase in larval collection, specifically in partial harvests, indicating that prolonged (multi-day) winter weather is necessary to terminate questing. some larvae may have been protected within insulative layers/gaps in the more complex vegetative/stand structure of partial harvests than in more open clearcuts. eventually, sustained below-freezing temperatures and snow cover were lethal to questing ticks in all plots. preferential habitat use by moose is well documented in northeastern forests (scarpitti et al. 2005, wattles and destefano 2013), as is selective use of regenerating forest habitat during the autumn questing and spring drop-off seasons of winter ticks (healy et al. 2018). open, regenerating habitat presumably provides higher relative survival of larvae that decline in abundance and survival as canopy cover exceeds 60% (drew and samuel 1986, aalangdong 1994, terry 2015, addison et al. 2016). abnormally dry and drought-like conditions in late summer and early autumn can measurably reduce larval survival (dunfey-ball 2017), but less so in closed canopy habitat (addison et al. 2016). partial harvests arguably provide an optimal mix of foraging (open) and bedding (canopy) habitat for moose, an optimal mix of microhabitats to sustain egg and larval abundance of winter ticks in a range of environmental conditions, and subsequently, an optimal transmission nidus that sustains winter tick infestation of moose. although a moose-tick encounter rate was not measured, the stable abundance measured throughout the questing period is potentially useful to estimate the final infestation (index) at the termination of questing. infestation is measured on the shoulder and rump of harvested moose (sine et al. 2009, bergeron and pekins 2014) in october in maine, new hampshire, and vermont to produce an annual harvest index that is correlated with the probability of winter tick associated mortality of calves (dunfey-ball 2017). however, a stronger relationship exists between a similar index measured on january-captured calves of known fate (ellingwood et al. 2019, jones et al. 2019). assuming the infestation rate is stable throughout the questing period (as reflected by the stable abundance measured here), the harvest index could be extrapolated to a final index by assuming two dates: 1) the start date of the questing period and 2) the date that questing terminates due to environmental conditions. the extrapolated final index could be substituted for the january index to better predict survival of calves, assuming that larvae and nymphs are not measurably reduced by grooming prior to early january; however, this assumption may be invalid as experimentally infested (larvae) captive moose groomed throughout autumn (addison et al. 2019). ongoing analyses are exploring the potential accuracy and usefulness of such an approach. the variability in tick abundance by cut and transect type not only reflects areas of lower and higher infestation risk, but also, that relative risk reflects individual differences in activity, foraging behavior, and habitat use by moose. likewise, the annual infestation on harvested moose varies considerably by sex and age (samuel and barker 1979, drew and samuel 1985, bergeron and pekins 2014), and for calves, mortality is directly related to the level of individual infestation (ellingwood et al. 2019). those calves surviving in an epizootic year presumably reflect local variance in tick abundance, alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 11 relative infestation risk, and individual habitat use within the epizootic area. using previous larval abundance estimates (bergeron and pekins 2014) in an agent-based model based upon availability of local regenerating (cut) habitat and its use by radio-collared moose, healy et al. (2020) predicted calf mortality similar to that measured in the field (jones et al. 2019). the strong influence of preferential habitat use on infestation was supported by this modeling exercise that restricted moose-tick encounters to cut habitat that was <20% of the home range of moose. the higher larval abundances reported here suggest that predictions of healey et al. (2020) were conservative and that proportionally small, yet high-use travel routes and foraging areas within cuts provide the nexus for high infestations on moose. interestingly, differences in moose and tick response to clear-cuts and partial harvests might lead to differences in the adjacent states of maine and new hampshire. forest harvest regulations enacted in the 1989 state practices act of maine effectively restricted size of clear-cuts in response to extensive salvage operations associated with the regional outbreak of spruce budworm (choristoneura spp.); ironically, moose expansion in the northeast was spurred by these operations (bontaites and gustafson 1993, wattles and destefano 2011, dunfey-ball 2017). however, timber removal has since increased not declined in maine because the footprint of forest harvesting has expanded as partial harvests have increased >90% (mfs 2016). these harvest regulations may have increased and sustained high availability of more preferred/optimal habitat and moose density, while inadvertently increasing local tick abundance, infestation rate of moose, and the probability of an epizootic during warming weather and environmental conditions that simultaneously benefit winter ticks. acknowledgements this research was made possible through funding from the wildlife restoration program grant no. f13af01123 (nh w-104-r-1) to n.h. fish and game department from the u.s. fish and wildlife service, division of wildlife and sport fish restoration with matching funds provided by the university of new hampshire. field and laboratory efforts were aided by many students and research technicians including b. bousquet, p. fitzgibbons, o. fortuna, p. massingham, o. mcgovern, a. miller, j. o’del, r. parker, c. simmons, and m. vogt. finally, e. addison provided his usual expert editing to improve this manuscript. references aalangdong, o. i. 1994. winter tick (dermacentor albipictus) ecology and transmission in elk island national park, alberta. m. s. thesis. university of alberta, edmonton, canada. addison, e. m., d. j. h. fraser, and r. f. mclaughlin. 2019. grooming and rubbing behavior by moose experimentally infested with winter ticks (dermacentor albipictus). alces 55: 23–35. _____, r. f. mclaughlin, p. a. addison, and j. d. smith. 2016. recruitment of winter ticks (dermacentor albipictus) in contrasting forest habitats, ontario, canada. alces 52: 29–40. bergeron, d. h., and p. j. pekins. 2014. evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire. alces 50: 1–15. _____, _____, h.f. jones, and w.b. leak. 2011. moose browsing and forest regeneration: a case study. alces 47: 39–51. _____, _____, and k. rines. 2013. temporal assessment of physical characteristics and reproductive status of moose in new hampshire. alces 49: 39–48. blyth, c. b. 1995. dynamics of ungulate populations in elk island national park. winter tick abundance and distribution – powers and pekins alces vol. 56, 2020 12 m. s. thesis. university of alberta, edmonton, canada. bontaites, k. m., and k. guftafson. 1993. the history and status of moose management in new hampshire. alces 29: 163–167. degraaf, r. m., m. yamasaki, w. b. leak, and j. w. lanier.1992. new england wildlife: management of forested habitats. general technical report ne-144. usda forest service, northeast experiment station, randor, pennsylvania, usa. drew, m. l., and w. m. samuel. 1985. factors affecting transmission of larval winter ticks, (dermacentor albipictus packard), to moose, alces alces l., in alberta, canada. journal of wildlife diseases 21: 274–282. doi: 10.7589/0090-3558-21.3.274 _____, and _____. 1986. reproduction of the winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64: 714–721. doi: 10.1139/z86-105 dunfey-ball, k. r. 2017. moose density, habitat, and winter tick epizootics in a changing climate. m.s. thesis. university of new hampshire, durham, new hampshire, usa. ellingwood, d., p. j. pekins, and h. jones. 2019. using snow urine samples to assess the impact of winter ticks on moose calf condition and surivival. alces 55: 13–21. _____, _____, _____, and a. r. musante. 2020. evaluating moose (alces alces) population response to infestation level of winter ticks (dermacentor albipictus). wildlife biology. doi: 10.2981/w16. 00619 ginsberg, h. s., and c. p. ewing. 1989. comparison of flagging, walking, trapping and collecting from hosts as sampling methods for northern deer ticks, ixodes dammini, and lone-star ticks, amblyomma americanum (acari: ixodidae). experimental & applied acarology 7: 313–322. doi: 10.1139/ z86-105 healy, c., p. j. pekins, s. attallah, and r. g. congalton. 2020. using agent-based models to inform the dynamics of winter tick parasitism of moose. ecological complexity 41: 100813. doi: 10.1016/j. ecocom.2020.100813 _____, _____, l. kantar, r. g. congalton, and s. atallah. 2018. selective habitat use by moose during critical periods in the winter tick life cycle. alces 54: 85–100. holmes, c. j., c. j. dobrotka, d. w. farrow, a. j. rosendale, j. b. benoit, p. j. pekins, and j. a. yoder. 2018. low and high thermal tolerance characteristics for unfed larvae of the winter tick dermacentor albipictus (acari: ixodidae) with special reference to moose. ticks and tick-borne diseases 9: 25–30. doi: 10.1016/j.ttbdis.2017.10.013 jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. _____, _____, _____, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’ neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi: 10.1139/ cjz-2018-0140 mcpherson, m., w. shostak, and w. m. samuel. 2000. climbing simulated vegetation to heights of ungulate hosts by larvae of dermacentor albipictus (acari: ixodidae). journal of medical entomology 37: 114–120. doi: 10.1603/0022-2585-37.1.114 musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. _____, _____, and _____. 2010. characteristics and dynamics of a regional moose alces alces population alces vol. 56, 2020 winter tick abundance and distribution – powers and pekins 13 in the northeastern united states. wildlife biology 16: 185–204. doi: 10.2981/09-014 new hampshire fish and game department (nhfg). 2015. new hampshire game management plan: 2016–2025. new hampshire fish and game department, concord, new hampshire, usa. piesman, j., j. g. donahue, t. n. mather, and a. spielman. 1986. transovarially acquired lyme disease spirochetes (borrelia burgdorferi) in field collected larval ixodes dammini (acari: ixodidae). journal of medical entomology 23: 219. doi: 10.1093/jmedent/ 23.2.219 powers, b. i. 2019. assessing the relationship of winter ticks, weather, and a declineing moose population in northern new hampshire. m.s. thesis. university of new hampshire, durham, new hampshire, usa. pybus, m. j. 1999. moose and ticks in alberta: a die-off in 1998/99. occasional paper no. 20. fisheries and wildlife management division, edmonton, alberta, canada. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. _____, w. m. 2007. factors affecting epizootics of winter ticks and mortality of moose. alces 43: 39–48. _____, and m. j. barker. 1979. the winter tick dermacentor albipictus (packard, 1869) on moose, alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. scarpitti, d. l., c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. sing, m.e., k. morris, and d. knupp. 2009. assessment of a line transect method to determine winter tick abundance on moose. alces 45: 143–146. terry, j. 2015. the habitat of winter ticks (dermacentor albipictus) in the moose (alces alces) range of northeast minnesota. m.s. thesis. university of minnesota, saint paul, minnesota, usa. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. _____, and _____. 2013. space use and movement of moose in massachusetts: implications for conservation of large mammals in a fragmented environment. alces 49: 65–81. yoder, j. a., p. j. pekins, h. f. jones, b. w. nelson, a. l. lorenz, and a. j. jajack. 2016. water balance attributes for offhost survival in larvae of the winter tick (dermacentor albipictus; acari: ixodidae) from wild moose. international journal of acarology 42: 26–33. doi: 10.1080/01647954.2015.1113310 _____, _____, a. l. lorenz, and b. w. nelson. 2017a. larval behaviour of the winter tick, dermacentor albipictus (acari:ixodidae): evaluation of co2 (dry ice), and shortand long-range attractants by bioassay. international journal of acarology 43: 187–193. doi: 10.1080/01647954.2016.1275791 _____, _____, b. w. nelson, c. r. randazzo, and b. p. siemon. 2017b. susceptibility of winter tick larvae and eggs to entomopathogenic fungi – beauveria bassiana, beauveria caledonica, metarhizium anisopliae, and scopulariopsis brevicaulis. alces: 53: 41–51. alces37(1)_19.pdf alces vol. 44, 2008 young and boertje – recovery of bull:cow ratios 65 recovery of low bull:cow ratios of moose in interior alaska donald d. young jr. and rodney d. boertje alaska department of fish and game, 1300 college road, fairbanks, ak 99701-1599, usa, e-mail: don.young@alaska.gov abstract: during 1996–1999, hunters killed an estimated 24–30% of the pre-hunt bull moose (alces alces) in game management unit 20a. as a result, the 1999 post-hunt bull:cow ratios declined to 24:100, well below the management objective of 30:100. during 2000 and 2001 we shortened the hunting season from 25 to 20 days to reduce the harvest of bull moose, but kill rates of bulls remained high (23–27%) and ratios remained unacceptably low (22–26 bulls:100 cows). subsequently, to recover bull:cow ratios to 30:100, hunters were restricted unit-wide to taking bulls with 1) spike-fork antlers, 2) antlers ≥50 inches wide, or 3) ≥3 brow tines on ≥1 antler. these restrictions were in place from 20022007, but results occurred rapidly. after only 2 years of antler restrictions, hunters killed an average of 36% fewer bulls compared with the previous 2-year average harvest rate ( x = 715 during 2000–2001 and 455 during 2002–2003). comparing these same 2-year periods, average kill rates of bulls declined from 25% to 12% of the pre-hunt bull population, average number of hunters declined 24% (1,568 to 1,187), and the average hunter success rate declined from 34% to 29%. bull:cow ratios increased from 26:100 to 32:100 after 2 years of antler restrictions. with an additional 2 years (2004–2005) of antler restrictions and high harvest of cow moose, bull:cow ratios reached 38:100. modeling indicated that the bull:cow ratio would have stabilized at 33:100 without the high harvest of cows. the recovery of bull:cow ratios to our objective of 30:100 with 2 years of antler restrictions allowed 1) bull seasons to be lengthened from 20 to 25 days beginning in 2004 and, 2) a limited number of drawing permits for any bull during 2006–2007. elsewhere, similar selective harvest strategies should also allow recovery of bull:cow ratios, unless the total kill rate of bulls is higher than estimated here. alces vol. 44: 65-71 (2008) key words: alaska, alces alces, antler restrictions, bull:cow ratios, game management unit 20a, recovery, selective harvest strategy. historically, low bull:cow ratios (10:100) of moose (alces alces) resulted where hunter access was good, and hunting was restricted largely to any bull (rausch et al. 1974). low bull:cow ratios did not occur in remote inaccessible areas (60–80:100, gardner 2002), so we infer that humans, not wolves (canis lupus) and bears (ursus americanus and ursus arctos), caused the skewed ratios by selectively killing bull rather than cow moose. to recover low bull:cow ratios, selective harvest strategies (shs) were first implemented in british columbia in 1980 (child 1983, child and aitken 1989). these shs were based on limiting hunters to particular bulls with regulations on antler architecture. in alaska, shs were first initiated on the kenai peninsula in 1987 (schwartz et al. 1992). selective harvest strategies spread rapidly to other alaska roadside areas during 1988–1993 because of low bull:cow ratios. this is the first paper that describes successful recovery of bull:cow ratios in interior alaska. high kill rates of bull moose by hunters in game management unit 20a (unit 20a) during 1996–1999 resulted in post-hunt bull:cow ratios declining below the management objective of 30:100. during 2000–2001 recovery of bull:cow ratios – young and boertje alces vol. 44, 2008 66 we shortened the hunting season to reduce the harvest of bull moose, but bull:cow ratios remained low. in 2002, hunters were restricted unit-wide to taking bulls with 1) spike-fork antlers, 2) antlers ≥50 inches wide, or 3) ≥3 brow tines on ≥1 antler. we report the effects of these antler-based shs on 1) total kill of bull moose by hunters, 2) kill rate of pre-hunt bull moose numbers, 3) hunter participation, 4) hunter success rate, 5) antler size of harvested bulls, and 6) bull:cow ratios. study area our study area encompassed unit 20a immediately south of fairbanks and across the tanana river. the study area is in interior alaska and is centered on 64°10′n latitude and 147°45′w longitude. the study area encompasses 17,601 km2, but only 13,044 km2 contains topography and vegetation used characteristically by moose. gasaway et al. (1983), boertje et al. (1996), and keech et al. (2000) described the physiography, habitat, climate, and factors limiting moose during 1963–1997. young and boertje (2004) described hunter access, moose seasons, and bag limits from the 1960s through the early 2000s, moose population status from 1997-2003, and the use of calf hunts to increase yield. young et al. (2006) described impediments and achievements of managing unit 20a moose for elevated yield. moose in unit 20a (1997–2005) exhibited the lowest nutritional status documented for noninsular, wild moose in north america (boertje et al. 2007). boertje et al. (2009) described how predation and reproduction affected the harvest of moose during 1996–2007. methods estimating harvest and hunter statistics we monitored reported moose harvest, hunter participation, hunter success rates, and antler characteristics of the harvest using a mandatory harvest report card system (schwartz et al. 1992). to estimate the number of moose killed by hunters, we multiplied the reported harvest by 1.34 (47/35) based on the reported harvest of 35 radio-collared moose and an additional 12 radio-collared moose that died from unreported harvest (6) and wounding loss (6) in unit 20a during 1996–2006 (boertje et al. 2009). we focused this study on the kill of bull moose; bulls included all males ≥15 months of age. we derived the pre-hunt bull kill rate based on pre-hunt bull numbers (gasaway et al. 1992). we estimated the prehunt bull numbers each year by adding the total estimated kill of bulls by hunters, which occurred in september, to the aerial estimates of november bull numbers. we derived hunter success rate using the reported harvest. we primarily compared results from the 2 years prior to shs (2000 and 2001) versus the first 2 years of shs (2002–2003) because those 4 years had 20-day hunting seasons, whereas all other years (1996–1999 and 2004–2005) had 25-day seasons. we used student’s t-tests to compare total kill, kill rate, hunter participation, hunter success rate among years, and antler size in the harvest. we assumed significance at α = 0.05. estimating moose population characteristics we flew moose population estimation and composition surveys (1999–2005) each november, except in 2002 due to poor survey conditions. to calculate moose numbers, we used spatial statistics (kellie and delong 2006, ver hoef 2008) and a sightability correction factor of 1.21 (boertje et al. 2009). to estimate the finite annual population growth rate (λ), we fitted population estimates during 1996–2004 with a trendline through parametric empirical bayes estimates (ver hoef 1996:1048). we used the student’s ttest to compare mean bull:cow ratios prior to and after the initiation of shs. we assumed significance at α = 0.05. alces vol. 44, 2008 young and boertje – recovery of bull:cow ratios 67 modeling to assess effect of cow moose hunts on bull:cow ratios we used a deterministic model using microsoft® office excel 2003® software to compare bull:cow ratios with and without harvests of cow moose. cows included all females ≥15 months of age. input variables included annual estimates of the number of bulls and cows killed by hunters, annual survival rates of bulls and cows, and calves:100 cows observed during november surveys. also, during the initial year we input the number of bulls, cows, and calves. we backdated the model to the year 1991 and varied survival rates until output values were similar to estimates obtained from annual surveys. output values included annual post-hunt moose numbers, bull:cow ratios, and λ. in the initial model we used empirical estimates of 98, 161, 698, and 794 cow moose killed by hunters during 2002–2005. we then ran the same model using a simulated adult kill of 10 cows (i.e., estimated illegal kill) during 2002–2005. the result was paired bull:cow ratios with and without harvests of cows during the 4 years of unit-wide shs. results harvest statistics the mean estimated kill of bull moose by hunters declined 36% after shs were initiated (table 1). specifically, the mean kill of 455 bulls taken during 2002–2003 was lower (p = 0.003, df = 2) than the mean kill of 715 bulls taken during the 2 years prior to shs (2000–2001). mean kill of bulls by hunters was not lower during 2004–2005 ( x = 577, p = 0.079, df = 2) when the shs season was lengthened from 20 to 25 days. the mean estimated kill rate of 12% of the pre-hunt bull population during 2002–2003 was lower (p = 0.029, df = 2) than the 2-year mean 25% kill rate observed prior to shs (table 1). kill rates remained lower ( x = 14%, p = 0.026, df = 2) after the shs season was lengthened from 20 to 25 days. hunter participation and success the mean reported number of hunters declined (24%, p = 0.002, df = 2) after shs were initiated (table 1). mean reported hunter success rates also declined (p = 0.034, df = 2) from 34% to 29% after shs were initiated. success rates remained lower (x = 25%, p = 0.004, df = 2) after the shs season was lengthened from 20 to 25 days. antler size of harvested bulls we observed a shift away from 30to 40-inch antlered bulls and toward ≥50-inch antlered bulls in the harvest after shs were implemented (table 1). the mean proportion of 30to 40-inch antlered bulls declined (p = 0.019, df = 2) in the harvest from 30% to 6%. conversely, the mean proportion of ≥50-inch antlered bulls increased (p = 0.023, df = 2) from 26% to 49% in the harvest. these trends continued into 2004–2005 when the shs season was extended 5 days (30to 40-inch antlered bulls: x = 7%, p = 0.019, df = 2; ≥50-inch antlered bulls: x = 52%, p = 0.009, df = 2). we observed no change in the proportions of <30-inch ( x = 22% vs. 23%, p = 0.607, df = 2) or 40to 50-inch antlered bulls (x = 21% vs. 19%, p = 0.417, df = 2) in the harvest. bull:cow ratio of the population the mean bull:cow ratio increased (p = 0.006, df = 4) by 11 bulls:100 cows after shs were implemented when comparing data from 1999–2001 (x = 24) versus 2003–2005 (x = 35, table 1). however, shs alone did not increase the bull:cow ratio because high harvests of cows during 2004–2005 also contributed to the increase in the bull:cow ratio (fig. 1). modeling indicated that without the harvest of cows, the bull:100 cow ratio would have increased to only 33 in 2005 rather than the 38 observed. regardless, when comparing the observed bull:cow ratios pre-shs (x =24) versus the simulated bull:cow ratios without the harvests of cows (x = 33, recovery of bull:cow ratios – young and boertje alces vol. 44, 2008 68 r ep or te d ha rv es t r at e (% ) b y an tle r w id th (i nc he s) y ea r se as on le ng th / b ag lim ita n o. b ul ls : 10 0 co w s 90 % c i pr ehu nt b ul l po pu la tio nb r ep or te d ha rv es t to ta l k ill o f bu lls c pr ehu nt k ill ra te (% ) o f bu lls r ep or te d no . of h un te rs r ep or te d su cc es s ra te d (% ) <3 0 30 –4 0 40 –5 0 50 + u nk 19 96 25 d ay /a ny bu ll 39 3, 28 9 60 4 80 9 24 .6 1, 63 6 36 .9 23 29 21 25 3 19 97 25 d ay /a ny bu ll 33 3, 45 4 62 0 83 1 24 1, 59 5 38 .9 22 28 25 22 4 19 98 25 d ay /a ny bu ll 31 2, 99 5 60 8 81 5 27 .2 1, 65 2 36 .8 17 28 23 27 5 19 99 25 d ay /a ny bu ll 24 17 –3 1 3, 01 0 66 9 89 7 29 .8 1, 57 4 42 .5 19 24 24 30 3 20 00 20 d ay /a ny bu ll 22 15 –2 9 2, 66 8 53 4 71 5 26 .8 1, 58 4 33 .7 23 27 22 26 2 20 01 20 d ay /a ny bu ll 26 18 –3 4 3, 06 9 53 4 71 5 23 .3 1, 55 1 34 .4 21 33 20 26 1 20 02 20 d ay sf/ 50 –e –e 3, 41 4f 35 0 46 9 13 .7 1, 18 5 29 .5 25 5 17 52 1 20 03 20 d ay sf/ 50 32 25 –3 9 3, 97 5 32 8 44 0 11 .1 1, 18 9 27 .6 21 7 21 45 6 20 04 25 d ay sf/ 50 35 27 –4 3 3, 90 1 40 0 53 6 13 .7 1, 63 8 24 .4 16 7 25 49 3 20 05 25 d ay sf/ 50 38 33 –4 2 4, 25 2 46 1 61 8 14 .5 1, 81 6 25 .4 11 6 27 54 3 a “ a ny b ul l” b ag li m it oc cu rr ed in 1 4, 32 5 km 2 ( 81 % ) o f t he 1 7, 60 1 km 2 i n g am e m an ag em en t u ni t 2 0a . w e de si gn at ed u ni tw id e se le ct iv e ha rv es t s tr at egi es a s “s -f /5 0” , w hi ch in cl ud ed re st ri ct in g ha rv es t t o bu lls w ith 1 ) s pi ke -f or k an tle rs , 2 ) a nt le rs ≥ 50 in ch es w id e, o r 3 ) ≥ 3 br ow ti ne s on ≥ 1 an tle r. b p re -h un t b ul l p op ul at io n eq ua le d nu m be r o f b ul ls e st im at ed fr om n ov em be r s ur ve ys p lu s es tim at ed to ta l k ill o f b ul ls b y hu nt er s. c t ot al k ill o f b ul ls e qu al ed re po rt ed h ar ve st o f b ul ls x 1 .3 4, b as ed o n a ra di oco lla re d sa m pl e of m oo se k ill ed b y hu nt er s (4 7/ 35 ). d r ep or te d su cc es s ra te e qu al ed re po rt ed h ar ve st d iv id ed b y re po rt ed n um be r o f h un te rs . e n o da ta . f d at a on p os thu nt b ul l n um be rs w er e in te rp ol at ed fr om a dj ac en t y ea rs . ta bl e 1. c ha ra ct er is tic s of th e bu ll m oo se (≥ 15 m on th s of a ge ) p op ul at io n, re po rt ed h ar ve st a nd k ill o f b ul ls b y hu nt er s, a nd h un tin g re gu la tio ns fo r b ul ls in g am e m an ag em en t u ni t 2 0a , i nt er io r a la sk a, 1 99 6– 20 05 . alces vol. 44, 2008 young and boertje – recovery of bull:cow ratios 69 fig. 1), bull:cow ratios remained different (p = 0.002, df = 4). discussion using antler-based shs, bull:cow ratios increased from about 26 bulls:100 cows to above our objective of 30 bulls:100 cows in 2 years, 2002–2003. we documented that shs reduced kill and pre-hunt kill rates of bulls, hunter participation, and hunter success rates. after the initiation of shs, bull:cow ratios increased to a 3-year average of 35 bulls:100 cows, but the increase above 33 bulls:100 cows apparently resulted from liberal harvests of cows during the last 2 years (fig. 1). when comparing 3-year averages preand post-shs, bull:cow ratios increased from 24 to 35 with ratios as high as 38. with shs on the kenai peninsula, average bull:cow ratios increased from 16 to 25 with ratios as high as 29 (schwartz et al. 1992). schwartz et al. (1992) reported a 27% decline in the mean harvest of bull moose when comparing 5-year periods pre(x = 636) and post-shs (x = 466); whereas, we observed a 36% decline in the mean harvest of bulls when comparing 2-year periods pre(2000–2001) and post-shs (2002–2003). after shs were implemented on the kenai peninsula, the distribution of the harvest of bulls shifted toward yearlings and away from 30to 40-inch antler class with little change in the proportion of bulls ≥50-inches (schwartz et al. 1992). we also observed a shift away from bulls in the 30to 40-inch antler class. however, we also observed a shift toward bulls with ≥50-inch antlers, presumably because our higher bull:cow ratio had resulted in a larger proportion of large antlered bulls in the population. unlike the kenai study, we observed little change in the proportion of ≤30-inch bulls (i.e., yearlings). we speculate that hunters had difficulty identifying yearling bulls because of poor antler development in unit 20a. for example, during a flight prior to the hunt in 2007, we observed 6 (22%) of 27 known-age, radio-collared yearlings with antlers consisting of small spikes (2–8 cm). poor antler development among yearlings was likely related to the low nutritional status of this population (boertje et al. 2007). the rapid recovery of bull:cow ratios allowed us to elevate the harvest of bulls after 2 years. recovery of bull:cow ratios to 30:100 was a secondary management objective during this study. the primary objective was to improve nutritional status of moose by encouraging greater harvest (boertje et al. 2007). therefore, during 2004–2007, when the post-hunt bull:cow ratios exceeded our management objective of 30 bulls:100 cows, we encouraged greater harvest of bulls using 2 methods. first, 5 days were added to the hunting season with shs during 2004–2005. subsequently, we issued a limited number of drawing permits for any bull during 2006–2007 (schwartz et al. 1992). we recognize that increasing harvest of cows, not bulls, was the most effective method to decrease population size, but hunts targeting cows were difficult to implement and far more controversial than hunts targeting bulls (young and boertje 2004, young et al. 2006). we know that low bull:cow ratios result from hunters favoring bulls versus cows, based on data from favored roadside hunting areas (10 bulls:100 cows, rausch et al. 1974). in contrast, in remote, lightly hunted areas, we observed 60–80 bulls:100 cows. 29.5 33.0 29.3 32.3 33.3 33.5 36.1 38.0 20 25 30 35 40 2002 2003 2004 2005 b ul ls :1 00 c ow s with cow harvests without cow harvests fig. 1. estimated post-hunt bull:cow ratios with and without cow moose harvests, game management unit 20a, interior alaska, 2002–2005. recovery of bull:cow ratios – young and boertje alces vol. 44, 2008 70 we demonstrate that shs were a key factor allowing bull:cow ratios to increase from 26:100 to 32:100 in 2 years. bull:cow ratios should increase elsewhere where similar shs are implemented, unless kill rates of bulls are higher than estimated here. acknowledgements many individuals contributed to the work reported in this manuscript and we thank them. those deserving special thanks include skilled pilots, generous volunteers, and dedicated alaska department of fish and game (adf&g) wildlife biologists and technicians that flew many arduous hours conducting moose survey flights. we particularly wish to thank adf&g publications technician laura mccarthy for her timely and much appreciated technical assistance. this study was funded by adf&g and federal aid in wildlife restoration. references boertje, r. d., m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73(3): in press. _____, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. _____, p. valkenburg, and m. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal of wildlife management 60: 474–489. child, k. n. 1983. selective harvest of moose in omineca: some preliminary results. alces 19: 162–177. _____, and d. a. aitken. 1989. selective harvests, hunters and moose in central british columbia. alces 25: 81–97. gardner, c. l. 2002. unit 20e moose. pages 406–429 in c. healy, editor. moose management report of survey and inventory activities 1 july 1999–30 june 2001. project 1.0. juneau. alaska, usa. (accessed february 2008). gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120: 1–59. _____, r. o stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84: 1-50. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450–462. kellie, k. a., and r. a. delong. 2006. geospatial survey operations manual. alaska department of fish and game, fairbanks, alaska, usa. rausch, r. a., r. j. somerville, and r. h. bishop. 1974. moose management in alaska. naturaliste canadien 101: 705–721. schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula. alces 28: 1–13. ver hoef, j. m. 1996. parametric empirical bayes methods for ecological applications. ecological applications 6: 1047–1055. _____. 2008. spatial methods for plotbased sampling of wildlife populations. environmental and ecological statistics 15: 3–13. young, d. d., jr., and r. d. boertje. 2004. initial use of moose calf hunts to increase alces vol. 44, 2008 young and boertje – recovery of bull:cow ratios 71 yield, alaska. alces 40: 1–6. _____, _____, c. t. seaton, and k. a. kellie. 2006. intensive management of moose at high density: impediments, achievements, and recommendations. alces 42: 41–48. 44_front_cover v2.pdf 141 distinguished moose biologist award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public's attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (> alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual's interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3-8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special "distinguished moose biologist" presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http: //bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca instructions for contributors to alces sentence, in which case it is spelled out. italics should only be used in the text for scientific names and statistical symbols. use the name-and-year system to cite published literature. cite references chronologically in the text. references – use large and small capitals for author’s last names and initials. do not use any abbreviations in the references. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. titles and all parts of tables must be typed doublespaced. tables must be constructed to fit the width of the page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use numerical superscripts to identify footnotes or asterisks for probabilities. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table columns must be generated with tab settings or a table editor. do not use spaces (i.e., the space bar). illustrations type figure captions on a separate page. identify each illustration by printing the author’s name and the figure number on the back in soft pencil. if necessary, also indicate the orientation of the illustration on the back. each illustration (either a photograph or linedrawn figure), must be of professional graphics quality, and reduced to fit into the area of either 1 (67 mm) or 2 (138 mm) columns of text by the author(s). letters and numbers on reduced figures must remain legible and be no less than 1.5 mm high after reduction. the same size and font of lettering should be used for all figures in the manuscript. photographs must be of high contrast and printed with a matte finish. typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original electronic graphics files or bitmap images (preferably as tagged image file format files) in an ibm-compatible format on 9-cm (3.5-inch) diskette or cd-rom. send manuscripts to: gerald redmond, submissions editor maritime college of forest technology hugh john flemming forestry centre 1350 regent street fredericton, new brunswick canada e3c 2g6 e-mail: gredmond@mcft.ca telephone: (506) 458 5128 fax: (506) 458 0652 editorial policy alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented at the annual north american moose conference and workshop, but works may be submitted directly to the editors at any time. reviewers judge submitted manuscripts on data originality, ideas, analyses, interpretation, accuracy, conciseness, clarity, appropriate subject matter, and on their contribution to existing knowledge. page charges current policies and charges are explained in a covering letter and invoice sent to authors with galley proofs. manuscript preparation authors should follow “manuscript guidelines for contributors to alces”, by rodgers et al. appearing in alces, vol. 34 (1): 1998 (available from the co-editors and associate editors). updates are posted on the alces web page; http: //bolt.lakeheadu.ca/~alceswww/alces.html. copy – please provide an electronic copy of the manuscript in ms word to the submissions editor. this copy should maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. double-space and leftjustify all text. except for the first page, number all pages consecutively, including tables and figure captions. revisions should be handled similarly. corresponding author do not use a title page. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlespaced in the upper left corner of the first page. if available, the author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (<10 words) begins left justified on the next line. type the title in upper-case bold letters. do not use abbreviations or scientific names in the title. abstract & key words following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. do not use abbreviations or literature citations. type alces vol. 00: 000 000 (0000), right justified on the line following the abstract. after leaving a single blank line, provide 6-12 key words in alphabetical order. footnotes use only in tables and at the bottom of the first page to provide the present address of an author when it differs from the address at the time of the study. style accompany the first mention of a common name with its scientific name. do not use scientific names for the names of domesticated animals or cultivated plants. use système international d’unités (si) units and symbols. use digits for numbers unless the number is the first word of a 207 editorial review committee our thanks to the following individuals who served as referees for alces volume 46. each paper was reviewed by at least 2 referees who judged its appropriateness for publication and provided editorial assistance. edward addison ecolink science, aurora, on cedric alexander vermont fish and wildlife, st. johnsbury, vt warren ballard texas tech university, lubbock, tx karen beazley dalhousie university, halifax, ns alan bisset ontario ministry of natural resources (retired), kenora, on james bridgland parks canada, ingonish beach, ns hughie broders saint mary’s university, halifax, ns karen clyde yukon fish & wildlife management board, whitehorse, yt terry creekmore wyoming game and fish department , laramie, wy andy edwards 1854 treaty authority, duluth, mn william faber central lakes college, brainerd, mn mike gillingham university of northern british columbia, prince george, bc ian hatter ministry of environment, victoria, bc kris hundertmark university of alaska fairbanks, ak nic larter department of environment & natural resources, fort simpson, nt martha minchak minnesota department of natural resources, duluth, mn brent patterson ontario ministry of natural resources, peterborough, on peter pekins university of new hampshire, durham, nh bill peterson minnesota department of natural resources (retired), grand marais, mn rolf peterson michigan technological university, houghton, mi margo pybus alberta sustainable development, edmonton, ab roy rea university of northern british columbia, prince george, bc william samuel university of alberta (retired), edmonton, ab david scarpitti mass. division of fisheries and wildlife, westborough, ma tim timmermann ontario ministry of natural resources (retired), thunder bay, on alces 44, 2008 a journal devoted to the biology and management of moose edward m. addison ecolink science vince f. j. crichton manitoba conservation murray w. lankester lakehead university (retired) brian e. mclaren lakehead university printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 kristine m. rines new hampshire fish and game edmund s. telfer canadian wildlife service richard m. p. ward yukon department of renewable resources associate editors chief editor peter j. pekins university of new hampshire submissions editor gerald w. redmond maritime college of forest technology business editor arthur r. rodgers ontario ministry of natural resources 136 distinguished moose biologist past recipients 1998 peter a. jordan, university of minnesota, st. paul, minnesota. 1997 margareta stéen, swedish university of agricultural sciences, uppsala, sweden. 1996 vic van ballenberghe, u.s. forest service, anchorage, alaska. 1995 not presented 1994 james m. peek, university of idaho, moscow, idaho. 1993 murray w. lankester, lakehead university, thunder bay, ontario. 1992 not presented 1991 charles c. schwartz, alaska dept. of fish and game, soldotna, alaska. 1990 rolf peterson, michigan technological university, houghton, michigan. 1989 warren b. ballard, alaska dept. of fish and game, nome, alaska. 1988 vince f. j. crichton, manitoba dept. of natural resources, winnipeg manitoba. and michel crête, ministère du loisir, de la chasse et de la péche, service de la faune terrestre, québec, pq. 1987 w. c. (bill) gasaway, alaska dept. of fish and game, fairbanks, alaska. 2011 kjell danell swedish university for agricultural studies uppsala, sweden 2010 michael w. schrage fond du lac resource management division, cloquet, minnesota. 2009 kenneth n. child, prince george, british columbia. 2007 kris j. hundertmark, university of alaska fairbanks, fairbanks, alaska. 2006 kristine m. rines, new hampshire fish and game department, new hampton, new hampshire. 2005 w. m. (bill) samuel, university of alberta, edmonton, alberta. 2004 w. eugene mercer, wildlife division, st. john's, newfoundland. 2003 arthur r. rodgers, ontario ministry of natural resources, thunder bay, ontario. 2002 bernt-erik sæther, norwegian university of science and technology, trondheim, norway. 2001 r. terry bowyer, university of alaska, fairbanks, alaska. 2000 gerry m. lynch, alberta environmental protection, edmonton, alberta. 1999 william j. peterson, minnesota department of natural resources, grand marais, minnesota. 137 1986 h. r. (tim) timmermann, ontario ministry of natural resources, thunder bay, ontario. 1985 ralph ritcey, fish and wildlife branch, kamloops, british columbia. 1984 edmund telfer, canadian wildlife service, edmonton, alberta. 1983 albert w. franzmann, alaska division of fish and game, soldotna, alaska. 1982 a. (tony) bubenik, ontario ministry of natural resources, maple, ontario. 1981 patrick d. karns, minnesota division of fish and wildlife, grand rapids, minnesota. and al elsey, ontario ministry of natural resources, thunder bay, ontario. in 1974, prior to the establishment of the distinguished moose biologist award, the group recognized the pioneering moose research of the late laurits (larry) krefting, u.s. fish and wildlife service, with an individual award. alces37(1)_123.pdf p1-8_4113.pdf alces vol. 41, 2005 poole and stuart-smith moose winter habitat 1 fine-scale winter habitat selection by moose in interior montane forests kim g. poole1 and kari stuart-smith2 1aurora wildlife research, 2305 annable rd., nelson, bc, canada v1l 6k4, e-mail: kpoole@ aurorawildlife.com; 2tembec inc., british columbia division, p.o. box 4600, cranbrook, bc, canada v1c 4j7, e-mail: kari.stuart-smith@tembec.com alces alces salix cornus stolonifera amelanchier alnifolia key words: alces alces alces alces moose winter habitat – poole and stuart-smith alces vol. 41, 2005 2 study area 2 pseudotsuga menziesii larix occidentalis pinus contorta populus tremuloides tsuga heterophylla thuja plicata populus balsamifera betula papyrifera picea glauca alces vol. 41, 2005 poole and stuart-smith moose winter habitat p. engelmannii abies lasiocarpa methods fieldwork n moose winter habitat – poole and stuart-smith alces vol. 41, 2005 4 statistical analyses n r r2 p results salix cornus stolonifera amelanchier alnifolia betula alnus shepherdia canadensis 0 10 20 30 40 50 60 70 80 w ill ow sp p. ro do gw oo d sa sk at oo n bi rc h as pe n pi ne su ba lp in e fir al de r s pp . so op ol al lie o th er p ro p o rt io n o f to ta l (% ) spillimacheen upper elk flathead n n n alces vol. 41, 2005 poole and stuart-smith moose winter habitat 5 r p r p r2 2 p r2 2 p r2 2 p r2 2 p r2 2 p p 0.002 0.024 0.004 n n n p moose winter habitat – poole and stuart-smith alces vol. 41, 2005 discussion management implications alces vol. 41, 2005 poole and stuart-smith moose winter habitat acknowledgements references balsom, s., w. b. ballard, whitlaw 140. boyce, m. s cervus elaphus braumandl urran burnam nderson. 2002. coady, eon huggard hundertmark berhardt r. e. ball elsall langley mackie ac amlin l. dusek moose winter habitat – poole and stuart-smith alces vol. 41, 2005 matchett meidinger ojar peek in pierce eek poole tuart-smith renecker chwartz in sas institute schwartz in a.w. sweanor andegren tabachnik idell telfer thompson tewart in ukelich tyers van dyke welsh orrison swald, homas westworth, d., l. brusnyk, j. roberts h. veldhuzien << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice alces vol. 44, 2008 butler et al. – grain overload effects on moose 73 grain overload and secondary effects as potential mortality factors of moose in north dakota erika a. butler1, william f. jensen1, roger e. johnson2, and jason m. scott3 1north dakota game and fish department, 100 north bismarck expressway, bismarck, nd, united states 58501; 2north dakota game and fish department, 7928 45th street ne, devils lake, nd, united states 58301; 3north dakota game and fish department, 2305 elm street, fargo, nd, united states 58102 abstract: the intent of this article is to alert biologists of a potential mortality factor of moose in agricultural areas. it has long been recognized that ruminants switching from a natural diet of browse (a cellulose-based diet) to one of more readily digestible carbohydrates (a starch-based diet), such as corn and wheat, are predisposed to developing conditions such as enterotoxemia, polioencephalomalacia, acute rumenitis, liver abscesses, laminitis, and to sudden death. these are often secondary to grain overload (acute acidosis) and are frequently documented in cattle and sheep which are moved from pasture to feedlot. necropsies of 4 moose in north dakota were not entirely conclusive, but suggested that grain overload occurred and was a cause of mortality. necropsy findings that supported grain overload as a contributing factor to death included acute rumenitis, isolation of clostridium perfringens coupled with hemorrhagic enteritis, chronic laminitis, and polioencephalomalacia. four likely scenarios exist in which grain overload occurs in north dakota moose including consumption of planted crops such as corn and wheat, access to bait piles mainly intended for deer, access to cattle feeding sites, and access to recreational feeding sites. these findings have important implications for the regulation of baiting and recreational feeding practices in north dakota and elsewhere in moose range of similar situation. alces vol. 44: 73-79 (2008) key words: agriculture, alces alces, feeding, grain overload, moose, mortality, rumenitis. moose (alces alces) in north dakota have traditionally occupied forested areas in the northeast and north-central portions of the state. moose were first recorded during white-tailed deer (odocoileus virginianus) surveys in the winter of 1969-70. in 1979 the first modern moose hunting season allowed the harvest of 15 moose in cavalier, pembina, and walsh counties. recently, based upon winter aerial survey data for white-tailed deer, moose have expanded their range south and west into agricultural areas. areas of the state open to moose hunting include the eastern one-third of the state south to the south dakota border, and much of the northern one-third of the state. in 2008, 141 moose licenses were issued in north dakota. the red river valley (glacial lake agassiz plain) was historically tall grass prairie. due to the high productivity of the soil, flat terrain, and ease of conversion to agricultural use, <2% of the surface area remains as native prairie (jensen 2001). common crops in these areas include corn, wheat, sugar beets, potatoes, barley, beans, sunflowers, and soybeans (fig. 1). reports of moose feeding in these agricultural fields, both planted and plowed, have become increasingly common over time. long recognized as a concern with domestic ruminants, grain overload (i.e., rumen overload, carbohydrate overload, acute overeating, ruminal acidosis) results from ingestion of toxic amounts of highly fermentable and readily digestible carbohydrates such as grain. this frequently occurs when ruminants abruptly switch from a diet of natural browse (a cellulose-based diet) to grain overload effects on moose – butler et al. alces vol. 44, 2008 74 one of readily digestible carbohydrates (a starch-based diet) causing a change in the rumen microbial population; the number of gram positive bacteria increase markedly while gram negative bacteria decline. the gram positive bacteria produce excessive lactic acid which lowers the rumen ph to <5. this increase in acidity destroys protozoa, cellulolytic organisms, and lactate-utilizing organisms all of which normally inhabit the rumen, and impairs rumen motility. clinical signs of acute overload include indigestion, rumen stasis, acute ruminitis and acidosis, dehydration, toxemia, incoordination, collapse and recumbancy, and frequently death. if the animal survives the initial episode of grain overload, secondary effects include, but are not limited to, enterotoxemia, polioencephalomalacia, liver abscesses, and chronic laminitis (merck veterinary manual 2008). however, other potential causes for these conditions also exist and are presented in the discussion. ruminitis and rumen scarring have been reported in a supplementary fed captive deer herd (woolf and kradel 1977) and grain overload has been identified as a primary diagnosis in captive elk (cervus elaphus) of all ages submitted to pathology labs (woodbury et al. 2005). rumenitis was diagnosed in 30 of 108 free-ranging white-tailed deer examined in saskatchewan; rumenitis and rumen overload were determined to be the causes of death of 5 of these deer (wobeser and runge 1975). a polioencephalomalacia-like disease, possibly secondary to grain overload, has been reported in wild pronghorn (antilocapra americana) from saskatchewan (wobeser et al. 1983) and north dakota (w. jensen, north dakota game and fish department, pers. comm.), wild white-tailed deer in minnesota (kurtz and karns 1969) and south dakota (reed et al. 1976), and a wild mule deer (odocoileus hemionus) from south dakota (reed et al. 1976). no documentation of grain overload in moose was found, though moose might be especially vulnerable to this condition as their mean retention time for digestion is the longest documented for cervids, and is surpassed by only 1 member of the order artiodactyla, the asian water buffalo (bubalus bubalis; stevens 1998). methods reports of dead or sick moose throughout the state are generally investigated by the north dakota game and fish department (ndgf) wardens or biologists. when possible, either the entire carcass is submitted to the north fig. 1. corn and wheat production throughout the state of north dakota. compiled and published by the usda, national agricultural statistics service, north dakota field office (source: http://www.nass.usda.gov/statistics_by_state/ north_dakota/index.asp). alces vol. 44, 2008 butler et al. – grain overload effects on moose 75 dakota state university veterinary diagnostic laboratory (vdl) for necropsy, or samples of major organs and serum are collected for submission. this study is a retrospective review of necropsy reports during a period of time when the ndgf did not have a veterinarian on staff. the results of 4 necropsies performed from 1991-2006 suggested that grain overload and its secondary effects could have contributed to moose mortality in north dakota. each of the 4 cases discussed below had the entire carcass delivered to the vdl for necropsy. grain overload was not considered a differential diagnosis by ndgf employees until results of the necropsies were received. consequently, some pertinent information is lacking, and/or difficult to interpret from the available necropsy reports. further, the vdl may have believed that certain animals were from captive facilities. results the locations of the 4 moose mortalities were in areas of substantial corn and wheat production (fig. 1 and 2). all 4 cases died or exhibited clinical signs during the north dakota deer archery season which runs from september–january. case c was euthanized during the deer gun season which occurs for 17 days in november. no hunting-related injuries were documented in any case. in october 1991 the ndgf received multiple calls regarding case a, an adult cow moose east of sheldon. reports indicated that case a had extremely overgrown hooves (both the fore and hind limbs). on 21 october it was decided to euthanize case a as her travel had become extremely restricted. case a was in good body condition and was presented at necropsy with moderate postmortem change. it was noted that all 4 digits had severely overgrown claws. the hind and front claws were approximately 30 cm and 20 cm longer than normal, respectively. separation of the dorsal claw from the third phalanx was evident and proliferation of fibrous tissue was present. severe congestion was noted around the distal aspects of the third phalanx in both front digits. the rumen and reticulum contained large amounts of grain that appeared to be corn. hyperemia of the ruminal mucosa was noted. serology for both bluetongue and epizootic hemorrhagic disease virus was negative. a diagnosis of severe, chronic laminitis in all 4 feet was made. the pathologist commented that this lesion in domestic livestock is often associated with an episode of acute grain engorgement, most likely occuring several months earlier. other causes of chronic laminitis, however, could not be ruled out. on 6 december 2002 case b, a young bull moose, was found dead near kelso. the reporting party stated that the moose had been seen in that area for a few days. there were no apparent gunshot wounds. case b was in good body condition at necropsy. there were 1-2 liters of free serosanguinous fluid within the abdomen. the rumen contents included grain of unrecorded type and grass. several feet of the jejunum were flaccid and contained a thick bloody fluid and there was extensively petechiated hemorrhage in the mesentery in this area. histologically, sections of intestine contained multifocal congestion in mucosa and submucosa with occasional mucosal hemorrhage. mild lymphoplasmacytic inflammation was found in the mucosa. clostridium perfringens was isolated from the intestine. the pathologist indicated that these findings were consistent with a diagnosis of enterotoxemia. on 16 november 2004 case c, a bull moose, was reported recumbent and kicking in a field near gardner; the district warden was sent to investigate. the warden reported that he had seen this bull the previous day and it had appeared healthy, but when the moose was approached it was unable to rise. it was noted that a large amount of feces appearing to contain blood surrounded the moose. the reporting party claimed that a neighbor had been feeding the moose all summer. case grain overload effects on moose – butler et al. alces vol. 44, 2008 76 c was shot and transported to the vdl. at necropsy, case c was found to be in good body condition with moderate postmortem autolysis. the contents and gross appearance of the forestomachs and abomasum were unremarkable. a large mass of hemorrhage was present within the tissue at the base of the mesentery and the small intestine contained frank blood. no mucosal lesions were present and the colon appeared normal. congestion of the kidney and adrenal glands was noted. histological examination revealed multifocal intraepithelial neutrophilic aggregates (microabscesses) in the ruminal epithelium and diffuse autolysis of all sections of the intestines. fluorescent antibody examinations for bovine herpes virus 1, bovine viral diarrhea virus, bovine respiratory syncytial virus, and parainfluenza 3 virus were negative. bacteriology isolated escherichia coli, alpha streptococcus, and clostridium perfringens from the liver, and e. coli, alpha streptococcus, bacillus sp., and c. perfringens from the intestine. a diagnosis of hemorrhagic enteritis (with c. perfringens) and acute rumenitis was made. the acute rumenitis was suspected to be biochemical in origin, with grain overload a strong possibility. on 3 january 2006 case d, an adult female moose, was seen circling and stumbling with a pronounced head tilt in cass county, and was subsequently shot by a warden. case d was in fair body condition at necropsy. in addition to 2 live nematodes removed from the abdominal cavity, a massive intrameningeal blood clot was seen ventral to the brain at the level of the thalamus and midbrain. the intramenineal hemorrhages (not related to euthanasia) was associated with focally extensive polioencephalomalacia. on histopathology, a fig. 2. locations of moose mortalities in north dakota which were possibly due to grain overload or its secondary effects. alces vol. 44, 2008 butler et al. – grain overload effects on moose 77 single area of the meninges showed a massive blood clot, while the subjacent gray matter was markedly rarefied. discussion laminitis, such as the severe chronic case documented in case a, is a common chronic sequel to previous grain overload in both domestic ruminants and horses. less commonly, it can be secondary to postparturient metritis, endotoxemia, colic, enteritis (merck veterinary manual 2008), copper deficiency (flynn et al. 1977), and epizootic hemorrhagic disease (prestwood et al. 1974). while serology for both bluetongue and epizootic hemorrhagic disease was negative, this animal did not have its copper levels tested. the large amount of corn present in the ingesta proves that large quantities were available to this animal suggesting that grain overload could have initiated laminitis several months earlier, however, other causes of laminitis cannot be ruled out. the gross and histological intestinal lesions coupled with the isolation of clostridium perfringins from the intestines of case b were highly suggestive of enterotoxemia. enterotoxemia is a feed-related condition that often occurs in late winter when animals have not had access to grain for an extended period of time or with high intake of lush, green grasses (merck veterinary manual 2008). there are multiple types of clostridium perfringens, however, it is generally clostridium perfringens types d and c that are responsible for enterotoxemia. pcr on the clostridium isolate to determine its type was not performed. grain overload is considered a common cause of enterotoxemia, though other less likely causes such as e. coli infection in swine have been identified (merck veterinary manual 2008). the presence of grain in the rumen indicates that this moose also had access to grain and supports the possibility that case b suffered from fatal enterotoxemia due to grain overload. the history of case c being observed by the warden as healthy in a field the day prior to its euthanasia, its clinical signs of recumbancy and kicking, and the postmortem findings of acute rumenitis and hemorrhagic enteritis are highly indicative of grain overload. like most cases of polioencephalomalacia, the pathology observed in case d was most likely due to a thiamine deficiency. thiamine deficiencies are often secondary to ruminal acidosis, as one of the microorganisms which proliferate with grain intake produces thiaminase ii, an enzyme which catalyzes the cleavage of thiamine (merch veterinary manual 2008). other causes for thiamine deficiency include water deprivation and hypernatremia (padovan 1980), diets high in sulfate salts (raisbeck 1982), acute lead poisoning (merck veterinary manual 2008), and thiaminase toxicity, caused by the ingestion of certain plants (evans et al. 1975). while the exact cause of case d’s polioencephalomalacia remains unknown, it is possible that it was due to grain overload. based on these 4 cases, it appears that grain overload may be a contributing mortality factor of moose in north dakota. numerous scenarios which could result in grain overload exist. the 4 most likely include the consumption of planted crops such as corn and wheat, access to bait piles mainly intended for deer, access to cattle feeding sites, and access to recreational game feeding sites. agricultural crops, especially corn and soybean, are common in the counties where the moose mortalities occurred, while cattle are less common than in other areas of the state (fig. 1). all 4 of these cases died in mid-fall or winter during the archery season. case c was also found during the deer-gun season and was an acute case of rumenitis. its euthanasia occurred in the middle of november when most crops had been harvested. this suggests that if grain overload was the cause of the rumenitis, the grain was most likely accessed at a baiting site or possibly a recreational feeding site. according to the ndgf district warden, reports of moose feeding with or harassing cattle grain overload effects on moose – butler et al. alces vol. 44, 2008 78 in this area are extremely uncommon. case a was an extremely chronic case of laminitis, indicating that if grain overload had occurred, it was quite some time prior to its euthanasia in october. when the possible grain overload occurred in cases b and d is unknown. polioencephalomalacia and enterotoxemia are sub-acute results of grain overload. however, few, if any crops, remained in the field at the time of their death indicating that bait or feed was the most likely source. the fact that deer hunting season was also open supports the possibility of availability and access to bait. these findings could have important implications for baiting and recreational feeding regulations in north dakota. if agricultural crops are contributing to grain overload and its secondary effects in moose in north dakota, its incidence could be expected to rise given the political and economic push for increased biofuel and ethanol production. the ndgf expects a fair amount of conservation reserve program land, highly erodible land taken out of crop production, to be converted to corn fields and for corn production to increase in existing fields as the value of this commodity continues to rise. based on these necropsy findings, grain overload and its secondary effects should be monitored in north dakota’s moose population, as well as in other agricultural areas or regions where baiting and recreational feeding are common. it is extremely difficult to attribute the mortality of free-ranging wildlife to grain overload. therefore, whenever possible, entire carcasses or appropriate samples, including fixed and fresh specimens of liver, rumen, abomasum, omasum, reticulum, small intestine, large intestine, and brain, and when possible, whole blood, serum, feces, and rumen contents, should be submitted to a diagnostic lab for investigation. acknowledgments we would like to thank the ndgf wardens and field staff for their assistance in handling and transporting moose and other wildlife to the vdl in fargo for necropsies. we thank the vdl for conducting these necropsies. we thank the editor and two anonymous reviewers for their comments and suggestions. finally, we thank the citizens of north dakota for reporting sick and dead wildlife in a timely manner so that we may learn the causes of these illnesses and diseases. references evans, w. g., i. a. evans, d.j. humphreys, b. lewis, w. e. j. davies, and r. f. e. axford. 1975. introduction of thiamine deficiency in sheep with lesions similar to those of cerebrocortical necrosis. journal of comparative pathology 85: 253-265. flynn, a., a.w. franzmann, p.d. arneson, and j.l. oldemeyer. 1977. indications of a copper deficiency in a subpopulation of alaskan moose. journal of nutrition 107: 1182-1189. jensen, w.f. 2001. lewis and clark in north dakota: wildlife then and now. north dakota outdoors june: 10-19. kurtz, h. j., and p. d. karns. 1969. polioencephalomalacia in a white-tailed deer (odocoileus virginianus borealis). veterinary pathology 6: 475-480. merck veterinary manual. april 10, 2008. whitehouse station, new jersey. (accessed june 2008). padovan, d. 1980. polioencephalomalacia associated with water deprivation in cattle. cornell veterinarian 70: 153-159. prestwood, a.k, t.p. kistner, f.e. kellog, and f.a. hayes. 1974. the 1971 outbreak of hemorrhagic disease among whitetailed deer of the southeastern united states. journal of wildlife diseases 10: 217-224. raisbeck, m. f. 1982. is polioencephalomalacia associated with high sulfate diets? journal of the american veterinary medical association 180: 1303-1304. alces vol. 44, 2008 butler et al. – grain overload effects on moose 79 reed, d. e. h. shave, m.e. bergeland, and c.e. gates. 1976. necropsy and laboratory findings in free-living deer in south dakota. journal of the american veterinary medical association 169:975-979. stevens, c. e. 1998. comparative physiology of the vertebrate digestive system. cambridge university press, cambridge, new york. usda, national agricultural statistics service, north dakota field office, fargo, north dakota. (accessed june 2008). wobeser, g., and w. runge. 1975. rumen overload and ruminitis in white-tailed deer. journal of wildlife management 39: 596-600. _______, p. y. daoust, and h.m. hunt. 1983. polioencephalomalacia-like disease in pronghorns (antilocapra americana). journal of wildlife diseases 19: 248252. woodbury, m. r., j. berezoski, and j. haigh. 2005. a retrospective study of the causes of morbidity and mortality in farmed elk (cervus elaphus). the canadian veterinary journal 46: 1108-1121. woolf, a., and d. kradel. 1977. occurrence of rumenitits in a supplementary fed white-tailed deer herd. journal of wildlife diseases 13: 281-285. alcessupp1_29.pdf 4014.p65 alces vol. 40, 2004 routledge and roese moose winter diet 95 moose winter diet selection in central ontario robert g. routledge1,2 and john roese1 1department of biology, lake superior state university, sault ste. marie, mi 49783, usa; 2ontario ministry of natural resources, 1235 queen street, sault ste. marie, on, canada p6a 2e5, e-mail: rob.routledge@mnr.gov.on.ca abstract: this paper documents moose (alces alces) winter diets in the northern portion of the great lakes-st. lawrence forest region of ontario. seventeen of 20 species available along 2,890m of moose foraging path were browsed. striped maple (acer pensylvanicum), eastern hemlock (tsuga canadensis), balsam fir (abies balsamea), and red maple (acer rubrum) comprised a combined 74 and 56 % of the browse dry weight consumed and available, respectively. moose used balsam fir, eastern hemlock, and red maple proportionally more than their availability, used sugar maple (acer saccharum), yellow birch (betula alleghaniensis), and mountain ash (sorbus americana) proportionally less than their availability, and used striped maple, beaked hazel (corylus cornuta), and mountain maple (acer spicatum) proportional to their availability. the important contribution of striped maple and eastern hemlock to moose diets contrast with other studies. these results may be used to assist in the evaluation of moose winter habitat. alces vol. 40: 95-101 (2004) key words: alces alces, diet selection, great lakes-st. lawrence forest, moose, northern hardwoods, ontario, winter food habits dietary information about moose (alces alces) during winter can be used to explain local distribution patterns (telfer 1978), to evaluate habitat suitability (crête and jordan 1982, crête 1989, crête and courtois 1997), and to provide information for habitat e n h a n c e m e n t ( p e e k e t a l . 1 9 7 6 , lautenschlager et al. 1997). moose winter food habits have been documented widely across their primary range in the boreal forest region (crête and bedard 1975, cumming 1987, crête 1989). composition of moose winter diets have also been reported in the southern extent of their range in eastern north america, represented by the transitional mixed coniferous-deciduous forest of the great lakes-st. lawrence (gl-sl) and acadian forest regions (rowe 1972), from maine (ludewig and bowyer 1985, lautenschlager et al. 1997) and southwestern quebec (joyal 1976, crête and jordan 1982) to northeastern minnesota (peek et al. 1976). the purpose of this paper is to report results from a survey of the winter food habits and browse preferences of moose in the northern portion of the gl-sl forest region of ontario. the unusually low snow depth for this area (< 20 cm; environment canada 1995) combined with a low moose density (0.18 per km2; scott jones, ontario ministry of natural resources, personal communication), provided an opportunity to determine moose winter browse preferences without the constraint of deep snow conditions in an area with minimal browsing activity in past years. study area a moose yarding area was located 19 km north of sault ste. marie, ontario (46°42'n, 84°24'w) in mid-january 1995. the area varied in elevation from 200 to 230m and was dominated by mature sugar maple (acer saccharum), with yellow birch moose winter diet – routledge and roese alces vol. 40, 2004 96 (betula alleghaniensis), red maple (acer r u b r u m ) , e a s t e r n h e m l o c k ( t s u g a canadensis), and white spruce (picea glauca) occurring at a lower density. mean daily temperatures during january and february 1995 were –6 and –12°c, respectively, ranging from –30 to +4°c (environment canada 1995). methods moose were tracked in the snow to determine their diet composition from midjanuary through mid-february 1995 by counting all browsed twigs along a minimum 50 m of foraging path at each foraging site (wetzel et al. 1975, histol and hjeljord 1993). tracks were intercepted in the yarding area by walking (> 800 m) in a southerly direction from a forest access road (thielman road). if multiple foraging sites were located along the same track, they were separated by > 400 m. all tracks surveyed were clearly identified as being made by one individual. browsed and unbrowsed twigs, between snow surface and 3m in height, were counted in 4m2 circular plots (radius = 1.13 m) located at 10m intervals along the foraging paths to measure browse availability ( ≥ 6 plots per path). diameters at points of browsing (dpb) and diameters of unbrowsed twigs (dut) were randomly measured for each plant species represented by browsed twigs. duts were measured at the proximal end of twigs > 5 cm in length where they forked from a larger branch (crête 1989). the portion of the twig distal to the dpb or dut, which could contain more than the current annual growth (cag), was converted to biomass using equations for regression lines relating fresh diameter (y) and dry weight (x) of twigs (telfer 1969, potvin 1981). twigs were collected in the study area representing the range of dpbs and duts for commonly browsed species. published equations were used for species minimally browsed (table 1). diet composition was estimated by converting twig counts along foraging paths to dry weight. browse availability was estimated by converting counts of browsed and unbrowsed twigs in the circular plots to dry weight and summing browsed and unbrowsed dry weight. bonferroni confidence intervals were used to determine which species were consumed relative to their availability (byers et al. 1984). a measure of browsing intensity was calculated for each species by dividing the biomass consumed into the biomass available at each foraging site for which a particular species was present. results diet composition was determined from 9,183 instances of use (iu) along 2,890 m of foraging path (mean ± sd, 206.4 ± 102.7 m) at 14 foraging sites. browse availability was determined from 289 4-m2 plots. regression equations relating fresh twig diameter and dry weight of twigs were computed for the 9 most important browse species (table 1). published equations were used for chokecherry (prunus virginianus), northern red oak (quercus rubra), paper birch (betula papyrifera), and serviceberry (amelanchier sp.) (table 1). due to low availability and use, the latter species were grouped together for analyses. eastern white cedar (thuja occidentalis), fly honeysuckle (lonicera canadensis), canada yew (taxus canadensis), and willow (salix spp.) were not included in the analyses because they were uncommon and browsed minimally. ash (fraxinus sp.), red elderberry (sambucus pubens), and white spruce were removed because they were not consumed. mean dpbs were significantly larger than duts for all species (t-tests; p < 0.01) except for mountain ash, where the dut was significantly larger (t2,60 = 4.57; p < 0.01), and for yellow birch where the alces vol. 40, 2004 routledge and roese moose winter diet 97 table 1. regressions 1 relating fresh twig diameter (mm) and twig dry weight (g) and mean duts and dpbs for common browse species. dut and dpb did not differ (t2,43 = 1.15; p > 0.05) (table 1). striped maple, eastern hemlock, balsam fir, and red maple comprised 73.7% of the browse dry weight consumed at 28.8, 17.1, 16.7, and 11.1%, respectively, and 55.7% of the browse dry weight available (table 2). sugar maple, striped maple, mountain maple, and balsam fir were present at 12 or more of the 14 foraging sites (table 2). an average of 7 browse species was present (range 3 – 11) over the 14 foraging sites. stem breakage was recorded frequently on striped maple and less commonly on balsam fir. balsam fir, red maple, and eastern hemlock were used 2 – 14% more than expected (p < 0.05); sugar maple, mountain ash, and yellow birch were used 4 – 9% less than expected (p < 0.05); and striped maple, 1 prediction equation: (ln weight) = a + b×(ln diameter); weight = a + b×diameter for serviceberry. 2 standard error of estimate divided by mean of dependent variable, twig diameter (zar 1996:328). 3 diameter at point of browsing (dpb) and diameter of unbrowsed twig (dut). dpb significantly different from dut for all species (t-tests; p < 0.01) except for yellow birch (p > 0.05). 4 based on air-dry twig diameter (telfer 1969). 5 crête (1989). 6 amelanchier sp.; equation also used for prunus virginianus (lyon 1970). beaked hazel, and mountain maple were browsed in proportion to their availability (p > 0.05) (table 2). average browsing intensities ranged as high as 68, 71, and 97% of the available biomass for balsam fir, red maple, and eastern hemlock, respectively, to as low as 6% for yellow birch (table 2). browsing intensities for eastern hemlock, red maple, balsam fir, striped maple, beaked hazel, mountain maple, and sugar maple were highly consistent (r > 0.98, p < 0.0001) from one foraging site to the next (table 2). at all seven sites where eastern hemlock was present, ≥ 93.3% of available biomass was removed regardless of its absolute (0.3 – 5.9 kg ha-1) or relative (1.8 – 13.8%) availability. fifty percent or less of available sugar maple browse, ranging from 0.5 – 7.9 species a b r2 see/y2 n dpb (sd)3 dut (sd) 3 balsam fir -3.33 4.30 0.641 0.179 56 1.81 (0.78) 1.30 (0.47) beaked hazel -4.12 3.64 0.677 0.205 60 2.69 (0.98) 1.55 (0.62) eastern hemlock -1.93 3.72 0.657 0.215 43 1.89 (0.81) 0.75 (0.27) mountain ash -7.04 4.86 0.614 0.065 80 3.76 (1.00) 5.35 (1.75) mountain maple -4.64 3.70 0.686 0.093 93 3.64 (0.95) 2.19 (0.67) red maple -4.09 3.04 0.866 0.084 96 3.57 (0.99) 2.46 (0.88) striped maple -4.97 3.94 0.639 0.090 100 4.10 (1.29) 2.56 (0.62) sugar maple -4.86 3.51 0.773 0.068 47 2.66 (0.67) 1.84 (0.62) yellow birch -2.92 2.64 0.808 0.148 83 2.31 (0.63) 2.03 (0.78) northern red oak 4 -3.00 2.80 0.988 20 paper birch 5 -2.66 2.21 0.750 99 serviceberry 6 -0.90 0.56 >0.750 50 moose winter diet – routledge and roese alces vol. 40, 2004 98 kg ha-1, was consumed in 13 of the 14 sites. discussion striped maple, a frequent understory component in mature hardwood stands representative of the study area (rowe 1972, rutkowski and stottlemeyer 1993), comprised nearly one-third of all browse consumed and had a relatively high mean browsing intensity. only trace amounts were available and consumed in southwestern quebec, an area also lying within the glsl forest region (joyal 1976, crête and jordan 1982). the large contribution of striped maple to moose diets may be partly attributed to the ability of moose to readily break striped maple stems to access twigs that would otherwise be out of their reach. balsam fir and mountain maple ranked third and fifth, respectively, in their contribution to moose diets, as was found in southwestern quebec (joyal 1976, crête and jordan 1982) and northeastern minnesota (peek et al. 1976). red maple was preferred during this study, in southwestern quebec (joyal 1976, crête and jordan 1982), and northeastern minnesota (peek et al. 1976). balsam fir and red maple were browsed much more intensively than mountain maple, indicating a greater preference. eastern hemlock had a low availability, yet ranked second in contributing to moose diets and was the most intensively browsed species. trace amounts of eastern hemlock have been reported in moose winter diets in maine (ludewig and bowyer 1985, lautenschlager et al. 1997) and algonquin park, ontario (peterson 1955). extensive winter browsing by white-tailed deer (odocoileus virginianus) on eastern hemlock is well documented (anderson and loucks 1979, frelich and lorimer 1985). 1 foraging site frequency. 2 proportion of a species’ available biomass consumed (sample size is the number of foraging sites1 the species was present at). 3 selection of browse more (+) or less than (–) availability (p < 0.05), or equal (=) to availability (p > 0.05). 95% bonferroni confidence interval lengths are in parentheses. 4 includes prunus virginianus, amelanchier sp., quercus rubra, and betula papyrifera. *p < 0.001, **p < 0.0001, ns = not significant. table 2. diet composition, browse availability, and browsing intensity. % (sd ) r east ern h emlock 775 7 17.1 (1.7) 3 + 4.4 0.3 – 5.9 96.8 (1.8) 0.997** balsam f ir 2560 12 16.7 (1.7) + 13.9 1.3 – 6.6 67.7 (29.0) 0.935** red m ap le 982 7 11.1 (1.5) + 8.7 0.3 9.5 71.4 (32.0) 0.984 * st rip ed m ap le 1134 14 28.8 (2.1) = 28.7 1.5 – 15.0 59.0 (30.0) 0.999** m ount ain m ap le 446 12 7.3 (1.2) = 7 0.1 – 9.2 29.7 (37.1) 0.994** beaked h az el 825 7 6.9 (1.2) = 7.4 0.2 – 8.2 33.4 (43.4) 0.898** sugar m ap le 1518 14 5.2 (1.0) – 10 0.1 – 7.9 30.9 (21.8) 0.883** m ount ain a sh 196 5 1.5 (0.6) – 10.8 0.1 – 15.1 28.4 (40.8) 0.710 ns yellow birch 147 11 1.0 (0.5) – 5 0.1 – 6.0 6.2 (12.0) 0.321 ns o t her 4 600 4.3 (0.9) 4 m ean brow sing int ensit y 2 range kg p er habrow se sp ecies % of t ot al biomass consumed % of t ot al biomass a vailable inst ances of u se n 1 alces vol. 40, 2004 routledge and roese moose winter diet 99 since the southern extent of moose range in eastern north america overlaps the northern limits of the range of eastern hemlock, the impacts of long-term browsing by sympatric moose and white-tailed deer on eastern hemlock regeneration may be critical. eastern hemlock saplings provide winter food for white-tailed deer, while mature eastern hemlock stands provide important winter cover (euler and thurston 1980). where locally abundant, striped maple, balsam fir, eastern hemlock, and red maple should be considered important winter browse species in gl-sl forests. sugar maple, beaked hazel, mountain maple, and mountain ash were not browsed intensively, and should be considered marginal browse species. not surprisingly, yellow birch was under-used, probably because it contains a secondary compound that reduces the rate o f f e r m e n t a t i o n b y r u m e n m i c r o b e s (belovsky 1981). although found to be important in other studies of winter diets (e.g., peek et al. 1976, snyder and janke 1976, cumming 1987), mountain ash was under-used. the large availability of mountain ash, the third highest amongst all browse species, can be attributed to the large mean dut recorded, which was the only dut larger than the mean species dpb. many mountain ash twigs, particularly the leaders, were large, possibly deterring moose from selecting them. moose may browse twigs at increasingly larger diameters, maximizing biomass consumption, to the point where quality (i.e., digestibility) eventually becomes compromised (vivas and saether 1987). this effect may not have been as noticeable if duts were measured at the proximal end of the cag (e.g., ditchkoff and servello 1998, shipley et al. 1998) as opposed to the proximal end of twigs > 5 cm in length where they forked from a larger branch (crête 1989). the mean dpb was larger than the dut for other browse species indicating that availability was underestimated even though more than the cag was likely sampled by measuring duts using crête’s method (crête 1989). this indicates that moose may have been removing considerably more than the cag (telfer 1969, shipley et al. 1998). the study area was in mature forest where the length of cag and subsequent biomass of browse species would be less than in recently disturbed areas where the plants would be exposed to full sunlight. under this latter situation the cag on twigs may be much larger in effect decreasing the amount of twig in excess of the cag that may be browsed. in mature forest, species such as striped maple, may produce minimal stem growth annually (hibbs et al. 1980), increasing the likelihood that moose will consume multiple year’s growth. white spruce was readily available to moose but was not browsed. trace amounts of white spruce have been recorded in winter moose diets in maine (ludewig and bowyer 1985) and on isle royale even though it had sufficient nutritional quality to be heavily browsed (belovsky 1981). snyder and janke (1976) reported greater densities of white spruce regeneration on isle royale i n a r e a s o f m o o s e b r o w s i n g v e r s u s unbrowsed areas, because competing species such as balsam fir were heavily used. browse preferences of moose have generally been expressed through browse use and availability estimates from plots placed throughout wintering areas following the winter-use period and before leaf flush (e.g., crête and bedard 1975). browse preferences determined using the methods in this study may better reflect preferences because only potential browse species encountered by moose were sampled. however, van vreede et al. (1989) found little difference in forage preference indices for white-tailed deer in oklahoma when either moose winter diet – routledge and roese alces vol. 40, 2004 100 technique was used to estimate forage use and availability. as browse availability was not estimated independent of moose foraging paths, it cannot be determined if the method in this study would yield results similar to random plots placed throughout the wintering area. acknowledgements we appreciate lake superior state university’s department of biology for providing equipment for the study. review comments were kindly provided by jeanfrancois robitaille, ian thompson, and two anonymous referees. references anderson, r. c., and o. l. loucks. 1979. w h i t e t a i l d e e r ( o d o c o i l e u s virginianus) influence on structure and composition of tsuga canadensis forests. journal of applied ecology 16:855861. belovsky, g. e. 1981. food plant selection by a generalist herbivore: the moose. ecology 62:1020-1030. byers, c. r., r. k. steinhorst, and p. r. krausman. 1984. clarification of a technique for analysis of utilization-availability data. journal of wildlife management 48:1050-1053. crête, m. 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67:373-380. _____, and j. bedard. 1975. daily browse consumption by moose in the gaspe peninsula, quebec. journal of wildlife management 39:368-373. _____, and r. courtois. 1997. limiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal of zoology 242:765-781. _____, and p. a. jordan. 1982. population consequences of winter forage resources for moose, alces alces, in southwestern quebec. canadian field-naturalist 96:467-475. cumming, h. g. 1987. sixteen years of moose browse surveys in ontario. alces 23:125-155. ditchkoff, s. s., and f. a. servello. 1998. litterfall: an overlooked food source for wintering white-tailed deer. journal of wildlife management 62:250-255. environment canada. 1995. monthly meteorological summary. summaries obtained for december 1994 and january and february 1995. documents on file sault ste. marie public library, sault ste. marie, ontario, canada. euler, d., and l. thurston. 1980. characteristics of hemlock stands related to deer use in east-central ontario. journal of applied ecology 17:1-6. frelich, l. e., and c. g. lorimer. 1985. current and predicted long-term effects of deer browsing in hemlock forests in michigan, usa. biological conservation 34:99-120. hibbs, d. e., b. f. wilson, and b. c. fischer. 1980. habitat requirements and growth of striped maple (acer pensylvanicum l.). ecology 61:490496. histol, t., and o. hjeljord. 1993. winter feeding strategies of migrating and nonmigrating moose. canadian journal of zoology 71:1421-1428. joyal, r. 1976. winter foods of moose in la verendrye park, quebec: an evaluation of two browse survey methods. canadian journal of zoology 54:17641770. lautenschlager, r. a., h. s. crawford, m. r. stokes, and t. l. stone. 1997. forest disturbance type differentially affects seasonal moose forage. alces 33:49-73. ludewig, h. a., and r. t. bowyer. 1985. overlap in winter diets of sympatric alces vol. 40, 2004 routledge and roese moose winter diet 101 moose and white-tailed deer in maine. journal of mammalogy 66:392-395. lyon, l. j. 1970. lengthand weightdiameter relations of serviceberry twigs. journal of wildlife management 34:456460. peek, j. m., d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. potvin, f. 1981. constructing dry weightdiameter curves for browsed twigs. journal of wildlife management 45:276279. rowe, j. s. 1972. forest regions of canada. canadian forest service publication 1300. ottawa, ontario, canada. rutkowski, d. r., and r. stottlemyer. 1993. composition, biomass and nutrient distribution in mature northern hardwood and boreal forest stands, michigan. american midland naturalist 130:13-30. shipley, l. a., s. blomquist, and k. danell. 1998. diet choices made by free-ranging moose in northern sweden in relation to plant distribution, chemistry, and morphology. canadian journal of zoology 76:1722-1733. snyder, j. d., and r. a. janke. 1976. impact of moose browsing on borealtype forests of isle royale national park. american midland naturalist 95:79-92. telfer, e. s. 1969. twig weight-diameter relationships for browse species. journal of wildlife management 33:917921. _____. 1978. cervid distribution, browse and snow cover in alberta. journal of wildlife management 42:352-361. _____, and a. cairns. 1978. stem breakage by moose. journal of wildlife management 42:639-642. van vreede, g., l. c. bradley, f. c. bryant, and t. j. deliberto. 1989. evaluation of forage preference indices for white-tailed deer. journal of wildlife management 53:210-213. vivas, h. j., and b.-e. saether. 1987. interactions between a generalist herbivore, the moose alces alces, and its food resources: an experimental study of winter foraging behaviour in relation to browse availability. journal of animal ecology 56:509-520. wetzel, j. f., j. r. wambaugh, and j. m. peek. 1975. appraisal of white-tailed deer winter habitats in northeastern minnesota. journal of wildlife management 39:59-66. zar, j. h. 1996. biostatistical analysis. third edition. prentice hall, upper saddle river, new jersey, usa. 137 53rd north american moose conference and workshop carrabassett valley (sugarloaf), maine june 10-14th, 2019 the 53rd north american moose conference and workshop was held in carrabassett valley, maine at the sugarloaf resort on june 10-14th, 2019. the maine department of inland fisheries and wildlife (conference organizer lee kantar) hosted this annual gathering of biologists, managers, researchers, students and stakeholders who gather to present information on moose research and management. the conference and… the theme “the research and management nexus: integration and synergy” represents both the power of harnessing research and management as well as a reminder of the importance of understanding and effectively melding these two aspects of wildlife biology. moose management relies on understanding population dynamics, habitat relationships, impacts of disease and a host of other complicating elements. research dedicated to quantifying and elucidating these complexities is vital to the manager who must provide empirical evidence and the “best available” data to make informed decisions on behalf of the public interest. managers often find themselves challenged by perceptions and observations that, without evidence to the contrary, weakens their credibility and ability to adapt to short and long-term changes. researchers must also consider the needs of the manager, the social-political pressures and demands that complicate management decisions. research projects that address the needs of managers as well as regional phenomena serve to strengthen area knowledge while bolstering the collective wisdom of agencies responsible for managing moose. managers gain immensely from regional research and researchers gain as well from understanding the pitfalls and benefits that managers face in the absence/presence of current information. maine hosted over 92 delegates (plus family members) and mdifw staff. this included tribal representation, many canadian provinces and a wide representation of delegates from across moose range in the us. in lieu of a traditional workshop, the conference hosted an evening q & a for the public with renowned moose biologists from across north america. an expert panel presentation and interactive discussion of chronic wasting disease highlighted the current status, knowledge, and developments of this wide-ranging, troublesome, and minimally understood disease. papers were organized into 8 sessions: “unique perspectives in provincial moose management”, “management approaches”, “habitat”, “techniques”, “harvest and predation”, “diseases”, “winter ticks” and “management issues” with informative posters also available to participants throughout the meeting. dr. roy rea, the 2018 recipient of the distinguished moose biologist award, provided the traditional dmb presentation “use of the montréal process for constructing criteria and indicators for the conservation and sustainable management of moose”. participants chose from one of four field trip options that included a drive to the top of quill hill for a discussion on wind power and impacts to wildlife; a high elevation hike up bigelow mountain in the public reserve lands; a scenic road trip to look at mitigation strategies to minimize moose-vehicle collisions; and a hike into the mt. abraham area to discuss high elevation moose habitat and moose hunting. the annual business meeting occurred at the end of the week with review of publication items and finances of the journal alces, discussion of future meetings, and other moose issues. the annual banquet featured amazing views of the bigelow mountain range, music by vermont’s only davey davis, and the silent and not-so-silent auctions were successful in providing important funding for future travel awards. recipients of travel awards and the many folks on the local committee were duly recognized followed by presentation of the 2019 distinguished moose biologist award to lee kantar, state moose biologist for the maine department of inland fisheries and wildlife. 138 delegates were provided with canvas gift bags featuring the 53rd north american moose conference logo holding logo t-shirts and hats, logo insulated mug, and a bronze moose head magnet created and made in maine by dave hentosh of smoldering lake outfitters. chair: lee kantar, maine department of inland fisheries and wildlife host: maine department of inland fisheries and wildlife location: carrabassett valley, maine date: june 10-14th, 2019 number of delegates/ participants: 92+ 143 distinguished moose biologist – award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public’s attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (≥10 in journals including alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces – a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. 6. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual’s interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3–8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special “distinguished moose biologist” presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http:// bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca http://bolt.lakeheadu.ca/~alceswww/alces.html http://bolt.lakeheadu.ca/~alceswww/alces.html mailto:art.rodgers@ontario.ca alces37(1)_217.pdf alces37(1)_175.pdf alcessupp1_84.pdf 4309.pdf alces vol. 43, 2007 musante et al. metabolic impacts of winter ticks 101 metabolic impacts of winter tick infestations on calf moose anthony r. musante, peter j. pekins, and david l. scarpitti department of natural resources, university of new hampshire, durham, nh 03824, usa abstract: moose (alces alces) are susceptible to late winter mortality from infestation of winter ticks (dermacentor albipictus) throughout much of north america. calves, perhaps more so than other ages of moose, likely experience chronic, and eventually acute anemia from blood removal by adult female ticks that peaks during weeks 4 – 6 of the 8-week engorgement period. we modeled the potential metabolic impact on protein and energy balance of moose calves associated with blood loss during four levels, low to severe, of winter tick infestation. our conservative estimates indicated that total blood loss in weeks 4 – 6, as a percent of total blood volume, ranged from 27 to 48% and 64 to 112% during moderate (30,000 ticks) and severe (70,000 ticks) infestations, respectively. the percent of the daily metabolizable energy requirement needed to replace daily blood loss during weeks 4 – 6 was 4.9 – 8.2% and 11.4 – 19.2% during moderate and severe infestations, respectively. the protein protein loss during weeks 4 – 6 was 29 – 49% and 68 – 114% of the daily protein requirement in continuous weeks. energy costs associated with compensating for blood loss would likely elevate the increased physiological stress related to concurrent anemia. severely infested calves are obviously susceptible to late winter mortality, and the impact of moderate infestations would be exacerbated by secondary parasitic infestations, severe winters, and poor body condition. alces vol. 43: 101-110 (2007) key words: alces alces, anemia, blood loss, dermacentor albipictus, energy, infestation, moose calves, mortality, protein, winter ticks moose (alces alces), elk (cervus elaphus), and white-tailed deer (odocoileus virginianus) are the three main hosts of the winter tick (dermacentor albipictus) in north america. moose are the most severely affected from infestations of winter ticks that cause hair loss and damage (mclaughlin and addison 1986, samuel 1991), excessive grooming (mooring and samuel 1999), chronic weight loss (glines and samuel 1989, addison et al. 1994), and reduced growth and fat stores (mclaughlin and addison 1986). no reduction in food intake (addison and mclaughlin 1993) and only one case of tick-induced anemia have been reported for captive moose (glines and samuel 1989, addison et al. 1998b); however, these consequences are more likely to occur in wild moose. substantial mortality associated with severe tick infestations has occurred throughout much of north american moose range (samuel 2004). winter tick related mortality was responsible for 41% (n = 16) of radio-marked mortality in new hampshire with calves representing 88% of deaths (musante 2006), although all age classes of moose have been associated with winter tick-related mortality (samuel and barker 1979, pybus 1999, samuel and crichton 2003). presumably, the large volume of blood loss associated with severe tick infestations further reduces nutritional status during march – april when cit (glines and samuel 1989, samuel 2004). calves are likely the most susceptible cohort metabolic impacts of winter ticks musante et al. alces vol. 43, 2007 102 and anemia is suspected as the primary factor of mortality. ing a 9 – 10 week period in late february to mid-may, peaking in late march – early april feeding for several species of adult female ixodid (hard body) ticks ranges from 6 to 13 days and blood removal can equal 3.0 – 7.5 times their engorged body weight (sonenshine engorgement is completed within several days with a large volume of blood loss during the last 24 – 36 hours of feeding. blood concentration in the cattle tick (boophilus microplus) is greatest during the last hours sumed a concentrated blood meal twice their own weight (seifert et al. 1968). blood removal by ectoparasites has been roby et al. 1992, simon et al. 2003), reptiles (wikelski 1999), small mammals (khokhlova et al. 2002), and livestock (seifert et al. 1968, springell et al. 1971, corrier et al. 1979, norval et al. 1988). these studies indicated that blood consumption by parasites has varying effects on blood protein, weight gain, behavior, productivity, and metabolic rate of hosts. studies have been conducted involving energetic consequences of tick-induced hair loss and grooming in moose (mclaughlin and addison 1986, samuel 1991, mooring and samuel 1999); however, little research exists on the relationships among tick infestation, blood loss, and metabolic balance in wild ungulates. the link between tick infestation and mortality of moose calves is evident; however, the physiological impact on their energy and protein balance has not been estimated quantitatively. such estimates are useful to understand and predict mortality associated with winter tick infestations. the objective of this study is to estimate the impact of blood loss from winter tick infestations on energy and protein balance of moose calves. methods metabolic impacts of blood loss were estimated with models that incorporated variable calf weights, levels of tick infestation, weight of engorged ticks, and timing of feeding. calf weight in march – april was set at 150 and 175 kg (samuel 2004), although weights of 11-month-old captive calves may exceed 200 kg (addison et al. 1994, broadfoot et al. 1996). tick infestation level was set as 10,000 (low), 30,000 (moderate), 50,000 (high), and 70,000 (severe) ticks (w.m. samuel, university of alberta, personal communication). the number of adult females at each infestation level was estimated as 25.6% of the total tick load as measured on calves in march – april (samuel 2004). larvae, nymphs, and adult males were not considered in the analysis, because adult males consume relatively little blood compared to adult females (sonenshine 1991); and unfortunately, similar data about immature life stages is limited (w.m. samuel, university of alberta, personal communication). mean engorged weights of adult female ticks have been estimated at 0.61 (glines 1983) and 0.85 g (addison et al. 1998a); we used the same conservative estimate of 0.50 g as samuel (2004) to account for ticks not fully engorged due to early removal by grooming all adult female ticks fed on this amount of blood. total amount of blood loss per adult weight because undigested blood can be 2x 1991); therefore, blood loss was estimated as 2 and 3x engorged weight. using moose calves experimentally infested with 30,000 that tick drop-off occurred primarily from march to april, peaking between 20 march and 6 april. in the current study the drop-off alces vol. 43, 2007 musante et al. metabolic impacts of winter ticks 103 period and total blood loss was estimated over an 8-week period between 1 march and 25 april; proportional blood loss was estimated as 15% in weeks 0 – 2, 25% in weeks 2 – 4, 50% in weeks 4 – 6, and 10% during weeks 6 – 8. vertebrate blood contains approximately 15% hemoglobin and 7% plasma proteins (sonenshine 1991). hemoglobin and total protein in blood of moose calves during late winter average 0.17 and 0.06 g/ml, respectively (franzmann and leresche 1978). we assumed a conservative value of 0.20 g protein/ml of blood, 4.3 kcal/g protein (schmidt-nielson 1997), and a metabolic ef1989); energetic cost of replacing blood was protein requirements were assumed as 168 and 189 g protein/day for 150 and 175 kg calves, respectively (schwartz et al. 1987, robbins 1993). total blood volume was estimated as 8% of body weight (see samuel 2004); calves weighing 150 and 175 kg had blood volumes of 12,000 and 14,000 ml, respectively. the daily metabolizable energy requirement for maintenance of a calf was assumed as 134 kcal/kg0.75/d (cool and hudson 1996), which equaled 5,743 and 6,447 kcal/d for a 150 and 175 kg calf, respectively. results total blood loss at the low infestation level (10,000 ticks) was estimated as 2,560 and 3,840 ml for 2 and 3x engorged weight, respectively; total blood loss at the severe infestation level (70,000) was 17,920 and 26,880 ml, values exceeding total blood volume (fig. 1). percent total blood volume lost in 150 and 175 kg calves with low, moderate, high, and severe infestations ranged from 21 to 224% and 18 to 192%, respectively; percent daily blood loss ranged from 0.4 to 4.0% and 0.3 to 3.4%, respectively (table 1). percent total blood volume lost in 150 and 175 kg calves with low to severe infestations during weeks 4 – 6 ranged from 11 to 112% and 9 to 96%, respectively; percent daily blood loss ranged from 0.8 to 8.0% and 0.7 to 6.9%, respectively (table 1). the energy cost to replace blood loss ranged from 2,944 to 20,608 kcal at 2x engorged weight at the four levels of infestation, and 4,416 to 30,912 kcal at 3x engorged weight 0 5000 10000 15000 20000 25000 30000 10,000 30,000 50,000 70,000 tick infe station le v e l b lo o d l o ss (m l ) 2x blood-fed weight 3x blood-fed weight total blood volume (150 kg calf) fig. 1. total blood removal by adult female winter ticks at low to severe infestation levels on moose calves over the 8-week engorgement period. calf weight (kg) 150 175 infestation 100% 50% 100% 50% level week week week week 0-8 4-6 0-8 4-6 10,000 2x 21 (0.4) 11 (0.8) 18 (0.3) 9 (0.7) 3x 32 (0.6) 16 (1.1) 27 (0.5) 14 (1.0) 30,000 2x 64 (1.1) 32 (2.3) 55 (1.0) 27 (2.0) 3x 96 (1.7) 48 (3.4) 82 (1.5) 41 (2.9) 50,000 2x 107 (1.9) 53 (3.8) 91 (1.6) 46 (3.3) 3x 160 (2.9) 80 (5.7) 137 (2.4) 69 (4.9) 70,000 2x 149 (2.7) 75 (5.3) 128 (2.3) 64 (4.6) 3x 224 (4.0) 112 (8.0) 192 (3.4) 96 (6.9) table 1. total and daily percent blood volume of calf moose removed by engorging adult female ticks. infestation level, stage, engorged weight (2 and 3x), and calf weight were varied; total blood volume was estimated as 8% of body weight. metabolic impacts of winter ticks musante et al. alces vol. 43, 2007 104 budget for a 150 kg calf ranged from 0.9 to 9.6% for low to severe infestations over 8 weeks, and 1.8 to 19.2% during weeks 4 – 6; the estimates for a 175 kg calf were 0.8 – 8.6% and 1.6 – 17.1%, respectively (table 2). total protein lost during low-to-severe infestations ranged from 512 to 3,584 and 768 to 5,376 g at 2 and 3x blood-fed weight, moderate infestation level (30,000) and 2-3x blood-fed weight was 11.0 – 16.5 g during weeks 6 – 8 and 54.9 – 82.3 g during weeks 4 a severe infestation peaked during weeks 4 – 6 and exceeded the daily protein requirement of 150 and 175 kg calves; daily protein loss was 50 – 100% of the daily protein requirement during weeks 2 – 6 (fig. 4). as a percent of the daily protein requirement during weeks 4 – 6, protein loss of a 150 kg calf peaked at 33 – 49% at a moderate infestation level, and 76 – 114% at a severe infestation level; protein loss of a 175 kg calf peaked at 29 – 44% and 68 – 102% in moderate and severe infestations, respectively (fig. 4). discussion this exercise, performed with conservative estimates, indicated that blood loss associated with moderate to severe infestations of ticks has substantial impact on energy and protein balance in moose calves. the physiological impact of blood removal by adult female ticks extends for approximately 8 weeks, and calves likely experience chronic, and eventually acute anemia during peak engorgement by weeks 4 – 6 during early to mid-april. although anemia associated with blood removal by ticks is well recognized in cattle (francis 1960, o’kelly and seifert 1969, corrier et al. 1979), there is little to no evidence in captive moose (glines and samuel 1989, addison et al. 1998b). nevertheless, anemia is hypothesized in tick-infested wild moose. hemorrhagic anemia caused by parasites occurs when the balance between blood loss and production is not maintained, and calves are unable to compensate for blood loss. the protein loss estimates in this study indicate that the potential for hemorrhagic anemia is greatest in weeks 4 – 6 (fig. 4). as a consequence of their age and smaller body size, calves have higher metabolic demands than adults on a relative scale 0 5000 10000 15000 20000 25000 30000 35000 10,000 30,000 50,000 70,000 tick infe station le v e l e n er g y c o st (k ca l) 2x blood-fed weight 3x blood-fed weightdaily energy requirement (150 kg calf) fig. 2. total energy cost for moose calves to replace blood loss at low to severe infestation levels of adult female winter ticks over the 8-week engorgement period. calf weight (kg) 150 175 infestation 100% 50% 100% 50% level week week week week 0-8 4-6 0-8 4-6 10,000 2x 0.9 1.8 0.8 1.6 3x 1.4 2.7 1.2 2.4 30,000 2x 2.7 5.5 2.4 4.9 3x 4.1 8.2 3.7 7.3 50,000 2x 4.6 9.2 4.1 8.2 3x 6.9 13.7 6.1 12.2 70,000 2x 6.4 12.8 5.7 11.4 3x 9.6 19.2 8.6 17.1 table 2. the cost of replacing blood removed by engorging adult female ticks as a percent of the daily metabolizable energy requirement of moose calves. infestation level, stage, engorged weight (2 and 3x), and calf weight were varied; total blood volume was estimated as 8% of body weight. alces vol. 43, 2007 musante et al. metabolic impacts of winter ticks 105 (schwartz et al. 1991). stored body fat and protein allow moose to survive normal enrenecker 1998); however, calves are more susceptible to late winter mortality, because they have proportionally less body fat than adults (van ballenberghe and ballard 1998). of minimal nutritional value (schwartz and renecker 1998) and the relative energetic cost associated with compensating for blood loss is presumably higher for animals in poor condition, especially calves. moose calves severely infested or in weakened condition are probably unable to sustain the energetic demand for blood regeneration and consumption of adequate food resources. volume of daily blood loss was an important factor in the mortality of smaller, tick-infested livestock calves compared to larger surviving calves (corrier et al. 1979). tick-infested moose calves that are heavier possibly have a better likelihood of recovery and survival from blood loss. calves in poor condition may also experience more pronounced energy and which likely groom more and should remove a greater number of ticks. the daily percent loss of total blood volume during weeks 4 – 6 ranged from 2.0 to 3.4% and 4.6 to 8.0%, respectively, during moderate and severe tick infestations of 30,000 and 70,000 ticks (table 1). calves infested with 50,000 ticks would lose 1 – 2x their blood volume over the 8-week engorgement period. guidelines for blood collection of healthy animals on an adequate nutritional plane suggest that 10% of blood volume can be removed every 3 – 4 weeks or 1% daily for repeated bleeds at shorter intervals (morton et al. 1993). further, total blood loss as a percent of total blood volume during weeks 4 – 6 ranged from 27 to 48% and 64 to 112%, respectively (table 1). most animals experience hemorrhagic shock if 30 – 40% of blood volume is removed over a short period of time, and rowan 1989). the percent of the daily metabolizable energy requirement required to replace the average daily blood loss during the engorgement period ranged from 2.4 to 4.1% and 5.7 to 9.6% in moderate and severe infestations (table 2). however, the daily estimates were twice that during weeks 4 – 6 (table 2), and these additional energy costs would increase ter, accelerate nutritional decline and weight loss, and likely cause increased physiological stress related to concurrent anemia. calves normally experience a negative energy balance in winter when the metabolizable energy requirement exceeds forage intake energy; 40% in winter assuming availability of quality dry matter forage (2.2 kcal/g of metabolizable energy) and a daily consumption rate of 1% body weight (schwartz and renecker 1998). 0 1000 2000 3000 4000 5000 6000 10,000 30,000 50,000 70,000 tick infe station le v e l p ro te in l o ss (g ) 2x blood-fed weight 3x blood-fed weight daily protein requirement (150 kg calf) fig. 3. total protein cost for moose calves to replace blood loss to adult female winter ticks at low to severe infestation levels over the 8-week engorgement period. 0 25 50 75 100 125 150 175 200 week 0-2 week 2-4 week 4-6 week 6-8 engorgement period p ro te in (g ) 70,000 ticks 2-3x bloodfed weight 30,000 ticks 2-3x bloodfed weight 175 kg 150 kg 100% 50% daily p rotein requirement ated with blood loss at moderate and severe infestation levels of adult female winter ticks over the 8-week engorgement period. metabolic impacts of winter ticks musante et al. alces vol. 43, 2007 106 calves infested with 70,000 ticks would lose an equivalent of 3 – 5 days of metabolizable energy requirement over the 8-week engorgement period. increased grooming and reduced feeding during march – april accentuate the negative energy balance at the end of winter (mclaughlin and addison 1986, samuel 1991, mooring and samuel 1999). although the role of protein metabolism may have the strongest cost associated with blood replacement adds to the negative impact of ticks. low, moderate, and heavy tick numbers had minimal effect on hematological parameters of captive moose maintained on a 16% protein diet (addison et al. 1998b); however, winter browse typically has 5 – 7% protein and is poorly digested (schwartz and renecker 1998). conversely, poor nutrition reduced hematological values of cattle lightly infested with boophilus microplus compared to those on an adequate diet (o’kelly et al. 1971), and tick-infested cattle had lower concentrations of hemotacrit, hemoglobin, serum albumin, and total protein than tick-free cattle (o’kelly and seifert 1969). alexander and kiesel (1965) report that blood loss coupled with low protein diet (8%) adversely affected weight gain, hemoglobin, and hematocrit in lambs; minimal effects occurred in lambs maintained on a 16% protein diet. thus, the negative impact and host-susceptibility of parasitism in ruminants are greater in malnourished animals; adequate nutrition and protein intake reduce impact (van houtert and sykes 1996, coop and kyriazakis 2001). loss and regeneration is probably the most critical physiological problem for calves. in severely infested moose calves occurred for 4 continuous weeks (weeks 2 – 6; fig. 4). losses were most pronounced during weeks 4 – 6; 29 – 49% in a moderate infestation and 68 – 114% in a severe infestation (fig. 4). calves infested with 70,000 ticks would lose an equivalent of 3 – 4 weeks of the daily protein requirement. moose are invariably ing winter (schwartz et al. 1988); therefore, compensation would be problematic, because the engorgement period occurs prior to spring green-up when quality and quantity of forage are limited. calves experiencing severe blood loss should be considered high-risk mortality from anemia and associated effects. their ability to survive tick infestations is probably the level of infestation. the estimated proportion of adult female ticks on adult moose during late winter is lower on cows (18.0%) and slightly higher on bulls (27.6%), relative to calves (25.6%) (samuel 2004). based on similar calculations as with associated with a severe infestation level is 4.5 – 6.7% for a cow (360 kg) and 6.3 – 9.5% for a bull (400 kg) during weeks 4 – 6. the during a severe infestation is 27.8 – 41.7% for a cow and 39.3 – 59.0% for a bull during weeks of the energy budget and protein requirement, are 50 – 70% less than those estimated for calves. these “baseline” estimates for by compounding the costs of pregnant (last trimester) cows and bulls suffering post-rut. calves are likely the most susceptible to annual winter tick-related mortality; however, adults in poor condition may be predisposed to mortality from heavy tick infestations or during tick epizootics, although adults should normally survive during years of average tick abundance. mortality associated with winter ticks was observed in radio-marked moose in new hampshire from 2002 to 2005 and was highest in april (75%) corresponding to weeks 4 – 6 of tick engorgement when blood loss was greatest and most concentrated (musante 2006). alces vol. 43, 2007 musante et al. metabolic impacts of winter ticks 107 although tick density was not measured on new hampshire calves, hair loss and damage was most severe on carcasses in 2002 when calf survival was lowest (0.49) and the highest percentage of tick-related mortality occurred; regional spikes in spring mortality (samuel and crichton 2003) and severe coat damage to non-study moose were concurrent. in addition to high and severe levels of infestation and tick-related hair loss/damage, the majority of the calves were emaciated with poor body fat and femur marrow fat ( x = 16.5%) indices, secondary infestations of lungworm, presumably dictyocaulus viviparous, and noticeable paleness of eye mucous membranes, which is a characteristic of anemia in domestic ruminants (kaplan et al. 2004). dictyocaulus viviparous is found commonly in the small bronchioles of lungs in elk calves, and when combined with severe weather conditions, poor host nutrition, or heavy winter tick infestations, has caused morbidity or death of elk (worley 1979, thorne et al. 2002). although d. viviparous is generally not believed related to morbidity of moose (lankester and samuel 1998), prevalent infections have been described in calves and yearlings (pybus 1990). calf mortalities in maine during late winter 1995 had infestations of d. viviparous and winter ticks (k. morris, maine if&w, personal communication). while this parasite is probably not the primary cause of death, combined infestations of lungworms and winter ticks may be more detrimental than tick infestations alone. in conclusion, this exercise indicated that blood loss to winter ticks alters protein and energy metabolism of moose calves substansurvival. our models would underestimate the effects associated with more synchronous and concentrated blood loss. severely infested calves are more susceptible to late winter mortality; however, the effect of a moderate parasitic infestations, severe winters, and poor body condition. the effect of chronic blood loss is exacerbated by a diet of low quality and digestible protein, and as a result, many calves probably are unable to adequately replace blood and protein loss and become acutely anemic. therefore, it is evident from our conservative estimates and pattern of calf mortality in this study and others, that winter tick infestations and primarily tick epizootics have considerable potential to reduce winter ics of a local or regional moose population. acknowledgements funding for this research was provided partment from dedicated funds generated through the sale of moose hunt applications and permits. this study was possible due to the cooperation of numerous commercial landholders and local residents in the north country who granted access to their property. we are sincerely grateful to the biologists and new hampshire who were vital to the project’s success. in particular we would like to thank kristine rines/nhfg moose biologist, who addison, who reviewed this manuscript and provided constructive comments. references addison oachim, r. f. mclaughlin raser. 1998a. ovipositional development and fecundity of dermacentor albipictus (acari: ixodidae) from moose. alces 34:165-172. _____, and r. f. mclaughlin. 1993. seasonal variation and effects of winter ticks (dermacentor albipictus) on consumption of food by captive moose (alces alces) calves. alces 29:219-224. roadfoot. 1994. metabolic impacts of winter ticks musante et al. alces vol. 43, 2007 108 growth of moose calves (alces alces americana) infested and uninfested with winter ticks (dermacentor albipictus). 1476. _____, _____, and _____. 1998b. effects of winter tick (dermacentor albipictus) on blood characteristics of captive moose (alces alces). alces 34:189-199. alexander iesel. 1965. the effect of blood loss on weight gain, hemoglobin and hematocrit in lambs fed different levels of protein. auburn veterinarian 21:114-117, 129. blaxter, k. l. 1989. energy metabolism in animals and man. cambridge university press, new york, new york, usa. broadfoot oachim, e. m. addison, and k. s. mac onald. 1996. weights and measurements of selected body parts, organs, and long bones of 11month-old moose. alces 32:173-184. cool udson. 1996. requirements for maintenance and live weight gain of moose and wapiti calves during winter. rangifer 16:41-45. coop, r. l., and i. kyriazakis. 2001. influence of host nutrition on the development and consequences of nematode parasitism in ruminants. trends in parasitology 29:479-488. corrier izcaino, m. terry, a. betancourt, k. l. kutter, c. a. carson, g. trevino, and m. ristic. 1979. mortality, weight loss and anemia in bos taurus calves exposed to boophilus microplus ticks in the tropics of columbia. tropical animal health and production 11:215-221. rew, m. l., and w. m. samuel. 1989. instar development and disengagement rate of engorged female winter ticks, dermacentor albipictus (acari: ixodidae), following singleand trickleexposure of moose (alces alces). experimental and applied acarology 6:189-196. francis growth-rate of cattle. proceedings of the australian society of animal production 3:130. franzmann, a. w., and r. e. leresche. 1978. alaskan moose blood studies with emphawildlife management 42:334-351. glines, m. v. 1983. the winter tick, dermacentor albipictus (packard, 1896): its life history, development at constant temperatures, and physiological effects on moose, alces alces l. m.sc. thesis, university of alberta, edmonton, alberta, canada. _____, and w. m. samuel. 1989. effect of dermacentor albipictus (acari: ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. experimental and applied acarology 6:197-213. gold ahlsten. 1983. effects of parasitic flies (protocalliphora spp.) on nestlings of mountain and chestnutbacked chickadees. wilson bulletin 95:560-572. kaplan urke, t. h. terrill, iller, w. r. getz, s. mobini, e. valencia illiams, l. h. williamson, m. larsen, and a. f. vatta. 2004. validation of the famacha© eye color chart for detecting clinical anemia in sheep and goats on farms in the southern united states. veterinary parasitology 123:105-120. khokhlova, i. s., b. r. krasnov, m. kam, n. i. burdelova egen. 2002. energy cost of ectoparasitism: the flea xenopsylla ramesis on the desert gerbil gerbillus dasyurus. london 258:349-354. lankester, m. w., and w. m. samuel. 1998. pests, parasites and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washingalces vol. 43, 2007 musante et al. metabolic impacts of winter ticks 109 mcguill, m. w., and a. n. rowan. 1989. biological effects of blood loss: implications for sampling volumes and techniques. institute for laboratory animal research news 31:5-18. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)-induced winter hair-loss in captive moose (alces alces 22:502-510. mooring, m. s., and w. m. samuel. 1999. premature winter hair loss in free-ranging moose (alces alces) infested with winter ticks (dermacentor albipictus) is correlated with grooming rate. canadian morton bbot, r. barclay, b. s. close, r. ewbank ask, m. heath, s. mattic, t. poole eamer outhee, a. thompsan, b. trussel, c. west, and m. ennings. 1993. removal of blood from laboratory animals and birds. laboratory animals 27:1-22. musante, a. r. 2006. characteristics and dynamics of a moose population in northern new hampshire. m.sc. thesis, new hampshire, usa. norval, r. a. i., r. w. sutherst urki, ibson err. 1988. the effect of the brown ear-tick rhipicephalus appendiculatus on the growth of sanga and european breed cattle. veterinary parasitology 30:149-164. o’kelly eebeck, and p. h. springell. 1971. alterations in host metabolism by the specific and anorectic effects of the cattle-tick (boophilus microoplus). ii. changes in blood composition. aus24:381-389. _____, and g. w. seifert. 1969. relationships between resistance to boophilus microplus, nutritional status, and blood composition in shorthorn x hereford sciences 22:1497-1506. pybus pulmonary parasites of wild cervids in _____. 1999. moose and ticks in alberta: a dieoff in 1998/99. occasional paper number 20, fisheries and wildlife mancanada. robbins, c. t. 1993. wildlife feeding and nutrition, second edition. academic press, new york, new york, usa. roby rink, and k. wittmann. 1992. effects of blowfly parasitism on eastern bluebird and tree swallow nestlings. wilson bulletin 104:630-643. samuel, b. 2004. white as a ghost: winter ticks & moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m. 1991. grooming by moose (alces alces) infested with the winter tick, dermacentor albipictus (acari): a mechanism for premature loss of win69:1255-1260. _____, and m. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869) on moose alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15:303-348. _____, and v. crichton. 2003. winter ticks and winter-spring losses of moose in western canada, 2002. the moose call 16:15-16. elch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27:169-182. schmidt-neilsen, k. 1997. animal physiology: adaptation and environment. fifth edition. cambridge university press, new york, new york, usa. metabolic impacts of winter ticks musante et al. alces vol. 43, 2007 110 schwartz, c. c., m. e. hubbert, and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. 33. _____, _____, and _____. 1991. energy wildlife management 55:391-393. _____, w. l. regelin, and a. w. franzmann. of wildlife management 51:352-357. _____, and l. a. renecker. 1998. nutrition and energetics. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution seifert, g. w., p. h. springell tatchell. 1968. radioactive studies on the feeding of larvae, nymphs, and adults of the cattle tick, boophilus microplus (canestrini). parasitology 58:415-430. simon homas peakman, londel, p. perret, and m. m. lambrechts. 2003. impact of ectoparasitic blowfly larvae (protocalliphora spp.) on the behavior and energetics of nestling 76:402-410. sonenshine volume i. oxford university press, new york, new york, usa. springell kelly, and r. m. seebeck. 1971. alterations in host metabolism by the specific and anorectic effects of the cattle tick (boophilus microplus). iii. metabolic implication of blood volume, body water, and carcass composition sciences 24:1033-1045. thorne, e. t., e. s. williams, w. m. samuel, and t. p. kistner parasites. pages 351-387 in american elk: ecology and management. smithsonian institution press, washingvan ballenberghe, v., and w. b. ballard. 1998. population dynamics. pages 223-245 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washingvan houtert ykes. 1996. implications of nutrition for the ability to withstand gastrointestinal nematode sitology 26:1151-1168. wikelski, m. 1999. influences of parasites and thermoregulation on grouping tendencies in marine iguanas. behavioral ecology 10:22-29. worley diseases of elk in northern rocky mountain region: a review. pages 206-211 in editors. north american elk: ecology, behavior, and management. university of wyoming, laramie, wyoming, usa. alces vol. 44, 2008 rea and gillingham – shoot morphometric estimates 21 effects of plant compensation across sites on regression estimates of shoot biomass and length roy v. rea and michael p. gillingham natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia, canada v2n 4z9, email: reav@unbc.ca abstract: regression estimates for determining browse shoot biomass from bite diameters and shoot basal diameters are commonly used to estimate biomass consumption and the impacts that herbivores have on range resources. such estimates tend to be based on equations built from data taken across the continuum of shoot morphometries present on plants within a given study area. how these morphometric relationships differ between the shoots of undamaged and damaged (e.g., following browsing, shoot breakage, or brush-cutting) plants is unclear. to assess the effects of plant compensation and the importance of site on shoot morphometrics for scouler's willow (salix scouleriana), we clipped and measured current annual shoots at 5 sites in central british columbia. each site had been previously brush-cut and current annual shoots were collected from both brush-cut and control willows. for each treatment and site, we developed separate regressions to predict shoot weight from length, weight from basal diameter, and length from basal diameter. comparisons of individual regressions indicated that different regressions, or even different forms of regressions (i.e., power function versus linear), are needed to accurately predict shoot weight and length depending on whether or not plants are producing compensatory or non-compensatory shoots. for some willows in the same treatment category (brush-cut versus uncut), the appropriate regressions differed among some sites. these results suggest that the effects of plant compensation following mechanical damage have important implications to the extrapolation and interpretation of shoot morphometric relationships, and thus, biomass estimates across different study areas. alces vol. 44: 21-30 (2008) key words: biomass estimation, browse, compensatory growth, mechanical brushing, plant response, regression analysis, salix scouleriana in the absence of direct observations and measurements, determining biomass consumption of browse shoots by ungulates is difficult and time consuming (provenza and urness 1981). one method of estimating biomass removal is to develop regression equations for shoot biomass based on the diameter and other morphometric parameters of the current annual shoot (telfer 1969a, lyon 1970, provenza and urness 1981). in this way, shoot biomass beyond the point of browsing (consumption) can be estimated (ferguson and marsden 1977, provenza and urness 1981, maccracken and van ballenberghe 1993) in a non-destructive manner (thilenius 1990). likewise, availability of browse, carrying capacity of ranges (telfer 1969a), and animal stocking rates (ruyle et al. 1983) can be estimated using similar equations that predict biomass from measurements taken at the basal diameter of shrub and tree shoots. regression equations for estimating shoot biomass and length from other shoot attributes have been developed for several browse plants commonly consumed by moose (alces alces l.; e.g., telfer 1969b, thilenius 1990, maccracken and van ballenberghe 1993). these equations, however, have not accounted for variations in shoot architecture resulting from exaggerated vegetative shoot growth on plants compensating for various forms of mechanical damage such as browsshoot morphometric estimates – rea and gillingham alces vol. 44, 2008 22 ing, breakage, and cutting. in this paper, we investigated whether equations predicting biomass and shoot length for scouler’s willow (salix scouleriana barr.) varied among plants that were compensating for mechanical damage from brush-cutting between 2 and 3 years after cutting and undamaged plants at 5 sites in central british columbia. study area our study area consisted of 5 sites that were clear-cut logged (15-40 ha in size) and then planted with lodgepole pine (pinus contorta dougl. ex loud. var. latifolia engelm.) near vanderhoof, british columbia, canada (lat 54°01’n, long 124°00’w). all sites were characterized by open stands of lodgepole pine with poorly developed shrub and herb layers, and a well-developed moss layer dominated by lichens; soils on all sites were clay and/or sandy loam (rea 1999). methods site histories we selected 5 sites where brush-cutting had been conducted to determine the effects of mechanical damage on willow shoot morphometry 2 and 3 years after brush-cutting. three of the sites (layton, buck, and sackner) were clear-cut logged 12-15 years prior to our study; these sites were then brush-cut during the 1993 growing season (june-september) and sampled 3 years later (winter 1995-96). the other 2 sites (sawmill and huckleberry) were clear-cut logged 9-11 years prior to the beginning of the study, were brush-cut during the 1995 growing season, and sampled 2 years later in winter 1996-97. during brush-cutting operations in 1993 and 1995, all above-ground biomass, except ~10 cm of stump tissue, was removed from willows and all other deciduous shrubs and trees at each site. wildlife strips (sensu santillo 1994; areas established for wildlife food and cover after clear-cut logging but prior to brush-cutting treatments) at each site were not brush-cut and contained willows about 4-5 m tall at the beginning of this study; willows that had been brush-cut on these sites were about 1-2 m tall. all sites had a long history of browse utilization by moose and deer (odocoileus spp.). additionally, free-range cattle (bos taurus l.) utilized the buck and sackner sites in summer. current annual shoots during the winter of 1995-1996, we randomly selected 6 scouler’s willow plants from brush-cut areas and 6 from the wildlife strips (controls) on each of the 3 plantation sites that had been brush-cut in 1993. we similarly selected willows at each of the 2 sites brush-cut in 1995 in the winter of 1996-1997. once willows were selected, we clipped shoots accessible above the snowpack. shoots were systematically collected at different clipping intensities (as part of a larger study, rea 1999, rea and gillingham 2001) from willows in the layton, buck, and sackner sites at the time we selected the plants during the winter of 19951996, and from the sawmill and huckleberry sites during the winter of 1996-1997. we collected all shoot samples while plants were dormant in mid-winter by clipping shoots at the current annual growth scar. we sealed all collected shoots in plastic freezer bags in the field to inhibit water loss during transportation back to our laboratory at the university of northern british columbia. all shoots were weighed to the nearest mg and measured for length (cm) and basal diameter (mm). when >30 shoots were collected from a particular willow, we randomly sub-sampled 30 shoots for morphometric measures. regression analyses we began by examining the fit of 4 linear and non-linear regressions for each treatment (brush-cut versus uncut) at each of the 5 sites; we considered linear (y = a + bx), power (y = a + bxc), and exponential (y = aebx and y = a + becx) regression models. following alces vol. 44, 2008 rea and gillingham – shoot morphometric estimates 23 the recommendation of verwijst (1991) for biomass estimation, we did not use any logtransformed variables in any regression model. for each site and treatment, we developed separate regressions for: 1) shoot weight (y) based on shoot length (x), 2) shoot weight (y) based on shoot basal diameter (x, at the point of the growth scar), and 3) shoot length (y) based on shoot basal diameter (x). in choosing the best regression for each set of data, we considered r2 values (r2 for linear regression) and the fit of residuals. linear models were selected if the residuals did not justify a nonlinear relationship. in all but 1 of the nonlinear relationships, the power function was the best fit; because the power function was a very close second to the exponential model in the single other case, we chose to use the power function to simplify the comparison with other non-linear predictions. we fit all nonlinear models with proc nlin (version 9.1, sas institute 2003); linear regressions were fit with the regression procedure (reg) in stata (version 9.2, statacorp. 2007). confidence intervals around individual regression parameters were estimated by asymptotic approximations in the respective packages. we considered morphometric relationships to be different between treatments and/or among sites if the form of the regression was different (i.e., linear versus power), or if the confidence intervals around individual parameters of regressions of the same form did not overlap. we did not apply bonferroni corrections to the confidence intervals because individual regressions with non-overlapping simple confidence intervals would yield different biomass estimates. we considered an α of 0.05 throughout our analyses. results shoot weight from shoot length all regression estimates of shoot weight from shoot length were best fit with power functions. in addition, there were no differences among regressions in shoot weight (y) predicted by shoot length (x) for brush-cut willows across all sites (table 1). there were, however, differences in regression equations for uncut plants among sites (i.e., one or more parameters in the power functions were significantly different from each other, table 1). these differences included regressions for uncut willows at the 2-year post-cutting sites (i.e., huckleberry uncut versus sawmill uncut) and at the 4-year sites (e.g., buck uncut versus sackner uncut). there were also numerous differences among regression equations developed for shoot weight versus shoot length when shoots from brush-cut and uncut plants were compared (table 1, fig. 1). although the parameter that varied was not consistent among comparisons, any equation that varied significantly in any parameter would yield a significantly different prediction. shoot weight from shoot basal diameter all regression estimates of shoot weight from shoot basal diameter were also best fit with a power function. with the exception of 1 case (huckleberry versus buck) that represented a difference in year-since-brushcutting, we detected no difference in the form or parameters of the regression equations that explained the relationship of shoot weight (y) to basal diameter (x) for shoots of brush-cut plants. there was less consistency in the regression parameters of the power functions among uncut treatments (table 2). although the 2, 2-year sites (i.e., huckleberry and sawmill) were not different, there were differences in equations between 4-year sites (e.g., table 2: buck versus sackner and sackner versus sawmill). relationships for uncut plants also differed significantly between 4-year, postcutting sites (table 2: layton and sackner). there were no differences in equations for shoots of brush-cut and uncut plants growing on sites that were sampled 2 years after shoot morphometric estimates – rea and gillingham alces vol. 44, 2008 24 site buck huckleberry layton sackner sawmill site treatment br un br un br un br un br un buck br — b un — b c c c c c huckleberry br — b c a b c b c a b c un — a c b c layton br — a b un — a b c b c b c sackner br — a b c un — b c sawmill br — b c un — table 1. comparison of coefficients for shoot weight (y) versus shoot length (x) regressions. all regressions were best fit with a power function (y = a + bxc). brush-cut (br) and uncut (un) treatments were compared for each of 5 sites in central british columbia. letter entries in the table represent significant differences in the parameters (a, b, and c)a in the power function between treatments and among sites. because the table is symmetrical, only the cells above the diagonal (—) are completed. cells with no entries above the diagonal indicate that the corresponding regressions were not different from each other. aa = the intercept of the power function equation. b = the slope of the power function equation. c = the exponent of the power function equation. note: the appearance of a letter in the table indicates a significant difference in either the intercept (a), slope (b), or exponent (c) between the two equations being compared. a current annual shoot length (mm) 0 20 40 60 80 100 120 140 c ur re nt a nn ua l s ho ot w ei gh t ( g) 0 5 10 15 20 25 30 b current annual shoot length (mm) 0 20 40 60 80 100 120 140 c ur re nt a nn ua l s ho ot w ei gh t ( g) 0 5 10 15 20 25 30 y = 0.769+0.0003x2.38 y = 0.231+0.002x2.02 fig. 1. observed and predicted values for current annual shoot weight versus shoot length for brush-cut (a; n = 191) and uncut (b; n = 240) willows on the huckleberry site. alces vol. 44, 2008 rea and gillingham – shoot morphometric estimates 25 cutting (table 2). similarly, the relationship between shoot weight and basal diameter did not vary between cut and uncut plants within the same site. shoot length from shoot basal diameter unlike the relationships between shoot weight versus shoot length and shoot weight versus basal diameter, shoot length could not be predicted from basal diameter by a single equation form (table 3, fig. 2). again, the effect of brush-cutting appeared more important than site effects in that all brush-cut treatments did not differ in equation form (table 3: linear). there were, however, differences in the slope (e) for brush-cut treatments within and among treatments 2 and 3 years post-cutting. in those instances in which power functions were better fits than linear regressions, it was always for uncut treatments, although there was no consistency within and among 2and 3-year sites. finally, there were many differences within sites between cut and uncut treatments regardless of the number of years since cutting (table 3, fig. 2). discussion a fundamental difference appears to exist between the shoot morphometrics of brushcut plants and those of uncut plants in which the growth form of compensatory shoots appears more consistently predictable than that of shoots from undamaged plants. this phenomenon appears to be true both within and between sites regardless of the time since brush-cutting. our results further suggest site buck huckleberry layton sackner sawmill site treatment br un br un br un br un br un buck br — c c c un — huckleberry br — b c un — c b c c layton br — c un — b c b c c c sackner br — b c un — c sawmill br — un — table 2. comparison of coefficients for shoot weight (y) versus shoot basal diameter (x) regressions. all regressions were best fit with a power function (y = a + bxc). brush-cut (br) and uncut (un) treatments were compared for each of 5 sites in central british columbia. letter entries in the table represent significant differences in the parameters (a, b, and c)a in the power function between treatments and among sites. because the table is symmetrical, only the cells above the diagonal (—) are completed. cells with no entries above the diagonal indicate that the corresponding regressions were not different from each other. aa = the intercept of the power function equation. b = the slope of the power function equation. c = the exponent of the power function equation. note: the appearance of a letter in the table indicates a significant difference in either the intercept (a), slope (b), or exponent (c) between the two equations being compared. shoot morphometric estimates – rea and gillingham alces vol. 44, 2008 26 that when considering undamaged plants, the relationship between shoot weight and length may be more influenced by site than other morphometric relationships. although all regressions for predicting shoot weight from shoot length and weight from basal diameter were best fit to a power function, linear equations were better suited to predict shoot length from basal diameter for all brush-cut plants. the regressions used to predict weight from length of shoots taken from uncut willows on the buck and layton sites were also better described with linear equations, whereas the shoots of uncut plants on the remaining 3 sites were better characterized by a power function. our results are based on a relatively small sample size of twigs and plants. if larger samples resulted in more within-site variability, then the confidence intervals around the parameters would be wider and perhaps fewer significant differences would be detected between treatments and among sites. we would expect that a small sample size, site buck huckleberry layton sackner sawmill br un br un br un br un br un site treatment (linear) (linear) (linear) (power) (linear) (linear) (linear) (power) (linear) (power) buck br — e * e d e * * un — d e * e e * e * huckleberry br — * e d e * e * un — * * * * layton br — d e e * e * un — d e * e * sackner br — * * un — * c sawmill br — * un — table 3. comparison of coefficients for shoot length (y) versus shoot basal diameter (x) regressions. some regressions were best fit with a power function (y = a + bxc) while other regressions were linear (y = d + ex). brush-cut (br) and uncut (un) treatments were compared for each of 5 sites in central british columbia. letter entries in the table represent significant differences in the parameters (power: a, b, c; linear: d, e)a in the regressions between treatments and among sites (note: no a or b parameters were significantly different from each other). an * indicates that differences existed because the same form of regression could not be fit to the corresponding entries in the table. because the table is symmetrical, only the cells above the diagonal (—) are completed. cells with no entries above the diagonal indicate that the corresponding regressions were not different from each other in form or parameters. aa = the intercept of the power function. b = the slope of the power function. c = the exponent of the power function. d = the intercept of the linear equation. e = the slope of the linear equation. note: the appearance of a letter in the table indicates a significant difference in the parameter between the two equations being compared. alces vol. 44, 2008 rea and gillingham – shoot morphometric estimates 27 however, would increase and not decrease the variation in the regressions. our results do not suggest a distinct pattern between the way in which equations differed from one another relative to site or time since brush-cutting (2 versus 3 years post-cutting). there were differences in the equations between plants growing on sites that had been brush-cut 2 years versus 3 years earlier. but this was also true when comparing within year since cutting and across sites. therefore, we make no generalizations regarding site and year effects. the fact that predictive equations for predicting shoot biomass of brush-cut plants did not differ between sites and year since cutting, and all other comparisons demonstrated significant differences, suggests a consistency in the relationship of shoot mass to length and basal diameter of compensatory shoots not found in the shoots of undamaged plants (ferguson and marsden 1977) and is, to our knowledge, previously unreported. however, it is unclear why predictive equations of biomass from the length and diameter of larger shoots would be more consistent across sites and year-since-treatment than predictive equations generated from the same parameters on non-compensatory shoots. perhaps apical and lateral buds of winter shoots exhibit consistency in size and mass and influence morphometric relationships disproportionately more for smaller and moderately sized shoots arising from undamaged plants than for heavier shoots. such relationships are not necessarily true outside of the winter dormant period (schewe and stewart 1986). season and year of shoot collection (telfer 1969a, schewe and stewart 1986, thilenius 1990), plant species differences (telfer 1969b, potvin 1981, maccracken and van ballenberghe 1993), site/microsite and aspect (lyon 1970, peek et al. 1971, ruyle et al. 1983, schewe and stewart 1986), plant size/ age (lyon 1970, peek et al. 1971), and shoot age and position on the plant (telfer 1969a, lyon 1970, ferguson and marsden 1977) are known to influence predictive equations of 1 shoot attribute from another. however, no such claims have been made for the influence of compensatory growth on such equations. ruyle et al. (1983) found that the form of quadratic equations used to predict oven-dried shoot weight from other shoot attributes varied by the total number of kg of snowberry plants utilized in pastures by sheep. maccracken and van ballenberghe (1993) speculated that shoot size and age could significantly influence the character, and thereby, the utility of the regression equation. peek et al. (1971) speculated more specifically that browsing pressure was likely to account for variation in a basal diameter of current annual shoot (mm) 0 2 4 6 8 10 c ur re nt a nn ua l s ho ot w ei gh t ( cm ) 0 5 10 15 20 25 30 b basal diameter of current annual shoot (mm) 0 2 4 6 8 10 c ur re nt a nn ua l s ho ot l en gt h (c m ) 0 20 40 60 80 100 120 140 y = 0.060+0.026x3.32 y = -32.46+17.88x fig. 2. observed and predicted values for current annual shoot weight versus current annual shoot basal diameter (a; n = 191), and length of current annual shoot versus current annual shoot basal diameter (b; n = 191) for brush-cut willows at the huckleberry site. shoot morphometric estimates – rea and gillingham alces vol. 44, 2008 28 equations developed for mountain ash. our results seem to support such speculation and suggest that attempting to predict 1 attribute from another without accounting for shoot response to damage, could result in less accurate predictions than if separate regressions were developed for sites containing different treatment histories. our findings also indicate that using regression models to predict 1 shoot attribute from another should include some attention to site (lyon 1970, peek et al. 1971, ruyle et al. 1983), and more importantly, to shootspecific details. both the intercept and form of predictive equations for 1 shoot attribute based on another varied depending on whether or not shoots were compensatory and on the site at which the parent plant was growing. because plant compensation appears to be at least partially responsible for variation in shoot morphometric relationships, we suggest that the development of separate equations for shoots of compensatory and non-compensatory plants from different sites is likely to increase efficiencies in the field and increase predictive power more so than simply increasing sample sizes in an attempt to reduce variability (peek et al. 1971). estimates of shoot weight from basal diameter are often used by rangeland managers to approximate available and/or browsed biomass (e.g., ferguson and marsden 1977, provenza and urness 1981, maccracken and van ballenberghe 1993). equations we developed to predict shoot weight from basal diameter were consistent in form and parameters for brush-cut, but not uncut willows. estimating shoot biomass from shoot basal diameter with the use of our predictive equation for the shoots of brush-cut plants from the buck site reveals that a typical shoot with a basal diameter of 5 mm would weigh 4.79 g, whereas a shoot from an uncut plant on the same site with a basal diameter of 5 mm would weigh 3.50 g. estimating 100 such shoots per plant and 100 such plants per hectare, reveals that a difference of nearly 13 kg of browse per ha could go unaccounted for if prediction equations ignored differences between plants producing compensatory or non-compensatory shoots. increases in the number of shoots per plant or plants per hectare exaggerate such discrepancies. the degree to which predictive equations tested here varied between brush-cut and uncut plants underscores the need for managers to begin to account for whether or not plants used in building such equations are compensating from damage. although brush-cutting appears to represent an extreme form of damage not likely to occur in nature, willows scoured by ice flows and broken by snow press (danell et al. 1987) and browsers (telfer and cairns 1978) can incur similar magnitudes of damage. in fact, browse surveys are often conducted in areas influenced by anthropogenic activities such as brush-cutting and logging where interest in browse availability and the utility of such areas for rangeland use is commonly expressed (shafer 1963, rea and gillingham 2001). even so, moderate forms of damage in more remote areas can cause plants to respond with vigorous vegetative regeneration (danell et al. 1985) that is likely to influence attributes used in regression equations (telfer 1969a). regardless of the damage agent involved or to what degree compensation proceeds, implementing sampling designs that examine plant compensation as well as site effects will allow researchers and managers to better account for the range of variation in shoots growing on differently treated plants on different sites and, as a result, increase the accuracy of their predictions. acknowledgements we would like to thank r. brown, h. cedervind, b. clayton, v. corbett, m. deli, s. gibson, c. smith, and j. wiersma for their help in the field and in processing willow shoots in the lab. financial assistance for this project was provided by forest renewal british coalces vol. 44, 2008 rea and gillingham – shoot morphometric estimates 29 lumbia (fr-96/97-093). the prince george regional office of the bc forest service and the university of northern british columbia also contributed financial aid; in-kind contributions from the vanderhoof district of the bc forest service are acknowledged. references danell, k., t. elmqvist, l. ericson, and a. salomonson. 1987. are there general patterns in bark-eating by voles on different shoot types from woody plants? oikos 50:396-402. _____, k. huss-danell, and r. bergström. 1985. interactions between browsing moose and two species of birch in sweden. ecology 66:1867-1878. ferguson, r. b., and m. a. marsden. 1977. estimating overwinter bitterbrush utilization from twig diameter-length-weight relations. journal of range management 30:231-236. lyon, j. l. 1970. lengthand weight-diameter relations of serviceberry twigs. journal of wildlife management 34:456-460. maccracken, j. g., and v. vanballenberge. 1993. mass-diameter regressions for moose browse on the copper river delta, alaska. journal of range management 46:302-308. peek, j. m., l. w. krefting, and j. c. tappeiner. 1971. variation in twig diameterweight relationships in northern minnesota. journal of wildlife management 35:501-507. potvin, f. 1981. constructing dry weight-diameter curves for browsed twigs. journal of wildlife management 45:276-279. provenza, f. d., and p. j. urness. 1981. diameter-length-weight relations for blackbrush (coleogyne ramosissima) branches. journal of range management 34:215-217. rea, r. v. 1999. response of scouler’s willow (salix scouleriana) to mechanical brushing: implications to the quality of winter browse for moose (alces alces). m.sc. thesis. university of northern british columbia, prince george, british columbia, canada. _____, and m. p. gillingham. 2001. the impact of the timing of brush management on the nutritional value of woody browse for moose alces alces. journal of applied ecology 38:710-719. ruyle, g. b., j. e. bowns, and a. f. schlundt. 1983. estimating snowberry (symphoricarpos oreophilus) utilization by sheep from twig diameter-weight relations. journal of range management 36:472474. santillo, d. j. 1994. observations on moose, alces alces, habitat and use on herbicidetreated clearcuts in maine. canadian field naturalist 108:22-25. sas institute 2003. sas institute inc., version 9.1, cary, north carolina, u.s.a. schewe, a. m., and j. m. stewart. 1986. twig weight-diameter relationships for selected browse species on the duck mountain forest reserve, manitoba. canadian journal of forest research 16:675-680. shafer, e. l. 1963. the twig-count method for measuring hardwood deer browse. journal of wildlife management 27:428-437. statacorp. 2007. stata version 9.2, college station, texas, u.s.a. telfer, e. s. 1969a. twig weight-diameter relationships for browse species. journal of wildlife management 33:917-921. _____. 1969b. weight-diameter relationships for 22 woody plant species. canadian journal of botany 47:1851-1855. _____, and a. cairns. 1978. stem breakage by moose. journal of wildlife management 42:639-642. thilenius, j.f. 1990. dimensional weights and forage of barclay willow and sweetgale on moose ranges in the wetlands of the copper river delta, alaska. forest ecology and management 33:463-483. verwijst, t. 1991. logarithmic transformashoot morphometric estimates – rea and gillingham alces vol. 44, 2008 30 tion in biomass estimation procedures: violation of the linearity assumption in regression analysis. biomass and bioenergy 1:175-180. alces vol. 34 (1), (1998) i harold cumming, longtime ontario wildlife biologist, lakehead university professor, and first editor of alces (originally published as the proceedings of the north american moose conference and workshop) died aug. 18, 2011 at the age of 82. harold was educated at the university of toronto, michigan state university, and received his phd in 1966 from the university of aberdeen in scotland. harold was a deeply respected wildlife biologist and professor described by a former employee as a mentor who made biologists and students better observers, thinkers, and managers. harold was employed initially with the ontario department of lands & forests as a district biologist in geraldton in 1953-58. he then took a paid leave to undertake phd studies on the ecology of roe deer at aberdeen university in scotland in 1959-1963. returning to ontario, he became a provincial big game biologist game management section, wildlife branch, lands & forests, ministry of natural resources in maple, ontario from 1963-1971. his primary responsibilities included moose, deer, and caribou management. among his many accomplishments in moose management were coordination of the provincial moose inventory (aerial surveys), centralizing the harvest in memoriam harold greenfield cumming assessment program in 1968, making recommendations for annual hunting seasons, and coordinating browse surveys. harold understood the value of obtaining hunter cooperation to increase biological data, wisely starting the highly successful “moose hunter crest” program in 1967 to boost voluntary submission of lower moose jaws. harold also set deer seasons and helped to establish the provincial deer range management program. in 1972 harold moved to thunder bay and began a second career as professor in the school of forestry at lakehead university where he taught wildlife management to forestry students for more than 20 years (1972-1993). he also maintained a strong research program with much focus on the effect of moose browsing on forests and the impact of herbicides on moose habitat. several related publications by him and his students serve as the foundation for continued research today. during his latter tenure he concentrated on woodland caribou studies in the lake nipigon area of northwestern ontario. several of his students completed ms degrees in moose and caribou habitat-related research. harold was instrumental in developing alces into a respected, peer-reviewed international sci alces vol. 34 (1), (1998) ii entific journal by serving as the first chief editor of the annual proceedings of the north american moose conference and workshop 1978-82. he formed the first editorial committee to help arrange the timely publication and distribution of the annual proceedings and the successor journal alces by lakehead university in thunder bay, ontario. previously, each hosting jurisdiction was responsible for editing and publishing the annual issue and most took several years to fund and complete this task. thus, through harold’s leadership, hosting jurisdictions were no longer expected to produce the annual proceedings as production costs, page charges, and distribution arrangements were established at lakehead university and alces became financially independent. in an arrangement that continues to this day, harold had the lakehead university bookstore print and distribute alces through a special fiscal account. further, he wisely distributed sets of previous issues to three abstracting services to help advertise the availability of alces publications beyond the north american moose working group. harold’s efforts were instrumental at a pivotal time in establishing the long-term commitment at lakehead university to advance alces, a journal devoted to the biology and management of moose. his professional colleagues and students, and the current editorial board at alces recognize the outstanding contributions of harold cumming to the moose world and are deeply saddened by his passing. instructions for contributors to alces sentence, in which case it is spelled out. italics should only be used in the text for scientific names and statistical symbols. use the name-and-year system to cite published literature. cite references chronologically in the text. references – use large and small capitals for author’s last names and initials. do not use any abbreviations in the references. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. titles and all parts of tables must be typed doublespaced. tables must be constructed to fit the width of the page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use numerical superscripts to identify footnotes or asterisks for probabilities. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table columns must be generated with tab settings or a table editor. do not use spaces (i.e., the space bar). illustrations type figure captions on a separate page. identify each illustration by printing the author’s name and the figure number on the back in soft pencil. if necessary, also indicate the orientation of the illustration on the back. each illustration (either a photograph or linedrawn figure), must be of professional graphics quality, and reduced to fit into the area of either 1 (67 mm) or 2 (138 mm) columns of text by the author(s). letters and numbers on reduced figures must remain legible and be no less than 1.5 mm high after reduction. the same size and font of lettering should be used for all figures in the manuscript. photographs must be of high contrast and printed with a matte finish. typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original electronic graphics files or bitmap images (preferably as tagged image file format files) in an ibm-compatible format on 9-cm (3.5-inch) diskette or cd-rom. send manuscripts to: gerald redmond, submissions editor maritime college of forest technology hugh john flemming forestry centre 1350 regent street fredericton, new brunswick canada e3c 2g6 e-mail: gredmond@mcft.ca telephone: (506) 458 5128 fax: (506) 458 0652 editorial policy alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented at the annual north american moose conference and workshop, but works may be submitted directly to the editors at any time. reviewers judge submitted manuscripts on data originality, ideas, analyses, interpretation, accuracy, conciseness, clarity, appropriate subject matter, and on their contribution to existing knowledge. page charges current policies and charges are explained in a covering letter and invoice sent to authors with galley proofs. manuscript preparation authors should follow “manuscript guidelines for contributors to alces”, by rodgers et al. appearing in alces, vol. 34 (1): 1998 (available from the co-editors and associate editors). updates are posted on the alces web page; http: //bolt.lakeheadu.ca/~alceswww/alces.html. copy – please provide an electronic copy of the manuscript in ms word to the submissions editor. this copy should maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. double-space and leftjustify all text. except for the first page, number all pages consecutively, including tables and figure captions. revisions should be handled similarly. corresponding author do not use a title page. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlespaced in the upper left corner of the first page. if available, the author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (<10 words) begins left justified on the next line. type the title in upper-case bold letters. do not use abbreviations or scientific names in the title. abstract & key words following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. do not use abbreviations or literature citations. type alces vol. 00: 000 000 (0000), right justified on the line following the abstract. after leaving a single blank line, provide 6-12 key words in alphabetical order. footnotes use only in tables and at the bottom of the first page to provide the present address of an author when it differs from the address at the time of the study. style accompany the first mention of a common name with its scientific name. do not use scientific names for the names of domesticated animals or cultivated plants. use système international d’unités (si) units and symbols. use digits for numbers unless the number is the first word of a 142 editorial review committee our thanks to the following individuals who served as referees for alces volume 43. each paper was reviewed by at least 2 referees who judged its appropriateness for publication and provided editorial assistance. ed addison ecolink science, aurora, on cedric alexander vermont fish and wildlife, st. johnsbury, vt ken child northern region of bc hydro, prince george, bc vince crichton manitoba conservation, winnipeg, mb christian dussault ministère des ressources naturelles et de la faune du québec, québec, pq gordon eason ontario ministry of natural resources, wawa, on william faber central lakes college, brainerd, mn michael gillingham university of northern british columbia, prince george, bc mary hindelang michigan technological university, houghton, mi robert j. hudson university of alberta, edmonton, ab steve kilpatrick wy game and fish, jackson, wy murray lankester lakehead university (retired), thunder bay, on gerry lynch alberta environmental protection (retired), edmonton, ab brian mclaren lakehead university, thunder bay, on martha minchak minnesota department of natural resources, duluth, mn karen morris maine department of inland fisheries & wildlife, bangor, me brent patterson ontario ministry of natural resources, peterborough, on peter pekins university of new hampshire, durham, nh bill peterson minnesota department of natural resources (retired), grand marais, mn kim poole aurora wildlife research, nelson, bc derek quann parks canada, ingonish beach, ns gerry redmond maritime college of forest technology, fredericton, nb kris rines new hampshire fish and game department, new hampton, nh bruce roberts natural resources canada, corner brook, nl art rodgers ontario ministry of natural resources, thunder bay, on bill samuel university of alberta, edmonton, ab david scarpitti mass. division of fisheries and wildlife, westborough, ma robert serrouya columbia mountains caribou project, revelstoke, bc ed telfer canadian wildlife service (retired), edmonton, ab tim thomas wy game and fish, jackson, wy eric wald yukon delta wildlife refuge, bethel, ak rick ward yukon department of environment, whitehorse, yt gary wobeser univ. saskatchewan, saskatoon, sk donald young alaska department of fish and game, fairbanks, ak 204 kenneth n. child distinguished moose biologist 2009 recipient the distinguished moose biologist award was presented to kenneth n. child at the 44th north american moose conference and workshop, held at the university of idaho, in pocatello, idaho, usa, 14-17 june 2009, in recognition of his contribution to moose management. ken is a graduate of carleton university, ottawa (hons. b. sc. 1967), and the university of victoria, victoria, bc (m. sc. 1970). he spent 19 years (1973-1992) with the bc ministry of environment as a regional wildlife biologist in prince george responsible for the management of all big game species, but specialized in the management of moose. he retired from government service in 1992 and joined b.c. hydro corporation as the regional environmental co-ordinator for delivery of an environmental management system (iso 14000) for the production and transmission of electrical power in the province. during his years with bc hydro, ken remained involved in wildlife in the establishment and field delivery of major compensation programs for fish and wildlife resources in the both the peace-williston and columbia reservoirs and generating facilities. ken has published numerous peer-reviewed and popular articles and technical reports on a number of big game species in the canadian journal of zoology, journal of mammalogy, canadian veterinary journal, wildlife review, canadian field naturalist, and 12 papers in alces. he is also a contributing author in ecology and management of the north american moose (chapter 8, incidental mortality). he is currently volunteering his services on a moose-train collision working group with local stakeholders, government biologists, and the canadian national railway to address the mitigation of moose-train mortality along railway corridors in central british columbia. ken, as the regional wildlife biologist for the omineca subregion in british columbia, introduced a selective harvesting strategy for moose. after 28 years, the omineca continues to report the highest and most consistent harvest level of moose per km2 in the province: a testament to ken’s management efforts in the early 80’s. ken has been an active participant in the north american moose conference and workshop. he was a regular attendee and contributor to workshops and chaired the 19th workshop in prince george (1983) and co-chaired the 43rd workshop at the university of northern british columbia, prince george (2007). he plans to continue his research on moose at his leisure. ken has served as a peer reviewer for alces on numerous occasions over the years and even when in the employ of bc hydro. the north american moose conference and workshop is pleased to recognize ken child as the recipient of the distinguished moose biologist award for 2009 for his contributions to moose management and the successful introduction of selective harvesting strategies for moose in the omineca. p121-128_4109.pdf alces vol. 41, 2005 kreeger et al. – health assessment 121 health assessment of shiras moose immobilized with thiafentanil terry j. kreeger1, william h. edwards1, eric j. wald2, scott a. becker3, douglas brimeyer4, gary fralick4, and joel berger5 1wyoming game and fish department, 2362 highway 34, wheatland, wy 82201, usa; 2university of wyoming, department of renewable resources, 1000 e. university ave., laramie, wy 82071, usa; 3university of wyoming, department of zoology and physiology, 1000 e. university ave., laramie, wy 82071, usa; 4wyoming game and fish department, 420 n. cache, jackson, wy 83001, usa; 5 abstract: seventy-three (30 male, 43 female) free-ranging adult shiras moose (alces alces shirasi) were captured in southeastern and northwestern wyoming, blood sampled, and radio-collared in 2004 and 2005. moose were darted from the ground and air using 10 mg thiafentanil. blood samples were analyzed for hematology, serum chemistry, cortisol, and bacterial and viral serology. selected serum chemical parameters and cortisol were analyzed as indicators of physical exertion or physiological stress and none of these parameters suggested that moose were stressed as a result of capture. hematologic parameters were considered within normal limits. moose were serologically negative for brucella, leptospira and bovine respiratory syncytial virus. fecal and ear swab analysis and examination of the moose indicated that they were relatively free of ectoand endoparasites. three moose died within 30 days of capture for reasons probably associated with the capture effort. alces vol. 41: 121-128 (2005) key words: a-3080, alces alces shirasi, cortisol, hematology, immobilization, moose, naltrexone, parasites, serum chemistry, thiafentanil, wyoming shiras moose (alces alces shirasi) are the smallest subspecies of north american moose found in parts of wyoming, colorado, utah, idaho, montana, alberta, and british columbia (bubenik 1998). mortality of shiras moose in northwestern wyoming has subjectively appeared to increase in recent years. very few carcasses have undergone extensive necropsy because of their condition when found and no tentative diagnoses have been made. to examine this phenomenon further, a multi-year study has been undertaken to capture, sample, and track moose in an effort to assess survival and mortality factors. additionally, we wished to evaluate the tion of shiras moose. thiafentanil is a potent opioid that has been used to capture shiras moose (mcjames et al. 1994), but no information regarding physiological parameters of captured moose has been reported while using this drug. therefore, the purpose of this report was to obtain hematologic and serum chemical values to evaluate the health of captured shiras moose, establish reference values for future data collection, and evaluate thiafentanil as a capture drug for moose. methods capture of moose took place in northwestern wyoming in jackson hole, in the vicinity of moran junction, during february 2004 and 2005 and in southeastern wyoming in the snowy range region of the medicine bow national forest in december 2004. the southeastern moose were captured as part of a habitat utilization study but samples were health assessment – kreeger et al. alces vol. 41, 2005 122 taken to compare to the northwestern population. capture techniques included darting from the ground and aerial darting from a helicopter. only adult female moose were captured in february 2004 using ground approach. subsequent captures of both sexes were from the ground and the air. the (bushnell, overland park, kansas, usa) to ensure range accuracy while using co2-powered, adjustable dart guns (dan-inject north america, fort collins, colorado, usa). the helicopter darting utilized a .22-caliber blank dart gun (model 193, pneu-dart, williamsport, pennsylvania, usa) with open sights. all guns mm barbed needles (pneu-dart, williamsport, pennsylvania, usa). a pre-loaded dose of 10 mg thiafentanil (a-3080®, wildlife pharmaceuticals, fort collins, colorado, usa) was used for all moose, based on previous reports for shiras moose (mcjames et al. 1994). when helicopter capture was employed, ground crews were often utilized to locate and collect biological samples from moose. induction times and recovery times were measured by digital stopwatches. once immobilized, technicians blindfolded, radio collared, (telonics, inc., mesa, arizona, usa; advanced telemetry systems, isanti, minnesota, usa) and collected samples from moose. fecal samples and ear swabs were collected for parasitic evaluation, while blood samples were collected for: (1) serum chemical analyses (vetex, alfa wasserman, west caldwell, new jersey, usa); (2) hematologic oxford, connecticut usa); (3) cortisol concentration (immulite, diagnostic products corporation, los angeles, california usa); and (4) bacterial and viral serology. moose were given oxytetracycline antibiotics in the event that the dart caused infection (oxycure 200, vedco inc., st. joseph, missouri, usa). thiafentanil was antagonized with 300 mg naltrexone (trexonil®, wildlife pharmaceuticals, fort collins, colorado, usa) administered one-half intramuscularly and one-half subcutaneously. descriptive statistics were used to report means and standard errors along with upper analysis of variance was used in comparisons where appropriate. this study was approved by the university of wyoming animal care and use committee. results ten adult female moose in the northwest were immobilized in february 2004, 16 adult southeastern moose (5 male, 11 female) were captured in december 2004, and 47 adult moose (25 male, 22 female) in the northwest were captured in february 2005. not all analyses were conducted on all moose due to lost or poor quality blood samples. white blood counts (wbc) for the female moose captured in february 2004 were discarded due to laboratory error. physiological data among different groups could not be compared statistically because the conditions of capture and time to blood sampling could not be controlled. moose were pursued for 0.25 – 3.0 min before being darted and moose darted on the ground were usually blood sampled in < 5 min after induction whereas some moose darted from a helicopter were not located and sampled for > 60 min post induction. thus, only descriptive statistical data were reported based on sex and method and location of capture (tables 1 and 2). induction times for moose darted on the ground (2.4 ± 0.4 min) were generally lower than moose darted from helicopters (3.6 ± 0.2 min; table 3), but concentrations of the stress hormone, cortisol, did not appear to correlate with any pattern of capture method (table 4). immobilizations were characterized by moose remaining sternal, head upright, slight rigidity, and slight responsiveness to tactile stimulation. the mean recovery time (time from naltrexone administration to standing) for all groups was 2.9 ± 0.2 min. recoveries alces vol. 41, 2005 kreeger et al. – health assessment 123 table 1. serum chemical analyses of shiras moose chemically captured in wyoming. parameter (units) method1 sex (n) mean ± s.e. 95% c.l.2 albumin (g/dl) air male (27) 2.7 ± 0.1 2.5 – 2.8 ground male (3) 3.3 ± 0.2 2.5 – 4.1 air female (30) 3.1 ± 0.1 2.9 – 3.3 ground female (10) 3.5 ± 0.1 3.2 – 3.9 alkaline phosphatase (u/l) air male (27) 200.0 ± 15.3 168.4 – 231.4 ground male (3) 223.3 ± 50.0 8.4 – 438.2 air female (30) 258.2 ± 27.1 202.9 – 313.6 ground female (10) 257.1 ± 27.7 194.5 – 320.0 aspartate aminotransferase (u/l) air male (27) 64.7 ± 4.3 55.8 – 73.6 ground male (3) 65.0 ± 10.0 22.0 – 108.0 air female (30) 63.4 ± 2.6 58.2 – 68.6 ground female (10) 74.9 ± 4.1 65.6 – 84.1 blood urea nitrogen (mg/dl) air male (27) 4.1 ± 0.5 3.2 – 5.0 ground male (3) 4.7 ± 1.2 -0.5 – 9.8 air female (30) 3.6 ± 0.3 3.0 – 4.3 ground female (10) 4.2 ± 0.8 2.4 – 5.9 calcium (mg/dl) air male (27) 8.0 ± 0.3 7.4 – 8.6 ground male (3) 9.0 ± 0.2 8.1 – 9.9 air female (30) 8.8 ± 0.3 8.2 – 9.4 ground female (10) 10.1 ± 0.2 9.7 – 10.5 creatine kinase (u/l) air male (27) 125.5 ± 14.8 95.1 – 155.8 ground male (3) 103.3 ± 4.8 82.6 – 124.0 air female (30) 130.7 ± 15.1 99.8 – 161.6 ground female (10) 323.7 ± 112.0 70.4 – 577.0 gamma-glutamyl transferase (u/l) air male (27) 12.1 ± 1.6 8.8 – 15.4 ground male (3) 14.0 ± 2.1 5.0 – 23.0 air female (30) 14.1 ± 1.4 11.3 – 16.9 ground female (10) 16.1 ± 1.8 12.0 – 20.2 globulins (g/dl) air male (27) 3.4 ± 0.2 3.1 – 3.8 ground male (3) 3.7 ± 0.5 1.6 – 5.7 air female (30) 3.4 ± 0.1 3.1 – 3.7 ground female (10) 4.8 ± 0.4 4.0 – 5.6 lactate dehydrogenase (u/l) air male (27) 186.6 ± 12.5 160.9 – 212.3 ground male (3) 209.3 ± 18.0 132.0 – 286.7 air female (30) 202.6 ± 9.9 182.4 – 222.8 ground female (10) 236.7 ± 16.9 198.4 – 275.0 health assessment – kreeger et al. alces vol. 41, 2005 124 table 1. (continued...) serum chemical analyses of shiras moose chemically captured in wyoming. parameter (units) method1 sex (n) mean ± s.e. 95% c.l.2 magnesium (mg/dl) air male (27) 2.0 ± 0.1 1.8 – 2.2 ground male (3) 2.2 ± 0.2 1.4 – 3.1 air female (30) 2.2 ± 0.1 2.1 – 2.4 ground female (10) 2.7 ± 0.1 2.5 – 2.8 phosphorous (mg/dl) air male (27) 4.1 ± 0.2 3.6 – 4.6 ground male (3) 4.2 ± 0.6 1.6 – 6.9 air female (30) 4.0 ± 0.2 3.6 – 4.4 ground female (10) 4.2 ± 0.2 3.7 – 4.6 total protein (g/dl) air male (27) 6.0 ± 0.2 5.7 – 6.4 ground male (3) 6.9 ± 0.3 5.7 – 8.2 air female (30) 6.5 ± 0.2 6.1 – 7.0 ground female (10) 8.3 ± 0.2 7.8 – 8.8 1moose were darted with 10 mg thiafentanil either by ground personnel or from a helicopter. 2 table 2. hematologic analyses of shiras moose chemically captured in wyoming. parameter (units) method1 sex (n) mean ± s.e. 95% c.l.2 hematocrit (%) air male (25) 51.7 ± 1.0 49.7 – 53.7 ground male (3) 52.3 ± 1.2 47.2 – 57.4 air female (29) 52.7 ± 0.9 50.8 – 54.5 ground female (9) 52.2 ± 2.1 47.4 – 57.0 hemoglobin (g/dl) air male (25) 16.0 ± 0.4 15.1 – 16.9 ground male (3) 14.5 ± 1.0 9.9 – 19.0 air female (29) 16.9 ± 0.3 16.3 – 17.6 ground female (9) 16.4 ± 0.6 15.1 – 17.7 mean corpuscular hemoglobin concentration (g/dl) air male (25) 31.1 ± 0.9 29.2 – 33.0 ground male (3) 27.7 ± 1.5 21.3 – 34.1 air female (29) 32.3 ± 0.6 31.1 – 33.6 ground female (9) 31.6 ± 0.6 30.2 – 33.1 red blood count (x106 air male (25) 7.8 ± 0.2 7.5 – 8.1 ground male (3) 7.8 ± 0.1 7.2 – 8.4 air female (29) 7.9 ± 0.1 7.6 – 8.1 ground female (9) 7.4 ± 0.2 6.9 – 7.9 air male (25) 5355 ± 329 4677 – 6034 ground male (3) 6053 ± 1123 1219 – 10887 alces vol. 41, 2005 kreeger et al. – health assessment 125 parameter (units) method1 sex (n) mean ± s.e. 95% c.l.2 air female (28) 5801 ± 355 5074 – 6529 ground female (3) 4460 ± 574 1992 – 6928 air male (25) 1739 ± 134 1462 – 2016 ground male (3) 1566 ± 326 165 – 2968 air female (28) 1769 ± 110 1543 – 1995 ground female (3) 1765 ± 355 239 – 3291 air male (25) 3233 ± 229 2760 – 3706 ground male (3) 3930 ± 963 -217 – 8077 air female (28) 3582 ± 283 3000 – 4162 ground female (3) 2378 ± 483 300 – 4457 air male (25) 176 ± 17 140 – 212 ground male (3) 242 ± 45 48 – 436 air female (28) 182 ± 16 150 – 215 ground female (3) 73 ± 16 4 – 141 air male (25) 193 ± 40 111 – 276 ground male (3) 315 ± 183 -476 – 1106 air female (28) 266 ± 40 183 – 350 ground female (3) 243 ± 100 -187 – 674 platelets (x10-5 air male (25) 214 ± 17 179 – 249 ground male (3) 180 ± 43 -3 – 363 air female (29) 187 ± 13 159 – 214 ground female (3) 134 ± 26 22 – 247 table 2. (continued...) hematologic analyses of shiras moose chemically captured in wyoming. 1moose were darted with 10 mg thiafentanil either by ground personnel or from a helicopter. 2 were characterized by moose standing and calmly walking away. moose were negative for antigens against brucella, leptospira, infectious bovine rhinotracheitis virus, bovine viral diarrhea virus, syncytial virus. no southeastern moose had evidence of endoparasites, but 3 moose had a few dermacentor albipictus ticks present. fecal examination of northwestern moose indicated a low infection of nematodirus roundworms (< 8 eggs/gm) in 10 moose and trichuris in 1 moose. no moose had evidence of ear mites and some had a few dermacentor albipictus ticks. one female moose died 9 days post capture. there was no apparent cause of death and no problems associated with the capture event for this moose were noted. two males were found dead at 3 weeks post-capture with gross evidence of pneumonia in one (discolored lungs, adhesions) and malnutrition (depleted bone marrow) in the other. six other moose died > 4 weeks post capture; evidence suggested that 3 were possibly killed by wolves (canis lupus), mountain lion (felis concolor), health assessment – kreeger et al. alces vol. 41, 2005 126 and grizzly bear (ursus arctos), 1 was possibly due to natural causes (no evidence of predation found), and 2 were unknown due to scavenging of the carcasses. discussion serum chemical and hematologic values for shiras moose. the many variables associated with the collection of these data rendered statistical comparisons inappropriate both within this study as well as with other reports. data collection variables included sex, sample size, location, date, method of capture, and time from induction to sampling. this latter variable may have been the most troublesome because blood values, which may have changed in response to capture method for instance, may have reverted closer to baseline as time to sampling increased. nonetheless, serum chemical and hematologic values for shiras moose were subjectively similar to most values for other moose (franzmann et al. 1977, franzmann and leresche 1978, forbes et al. 1996). thiafentanil appeared to be an effective immobilizing drug for moose. induction times for moose darted on the ground (2.4 ± 0.4 min) were faster than moose darted on the ground from this same region with carfentanil and xylazine (4.4 ± 1.9 min; roffe et al. 2001). moose invariably became recumbent in a sternal position in a semi-rigid state (figure 1). this characteristic was desirable because moose that roll onto their sides often regurgitate and subsequently develop aspiration pneumonia (kreeger 2000). for example, the male that died from pneumonia 3 weeks post capture had rolled over from an initial sternal position and regurgitation was noted. the use of only thiafentanil in this drug regimen without the addition of tranquilizers, such as xylazine, supported previous studies, which showed that use of the opioids alone increased the probability of moose remaining sternal (kreeger 2000). the sternal position with the head raised also enhanced blood sampling and method1 sex (n) mean ± s.e. 95% c.l.2 induction (min) air male (25) 3.5 ± 0.3 2.9 – 4.0 ground male (3) 2.9 ± 0.8 -0.8 – 6.5 air female (25) 3.7 ± 0.4 3.0 – 4.4 ground female (11) 2.2 ± 0.3 1.5 – 2.9 recovery (min) air male (27) 2.6 ± 0.2 2.2 – 2.9 ground male (2) 2.1 ± 0.9 -9.0 – 13.2 air female (29) 2.7 ± 0.2 2.3 – 3.0 ground female (13) 4.2 ± 0.8 2.4 – 6.0 table 3. induction and recovery times of shiras moose chemically captured in wyoming. 1moose were darted with 10 mg thiafentanil either by ground personnel or from a helicopter. 2 method1 sex (n) mean ± s.e. 95% c.l.2 air male (27) 4.4 ± 0.3 3.8 – 5.1 ground male (3) 4.7 ± 0.2 3.8 – 5.6 air female (30) 4.6 ± 0.2 4.2 – 5.1 ground female (10) 4.5 ± 0.5 3.3 – 5.7 of shiras moose chemically captured in wyoming. 1moose were darted with 10 mg thiafentanil either by ground personnel or from a helicopter. 2 alces vol. 41, 2005 kreeger et al. – health assessment 127 radio collar attachment. it should be noted that opioids (carfentanil, thiafentanil) resulted in immobilization as opposed to anesthesia (kreeger et al. 2002). the prime characteristic of general anesthesia is loss of consciousness and this does not occur with opioids. moose (and other cervids) will respond to tactile stimulation (attaching ear tags, blood sampling, fecal sampling, and loud sharp noises). handlers should be aware of this phenomenon and either hobble the animal or be aware that it can jerk its head or feet and may even stand, although it will become recumbent again on its own. we measured cortisol concentrations to analyze any stress response to methods of capture. we hypothesized that moose chased and darted from helicopters would be more stressed than those darted from the ground. serum cortisol concentrations have been historically measured as an indicator of stress (matteri et al. 2000). when data from all groups were combined and compared, cortisol concentrations of moose darted from (p = 0.98) as moose darted from a helicopter between a balanced group (e.g., northwest female moose darted from the ground and air) were made, the concentrations were still the p = 0.95). cortisol concentrations for these moose were similar to unstressed alaskan moose (bubenik et al. 1994). the explanation for these data remains elusive. it was possible that cortisol concentrations in the supposedly stressed helicopter darted groups subsided to baseline before blood sampling occurred. however, this seemed unlikely because the cortisol response to stress (simulated by acth administration) in alaskan moose (bubenik et al. 1994) resulted in cortisol concentrations being elevated above controls for > 2 hr and all moose in the current study were sampled in < 2 hr. another explanation could be that the serum test employed in this study did not measure cortisol accurately. this also seemed unlikely because, even though not validated for moose cortisol, these commercial radioimmunoassay tests for cortisol have provided consistently appropriate results across species (kreeger et al. 1990, 1992; bubenik et al. 1994). it was also possible that these moose simply were not stressed by the capture methods, although moose are physiologically capable of generating classic endocrine stress responses under certain handling conditions (franzmann et al. 1975). we considered that the 3 moose that died within 30 days of being captured succumbed to some sequela of the capture event. the male moose that died from pneumonia was an obvious result of being captured. moose that become laterally recumbent under anesthesia often regurgitate rumen contents which are aspirated, resulting in pneumonia and death (kreeger 2000). the malnourished moose also probably died after being captured because moose in poor physical condition due to sickness, injury, or malnutrition are high immobilization risks and often die subsequent to capture (kreeger et al. 2002). the female moose that died of unknown causes was in good physical condition and laboratory analyses suggested no underlying pathogens but, because she died shortly after capture, it fig. 1. shiras moose demonstrating typical sternal posture resulting from immobilization with thiafentanil. this posture reduces the possibility of rumen regurgitation with subsequent aspiration and aids in blood sampling and radio-collar attachment. health assessment – kreeger et al. alces vol. 41, 2005 128 appeared that this event precipitated her death. moose that died > 30 days post-capture most likely died from reasons not directly related to the capture event, although this cannot be proven. mortality of shiras moose in northwest wyoming has increased in recent years due to unknown causes, which was in part why this current capture and collaring effort was initiated. the data gathered herein will provide a basis for future comparison and analysis for shiras moose. acknowledgements we wish to acknowledge the efforts of several wyoming game and fish department personnel in the capture effort and the wyoming state veterinary laboratory for diagnostic services. funding support came from the wyoming game and fish department, wyoming governors big game license coalition, animal damage management board, teton county conservation district, university of wyoming, wyoming department of transportation, and wildlife heritage foundation. references bubenik, a. b. 1998. evolution, taxonomy and morphology. pages 77-124 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. bubenik, g. a., c. c. schwartz, and j. carnes. 1994. cortisol concentrations in male alaskan moose (alces a. gigas) after exogenous acth administration. alces 30:65-69. forbes, l. b., s. v. tessaro, and w. lees. 1996. experimental studies on brucella abortus in moose (alces alces). journal of wildlife diseases 32:94-104. franzmann, a. w., a. flynn, and p. d. arneson. 1975. serum corticoid levels relative to handling stress in alaska moose. canadian journal of zoology 53:1424-1426. _____, _____, and t. n. bailey. 1977. serial blood chemistry and hematology values from alaskan moose. journal of zoo and wildlife medicine 8:27-37. _____, and r. e. leresche. 1978. alaskan moose blood studies with emphasis on condition evaluation. journal of wildlife management 42:334-351. kreeger, t. j. 2000. xylazine-induced aspiration pneumonia in shiras moose. wildlife society bulletin 28:751-753. _____, j. m. arnemo, and j. p. raath. 2002. handbook of wildlife chemical immobilization. international edition. wildlife pharmaceuticals incorporated, fort collins, colorado, usa. _____, u. s. seal, and e. d. plotka. 1992. influence of hypothalamic-pituitary-adrenocortical hormones on reproductive hormones in gray wolves (canis lupus). journal of experimental zoology 264:3241. _____, p. j. white, u. s. seal, and j. r. tester. 1990. pathological responses of red foxes to foothold traps. journal of wildlife management 54:147-160. matteri, r. l., j. a. carroll, and c. j. dyer. 2000. neuroendocrine responses to stress. pages 43-76 in g. p. moberg and j. a. mench, editors. the biology of animal stress. cabi publishing, new york, new york, usa. mcjames, s. w., j. f. kimball, and t. h. stanley. 1994. immobilization of moose with a-3080 and reversal with nalmefene hcl or naltrexone hcl. alces 30:21-24. roffe, t. j., k. coffin, and j. berger. 2001. survival and immobilizing moose with carfentanil and xylazine. wildlife society bulletin 29:1140-1146. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 180 distinguished moose biologist award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public's attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (>10 in journals including alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. 6. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual's interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3-8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special "distinguished moose biologist" presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http://bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca 177 previous meeting sites of the north american moose conference and workshop 1963 st. paul, minnesota 1964 st. paul, minnesota 1966 winnipeg, manitoba 1967 edmonton, alberta 1968 kenai, alaska 1970 kamloops, british columbia 1971 saskatoon, saskatchewan 8th 1972 thunder bay, ontario 9th 1973 québec city, québec 10th 1974 duluth, minnesota 11th 1975 winnipeg, manitoba 12th 1976 st. john’s, newfoundland 13th 1977 jasper, alberta 14th 1978 halifax, nova scotia 15th 1979 soldotna kenai, alaska 16th 1980 prince albert, saskatchewan 17th 1981 thunder bay, ontario 18th 1982 whitehorse, yukon territory 19th 1983 prince george, british columbia 20th 1984 québec city, québec 21st 1985 jackson hole, wyoming 22nd 1986 fredericton, new brunswick 23rd 1987 duluth, minnesota 24th 1988 winnipeg, manitoba 25th 1989 st. john’s, newfoundland 26th 1990 regina and ft. qu’apelle, saskatchewan 27th 1991 anchorage and denali national park, alaska 28th 1992 algonquin park, ontario 29th 1993 bretton woods, new hampshire 30th 1994 idaho falls, idaho 31st 1995 fundy national park, new brunswick 32nd 1996 banff national park, alberta 33rd 1997 fairbanks, alaska in conjunction with the 4th international moose symposium 34th 1998 québec city, québec 35th 1999 grand portage, minnesota 36th 2000 whitehorse, yukon territory 37th 2001 carrabassett valley, maine 38th 2002 hafjell, norway in conjunction with the 5th international moose symposium 39th 2003 jackson hole, wyoming 40th 2004 corner brook, newfoundland and labrador 41st 2005 whitefish, montana 42nd 2006 baddeck, nova scotia 43rd 2007 prince george, british columbia 44th 2008 6th international moose symposium, yakutsk, russia 45th 2009 pocatello, idaho 46th 2010 international falls, minnesota future meetings 47th 2011 jackson hole, wyoming 48th 2012 bialowieza, poland in conjunction with the 7th international moose symposium f:\alces\vol_39\p65\3817.pdf alces vol. 39, 2003 bottan et al. choice modelling and moose management 27 a choice modelling approach to moose management: a case study of thunder bay moose hunters brian bottan1, len hunt2,3, and wolfgang haider4 1539 mcintosh street, thunder bay, on, canada p7c 3a1; 2ontario ministry of natural resources, centre for northern forest ecosystem research, 955 oliver road, thunder bay, on, canada p7b 5e1; 4simon fraser university, 8888 university drive, burnaby, bc, canada v5a 1s6 abstract: we demonstrate the application of one type of model available for managers to better understand the people side of resource management. this choice modelling approach allows us to study issues such as the hunting site choices of moose hunters. to showcase the approach, we use a case study based on predicting the site choices of resident moose (alces alces) hunters from the thunder bay area. our case study shows that resident moose hunters of thunder bay prefer short travel distances, few encounters with other hunters, areas with better vehicular accessibility, more moose, more water, and shorter regenerating vegetation in harvested areas. we demonstrate the practical applicability of the model by examining a hypothetical scenario involving the issue of hunting site closures in areas with new forest cutovers. the results of this hypothetical scenario demonstrate that one can use the model to: (1) predict changes to moose hunting effort associated with a site restriction; and (2) estimate the economic losses that would arise to hunters from this restriction. a manager should seek both of these pieces of information before implementing a change such as a site restriction. alces vol. 39: 27-39 (2003) key words: choice model, economic value, experiment, human dimensions, hunter behaviour, moose hunting, preferences resource management has changed considerably over the past 20 years to embrace an ecosystem perspective (slocombe 1993, grumbine 1994). this shift in emphasis to a holistic view of forested environments has also encouraged the view that people are part of ecosystems. as such, it is more important than ever to manage resources with a mindful eye on the uses and desires of the public. for moose (alces alces) management, this creates a difficult problem for managers. on the one hand, there is a need to meet the demands of the hunting public. on the other hand, there is a need to control hunting pressure to ensure that moose populations are healthy and sustainable. to meet this careful balance, a moose manager requires effective information on both the desires of hunters and the reactions that hunters may exhibit towards changes that affect the moose hunting experience. a study in northwestern ontario by bottan et al. (2001) collected both sources of information. in this paper, we focus solely on the results of a choice modelling exercise designed to determine the factors that lead hunters to select different areas to hunt moose. we showcase the usefulness of this approach by discussing the model results and by presenting a fictitious example of how the model can be applied in a management context. in the example, we will show that one can use a choice model to estimate 3corresponding author. choice modelling and moose management bottan et al. alces vol. 39, 2003 28 changes in both hunting effort and economic values stemming from a management change. this modelling approach permits managers to forecast the likely effects of various management scenarios without actually implementing the scenarios on the landscape. in many situations, such as the call to limit hunting in areas with new cutovers, the approach offers information without possible confrontations with the hunting public. the choice model of northwestern ontario resident moose hunters emulates hunter behaviour on a hunting site scale that is finer than the wildlife management unit level. this choice of scale acknowledges that each management unit consists of a highly variable landscape that affords moose hunters with many different settings from which to choose a site to hunt. it is also important to emphasize that the study focuses solely on local moose hunters. it is expected that non-local and non-resident moose hunters will evaluate characteristics of a moose hunting site differently. therefore, we suggest that readers avoid the temptation of concluding that all moose hunters in ontario are captured by this study. the paper is organized as follows. the next section provides an introduction to choice modelling and a review of relevant choice model studies on moose hunting. this section is followed by a discussion of the methods used to collect data from hunters and to model the behavioural site choices of hunters. the third section discusses the results of the study, followed by the presentation of a fictitious scenario that will illustrate the application of the model results. finally, we conclude with a discussion that highlights key points from the paper. choice modelling basics and relevant studies choice models work from the simple premise that the behaviours of individuals convey important information. for example, a hunter believes that his/her chosen site will yield the greatest net benefits of all available sites. one method of measuring net benefits is through utility, which is a measure of happiness or aggregate preference. by using utility, we can assume that a hunter's site choice is governed by or mimics utility maximization (i.e., he/she selects the site with greatest utility). the utility of a hunting site is determined through the attributes (e.g., travel distance, moose abundance) that characterize that hunting site. to convert these attribute measures into utility, an individual must weight (i.e., parameterize) the attribute measures and combine these weighted attributes together in some fashion. a simple, but very popular, method to combine these weighted attributes is to add them together. this addition, which is consistent with information integration theory (anderson 1981), suggests that a high weighted score for one attribute may offset a low weighted score for another attribute. this means that the model explicitly permits individuals to trade-off desirable and undesirable attributes when making a choice decision. although an individual is always expected to choose the hunting alternative with maximum utility, researchers do not observe the utility measures from the hunters. as well, researchers accept that despite their efforts to learn about the process, they do not know and cannot model all aspects of the process that leads a hunter to select a hunting site. therefore, a researcher can only estimate a probability that a hunter would select a particular hunting site. it is under this foundation that choice modellers apply random utility theory (thurstone 1927). the researcher's task of estimating weights for each of the attributes is complicated by the uncertainty described above. alces vol. 39, 2003 bottan et al. choice modelling and moose management 29 to estimate the attribute weights, researchers must turn towards a statistical model, which requires assumptions about the uncertainty. a basic statistical model that is often used by choice modellers is the conditional multinomial logit regression model (see equation 1 below). in equation 1, the probability of individual n selecting alternative i from a set of c n alternatives equals the exponentiated attribute measures x in that are weighted by parameters ß i , which are estimated via maximum likelihood estimation. this exponential value is divided by the sum of the exponentiated values of all alternatives to produce the choice probability. the µ term, which relates to the variance of the utility scale, is not identifiable along with the ß estimates. however, a researcher can innocuously assume that the µ term equals 1 without consequence to the predictions from the model. (1) while the conditional multinomial logit is a very restrictive model, researchers often use this model because of its simplicity (louviere et al. 2000). the model is well suited to handle the discrete choices made by individuals for behaviours such as hunting site choice. the estimation of the model provides the weights (i.e., parameters) for the attributes that are necessary to calculate the probability that an individual will choose any alternative (i.e., a choice probability). as with any regression model, one can use the conditional multinomial logit regression model for forecasting. for moose hunting, the forecasts permit individuals to estimate how changes to one or more hunting sites (e.g., a site closure) may affect the choices for all hunting sites. choice models were originally estimated from actual choices (i.e., revealed preferences) made by individuals (e.g., past hunti n g t r i p s ) . h o w e v e r , l o u v i e r e a n d woodworth (1983) illustrated how researchers could also estimate these models from hypothetical choices (i.e., stated preferences). one may question the wisdom of conducting a study on what people say they will do rather than what they have done. there are, however, several reasons why a stated preference choice model may provide a better approach than would a revealed preference choice model (louviere et al. 2000). most of these reasons exploit the hypothetical nature of the stated preference choice model. for example, since the choice task is hypothetical, one can construct the choice task provided to individuals to follow an experimental design plan that contains good properties for statistical estimation. furthermore, one can stretch the range of attribute measures beyond existing levels to estimate how these levels may affect choices. in other words, we can use a stated preference choice model to evaluate conditions that do not currently occur on the landscape, but may occur as a result of management actions (e.g., a restriction on the use of all-terrain vehicles for hunting). resource economists have almost exclusively driven the application of choice models in outdoor recreation. this popularity among economists exists since choice models provide a convenient method to estimate changes to economic value for nonmarket goods such as hunting. for hunting, economists can use the forecasting ability of the model to estimate how a scenario (e.g., a site closure) may affect the value that hunters derive from hunting. the first applications of choice models and hunting were conducted on bighorn s h e e p ( o v i s c a n a d e n s i s) i n a l b e r t a (adamowicz et al. 1990, coyne and adamowicz 1992). other efforts on hunting by choice modellers include waterfowl (creel & loomis 1992), red deer (cervus nnj j n cjci e e ip ∈∀∈= ∑ = ,,)( 1 jjn iin ßx ßx µ µ choice modelling and moose management bottan et al. alces vol. 39, 2003 30 elaphus) (bullock et al. 1998), white-tailed deer (odocoileus virginianus) (schwabe et al. 2001), pronghorn (antilocapra americana) (boxall 1995), and general hunting (hausman et al. 1995). however, moose hunting has attracted the most interest among researchers (boxall et al. 1996, adamowicz et al. 1997, akabua et al. 1999, akabua et al. 2000, boxall and macnab 2000, haener et al. 2000, 2001). the above studies have uncovered several attributes deemed important by moose hunters when making a site choice, such as travel distance, evidence of moose, and encounter levels with other hunters. while most studies have found that vehicular accessibility was an important determinant of site choice, some studies suggest that poorer accessibility is preferred while others suggest it is not preferred. finally, a forest disturbance attribute has yielded mixed results in the various studies. in some instances, the authors concluded that the presence of logging reduced the site attractiveness for hunters (boxall and macnab 2000, haener et al. 2000). however, this result seems incongruent with the belief that hunters seek out logged areas to conduct their hunts. we feel that the problem in measuring the impact of forest disturbance on hunting site choice results from the poor descriptions of logged areas that other studies have applied. even when forest disturbance was measured in detail (akabua et al. 1999), the unit of analysis focused on a management unit level that was probably too coarse of a scale to model the importance of forest harvesting to moose hunters. in contrast, our study will overcome these previous limitations of research on moose hunting by examining the importance of forest harvest related site characteristics that are relevant to hunters. the inclusion of a description of the height of the regenerating vegetation should be more relevant to moose hunters than would be descriptions about the presence or absence of logged areas. sarker and surry (1998) also recommended that future social and economic research in ontario should concentrate on understanding the effects of forest management practices on the environmental settings preferred by moose hunters. methods in the fall of 1998, a mail survey of licenced moose hunters from the ontario ministry of natural resources' thunder bay district was undertaken. the initial survey was mailed to 1,000 randomly chosen hunters during the middle of the moose hunting season. this timing allowed for a better recall of hunting experiences by the moose hunters. survey implementation followed the total design method of dillman (1978) to maximize response rate. the total design method suggests that after the initial mail-out, a postcard reminder be mailed 1 week later, followed by another survey package to non-respondents 2 weeks after the postcard reminder. the response rate achieved was 63.5%, and we conducted no checks for non-response bias. in comparison, boxall and macnab (2000) reported a response rate of 49% for saskatchewan hunters who were also surveyed by mail. interested readers are referred to bottan (1999) and bottan et al. (2001) for detailed summaries of all survey results from the thunder bay respondents. a key aspect for conducting a stated preference choice modelling study is to determine a list of relevant attributes for the behaviour in question. when combined with an experimental design plan, it is also important to determine appropriate levels that the attributes may take. our list of hunting site attributes and attribute levels were developed after a careful review of the previously described literature, a focus group with hunters, and discussions with academics, resource management biologists, alces vol. 39, 2003 bottan et al. choice modelling and moose management 31 wildlife specialists, and foresters. table 1 describes the 7 attributes and associated levels used for this study. while many other attributes are likely to affect site choices by moose hunters (e.g., tag allocation), we attempted to simplify the choice task for the respondents by holding all regulations constant. to explore the potential demand for new hunting opportunities, one level from 4 attributes represented an environmental or social condition that seldom exists. based on these formal and informal discussions, we were confident that the choice experiment balanced the presentation of relevant information to hunters while minimizing the burden on the respondents. the survey task required respondents to choose one alternative from 2 hypothetical hunting alternatives and the option of not hunting (fig. 1). to properly estimate the attribute weights, the experimental design required us to obtain information from 27 different choice tasks like figure 1. each respondent received 1 of 3 survey versions that contained 9 of the 27 different hypothetical choice tasks. before respondents reached the choice task, each survey booklet contained attribute definitions and an example of how to answer the choice task. table 1. definition of attributes and associated attribute levels. attribute definition level distance the approximate 1-way distance (kms) 1 = 350km1 from the hunter’s home to the hunting area 2 = 250km 3 = 150km access approximate access conditions by a 2wd 1 = 70% of area by 2wd vehicle within the hunting area (all areas 2 = 50% of area by 2wd were assumed to be 4x4 accessible) 3 = 30% of area by 2wd encounters the number of encounters with other 1 = 4 or more other hunting parties hunting parties during a day’s moose 2 = 1-3 other hunting parties hunting within the area 3 = no other hunters1 lakes presence of lakes within hunting area 1 = many lakes 2 = few lakes 3 = no lakes1 moose evidence of moose seen during a day’s 1 = >3 moose per day1 moose hunting within the area based on 2 = 1 2 moose per day seeing or hearing moose or seeing fresh 3 = <1 moose per day sign such as tracks or droppings height height of regeneration growing in cutovers 1 = >2m within hunting area (meters) 2 = 1 2m 3 = <1m forest type predominant type of forest regeneration 1 = conifer growing in cutovers within hunting area 2 = hardwood 1denotes an atypical level for the attribute. choice modelling and moose management bottan et al. alces vol. 39, 2003 32 parameter estimates associated with the levels of an attribute do not statistically differ from zero, one can conclude that the attribute in question has no effect on site choice. in our model, at least one of the parameter estimates associated with any given attribute was statistically different from zero. one may notice that some attribute levels do not have parameter estimates. in two cases (i.e., distance and access), we estimated one single parameter estimate based on the quantitative values of the attribute levels. for the remaining attributes, which were specified at nominal levels only, parameter estimates could only be obtained for 2 of the 3 attribute levels, with the third level equal to the negative for a full discussion of stated preference choice models, experimental design, and attribute coding, the interested reader is referred to louviere et al. (2000) or bennett and blamey (2001), who provide a less technical discussion. results the data were analyzed with a conditional multinomial logit regression model using limdep 7.0 software (green 1998). table 2 presents the parameter estimates from this regression model along with asymptotic t-test values. the asymptotic ttests are large sample property t-tests that assess whether a parameter estimate differs significantly from zero. if all of the fig. 1. an example of a choice task provided to respondents. 24a. if you were to select a new hunting area, and these were the only two options available, which one would you choose on your next hunting trip, if either? features of hunting area area a area b 1-7 distance from home to hunting area (one way) hunting area accessibility by vehicle type: 2wd 4wd (or atv) frequency of encounters with other hunters presence of lakes moose population: evidence of forest characteristics cutovers: height of new growth predominant forest regeneration check one and only one box 150 kilometers 70% by 2wd 100% by 4wd no other hunters many lakes one moose every 2 or more days 3-6ft tall (1-2m) conifer neither site a or site b i will not go moose hunting 150 kilometers 50% by 2wd 100% by 4wd 1-3 other hunting parties many lakes 3 or more moose per day less than 3ft tall (<1m) hardwood alces vol. 39, 2003 bottan et al. choice modelling and moose management 33 sum of the other two parameter estimates. the quality of the regression model was assessed through an adjusted mcfadden's rho statistic. however, this statistic is not at all analogous to the well understood r2 term from linear ordinary least squares regression and the value of 0.15 for our study is very acceptable. table 2 also presents the partworth utility estimates for all attribute levels. the partworth utility represents the weighted contribution to the utility of an alternative that any attribute level provides. the partworth utilities are calculated by multiplying the coding for an attribute level by the relevant parameter estimate(s) (e.g., the partworth utility for the zero encounters level equals the negative sum of the parameter estimates from the 4 or more and 1 to 3 levels for encounters). in this sense, the partworth utilities are somewhat redundant, but we include them since they provide the best summary of the results. high partworth utilities increase the likelihood that a hunter would select a particular hunting site. while it is tempting to use the partworth utilities to pass judgment on the most important attributes, one must remember that the partworth utilities are likely to be affected by the range of levels associated with an attribute. for example, the partworth utilities for travel distance would probably be table 2. statistical model results and partworth utility estimates for attributes and levels. attribute level parameter estimate t-statistic partworth utility intercepts no hunting -0.2718** -8.27 -0.2718 generic not identifiable not identifiable 0.0000 travel distance linear estimate -0.0080** 27.47 not applicable 350 km not applicable not applicable -0.8002 250 km not applicable not applicable 0.0000 150 km not applicable not applicable 0.8002 encounters 4 or more -0.5444** -16.58 -0.5444 1-3 -0.0042 -0.13 -0.0042 0 not identifiable not identifiable 0.5486 accessibility linear estimate 0.0028* 2.00 not applicable 70% by 2wd not applicable not applicable 0.0561 50% by 2wd not applicable not applicable 0.0000 30% by 2wd not applicable not applicable -0.0561 lakes many lakes 0.2982** 9.33 0.2982 few lakes 0.1275** 3.99 0.1275 no lakes not identifiable not identifiable -0.4257 moose evidence 3 or more 0.3475** 11.01 0.3475 1 2 per day 0.1421** 4.51 0.1421 <1 per day not identifiable not identifiable -0.4896 regeneration height >2m -0.2359** 7.25 -0.2359 1-2m 0.0301 0.97 0.0301 <1 m not identifiable not identifiable 0.2058 vegetation conifer -0.0669* -2.03 -0.0669 hardwood not identifiable not identifiable 0.0669 * p<0.05; ** p<0.01. choice modelling and moose management bottan et al. alces vol. 39, 2003 34 much different if we chose levels of 50, 300, and 550km, respectively. the model appears to provide a good explanation of hunter preferences, as all partworth utilities follow a priori expectations. distance acts as a strong deterrent to the choice of a hunting site by a resident hunter from the thunder bay area. there is a significant positive relationship for a larger proportion of the hunting area being 2-wheel drive accessible, although that relationship is not as strong as the distance effect. as one might expect, the number of expected daily encounters with other hunting parties was negatively related to hunting site choice. sites in which many lakes were present yielded a positive preference, suggesting that respondents preferred to have an abundance of water present in the area they chose to hunt moose. intuitively, respondents were more likely to select an area if it had evidence of many moose. it was also revealed that areas with shorter heights of regenerating forest were preferred to areas that had regeneration heights exceeding 2 meters. lastly, respondents had a positive preference for hardwood as opposed to conifer vegetation that was regenerating in cutovers. fictitious forest management scenario this section will demonstrate two aspects about the managerial usefulness of a choice modelling approach. first, we illustrate how an individual can use the choice model results through a forecasting model to estimate the likely consequences of a change to the hunting environment on the distribution of hunting effort. second, we demonstrate how an individual can translate a change to a hunting environment into a change in economic value for hunting trips. many researchers and managers have proposed restricting access into new cutovers until suitable cover for moose is available to reduce moose vulnerability to hunters (eason et al. 1981, tomm et al. 1981, timmermann and gollat 1983, eason 1985, ferguson et al. 1989, rempel et al. 1997). while such a policy may achieve certain desirable ecological goals, the implications of such a policy change for moose hunters has never systematically been investigated. below, we use the results from table 2 to examine fictitious scenarios whereby one hunting site moves through 3 stages; from undisturbed, to a logged area that is open for hunting, and finally to an area that is closed to hunting. we purposely chose this fictitious scenario to demonstrate the usefulness of the model without becoming engaged in a debate about the assumptions we make regarding the scenarios. the scenarios we chose involved 6 hypothetical areas available to moose hunters along with the option of not hunting. table 3 describes these hunting areas by the attributes and attribute levels that we used to estimate our choice model. the bottom 3 rows of the table highlight the expected use of the respective areas by our thunder bay resident moose hunters. the choice probabilities (i.e., the last 3 rows in table 3) were calculated as follows. first, we replaced the verbal descriptions of each hunting site in table 3 by the partworth utilities from table 2. for the distance and accessibility attributes, the partworth utilities were obtained by multiplying the associated linear parameter estimate from table 2 by the difference between the value in table 3 and its mean value (i.e., 250 for distance and 50 for accessibility). second, for each hunting site, the partworth utilities were summed and the sum of the no hunting alternative was set to the partworth utility for the do not hunt alternative. third, we took the exponent of these summed values and summed all 7 of these values. finally, we divided the exponent sum for any alternative by the sum alces vol. 39, 2003 bottan et al. choice modelling and moose management 35 obtained from all 7 alternatives. the resulting proportions were converted into the percentages shown in table 3. in the before harvest scenario, the model predicted that about 19% of hunter effort would have occurred in site #6. after introducing the forest harvest in site #6, which would alter the regeneration height to less than 1 meter, the model predicted that hunter effort in site #6 would increase to around 27%. this predicted increase does not account for the fact that hunters may see more evidence of moose per day as a result of the forest harvest. while we did not consider this change to provide evidence of moose in our scenario, the user of this model is free to make whatever assumptions she/he likes about changes to attributes. it should also be noted that the relative changes to hunting sites in table 3 are identical among the unaffected hunting sites. this is a direct consequence of the independence of irrelevant alternatives (iia) property (luce 1959) of the multinomial logit model. while this rigid substitution pattern appears unrealistic, it is an empirical question whether this property holds for a given data set. the next scenario involves a closure of hunting in areas with new cutovers (i.e., site #6). the after closure of site #6 row in table 3 predicts how this closure may impact the use of the remaining 5 hunting areas along with the no hunting alternative. the table demonstrates that individuals can use a choice model to predict the impacts of management changes on the spatial distribution of hunting effort. furthermore, this forecasting model permits managers to investigate a suite of scenarios without having to implement the scenarios on the landscape. besides providing information about the redistribution of hunting effort, one can also use a choice model to determine the change in economic value of hunting associated with the site closure scenario presented above. we restrict our attention to the change in economic value that may arise from the hunting site closure after the forest harvest in site #6. if our model included some monetary attribute such as a fee or cost, we could directly estimate economic values. without a monetary attribute, we resort to an indirect method of valuation that employs the travel distance attribute table 3. simulation of closing a hunting site in an area with new cutovers. attribute site #1 site #2 site #3 site #4 site #5 site #6 not hunt one way travel distance (km) 150 175 190 165 180 160 encounters (per day) 4+ 4+ 4+ 4+ 4+ 4+ accessibility (% 2 wheel drive) 70 55 30 60 40 60 lakes few many none few few many evidence of moose (per day) <1 1-2 1-2 <1 3+ 1-2 regeneration height (m) >2 1-2 >2 1-2 1-2 >2 to <1 vegetation type conifer conifer conifer conifer conifer conifer predicted hunting effort (%) before harvest to site #6 9.63 22.02 6.77 10.84 21.00 19.29 10.45 after harvest of site #6 8.70 19.89 6.11 9.79 18.97 27.11 9.44 after closure of site #6 11.94 27.28 8.38 13.43 26.02 closed 12.95 choice modelling and moose management bottan et al. alces vol. 39, 2003 36 weight. in our scenario, we estimate that a hunter would have been willing to drive an additional 39.5km in 1-way travel distance to have avoided the restriction on hunting in site #6. this compensating km value was obtained by: (1) calculating the summed exponent values as described earlier for the scenarios with and without the site closure to site #6; (2) taking the natural logarithm for both of these summed exponent values; (3) subtracting these logarithm values; and (4) dividing this difference by minus one times the travel distance parameter estimate in table 2. the 39.5km travel distance is translated into dollars by multiplying this extra round trip distance by a suitable per km cost for operating a vehicle. even if we choose a reasonable value such as $0.35 per km, the loss per trip to the hunter would have equaled $27.66 for the round trip. we could also add to this amount, costs for the additional travel time associated with each trip multiplied by the value that hunters place on their travel time. clearly, this economic information would be of great importance to managers who must follow the careful balance of limiting hunting success, yet providing quality hunting opportunities. discussion lyon (1987: 289) suggested more than a decade ago that "when possible the relationships between participation, experience quality, and those site characteristics that can be managed, such as crowding, hunter success, and access, should be quantified and used to guide management decisions". by adopting a choice modelling approach, we have taken a step in that direction. more importantly, rather than investigating each environmental and social effect on hunting separately, the method permits a more holistic investigation that yields valuable estimates relating to use and to value associated with changes to hunting experiences. as with any modelling approach, the model does require validation with empirical data. our study has demonstrated that the behaviours of thunder bay area resident moose hunters are likely to be affected by a number of attributes. the model results illustrate that these hunters have preferences for shorter travel distances, fewer encounters with other hunters, greater vehicular accessibility, greater abundance of moose, more water, cutovers with short regenerating vegetation, and areas with hardwood tree species. besides identifying these preferences, the choice modelling approach provides a unifying method of linking behavioural theory to these preferences. the results of validated choice modelling studies may be used to forecast changes in hunting effort and economic values through a tradeoff approach espoused by the model. a fictitious forest management scenario was presented in this study to illustrate the ability of a choice model to answer two relevant questions to managers. first, we showed how the model could be used to estimate the expected redistribution of hunting effort arising from changes to the management of the resource. this information is important since managers need to be aware of the likely consequences of shifting hunting effort into other areas when deciding to restrict access or to change other management aspects in one or more hunting areas. managers could also use the approach to examine the tradeoffs that hunters may make between stricter regulations and better quality hunting experiences. second, we showed how one could use a choice model to estimate hunters' changes in economic values stemming from management changes. again this change in economic value provides managers with a better understanding of the costs that the hunting public would likely endure as a result of a specific management direction. alces vol. 39, 2003 bottan et al. choice modelling and moose management 37 one further positive aspect of the choice modelling approach based on hypothetical behaviours is that we may estimate the consequences of a wide suite of management scenarios without actually implementing these scenarios. besides the excessive cost of field experiments, many scenarios that managers wish to explore may invoke confrontation with hunters and their stakeholder representatives. therefore, the choice model permits resource managers to gauge the consequences of many scenarios without invoking a highly politicized response from the hunting public. in summary, our study provides some new human dimension information to managers. however, some caveats exist that reflect our inability to understand and to model the process that leads to hunting behaviours. for example, we examined hunting site choice in a static environment that does not take into consideration season, habits, or success. as well, we did not examine the relationships between regulations (e.g., tag allocation levels) and other hunting site attributes. finally, there may be several other attributes that influence hunting effort and the attribute levels specified in this study may not be suitable for every context (e.g., number of encounters on the opening week of the season). however, there is a tradeoff between model complexity and respondent burden, and we opted for data collection that would keep the response task as simple as possible for the hunters. we feel these caveats need to be understood by readers. however, we do not believe that these caveats take away from the overall positive contribution of our study. no one has the hubris to assume that they know all aspects of any biological or social process. we accept that our ability to understand hunting behaviour is incomplete and we provide much additional information to a growing body of literature. for example, our emphasis on the height of regenerating vegetation in cutovers, which has been ignored by past researchers, was found to be very important to moose hunters. additionally, the choice modelling perspective embraces researcher uncertainty directly into the model. it is exactly for this reason that the model is probabilistic rather than deterministic in its predictions. acknowledgements we would like to thank the ontario ministry of natural resources for funding this research. we would also like to thank the many individuals who assisted us in the development of this study and the hunters who kindly provided the necessary information to conduct this study. finally, we thank gord eason and torstein storaas for their constructive and thorough reviews of an earlier manuscript. any remaining errors are, of course, the responsibility of the authors. references adamowicz, w. l., s. jennings, and a. g. coyne. 1990. a sequential choice model of recreation behavior. western journal of agricultural economics 15:9199. , j. f. swait, p. c. boxall, j. j. louviere, and m. williams. 1997. perceptions versus objective measures of environmental quality in combined revealed and stated preference models of environmental valuation. journal of environmental economics and management 32:65-84. akabua, k. m., w. l. adamowicz, and p. c. boxall. 2000. spatial non-timber valuation decision support systems. forestry chronicle 76:319-327. , , w. e. phillips, and p. trelawny. 1999. implications of realization uncertainty on random utility models: the case of lottery rationed hunting. canadian journal of agriculchoice modelling and moose management bottan et al. alces vol. 39, 2003 38 tural economics 47:165-179. anderson, n. h. 1981. foundations of information integration theory. academic press, new york, new york, usa. bennett, r., and j. blamey. 2001. the choice modelling approach to environmental valuation. edward elgar, northampton, massachusetts, usa. bottan, b. j. 1999. exploring the human dimension of thunder bay moose hunters with focus on choice behaviour and environmental preferences. m.sc.f. thesis, lakehead university, thunder bay, ontario, canada. , l. m. hunt, w. haider, and a. r. rodgers. 2001. thunder bay moose hunters: environmental characteristics and choice preferences. cnfer technical report tr-007. ontario ministry of natural resources, thunder bay, ontario, canada. boxall, p. c. 1995. the economic value of lottery-rationed recreational hunting. canadian journal of agricultural economics 43:119-131. , w. l. adamowicz, j. f. swait, m. williams, and j. j. louviere. 1996. a comparison of stated preference methods for environmental valuation. ecological economics 18:243-253. , and b. macnab. 2000. exploring the preferences of wildlife recreationists for features of boreal forest management: a choice experiment approach. canadian journal of forest research 30:1931-1941. bullock, c. h., e. a. elston, and n. a. chalmers. 1998. an application of economic choice experiments to a traditional land use – deer hunting and landscape change in the scottish highlands. journal of environmental management 5:335-351. coyne, a. g., and w. l. adamowicz. 1992. modelling choice of site for hunting bighorn sheep. wildlife society bulletin 20:26-33. creel, m., and j. loomis. 1992. recreation value of water to wetlands in the san joaquin valley: linked multinomial logit and count data trip frequency models. water resources research 28:25972606. dillman, d. a. 1978. mail and telephone surveys: the total design method. john wiley and sons, toronto, ontario, canada. eason, g. 1985. overharvest and recovery of moose in a recently logged area. alces 21:55-75. , r. j. thomas, and k. oswald. 1981. moose hunting closure in a recently logged area. alces 17:111-125. ferguson, s. h., w. e. mercer, and s. m. oosenbrug. 1989. the relationship between hunter accessibility and moose condition in newfoundland. alces 25:36-47. green, w. h. 1998. limdep version 7.0. econometric software incorporated, plainview, new york, new york, usa. grumbine, e. 1994. what is ecosystem management? conservation biology 8:27-38. haener, m., p. c. boxall, and w. l. adamowicz. 2000. modeling recreation site choice: do hypothetical choices reflect actual behavior? american journal of agricultural economics 83:629642. , d. dosman, w. l. adamowicz, and p. c. boxall. 2001. can stated preference methods be used to value attributes of subsistence hunting by aboriginal peoples? a case study in northern saskatchewan. american journal of agricultural economics 83:629-642. hausman, j. a., g. k. leonard, and d. mcfadden. 1995. a utility-consistent, combined discrete choice and count data model: assessing recreational losses due alces vol. 39, 2003 bottan et al. choice modelling and moose management 39 to natural resource damage. journal of public economics 56:1-30. louviere, j. j., d. a. hensher, and j. f. swait. 2000. stated choice methods: analysis and applications. cambridge university press, new york, new york, usa. , and g. woodworth. 1983. design and analysis of simulated consumer choice or allocation experiments: an approach based on aggregated data. journal of marketing research 20:350367. luce, r. 1959. individual choice behavior. wiley, new york, new york, usa. lyon, j. r. 1987. basic and applied social research needs in wildlife management. pages 285-295 in d. j. decker and g. r. goff, editors. valuing wildlife: econ o m i c a n d s o c i a l p e r s p e c t i v e s . westview press, colorado, usa. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. sarker, r., and y. surry. 1998. economic value of big game hunting: the case of moose hunting in ontario. journal of forest economics 4:29-60. schwabe, k. a., p. w. schumann, r. boyd, and k. doroodian. 2001. the value of changes in deer season length: an application of the nested multinomial logit model. environmental and resource economics 19:131-147. slocombe, d. s. 1993. implementing ecosystem-based management. bioscience 43:289-303. thurstone, l. 1927. a law of comparative judgment. psychological review 34:273-286. timmermann, h. r., and r. gollat. 1983. age and sex structure of harvested moose related to season manipulation and access. alces 18:301-328. tomm, h.o., j. a. beck, and r. j. hudson. 1981. responses of wild ungulates to logging practices in alberta. canadian journal of forest research 11:606-614. alces vol. 45, 2009 hundertmark genetic diversity in introduced populations 137 reduced genetic diversity in two introduced and isolated moose populations in alaska kris j. hundertmark institute of arctic biology and department of biology and wildlife, university of alaska fairbanks, po box 757000, fairbanks, alaska 99775, usa. kris.hundertmark@alaska.edu abstract: i examined indices of genetic diversity in 2 isolated moose (alces alces) populations in alaska that were founded by low numbers of individuals to determine effects of founding and infer whether subsequent gene flow has occurred with surrounding moose populations. kalgin island is a small, predator-free island in cook inlet that was founded by 6 moose (3 females) in the late 1950s; its population has since undergone dramatic fluctuations. berners bay is an isolated population along the coast of southeastern alaska that was founded by 21 calves introduced in 1958-1960. genetic attributes of those populations were compared to a population in yukon flats in central alaska that served as an outbred control. indices from 11 microsatellite markers indicated substantial effects of founding and subsequent isolation. heterozygosity and allelic diversity, both of which are reduced by genetic bottlenecks, were significantly lower in the introduced populations than the yukon flats population. kalgin island diversity was significantly lower than that for berners bay, and was likely due to the smaller founding size and subsequent population fluctuations. neither introduced population exhibited evidence of gene flow from surrounding populations. managers should consider the isolation of those populations when assessing risks to population viability and crafting management strategies. alces vol. 45: 137-142 (2009) key words: alaska, alces alces, bottleneck, gene flow, insular, introduced population, moose. one of the primary concerns of conservation biology is the loss of genetic diversity through genetic drift in small populations. in cases where populations are isolated, thus preventing immigration from neighboring populations, genetic drift occurs at a maximum rate depending on population size. loss of diversity from drift is compounded in populations that are founded by low numbers of individuals due to demographic and genetic bottlenecks. loss of genetic diversity has been related to loss of fitness (reed and frankham 2003). in ungulates, studies have found correlations between indices of diversity and reduction in juvenile survival (coulson et al. 1999, mainguy et al. 2009, silva et al. 2009), variation in horn/antler growth (scribner and smith 1990, von hardenberg et al. 2007), and parasite resistance (coltman et al. 1999). thus, genetic diversity of populations should be a primary concern in ungulate management, particularly with small and isolated populations. valuable insight may be gained from studying wild populations with known demographic histories to determine effects of population size on genetic diversity. introduced populations often act as natural experiments in that regard, particularly when founding population size is known, as well as demographic trends since founding. i compared 2 small, isolated, introduced populations of moose (alces alces) in alaska to determine the effect of their respective demographic histories on indices of genetic diversity, and to infer the degree to which gene flow has affected diversity. i also compared those populations to an outbred moose population to demonstrate the extent to which the introduced populations have lost diversity. genetic diversity in introduced populations hundertmark alces vol. 45, 2009 138 study area moose populations in kalgin island and berners bay, alaska have similar histories. they both were established through introduction of individuals from southcentral alaska in the late 1950s. the population on kalgin island (60º 27’n, 152º 00’w) was founded by 6 moose (3 females) transported to the island in 1957-1959 (burris and mcknight 1973). the population in berners bay (58º 45’n, 134º 50’w) was founded by 15 calves in 1958 and 6 additional calves in 1960 (burris and mcknight 1973). moreover, both populations are seemingly isolated from neighboring moose populations. kalgin island is located in cook inlet which is characterized by strong tidal currents that have kept large mammals, including predators, from colonizing the island. nonetheless, the short distance from mainland to island (<10 km) has fueled speculation that gene flow is possible. berners bay is separated from neighboring moose populations by rugged coastline characterized by mature spruce-hemlock (picea sitchensistsuga heterophylla) forest that is avoided by moose (hundertmark et al. 1990). predators of moose occurring in berners bay are wolves (canis lupus), brown bears (ursus arctos), and black bears (u. americanus). the moose population on kalgin island has undergone dramatic fluctuations in population size due to density-dependent effects of habitat and periods of intense harvest, increasing in size to an estimated 212 individuals in 1982, declining to 8 in 1986, and increasing since then (bowyer et al. 1999). the berners bay population is stable and thought to be close to carrying capacity at 120-150 individuals (barten 2008). both populations support limited harvest. the moose population in yukon flats (66º 10’ n, 149º 00’ w) occurs in lowland boreal forest along the yukon river in central alaska. although the population exists in a large contiguous area of suitable moose habitat, it occurs at extremely low density (caikoski 2008) presumably due to predation (bertram and vivion 2002) and poaching of female moose (caikoski 2008). nonetheless, yukon flats is an open population as opposed to the presumably closed nature of the berners bay and kalgin island populations. although the yukon flats population was not the source of founders for either berners bay or kalgin island, it serves as a good example of an outbred alaskan moose population and should serve as a suitable control population in lieu of samples from south-central alaska. indeed, schmidt et al. (2009) found little difference in levels of diversity among 6 moose populations distributed widely within alaska, demonstrating that location of the population is less important than demographic history. methods samples for genetic analysis were acquired either as tissue from hunters (kalgin island and berners bay) or as blood samples from captured animals (yukon flats). samples for each population were collected within a single year. dna extraction and genotyping were conducted under contract in one of two laboratories: kalgin island and berners bay samples were analyzed at wildlife genetics international (nelson, british columbia, canada), whereas yukon flats moose were analyzed at the department of biological sciences, university of alberta (calgary, alberta, canada). two samples from yukon flats were also analyzed by wildlife genetics international to ensure consistency in allele calling and warrant comparison of genotypes from the two labs. i analyzed 19 samples from kalgin island, 8 from berners bay, and 28 from yukon flats. loci bl42, bm4513, bm888, bm1222, bm203, bm848 (bishop et al. 1994), fcb193 (buchanan and crawford 1993), rt5, rt9, rt24, and rt30 (wilson et al. 1997) were used to characterize genetic diversity. populations were tested to ensure compliance with hardyweinberg equilibrium using an exact chialces vol. 45, 2009 hundertmark genetic diversity in introduced populations 139 square test implemented in software genepop (raymond and rousset 1995). diversity was expressed as allelic richness (number of alleles per locus, a), observed heterozygosity (ho) and expected heterozygosity (he) based on allele frequencies assuming hardy-weinberg equilibrium. because estimates of allelic richness are related to sample size, we standardized our estimates by using rarefaction (kalinowski 2005) to express the expected number of alleles per locus based on a sample of 8 individuals from each population (the smallest sample size of our 3 populations). estimates of allelic richness were compared between population pairs using a sign test. estimates of inbreeding (fis) were calculated by genepop. population differentiation was assessed via nei’s unbiased genetic distance (nei 1978), pairwise fst, and comparison of allele frequencies of populations. significance of differences based on pairwise fst estimates was estimated from a permutation test conducted in software fstat (goudet 1995). significance of pairwise comparisons of allele frequencies was conducted as a chi-square test for each locus and significance values were combined via fisher’s method (fisher 1948) by genepop to compute a population-wide significance level. results and discussion all populations were in hardy-weinberg equilibrium (kalgin island: χ2 = 15.4, p = 0.75; berners bay: χ2 = 24.5, p = 0.14; yukon flats: χ2 = 18.8, p = 0.66). all loci were polymorphic in yukon flats whereas one locus (rt9) was monomorphic in both kalgin island and berners bay populations. all populations differed from each other in allelic richness (p < 0.001; table 1). the 3 populations shared at least one allele at each locus, but allele frequencies differed in pairwise comparisons (p < 0.0001). the extent of the reduction in diversity undergone by kalgin island and berners bay populations is illustrated by the occurrence of private alleles (those occurring in only one population; table 1), wherein yukon flats had 17 alleles not found in the other populations. kalgin island exhibited the least diversity, as measured by allelic richness and heterozygosity, followed by berners bay and yukon flats (table 1). inbreeding coefficients for all populations were very close to zero, indicating no evidence of inbreeding. all populations differed from each other in pairwise comparisons of genetic distance and fst (table 2); the largest differences were between the 2 introduced populations. it is striking that kalgin island and berners bay populations differed to such a degree considering that their founding individuals came from the same general area. sample size of the berners bay population was less than ideal but the samples were obtained opportunistically and there was a low probability of obtaining additional samples from the few hunters that harvest moose there. nonetheless, estimates of observed and expected heterozygosity, as well as fis and fst that are based on heterozygosity, are not related to sample size and would not be expected to change predictably with an increase population n a a8 private alleles ho he fis kalgin island 19 2.9 2.7 1 0.47 0.45 -0.01 berners bay 8 3.1 3.1 2 0.53 0.49 -0.03 yukon flats 28 5.5 4.2 17 0.67 0.64 -0.02 table 1. indices of genetic diversity for 11 microsatellite loci measured in 3 alaskan moose populations. kalgin island and berners bay populations show limited diversity due to founder events and lack of gene flow with neighboring populations. parameters are: n = sample size, a = allelic richness (alleles/locus), a8 = estimate of a standardized to a sample size of 8, ho = observed heterozygosity, he = expected heterozygosity, and fis = inbreeding coefficient. genetic diversity in introduced populations hundertmark alces vol. 45, 2009 140 in sample size. of the indices of diversity that i examined, only allelic richness is affected by sample size (kalinowski 2005, pruett and winker 2008), which is why i used a rarefaction method so that estimates of richness among populations could be compared. results from a simulation study indicated that mean estimates of ho and he did not change significantly for sample sizes ranging from 5-100 individuals, and that estimates were more consistent over a range of sample sizes for populations with low genetic diversity as compared with high diversity (pruett and winker 2008). clearly, moose populations in kalgin island and berners bay show severely reduced diversity relative to yukon flats, a result that arguably stems from genetic bottlenecks associated with introduction. moreover, kalgin island moose were significantly less diverse that berners bay moose; this difference is likely a function of the smaller founding size of the kalgin island population combined with its marked fluctuations. genetic differentiation among the 3 populations was much greater than that reported among 6 moose populations from alaska (fst range = 0.014-0.109; schmidt et al. 2009). that study found a remarkable lack of differentiation among moose across a large geographic scale, and was contrary to other studies (broders et al. 1999, wilson et al. 2003). the relative difference among the 3 study populations would be unexpected if the kalgin island and berners bay populations were open and exchanged individuals with neighboring populations. thus, these differences presumably indicate the strong effect of founding combined with genetic drift associated with isolation from neighboring populations. reduction in heterozygosity associated with founder effect can be calculated as: he = ho(1 – 1/2n). where: ho is the heterozygosity of the source population, he is the expected heterozygosity of the founded population at the time of founding, and n is the number of individuals introduced. using yukon flats as a proxy for the source population (ho = 0.67; table 1), kalgin island with a founding size of 6 would have he = 0.61 which is 30% greater than ho for kalgin (0.47; table 1). similarly, he for the initial population in berners bay would be expected to be 0.66 or 24% higher than observed (0.53; table 1). the differences between the current heterozygosity in the introduced populations and the expected heterozygosity based on number of founders can be explained by genetic drift occurring in the interim. moreover, the severity of a genetic bottleneck is directly related to the duration of the bottleneck (nei et al. 1975), suggesting that both populations grew slowly after founding, thus extending the length of the bottleneck; the kalgin island population was likely affected by a second bottleneck when the population declined abruptly to the founding size (8 individuals) in the 1980s. interpopulation distances reported for moose in canada (broders et al. 1999) were much lower than those reported here; distances were 0.013-0.298, however the latter value was essentially a comparison of 2 different subspecies representing moose from cape breton island, nova scotia (introduced from alberta) and moose from the avalon peninsula, newfoundland (introduced from nova scotia kalgin island berners bay yukon flats kalgin island 0.518 0.365 berners bay 0.301 0.31 yukon flats 0.197 0.151 table 2. indices of population differentiation based on 11 microsatellite markers measured in 3 alaskan moose populations. nei’s (1978) unbiased genetic distance is above the diagonal and fst is below the diagonal. based on fst estimates, all populations differ significantly (p < 0.001). alces vol. 45, 2009 hundertmark genetic diversity in introduced populations 141 and new brunswick). distances reported by wilson et al. (2003) were comparable to those reported here; however, the largest fst value (0.3013) occurred between populations in newfoundland and riding mountain national park, manitoba and was a comparison across a subspecies boundary. thus, the level of differentiation observed between the 2 introduced alaska populations was equal to or greater than that observed between subspecies elsewhere on the continent. comparisons of levels of diversity between studies employing different sets of molecular markers require caution because of the different levels of variation inherent in different loci. nonetheless, if loci are truly neutral and markers conform to the same model of mutation, estimates of population differentiation should be broadly comparable between studies. i have shown that 2 small, introduced moose populations in alaska underwent extreme reductions in genetic diversity associated with founding and subsequent genetic drift. it is highly unlikely that either population experiences gene flow with neighboring populations; otherwise, the level of diversity in the introduced populations would be greater and more similar with that of yukon flats. managers of the kalgin island and berners bay populations should consider the degree of isolation and paucity of genetic variation in those populations when assessing risks to population viability and crafting management strategies. as an example, recovery of diversity can probably be accomplished only with introduction of additional individuals rather than relying on immigration from surrounding populations. acknowledgements funding was provided by a grant from federal aid in wildlife restoration to the alaska department of fish and game. i thank n. barten and k. white for providing samples of berners bay moose. references barten, n. l. 2008. unit 1c moose. pages 27-52 in p. harper, editor. moose management report of survey-inventory activities 1 july 2005-30 june 2007. alaska department of fish and game, juneau, alaska, usa. bertram, m., and m. vivion. 2002. moose mortality in eastern interior alaska. journal of wildlife management 66: 747-756. bishop m. d., s. m. kappes, j. w. keele, r. t. stone, s. l. f. sunden, g. a. hawkins, s. solinas-toledo, r. fries, m. d. gross, j. yoo, and c. w. beattie. 1994. a genetic linkage map for cattle. genetics 136: 619-639. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73-89. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8: 1309-1315. buchanan f. c., and a. m. crawford. 1993. ovine microsatellites at the oarfcb11, oarfcb128, oarfcb193, oarfcb266 and oarfcb304 loci. animal genetics 24: 145. burris, o. e., and d. e. mcknight. 1973. game transplants in alaska. game technical bulletin 4. alaska department of fish and game, juneau, alaska, usa. caikoski, j. r. 2008. units 25a, 25b, and 25d moose. pages 617-647 in p. harper, editor. moose management report of survey-inventory activities 1 july 2005-30 june 2007. alaska department of fish and game, juneau, alaska, usa. coltman, d. w., j. g. pilkington, j. a. smith, and j. m. pemberton. 1999. parasite-mediated selection against inbred soay sheep in a free-living, island population. evolugenetic diversity in introduced populations hundertmark alces vol. 45, 2009 142 tion 53: 1259-1267. coulson, t., s. albon, j. slate, and j. pemberton. 1999. microsatellite loci reveal sex-dependent responses to inbreeding and outbreeding in red deer calves. evolution 53: 1951-1960. fisher, r. a. 1948. combining independent tests of significance. american statistician 2: 30. goudet, j. 1995. fstat (version 1.2): a computer program to calculate f-statisitcs. journal of heredity 86: 485-486. hundertmark, k. j., w. l. eberhardt, and r. e. ball. 1990. winter habitat use by moose in southeastern alaska: implications for forest management. alces 26: 108-114. kalinowski, s. t. 2005. hp-rare 1.0: a computer program for performing rarefaction on measures of allelic richness. molecular ecology notes 5: 187-189. mainguy, j., s. d. côté, and d. w. coltman. 2009. multilocus heterozygosity, parental relatedness and individual fitness components in a wild mountain goat, oreamnus americanus population. molecular ecology 18: 2297-2306. nei, m. 1978. estimation of heterozygosity and genetic distance from a small number of individuals. genetics 89: 538-590. _____, t. maruyama, and r. chakraborty. 1975. the bottleneck effect and genetic variability in populations. evolution 29: 1-10. pruett, c. l., and k. winker. 2008. the effects of sample size on population genetic diversity estimates in song sparrows melospiza melodia. journal of avian biology 39: 252-256. raymond m., and f. rousset. 1995. genepop (version 1.2): population genetics software for exact tests and ecumenicism. journal of heredity 86: 248-249. reed, d. h., and r. frankham. 2003. correlation between fitness and genetic diversity. conservation biology 17: 230-237. schmidt, j. i., k. j. hundertmark, r. t. bowyer, and k. g. mccracken. 2009. population structure and genetic diversity of moose in alaska. journal of heredity 100: 170-180. scribner, k. t., and m. h. smith. 1990. genetic variability and antler development. pages 460-473 in g. a. bubenik and a. b. bubenik, editors. horns, pronghorns and antlers. springer-verlag, new york, usa. silva, a. d., j.-m. gaillard, n. g. yoccoz, a. j. m. hewison, m. galan, t. coulson, d. allaine, l. vial, d. delorme, g. van laere, f. klein, and g. luikart. 2009. heterozygosity-fitness correlations revealed by neutral and candidate gene markers in roe deer from a long-term study. evolution 63: 403-417. von hardenberg, a., b. bassano, m. festabianchet, g. luikart, p. lanfranchi, and d. coltman. 2007. age-dependent genetic effects on a secondary sexual trait in male alpine ibex, capra ibex. molecular ecology 16: 1969-1980. wilson, g. a., c. strobeck, l. wu, and j. w. coffin. 1997. characterization of microsatellite loci in caribou (rangifer tarandus), and their use in other artiodactyls. molecular ecology 6: 697-699. wilson, p. j., s. grewal, a. rodgers, r. rempel, j. saquet, h. hristienko, f. burrows, r. peterson, and b. n. white. 2003. genetic variation and population structure of moose (alces alces) at neutral and functional dna loci. canadian journal of zoology 81: 670-683. f:\alces\vol_38\pagemaker\drk15 alces vol. 38, 2002 ben-david et al. stable isotopes in moose and caribou 219 utility of stable isotope analysis in studying foraging ecology of herbivores: examples from moose and caribou [alces 37(2): 421-434, 2001] merav ben-david1, einav shochat2, and layne g. adams2 1department of zoology and physiology, university of wyoming, laramie, wy 82071, usa; 2u.s. geological survey, biological resources division, 1011 tudor rd., anchorage, ak 99503, usa alces vol. 38: 219-220 (2002) key words: alaska, alces alces, amino acids, caribou, denali national park, moose, rangifer tarandus, seasonal diets, δ13c, δ15n erratum: due to a problem with the resolution of symbols used in figures 4 and 5 of the above paper (p. 427 of volume 37, issue number 2), those figures were not accurately reproduced in all printed copies of the alces journal. figures 4 and 5 are reprinted below using new symbols. please compare these figures with those published previously for accuracy. fig. 4. stable isotope ratios of blood cells from moose and caribou collected in denali national park and preserve, alaska, usa. black solid symbols represent late summer-autumn values for caribou in 1993 and moose in 1998. light shaded closed symbols represent winter values for caribou in 1993 and moose in 1998. open circles represent winter values for caribou in 1998. isotopic ratios of blood cells from moose and caribou were significantly different from each other in all seasons (k nearest neighbor randomization test, p < 0.001). values of δ15n were not significantly different between winter and summer for moose (anova, p = 0.4), or caribou (anova, p = 0.06). in both species, a significant enrichment of 0.5-0.6‰ in δ13c occurred in winter (anova, p = 0.03 for moose, and p < 0.001 for caribou). • • • • • •• • •• • • • • • • • • • •••• • • • ° ° °° °° °° ° ♦ ♦ ♦ ♦ ♦ ♦ ♦ ♦♦ ♦♦ ♦ ♦ ♦ ♦ ♦ ♦♦ ♦ ♦ ♦ ♦ ♦♦ ♦ ♦♦ ♦ ♦♦ -2 -1 0 1 2 3 4 5 6 -28 -26 -24 -22 -20 δ13c δ15n caribou moose stable isotopes in moose and caribou ben-david et al. alces vol. 38, 2002 220 fig. 5 – stable isotope ratios of blood cells from moose and caribou collected in denali national park and preserve, alaska, usa, in autumn and spring 1993 and 1998, plotted against the range of predicted isotopic ratios for herbivores (for symbol designations see figure 4). stable isotope ratios of moose and caribou were within the predicted range of values for each herbivore based on plant isotope ratios and trophic fractionation. caribou values did not register inside the area of overlap indicating that trees and shrubs contributed less to the diet of these herbivores than other components of their diet. • ••••••••• ••••••••• ••••••• ° ° °°°° °° ° ♦♦♦ ♦ ♦♦ ♦ ♦♦ ♦♦ ♦ ♦ ♦ ♦♦ ♦♦♦♦ ♦♦♦ ♦♦ ♦♦♦ ♦♦ -10 -5 0 5 10 15 20 -40 -30 -20 -10 δ13c δ15n herbivore feeding on trees, shrubs, and aquatic plants herbivore feeding on herbs, mushrooms and lichens herbivore feeding on herbs and 50% lichens f:\alces\supp2\pagema~1\rus29s. alces suppl. 2, 2002 anthropogenic effects – zablotskaya and zablotskaya 131 anthropogenic effects on moose populations in the southern taiga lidia v. zablotskaya and maria m. zablotskaya prioksko–terrasny biosphere reserve, 142474, danki, moscow region, russia abstract: this article focuses on variations in moose population densities and sex ratios, autoregulation of its population density, and related effects on the forest since a moose outbreak in 2 central parts of the east–european plain due to the appearance of early successional tree species, resulting from felling in the course of world war ii. alces supplement 2: 131-135 (2002) key words: anthropogenic effects, east–european plain, migration, moose population density, soils, southern taiga, winter forage investigation of the population dynamics of moose was conducted in 1949–1986 on the left bank of the oka river in the prioksko–terrasny reserve and surrounding forests. the left–bank south–facing slopes of the oka river have a varied vegetation. mixed and coniferous forests of various types predominate. by the early 1950s over 30% of the oka forests were young stands of aspen, birch, pine, and oak. young stands up to the age of 20 years covered over 14,000 ha; pine accounted for about 3,000 ha in old pine forests. the pine regenerated on a large scale, and there was a well–developed regrowth of juniper, mountain ash, and other trees. the flood plain of the oka and its tributaries had abundant willow thickets, and a rapid increase in moose numbers started there between 1950 and 1952, 8–10 years after mass felling (1942–1943). in 1960, moose density reached 97.7 animals per 1,000 ha (fig. 1). the rapid growth of moose populations was promoted by the abundance of winter forage, the low harvest of this species in russia over many years (no more that 2– 3% of the population), and migration of moose from more northern regions (zhirnov 1967). over the decade of 1950–1960, moose density increased 18–fold. after the peak of 1960, the population began to decline (zablotskaya 1964) until the 1980s. the wave of high density lasted 18–20 years (fig. 1), its highest level persisting for 9–10 years. in 1952–1955, when the population density of moose exceeded 25–30 moose/ 1,000 ha, the heavy injury by moose of early successional trees and shrubs became noticeable. in 1959, moose winter forage in the oka forest averaged only 256 kg per ha (koryakin 1961). in 1961, young pines, aspens, junipers, and willows injured by moose began to dry and perish. the reserves of primary winter forage for moose in the forests were practically destroyed (fig. 1). during these years moose appeared starved, and calves born late were weakened and died. the severe damage to trees and shrubs inflicted by moose led to a sharp increase in moose harvest. intensive harvest (up to 40% of the moose population) started during the 1966 season and continued for 4 years. harvesting moose in the reserve area was also permitted because of the vast destruction of pine. since autumn 1961 in the reserve (4,945 ha) and neighboring game management units, 230 – 370 animals were anthropogenic effects – zablotskaya and zablotskaya alces suppl. 2, 2002 134 the reduction of moose (fig. 3). in a number of regions, e.g., in mordovia (according to m. n. borodina, personal communication), the rise in the numbers of moose and the increase in wolf populations proceeded concurrently. mass breeding of moose strongly affected the regeneration and composition of forests. in the area of the oka forests at cutovers, dry pines were replaced by abundant regrowth of birch. in pine forests with herbs, the pine regrowth that died due to disturbance by moose was replaced by spruce regrowth. in pine forests with green moss, pine regrowth turned into shrubs because of constant browsing, this form persisting for over 30 years. despite the low numbers of moose, the oka forests can perish due to a lack of reliable regrowth. due to loss of winter forage, moose immediately browsed all young pines rising over the snow cover. we expect long–lasting effects of this moose foraging behavior on the regeneration of pine and juniper. destruction of the ancient oka forests can be prevented only by eliminating moose from their winter habitats in pine forests for at least 2 decades. the exceptional rise in numbers and expansion of range of such large mammals as moose is only possible in unbalanced forest habitats in the absence of appropriate harvest, handicapping the growth of moose populations as was the case in the 1940s and 1950s. in primary climax ecosystems no such reproduction of big ungulates in large areas can exist. not infrequently, local short–term rises in moose numbers are common natural components of succession and are delayed 8 – 10 years following regeneration of forest on cutovers and burns without substantial detriment to sylviculture. taking into account the imbalance of forest formations in russia over large areas, it is necessary to plan hunting of moose and other ungulates in order not to destroy forest ecosystems. in nonregulated moose hunting, there may be much stronger effects on forest ecosystems than the effects of felling. due to the lasting pattern of the effects of high waves of mass breeding of moose, these animals, through affecting vegetation, can influence the evolution of soils. the relationship between the dynamics of moose populations and human activity is of ancient origin. in fact, the paleo– and mesolithic camps of humans discovered in the northern half of the forest zone of eastern europe were commonly associated with sand pine forest terraces (bader 1970) in areas of latitudinal flow of rivers, due to massive, stable accumulations of moose in winter. as far back as the stone age to the felling of trees during world war ii and to our current organized harvesting of game, we affect our ecosystems for the future. references bader, n. o. 1970. mesolith. pages 90– 104 in stone age in the ussr territory. (in russian). chervonnyi, v. v. 1967. on the ecology, silvicultural significance and harvest of moose in karelia. pages 177–188 in biology and harvest of moose. (in russian). koryakin, d. a. 1961. the effect of moose on forest regeneration. proceedings of the prioksko–terrasny reserve 3:29–54. (in russian). yazan, y. p. 1964. population density and indices of moose fecundity of the pechora taiga. pages 101–111 in biology and harvest of moose. (in russian). zablotskaya, l. v. 1964. the experience of the regulation of moose in the prioksko–terrasny reserve and in the surrounding area. pages 156–173 in biology and harvest of moose. (in russian). alces suppl. 2, 2002 anthropogenic effects – zablotskaya and zablotskaya 135 . 1975. the cause of mortality of moose in different geographical regions. pages 105–129 in biology and harvest of moose. (in russian). zhirnov, l. v. 1967. migrations of moose in the european ussr. pages 80–104 in biology and harvest of moose. (in russian). p49-61_4107.pdf alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 49 hunting moose or keeping sheep? – producing meat in areas with carnivores jos m. milner1,2, erlend b. nilsen1,2, petter wabakken1, and torstein storaas1 1department of wildlife and forestry management, hedmark university college, evenstad, n-2480 koppang, norway. 2centre for ecological and evolutionary synthesis, department of biology, university of oslo, p.o. box 1050, blindern, n-0316 oslo, norway abstract: moose hunting is of considerable economic and social importance in much of scandinavia. in some parts, such as south-east norway, it is economically more important than sheep farming. we examine trends in moose harvesting and sheep production over a 12-year period in an area of increasing predator numbers and compare the meat yield before and after the re-establishment of wolves. the production of lamb meat at the county level declined, particularly from within the forest habitat, while the moose harvest showed only localized reductions. we also consider the scale of the economic loss caused by large carnivores and discuss management options for a future with carnivores. alces vol. 41: 49-61 (2005) key words: meat, wildlife harvesting, wolf predation unlike many parts of north america where the sale of game meat is restricted or prohibited, moose (alces alces) meat is a valuable commodity in scandinavia. since the 1970s the annual norwegian moose harvest has increased over 6-fold (statistics norway 2004a), due to changes in forestry practices and the introduction of a selective hunting regime (østgård 1987) in the near absence of large carnivores and with reduced competition from domestic cattle grazing (ahlen 1975). the current yield is around 35,000-40,000 moose per year with an estimated economic value of us$ 40-55 m from meat alone (see also storaas et al. 2001), making it by far the most economically important game species in scandinavia (mattsson 1990). at the lomeat and income in rural areas, and plays an important social and cultural role. although landowners do not legally own game animals on their land, they hold the right to hunt them and proceeds generated from hunting may of some large landowners. however, much of the meat is consumed privately and hunting rights are rarely sold for more than the meat value. furthermore, as there is no well developed scandinavian equivalent of the north contributes little towards local employment. consequently, much of the potential economic value of moose hunting is not realized. sheep production in norway has also increased since the 1970s but to a lesser extent and for different reasons. over this period, it has been government policy to support agriculture, including sheep farming, as a means of maintaining human settlements in rural norway and stabilizing food production (norwegian agricultural authority 2004, see also zimmermann et al. 2001). the introduction of production subsidies during an era when large carnivores were virtually extinct allowed for a rise in lamb production which peaked in the early 1980s (rogstad 2003, statistics norway 2004b). changing husbandry practices also meant that lamb production became concentrated on fewer, larger farms and became less labor hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 50 intensive with little shepherding (linnell et al. 1996, nersten et al. 2003, rogstad 2003). in general, ewes are over-wintered and lambed months of snow cover, and then released with their lambs to range freely in unenclosed forest and mountain pastures during the summer months (drabløs 1997). the current national production of lambs is about 1.4 million per year (statistics norway 2004b), with a meat value of us$ 120m approximately us$ 25 m (rogstad 2003, statistics norway 2004b). however, sheep production subsidies total approximately us$ income arising from subsidies (nersten et al. 2003). on an international scale, sheep farming in norway is a relatively small industry, providing about 9,000 full-time job equivalents nationwide, held on approximately 19,000 farms. although both moose hunting and sheep farming occur throughout large parts of norway, the main sheep farming districts are in the mountainous areas of western norway while moose hunting tends to be concentrated in the forested areas in the south-east and further north (fig. 1). concurrent with increases in the norwegian moose harvest and lamb production, there has been a change in attitudes towards large carnivores in europe and north america (linnell et al. 1996, bjerke et al. 1998, williams et al. 2002, ericsson and heberlein 2002). the norwegian government has explicitly stated the goal to maintain sustainable, breeding populations of four large carnivore species (miljøverndepartementet 2003-2004), following their near eradication due to human persecution over the last 150 years (swenson et al. 1995, wabakken et al. 2001, vilà et al. 2003). bears (ursus arctos) and wolves (canis lupus) have been protected in norway since 1973 and 1971 respectively, wolverine (gulo gulo) since 1973 in southern norway and 1981 in northern norway, and lynx (lynx lynx) since 1992 in southern norway (andersen et al. 2003). however, some controlled hunting of lynx and wolverine has been permitted under license since 1994. wolf numbers began increasing in southern and central scandinavia in 1991 and rose 10-fold during the following 10 years (wabakken et reproduction in norway occurred in 1997. one element of the policy to promote carnivores has been to pay compensation to farmers who lose domestic stock to carnivores (kaczensky 1996, linnell and brøseth 2003). number of large carnivores, it has become common practice to graze domestic sheep on unenclosed forest and mountain pastures without shepherding in summer (mysterud et al. 1996, linnell and brøseth 2003). consequently in some regions norwegian sheep farmers are now experiencing the highest losses of sheep per carnivore in europe (kaczensky 1996, linnell 2000). although a number of studies have investigated ways of linnell et al. 1996, mysterud et al. 1996, flaten and kleppa 1999, krogstad et al. 2000), the problem persists. furthermore, man is no scandinavia, as wolves also adversely affect local moose populations (gundersen 2003). here we present an exploratory analysis of trends in the relative size and economic importance of moose hunting and sheep farming since 1990 in the county of hedmark, southeast norway, where carnivore numbers have been increasing. we examine changes in trends since the re-establishment of large carnivores and quantify the economic loss incurred by local landowners and communities due to a reduced moose harvest and lamb production. we then go on to discuss land management options in an environment of increasing carnivore density and the appropriateness of continuing sheep farming in some areas. we do not attempt to demonstrate causal relationships alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 51 fig. 1. map of norway showing (a) the density of moose shot in 2002 (statistics norway 2004b) and (b) the number of sheep kept over winter in each municipality (norwegian agriculture authority 2004) and the location of hedmark county. (c) habitat zones and location of reproducing large carnivores in hedmark in 2003. no. over-wintering sheep per municipality no. over-wintering sheep per municipality no. over-wintering sheep per municipality hedmark county s w ed en hedmark county s w ed en a) b) c) mountain zone intermediate zone forest zone n fig. 1 no moose hunting allowed < 1.0 1.0 1.9 2.0 2.9 3.0 4.9 > 5.0 moose harvested per 10 km2 qualifying area hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 52 between moose, sheep, and predator numbers. study area the county of hedmark is in south-east norway on the swedish border and is composed of 22 municipalities (fig. 1) with a low and scattered human population density, averaging 6.8 persons / km2. the county covers about 27,000 km2 of which approximately 60% is boreal forest dominated by scots pine (pinus sylvestris) and norway spruce (picea abies), managed primarily for commercial timber production. hedmark has a relatively high moose density (>1 moose / km2 (gundersen 2003)). it is the most important moose hunting county in norway (fig. 1a), accounting for 20% of the national harvest (statistics norway 2004a). over 25,000 inhabitants are registered hunters. by contrast, sheep farming is not a major component of the local economy except in the north of the county. there are about 45,000 over-wintering adult female sheep in the whole 1b) which provide less than 600 full-time job equivalents (wabakken et al. 1996). but, more importantly for local rural politics, sheep farming provides income or part-time employment for over 1,100 households for whom having sheep may make the difference between keeping or abandoning the farm. although cattle farming also occurs throughout the county, there are about 7 times as many freeranging sheep as cattle. until recently there cattle (zimmermann et al. 2003) so public and media interest has focused on the issue of sheep farming. the county can be divided into 3 habitat zones (fig. 1c). the north of hedmark is a mountainous area, characterized by a high proportion of alpine vegetation above the treeline at about 900m above sea level. the south of hedmark is dominated by forest with over 70% forest cover and < 0.1 % of the land area above the tree-line. between these areas is an intermediate zone with a mixture of both forest (60% cover) and mountain habitats (table 1). moose density, indexed by harvest density, is highest in the forest zone and lowest in the mountain zone (fig. 1a), while sheep densities are considerably higher in the mountain zone and than in either the forest or intermediate zone (fig. 1b, fig. 3). hedmark is the only county in norway where breeding populations of all 4 species of large carnivores occur (fig. 1c) and has one of the largest numbers of carnivores. by proximity to the swedish border, with dispersing individuals, particularly bears and wolves, frequently crossing into hedmark. overall, carnivore numbers are highest in the forest zone. wolverines tend to occur most in the mountainous north while lynx numbers are highest in forest areas. both wolf and bear populations occur in the area east of the glomma river to the swedish border. breeding wolves have gradually been re-colonizing the area since 1997 but resident wolves are only found in the forest and intermediate zones. the norwegian parliament has declared a wolf conservation zone which was implemented in spring 2005, partly in southeastern hedmark, while a bear conservation zone in hedmark existed between 1993 and spring 2005 (miljøverndepartementet 2003-2004). however, these zones, which incorporate suitable habitat, are essentially political demarcations and dispersing individuals of both species can often be found in areas outside these zones. data & analyses local municipalities act as the executive game management authorities in norway (danielsen 2001). we therefore used data from hunters and sheep farmers collated at the municipality scale, by hunting teams, landowner and grazing organizations, municipality included the number and ageand sex-class of all moose shot during the hunting season, alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 53 table 1. size and vegetation of the habitat zones within hedmark county. qualifying area (qa) is based on the land considered suitable moose habitat for the purpose of hunting license allocation population size is the proportional change in the number of moose observed per hunter per day during the hunting season between the period 1990-1996 and 1997-2002. fig. 2. trends in (a) number of moose shot and (b) number of lambs produced in hedmark since 1970. a: introduction of selective hunting, b: introduction of sheep production subsidies, c: implementation of the brown bear conservation zone, d: re-establishment of breeding wolves in norway. habitat zone total area (km2) proportion forest proportion bog qualifying area (km2) qa /ta change in moose pop mountain 81,400 0.38 0.09 42,800 0.53 0.98 intermediate 171,300 0.6 0.13 124,750 0.73 1.2 forest 103,450 0.74 0.09 87,250 0.84 1.31 0 10,000 20,000 30,000 40,000 50,000 60,000 70,000 80,000 90,000 100,000 1970 1975 1980 1985 1990 1995 2000 2005 n o . la m b s 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 1970 1975 1980 1985 1990 1995 2000 2005 n o . m o o s e h a rv e s te d low carnivore numbers increasing carnivore numbers a b c c d d a) b) hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 54 fig. 3. trends by habitat zone in (a) number of moose shot, (b) number of ewes plus lambs released to hill and forest pastures in june, (c) percentage losses of ewes and lambs to large carnivores during summer, and (d) number of lambs per ewe in autumn. square symbol with dashed line: forest zone, triangle with dotted line: intermediate zone, diamond with solid line: mountain zone. the number of adult sheep and lambs released onto the unenclosed forest and mountain pastures during the summer, the number of adult sheep and lambs lost over the summer period, the number of compensation claims made by sheep farmers for different predators, and the number and size of compensation payments made. numbers of sheep were those of farmers belonging to grazing organizations only (about 90% of all sheep farmers in hedmark; e. maartmann, personal communications). we restricted our analysis to the time period 1990-2002 allowing a comparison of years before and after the re-colonization of wolves, pact on both sheep and moose in hedmark. all values are expressed in us$ assuming a constant exchange rate of 1 norwegian kroner values using the norwegian consumer price index (statistics norway, http://www.ssb. no/kpi/tab-01.html). the value of moose meat was assumed to be us$ 10.2 / kg throughout the study period, corresponding to a decrease in value in real terms from $13.4 / kg in 1990. average stripped moose carcass weights showed a density-dependent decrease during the study period and were taken as 70 kg prior to 1994 and 68 kg since 1994 for male calves, 70 kg and 64 kg for female calves, 149 kg and 139 kg for 1.5-year-old males, 142 kg and 131 kg for 1.5-year-old females, 221 kg and 200 kg for older males, and 183 kg and 176 kg for older females prior to and since 1994, 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 1990 1992 1994 1996 1998 2000 2002 m oo se sh ot /k m 2 0 2 4 6 8 10 12 1990 1992 1994 1996 1998 2000 2002 s h e e p + la m b s / k m 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 1990 1992 1994 1996 1998 2000 2002 n o . la m b s / e w e in a u tu m n 0 2 4 6 8 10 12 14 16 18 20 1990 1992 1994 1996 1998 2000 2002 % to ta l e w e + la m b s lo s t a) b) c) d) alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 55 respectively (statistics norway 2004a). the between a minimum of us$ 5 / kg in 1990 to a maximum of us$ 6.2 /kg in 2002 (statistics tion, meant a decrease in value from us$ 6.7 / ation in autumn lamb weights between years (steinheim et al. 2001, 2004), for simplicity we have assumed a constant stripped lamb carcass weight over time but varying between municipalities from 17 kg in the forest zone to 19.5 kg in the mountain zone (steinheim et al. 2001). we have assumed that all lambs rounded up in autumn are slaughtered rather than used for stock replacement. a generalized linear mixed modeling approach (mccullagh and nelder 1989), in which effects, was used to evaluate the trends which were investigated by habitat zone. the wald 2 distribution. results regional trends in moose hunting an average of 6,600 moose (range 4,958 8,215) have been shot annually in hedmark since 1990 (fig. 2a), yielding over 880 tonnes of meat per year. assuming all meat was sold, this represents a value of us$ 7.5 m 14.6 m per year. there has been considerable variation in the number of moose shot per year, both at the county level (fig. 2a) and when a comparison 2 12,267 = 244.4, p < 0.001; fig. 3a). inter-annual variation was greatest in the forest zone, showing a sharp decline in harvest yield in 1994-95 followed by a recovery to pre-decline levels. however, there was no evidence of a difference in the number of moose shot per year in the periods before (1990 1996) and after (1997 2002) the re-establishment of wolves in any habitat zone (mountain: 2 1,60 = 0.45, p 2 1,73 = 0.27, p 21,136 = 0.09, p = 0.77) and moose were not affected by any other large carnivore species. there were, nonetheless, clear differences in the number of moose shot /km2 qualifying area between habitat zones 2 2,277 = 42.83; p < 0.001), with considerably higher yields in the forest zone than in either the intermediate or mountain zone in all years. in moose densities between the habitats. furthermore, the number of moose seen per day by hunters increased over the study period in the forest and intermediate zones but not in the mountain zone (table 1). consequently, hunting yield than the presence of wolves, at the regional scale. regional trends in lamb production in hedmark has remained steady at around 46,000 ewes (range 43,673 48,375), while the total number of lambs produced and surviv2 1,247 = 7.20; p = 0.007), falling from a peak of over 80,000 in the early 1990s to 70,000 in 2000 (fig. 2b). the decline, from 1,280 tonnes of meat in 1990 to 1,130 tonnes in 2002 is equivalent to a 12% reduction in annual production, or a loss of 150 tonnes per year over 13 years, worth almost us$ 1m at 2002 prices. the decline in lamb yield corresponds with a decrease over time in the number of 2 1,247 = 23.29; p < 0.001) and a dramatic increase over time in the proportion of lambs going missing 2 1,247 = 82.23; p < 0.001; fig. 3c). compensation claims for both sheep and lambs lost to large carnivores in hedmark increased over 5-fold in real terms between 1990 and 2002, with claims paid amounting to tion, rising to us$ 1.7m in 2002 (fylkesmannen i hedmark, unpublished data). this was despite a decrease in compensation payment per head in both actual and real terms between 1990 and 2002. most of the compensation paid hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 56 out was for claims against bear (23%), lynx (19%), and wolverine (17%) predation, with only 6% due to wolves and a further 34% in the rate at which autumn lamb yield dezones, being considerably greater in the forest zone than in the mountains (habitat zone-year 2 2,243 = 15.83; p < 0.001). although all habitat zones have shown a steady increase in losses of both ewes and lambs since 1990 as carnivore numbers have increased 2 1,247 = 88.79; p < 0.001), proportional losses have been highest in the forest zone (habitat 2 2,243 = 8.35; p = 0.015; fig. 3c), reaching over 20% of lambs in 2002. this, combined with lower ewe productivity within the forest zone (average number of lambs per ewe in june is 1.48 in forest zone compared with 1.65 in mountain zone; 2 2,246 = 7.66; p = 0.022) and lower autumn weights of lambs that have spent the summer grazing in forest areas (steinheim et al. 2001), is reducing the viability of sheep production in the forest zone relative to other parts of the county. relative value of moose hunting and sheep production within the forest zone, the value of moose meat is nearly 2.5 times greater than the value of lamb meat (including compensation for lost animals) (fig 4). by contrast, in the mountain zone, the lamb meat produced has over 5 times the value of moose meat. the total value of moose and lamb meat together is considerably lower in the intermediate zone than in either of the other areas (averaging us$ 474 /km2, compared with us$ 859 /km2 and us$ 793 /km2 in the mountain and forest zones, respectively, over the last 5 years), with the value of moose meat being 1.5 times greater than lamb meat. however, income from sheep farming is considerably enhanced by government production subsidies. although subsidy payments have gradually been eroded since the 1970s, in 2002, total expenditure on sheep production subsidies amounted to approximately us$ 1.3 m in the forest zone, us$ 1.6 m in the intermediate zone and us$ 3.0 m in the mountain zone (fylkesmannen i hedmark, unpublished data). this has the effect of raising the total moose and lamb value to us$ 1,347 /km2, us$ 1,043 /km2, and us$ 613 /km2 in the mountain, forest, and intermediate zones, respectively, in 2002. of this, 90%, 41%, and 52%, respectively, was contributed by sheep enterprises. in all zones, total meat value decreased between 1990 and 2002 after adjusting for greater in the forest and intermediate zones than in the mountain zone (fig. 4). trends over source in each habitat type and consequently the forest and intermediate zones were affected by the relatively greater devaluation of moose meat than lamb meat. by contrast, the effect of increasing carnivore numbers in these zones was relatively minor because lamb meat was a less important component of the total meat value, and losses were generally compensated. consequently, the overall economics have not been strongly affected by increasing carnivore numbers. however, as discussed below, this may not be the case for individual landowners. local impact of carnivores impact on moose hunting yield is the wolf, but within hedmark, at the regional and even municipality scales, no effect of predation was apparent. however, in localized parts of the study area and for individual landowners within a wolf territory, wolves may have a profound economic impact. for example, gundersen (2003) suggests 27% of moose calves per year (24 31% ± 2 se) are killed by the koppang wolf pack, in central hedmark. this equates to approximately 100 moose alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 57 per year. to minimize any decline in moose numbers due to predation, the landowners in that area have voluntarily imposed a restriction on the number of moose shot per year (fig. 5). this has cost them over us$ 150 /km2 from the loss of meat sales alone. in addition, some landowners have experienced a considerable loss of rental income from the letting of cabins and small game hunting since the arrival of the wolf (c. mathiesen, personal communication). rising numbers of other carnivores have had a less dramatic effect on rental income because it is mainly wolves that pose a threat to hunting dogs. sheep and lamb numbers expressed at the regional scale also hide the trends at the scale of individual carnivore territories. brown bears kill more sheep in hedmark than other carnivores but the annual removal of some problem bears has not halted the increase in sheep losses (zimmermann et al. 2003). within wolf territories, many farmers have but high predation rates have forced some to abandon sheep production, either switching to cattle or alternative enterprises. however, the alternatives, one of which may be moose hunting (storaas et al. 2001), are somewhat limited and sheep farmers only rarely have hunting rights for the areas they graze. discussion the increase in carnivore numbers experienced throughout the 1990s in norway has stimulated a reappraisal of land management strategies and considerable debate about options for the future. it is clear that as carnivores have increased, sheep production in hedmark has declined, particularly where sheep are released into forest habitat for summer grazing. however, our data cannot demonstrate a causal link between the two. by contrast, the moose harvest has shown no such change due to a low bear density, the lower vulnerability of moose to carnivores such as lynx and wolverine, and a greater variability in population size and yield between years. although some commentators comparing the current moose hunting yield with the yield 10 years earlier note a drop which they attribute to wolves, our data suggest that at a regional scale this was largely a result of a decline in moose population size before the arrival of wolves. nonetheless, at the spatial scale of the wolf territory, some landowners have experienced a considerable economic loss due to reduced sustainable hunting yields (nilsen et al. 2005). this is also true of other game species and areas beyond hedmark (aanesland and holm 2003). in norway, natural summer mortality of lambs in the absence of carnivores is assumed to be around 4% (drabløs 1997), while lamb losses in some parts of the forest zone in southern hedmark are in excess of 20% (norwegian institute of land inventory, http://beite.nijos. fig. 4. value per km2 of moose and sheep meat (assuming all meat is sold) and compensation from sheep lost to large carnivores in three habitat zones for the period 1990-2002, adjusted 0 100 200 300 400 500 600 700 800 900 1000 1100 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 0 100 200 300 400 500 600 700 800 900 1000 1100 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 0 100 200 300 400 500 600 700 800 900 1000 1100 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 moose meat compensation lamb meat moose meat compensation lamb meat u s $ /k m 2 forest intermediate mountain hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 58 no/kart.htm). this is one of the highest rates of sheep loss per carnivore in europe (kaczensky 1996, linnell 2000) and raises ethical questions about the appropriateness of releasing lambs onto unenclosed land in summer. since it is government policy to maintain breeding populations of large carnivores, including demarcation of a zone for wolf reproduction, it can be expected that carnivores will remain in most of these areas. consequently, unless changes are made, sheep losses and declines in lamb production are likely to continue (sagor et al. 1997), particularly in forest habitat where carnivore numbers are higher and dense vegetation restricts prey vigilance. this presents a dilemma for the government which wants to promote both large carnivores and the rural population, without trading one off against the other. there are a number of ways in which sheep losses could be reduced (kaczensky 1996, linnell et al. wolf territories, few measures have been taken. this is primarily because radical changes to the current husbandry system of extensive and are likely to be costly (linnell and brøseth 2003). it appears that using shepherds or guarding dogs could prevent much of the predation but would increase production costs compared to current practice (krogstad et al. 2000, linnell 2000). furthermore, there is no tradition of using dogs in this way in norway and using children, as in former times, is no longer realistic without payment. if such measures are not implemented for economic reasons, it may be that in some habitats it is no longer appropriate to continue sheep production. if, for example, sheep zone, the primary economic loss to the area would be the loss of the production subsidy payments which amounts to about us$ 1.3 m per year. however, this represents an equal saving to the government. could this money instead be used to promote alternative economic activities associated with moose hunting and ecotourism or widen compensation schemes? here, the principal challenge would be to ensure that the individuals who were keeping sheep do not lose out. currently farmers are eligible for compensation for any of their domestic stock killed by large carnivores (kaczensky 1996). however, landowners are not eligible for compensation for losses of moose hunting yield. if the government wants to continue its policy of keeping rural areas populated, maybe it should consider widening the terms under which compensation is offered. however, it should be noted that compensation payments do not always improve tolerance towards carnivores, especially where emotional stress is caused (naughton-treves et al. 2003). furthermore, it would be undesirable if the effect was to shift compensation payments from many small farmers to fewer, relatively large landowners. a more appropriate approach may be for affected municipalities to receive some kind of economic impact of having large carnivores in the area. a second approach may be to explore ways of increasing the yield, and consequently income, from moose hunting, if a higher moose population density could be sustained. to do this, measures would have to be taken to prevent, alleviate, or pay for forest damage (johansson et al. 1988, fig. 5. number of moose shot per year within the koppang wolf territory showing a voluntary restriction on the size of the hunt since 2000 (from gundersen 2003). 0 20 40 60 80 100 120 140 1994 1995 1996 1997 1998 1999 2000 2001 n o. of m oo se sh ot calves yearlings cows bulls alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 59 storaas et al. 2001, gundersen 2003) or time and money invested in growing additional moose fodder such as willow. however, a high density moose population is also accidents (johansson et al. 1988, gundersen et al. 1998, storaas et al. 2001) and have adverse effects on biodiversity, so additional costs would be incurred in mitigation measures. thirdly, there may be opportunities to realize a greater proportion of the existing value of moose. although moose meat makes a considerable contribution to the household for many hunters (mattsson 1990), much of the value is never converted into cash. furthermore, in addition to the meat value, hunting has a recreational value (mattsson 1990) which is currently barely realized. consequently, and guiding businesses and promoting moose hunting to non-residents, as long as access to local hunters is not compromised. optimizing the allocation of moose hunting between individual hunters could also help maximize the value of moose (mattsson 1990). apart from hunting tourism, there appears to be scope, as yet unrealized, for promoting ecotourism, specializing in wildlife viewing or wolf-tracking. for example, in romania, the carpathian large carnivore project has demonstrated that considerable tourism revenue can be brought into an area by promoting its association with large carnivores (http://www. clcp.ro/etour/eco-prog.htm). in norway, while tourism in areas such as hedmark is marketed by focusing on outdoor pursuits, no mention is made of the large carnivores. it appears that in some parts of hedmark it may not be possible to maintain the status quo in sheep production for much longer. expansion of moose enterprises and eco-tourism may have the potential to provide some alternative income if moose management can be implemented appropriately and in such a way that farmers, as well as landowners, can acknowledgements we thank erling maartmann, jorunn stubsjøen, atle mysterud, odd reidar fremming, christian mathiesen, geir steinheim, and hege gundersen for useful comments, discussions, and access to unpublished data. jmm norwegian research council (nfr 96061). references aanesland, n., and o. holm. 2003. rovdyr og jaktinntekter. (carnivores and hunting income) report nr. 27, norges landbrukshøgskole, ås, norway. (in norwegian.) ahlén, i. 1975. winter habitats of moose and deer in relation to land use in scandinavia. swedish wildlife research supplement 9: 45-192. andersen, r., j. d. c. linnell, h. hustad, and s. breinard, editors. 2003. large predators and human society. a guide to co-existence in the 21th century. report nr. 25, nina, trondheim, norway. bjerke, t., o. reitan, and s. r. kellert. 1998. attitudes toward wolves in southeastern norway. society and natural resources 11: 169–178. danielsen, j. 2001. local community based moose management plans in norway. alces 37: 55-60. drabløs, d. 1997. the story of the norwegian sheep. anniversary review of the norwegian sheep and goat breeders 1947-1997. norwegian sheep and goat breeders, oslo, norway. ericsson, g., and t. a. heberlein. 2002. “jagare talar naturens sprak” (hunters of outdoor activities and attitudes toward wildlife among swedish hunters and the general public. zeitschrift für jagdwissenschaft 48: 301-308, suppl. s. flaten, o., and s. kleppa. 1999. en økonomisk analyse av forebyggende tiltak mot rovvilttap i saueholdet. (an economic hunting moose or keeping sheep – milner et al. alces vol. 41, 2005 60 analysis of protective measures to reduce sheep depredation.) report nr. 1999/1, norsk institt for landbruksøkonomisk forskning, oslo, norway. (in norwegian). gundersen, h. 2003. vehicle collisions and wolf predation: challenges in the management of a migrating moose population in southeast norway. ph.d. thesis, university of oslo, norway. _____, h. p. andreassen, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34: 385-394. johansson, p.-o., b. kriström, and l. mattson. 1988. how is the willingness to pay for moose hunting affected by the stock of moose? an empirical study of moosehunters in the county of vasterbotten. journal of environmental management 26: 163-171. kaczensky, p. 1996. large carnivore livetrondheim, norway. krogstad, s., f. christiansen, m. smith, o. c. røste, n. aanesland, r. h. tillung, and l. thorud. 2000. forebyggende tiltak mmot rovviltskader på sau: gjeting og bruk av vokterhund i lierne. (protective measures to reduce sheep depredation: shepherding and use of guarding dogs in lierne). nina fagrapport 041, trondheim, norway. (in norwegian with english summary). linnell, j. d. c. 2000. norwegian brown bears: holders of an unwanted world record. carnivore damage prevention news 1: 4-5. _____, and h. broseth. 2003. compensation for large carnivore depredation of domestic sheep 1994-2001. carnivore damage prevention news 6: 11-13. _____, m. e. smith, j. odden, j. e. swenson, and p. kaczensky. 1996. carnivores and sheep farming in norway. 4. strategies for the reduction of carnivore oppdragsmelding 443: 1-118. mattson, l. 1990. hunting in sweden: extent, economic values and structural problems. scandinavian journal of forest research 5: 563-573. mccullagh, p., and j. a. nelder. 1989. generalized linear models. chapman and hall, london, u.k. miljøverndepartementet. 2003-2004. rovvilt i norsk natur. – stortingsmelding 15. department of environmental protection, oslo, norway. (in norwegian). mysterud, i., a. o. gautestad, and i. mysterud. 1996. carnivores and sheep farming in norway. 6. comments on shepherding as preventive measure. report to department of biology, university of oslo, oslo, norway. naughton-treves, l., r. grossberg, and a. treves. 2003. paying for tolerance: rural and compensation. conservation biology 17: 1500-1511. nersten, n.k., a. hegrenes, o. sjelmo, and k. stokke. 2003. saueholdet i norge utvikling, politikk og virkemidler. note 2003-10, nilf, oslo, norway. (in norwegian). nilsen, e.b., t. pettersen, h. gundersen, j.m. milner, a. mysterud, e.j. solberg, h.p. andreassen, and n.c. stenseth. 2005. moose harvesting strategies in the presence of wolves. journal of applied ecology 42: 389-399. norwegianagricultural authority (statens landbruksforvaltning). 2004. årsrapport 2003 (annual report 2003). report nr. 3/2004. (in norwegian). østgård, j. 1987. status of moose in norway wildlife research supplement 1: 63-68. rogstad, b., editor. 2003. norwegian agriculture. status and trends 2003. norwegian agricultural economics research institute, oslo, norway. sagor, j.t., j.e. swenson, and e. roskaft. alces vol. 41, 2005 milner et al. hunting moose or keeping sheep 61 1997. compatibility of brown bear ursus arctos and free-ranging sheep in norway. biological conservation 81: 91-95. statistics norway (statistisk sentralbyrå). 2004a. jaktstatistikk 2002 (agricultural statistics 2002) oslo–kongsvinger. (in english and norwegian). _____. 2004b. jordbruksstatistikk 2002 (hunting statistics 2002) oslo–kongsvinger. (in english and norwegian). steinheim, g., y. rekdal, ø. holand, and t. ådnøy. 2001. produksjon av lammekjøtt på utmarksbeite: hvorfor varierer vektene så mye fra område til område, og fra år til år? (production of lamb meat on upland grazings: why do weights vary so much from area to area, and from year to year?). pages 33-39 in v. jaren and j. p. løvstad, editors. utmarksbeite og store rovdyr (upland grazing and large carnivores). report to norges forskningsråd, oslo, norway. (in norwegian). _____, r. b. weladji, t. skogen, t. ådnøy, a. o. skjelvåg, and ø. holand. 2004. climatic variability and effects on ungulate body weight: the case of domestic sheep. annales zoologici fennici 41: 525-538. storaas, t., h. gundersen, h. henriksen, and h. p. andreassen. 2001. the economic value of moose – a review. alces 37:97-107. swenson, j. e., p. wabakken, f. sandegren, a. bjärvall, r. franzén, and a. söderberg. 1995. the near extinction and recovery of brown bears in scandinavia in relation to the bear management policies of norway and sweden. wildlife biology 1: 11-25. vilà, c., a-k. sunderquist, ø. flagstad, j. seddon, s. björnerfeldt, i. kojola, a. casulli, h. sand, p. wabakken, and h. ellegren. 2003. rescue of a serverely bottlenecked wolf (canis lupus) population by a single immigrant. proceedings of the royal society of london b 270: 91-97. wabakken, p., e. maartmann, j. berg, and h. c. gjerlaug. 1996. forvaltning av fredet rovvilt i hedmark i 1995 bestandsregistrering, forebyggende tiltak, skadedocumentasjon og erstatniger. (management of protected carnivores in hedmark in 1995 population size, preventive measures, documentation of damage and compensation.) miljøvernavdelingen, fylkesmannen i hedmark, report 3/95. _____, h. sand, o. liberg, and a. bjärvall. 2001. the recovery, distribution, and population dynamics of wolves on the scandinavian peninsula, 1978-1998. canadian journal of zoology 79: 710-725. williams, c.k., g. ericsson, and t. a. heberlein. 2002. a quantitative summary of attitudes toward wolves and their reintroduction (1972–2000). wildlife society bulletin 30: 575–584. zimmermann, b., p. wabakken, and m. dö t t e r e r. 2001. human-carnivore interactions in norway: how does the re-appearance of large carnivores affect forest snow and landscape research 76 1/2: 137-153. _____, _____, and _____. 2003. brown bearzone in norway: are cattle a good alternative to sheep? ursus 14: 72-83. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 4201(1-11).pdf alces vol. 42, 2006 van ballenberghe predator control in alaska 1 predator control, politics, and wildlife conservation in alaska victor van ballenberghe department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775, usa abstract: lethal control programs aimed at reducing wolf (canis lupus) and bear (ursus arctos and u. americanus) numbers while attempting to increase densities of moose (alces alces) and caribou (rangifer tarandus) for hunters have occurred intermittently in alaska, usa, for the past 3 decades. these programs were accompanied by considerable controversy, much of it directed at methods of ing by private citizens. from 1976 to 1983, 1,300 wolves were taken in several areas of alaska by a combination of helicopter shooting and private trapping. adverse public reaction largely restricted wolf control from 1984-1994 when a snaring program again produced controversy and that control program was terminated. in 1997, a national research council review suggested numerous biological standards for alaska’s predator control programs. the review strongly endorsed the approach of conducting predator control as adaptive management. control proponents sponsored legislation in the 1990s that mandated intensive management of certain depleted populations of ungulates deemed important for consumptive use by humans. the primary management tool to increase such populations is predator control. intensive management also required setting population and harvest objectives for ungulates. these objectives often were based on historical highs that are now likely unattainable and almost certainly unsustainable. implementation of intensive management programs involving reductions of black bears and brown bears as well as wolves has now been approved in 5 areas of alaska totaling about 43,000 square miles with up to 610 wolves scheduled to be shot by april 2005. approval of additional programs is pending. controversy now is focused not merely on ethical objections to ungulate populations, protection of habitat integrity for ungulates, and population viability of predators. recommended biological standards and guidelines for justifying, implementing, monitoring, and evaluating control programs are not being applied alces vol. 42: 1-11 (2006) key words: alaska, alces alces, bears, canis lupus, caribou, moose, politics, predator control, rangifer tarandus, ursus arctos, ursus americanus, wolves lethal predator control aimed at reducing wolf (canis lupus) and bear (ursus arctos and u. americanus) populations while attempting to increase densities of ungulates for hunters has been a highly controversial issue in alaska, usa, for decades. much of the controversy centered on wolves and methods of control including the use of poison, bounties, aerial shooting by private pilots, helicopter shooting by state employees, and snaring. other issues including the quality of data used to justify, implement, monitor, and evaluate control programs were also part of the debate. in recent years the controversy broadened to include bears. approval of large-scale programs, now totaling about 43,000 square miles with up to 610 wolves to be shot by spring 2005, has raised several conservation concerns. these programs were adopted with weak implementation, monitoring, and evaluation protocols, no study plans, and no research components management law, population and harvest objectives have been set that, in many instances, predator control in alaska – van ballenberghe alces vol. 42, 2006 2 are based on historical population highs for ungulates that are now likely unattainable and almost certainly unsustainable. as a result, poorly designed control programs may forever chase unattainable objectives, and long-term conservation problems may outweigh shortterm gains. in 1995, governor tony knowles requested that the national academy of sciences conduct a review of past control programs and provide recommendations for future efforts. this review was conducted by the national research council (nrc 1997) and addressed biological and socioeconomic issues. the review contained 17 broad conclusions with 16 recommendations. of these, 8 recommendations applied to the biological aspects of the review. in addition, the review contained a section with decision-making guidelines. contained in the report were many recommended biological standards and guidelines. tempt to provide standards to guide alaska’s ensuring that sound science was incorporated in predator control programs. following the release of the nrc report, alaska’s department of fish and game (adfg) assembled a team to design a predator control program in the mcgrath area of interior alaska (adfg 2001). many of the nrc’s standards were incorporated in the team’s plan. shortly thereafter, frank murkowski was elected governor and the mcgrath program plus several additional areas were approved for control. these programs largely abandoned recommended standards and did not follow an adaptive management approach. programs and a return to the mcgrath model wherein the nrc’s recommended standards and guidelines were applied. a brief history following world war ii when alaska was still a u.s. territory, a federal poisoning and aerial shooting campaign began (harbo and dean 1983). by the mid-1950s, the program had greatly reduced wolf numbers in much of south-central and interior alaska. wolves persisted in some areas largely because the in the nelchina basin near glennallen, a 20,000 square mile area, only 1 wolf pack remained, reportedly spared for study. aerial shooting on the north slope reduced wolves to very low levels and they remained low for decades. after statehood in 1959, the controversy over poison was so intense that it was permanently banned by the new state legislature (harbo and dean 1983). aerial shooting and bounty payments, however, continued through the 1960s. large numbers of wolves were taken and densities remained low. after passage of the federal airborne hunting act in 1972 and termination of the bounty, wolf numbers increased as ungulate populations declined following irruptions in the 1960s (van ballenberghe 1985). in some instances, there were spectacular crashes evidently precipitated by severe winters and accelerated by predation caribou (rangifer tarandus) herd declined from 90,000 in 1962 to 8,000 in 1972. the tanana flats moose (alces alces) population south of fairbanks went from 23,000 to 2,800 during 1965-1975 (gasaway et al. 1983). faced with declining ungulate populations by the mid-1970s, hunters demanded wolf control in several areas and adfg responded by proposing helicopter-shooting programs. despite legal challenges, these programs accounted for 1,300 wolves at a cost of $824,000 between 1976 and 1983 (adfg 1983). by 1984, considerable public opposition largely terminated state-sponsored control programs, nevertheless, taking of wolves by private pilots, termed “land-and-shoot” hunting, continued. this served as de facto wolf control in certain areas where terrain features were suitable, and regulations requiring hunters to land before alces vol. 42, 2006 van ballenberghe predator control in alaska 3 as a new administration proposed more shooting of wolves from helicopters (franzmann 1993). governor walter hickel received more than 100,000 letters of protest. a wolf-snaring program emerged as a substitute to aerial shooting, but also provoked international protests as video footage documented wolves chewing their frozen feet caught in snares. in the 1990s, political involvement in control issues increased greatly. in 1994, hunting and trapping interests successfully lobbied the state legislature for an “intensive management” bill that mandated efforts to restore depleted ungulate populations to former levels of abundance. the bill’s clear intent was a strong emphasis on predator control. in 1996, however, a ballot initiative banning public land-and-shoot wolf hunting passed by a large margin. efforts by the legislature to resurrect the public’s use of airplanes to shoot wolves resulted in a public referendum in 2000 that again banned this practice. in 2003 and 2004, after a decade largely free of major predator control programs, a new state administration headed by governor frank murkowski approved 5 new programs involving the use of private pilots to shoot wolves from airplanes. affected areas total about 43,000 square miles with about 610 wolves scheduled to be shot initially and undetermined others to follow in subsequent years. in addition, hunting and trapping seasons, bag limits, and methods of take for wolves in these and most other areas of the state were liberalized. from mid-august to may, there are no trapping bag limits, and wolves can be pursued and shot from snowmachines. currently, hunters and trappers take about 1,500-1,700 wolves programs, from a total population crudely estimated at 7,500-11,000. black and brown bear populations also are scheduled for reduction in certain areas. in march 2004, the state board of game (bog), a 7-member body that promulgates hunting and trapping regulations and sets predator control policies, revised its bear conservation and management policy to include a section on predation. methods and means that the bog may consider include relocation, sterilization, use of electronic equipment for communication between hunters, sale of hides and skulls, trapping, baiting with human-derived foods as an aid to hunting, same-day airborne taking, and diversionary feeding. efforts to reduce bear numbers by lengthening autumn hunting seasons, opening spring seasons, increasing bag limits, and eliminating hunting tag fees have occurred during the past 2 decades in certain areas where bears were thought to prey on moose at high rate. in 2004 the bog approved baiting of brown bears as a predator control measure in one area. although baiting of black bears has long been legal, this is the brown bears. early standards and guidelines for predator control although poorly documented, the standards and guidelines used by adfg and the bog for predator control in several wolf control programs during 1976-1983 included preparation of “issue papers.” these consisted of reviews of the available data including predator population status and trend, harvest information for predators and prey, predatorprey ratios, and crude information on habitat where research resulted in ungulate populaaircraft, reliable population estimates for ungulates were generally unavailable. wolf population surveys in winter based on aerial track counts and observations of live animals were supplemented with trapper reports to estimate wolf numbers. at that time the bog adhered to a policy predator control in alaska – van ballenberghe alces vol. 42, 2006 4 prohibiting poison. this policy allowed private pilots to take wolves in certain areas under a permit system in accordance with federal airborne hunting act provisions, and directed adfg employees to take wolves by helicopter shooting where feasible (adfg 1983). bog policy prohibited total elimination of wolves leaving 20% of the pre-control wolf population. despite these guidelines, there were neither formal requirements in the predatorcontrol regulations requiring certain types of data or standards for data quality necessary to justify control, nor were there protocols for implementing, monitoring, or evaluating control programs. in the late 1980s, the bog adopted an “emergency” standard for justifying control programs. under this standard, wolf control would be infrequent and not applied unless prey populations were demonstrated to be at low densities and were unlikely to recover without control. protocols were established to determine if wolf predation was limiting ungulates rather than some other factor. control programs would cease when prey populations had recovered. the bog rescinded this standard by 1991 to accommodate proposed helicopter shooting of wolves under a zoning program as part of a strategic wolf management plan. in certain zones, ungulates would be managed at high densities and wolf numbers would be kept low. the intensive management statute, passed in 1994, mandated new standards for management of ungulates. these were based on restoring “depleted” populations to former levels of abundance, but depleted applied at any level of ungulate abundance, low, medium, or high, with the overall goal of increasing opportunity for hunters and to of human consumption. no attempt was made to understand the potential effects of habitat quality on moose numbers. under governor tony knowles, 3 broad standards were mandated for control programs. control would be based on sound science, would have broad public support. these standards precipitated debate on what constituted sound science and who determined science quality, and on methods of measuring public support. the national research council’s standards and guidelines the national research council review (nrc 1997) addressed two basic questions: 1. in attempts to understand interactions between moose and caribou and their habitats and predators, have appropriate types of data been gathered, and has enough been learned from past research to identify the information needed to enable us to predict quantitative responses of prey populations to predator control efforts? them? the committee reviewed past and present control programs, alaska’s biomes, people, and wildlife species of concern, predatorprey interactions, wolf and bear management implications of predator control, and decision making. the resulting report included 9 major biological conclusions and 8 recommendain the recommendations provide the basis for suggested standards and guidelines for predator control programs. these included: 1. wolves, bears, and ungulates should be managed with an adaptive management approach. 2. management actions should be planned alces vol. 42, 2006 van ballenberghe predator control in alaska 5 their outcome. control actions should to determine whether or not predictions are borne out and why. 3. managers should avoid actions with uninterpretable outcomes or low probability of achieving stated goals. 4. the status of predator and prey populations should be evaluated before predator reduction efforts occur. 5. better data on habitat quality should be collected and carrying capacity of the prey’s habitat should be evaluated. 6. changes in the population growth rate of prey and in hunter satisfaction should be monitored. 7. the scope of studies of predators and prey should be broadened and better data on bear ecology should be collected. 8. development of long-term data sets should continue and better data on long-term consequences of control should be collected. the carrying capacity of moose habitat should be further investigated. 10. decision-makers should be more sensitive to signs of over-harvest. 11. decision-makers should be more conservative in setting hunting regulations and designing control efforts. the nrc review also contained a section on decision-making that reiterated several of the standards and guidelines listed above and provided additional standards (nrc deciding whether or not to reduce predators was to identify reasons for wanting more ungulates. these include biological emergencies, subsistence emergencies, lifestyle and recreational hunting demands, and viewing ungulate numbers must be increased should be determined. population models and costnecessary to meet the projected demand and to estimate costs of predator reduction. once these issues have been addressed, ecological investigations should be conducted to assess the likelihood that predator reduction will achieve desired goals. necessary studies include: historic population trends of ungulates, current ungulate population trends, emigration studies, an evaluation of habitat conditions, studies of predator and identifying ecological consequences of predator control. and survival or decrease predation rates. these include habitat manipulation to improve the quantity, quality, or distribution of habitats; non-lethal control methods for predators including diversionary feeding, sterilization, and translocation; selective removal of individual animals or wolf packs; timing of removal methods to identify those that are most hutions to concentrate actions in critical areas effects on predator populations. finally, predator reductions must be show clear results. the report noted that most past programs resulted in unclear results. pre-treatment and post-treatment monitoring areas were not maintained, and weather conditions were often poorly measured. “wherever possible, predator control programs should be design to ensure that knowledge is one of 1997:130). predator control in alaska – van ballenberghe alces vol. 42, 2006 6 application of the nrc’s recommended standards, 2000-2001 nrc’s recommended predator control standards and guidelines came in 2000 and 2001 when alaska addressed a long-standing demand for wolf control by residents of mcgrath on the kuskokwim river in interior alaska. in 1995 the bog received reports from local residents that moose numbers had declined greatly from high levels in the 1970s and wolves were thought to be keeping moose numbers from increasing. preliminary data collected by adfg indicated a moose:wolf ratio of 12:1. the bog approved a control program to take 80% of the wolves in the area but the program was not implemented, nor were similar plans approved subsequently. governor tony knowles appointed a stakeholder’s group called the “adaptive wildlife management team” in 2000 to review the issues and to provide recommendations to the adfg commissioner. the team found that the moose population to support the harvest demand of 130-150 annually. adfg biologists estimated that 3,000-3,500 moose could provide the desired harvest and the team adopted this and the desired harvest as population and harvest goals (adfg 2001). the team recognized that there moose, quality of moose habitat in relation to moose body condition and pregnancy rates, movements of moose in the area, and more precise estimates of moose, wolf, and bear populations. adfg biologists prepared a detailed study plan that was peer reviewed by the team recommended a program of wolf and bear reduction involving wolf trapping by local residents followed by aerial shooting (adfg 2001). bear hunting by local residents would be encouraged if bear predation on neonate moose was found to be important. moose hunting seasons in a portion of the area would be closed until the moose population increased. studies and monitoring the entire program would be conducted in would reconvene periodically to review progress and suggest alternate approaches as necessary. adfg’s commissioner approved the plan early in 2001 with the provision that aerial shooting of wolves would be done by adfg employees using helicopters rather than by bog approved the plan, but before it could be implemented a moose census in autumn 2001 indicated 3,660 moose in the area versus the previous claim of 869. clearly, previous estimates were based on faulty censuses done under poor conditions and moose numbers plans to reduce predators were suspended in light of this new information. in general, many of the nrc’s recommendations were followed in designing this predator reduction was to begin immediately rather than be delayed pending additional data despite very limited information on key predation. and, wolf control, bear reduction, and moose hunting closures were to be simultaneously applied thereby confounding interpretation of results and complicating assessment of the relative importance of these limiting factors. predator control programs 2003-2004 frank murkowski was elected governor of alaska in november 2002 and shortly thereafter appointed 5 new members to the 7-member review the mcgrath program. in march 2003 the board approved a predator control program for the mcgrath area incorporating several important changes from the previous plan alces vol. 42, 2006 van ballenberghe predator control in alaska 7 (bog 2003a). aerial shooting of wolves by private pilots under permits issued by adfg replaced the proposed helicopter-shooting program conducted by adfg employees. about 35-45 wolves were thought to be in the control area and all were scheduled to be shot. bears were to be translocated after capture by adfg personnel. the adaptive wildlife management team was disbanded. subsequently, the wolf control area was doubled in size and the moose population objective was doubled with no in-depth assessment of habitat conditions or carrying capacity. the harvest objective for moose in the area was increased from 130-150 to 400-600. and the peer-reviewed study plan designed to guide research and monitoring was shelved. a second predator control program was approved in 2003 for the nelchina basin (game management unit 13, hereafter unit 13) (state of alaska 2004a). unlike other areas of concern, moose in unit 13 remained at moderate densities following declines from higher levels in the 1980s (bog 2003b). but the bog approved a control program under provisions in the intensive management statute to restore ungulate populations to former levels of abundance. about 140 wolves in the control area were to be shot by private pilots and moose hunting seasons would continue during the control program. in accordance with previous research indicating heavy bear predation on moose in this area, liberal bear hunting seasons and bag limits continued, numbers were approved. no study plan was required and no additional data collection conducted to obtain routine management information. limited data on habitat quality were available, indicating persistently heavy use of important browse species by moose in several areas, but carrying capacity was additional animals. during winter 2003-2004, 17 wolves were taken near mcgrath by aerial shooting with 11 more taken by trappers. private pilots took 127 wolves in unit 13. in spring 2003, 90 bears were translocated at mcgrath, with 35 additional bears moved in spring 2004. the bog approved two additional predator control programs in march 2004. these include an area in upper cook inlet near anchorage (unit 16b). moose numbers and harvests were thought to have declined during the past 10 years while wolf numbers increased (bog 2004a). no quantitative data were available on the effect of wolf predation on moose numbers. bears were suspected to be important predators of moose but no quantitative data were available. habitat conditions and carrying capacity were unknown. despite wolf control program using private pilots under permit to take about 80% of the wolves in the control area beginning in autumn 2004 (state of alaska 2004b). moose hunting seasons remained open and no further steps to reduce bear numbers were approved. a study plan was not required and no additional data collection routine management information. the second program approved in 2004 was in game management unit 19 (unit 19) in the central kuskokwim river area of interior alaska. moose numbers in this area apparently declined during the 1990s but crude estimates suggest moderate densities persist relative to other areas in interior alaska (bog 2004b). as is the case for unit 16b, no quantitative moose, on the effect of bear predation, or on moose habitat quality and carrying capacity. the bog approved a control program using private pilots to shoot wolves in this area beginning in autumn 2004 (state of alaska 2004c). moose hunting seasons were not closed. no further steps were approved to reduce bear numbers other than through continuation of liberal hunting seasons and bag limits. a study plan was not required and no additional predator control in alaska – van ballenberghe alces vol. 42, 2006 8 gathered for routine management. the bog approved an additional program in november 2004 (board of game 2004c). the program includes portions of two game management units, 12 and 20e, located in the eastern interior. wolves will be reduced in an area of about 6,600 square miles; brown bears will be reduced in a 2,700 square mile portion of the total area. wolves will be taken by public aerial shooting and bears will be baited. up to 60% of the bear population may be removed. research during the 1980s and limiting moose and caribou in this area (gasaway et al. 1992) and a wolf sterilization and implemented. as with the other programs, a study plan was not required, and there were no plans to collect additional data. the 5 areas approved by the bog for predator control in 2003-2004 (mcgrath, unit 13, unit 16b, unit 19, and units 12 and 20e) total about 43,000 square miles. private pilots with permits to shoot wolves may take up to 610 wolves in winter 2004-2005. this will be in addition to wolves taken in routine hunting and trapping seasons that in recent years accounted for 1,500-1,700 animals. how well do the predator control programs approved in 2003-2004 conform to the nrc’s recommended standards and guidelines? in from the process used in 2000-2001 to design a control program in the mcgrath area. for 16b, and units 12 and 20e) did not involve a citizen’s planning team. the unit 19 program was preceded by a team convened to review the issues, but the level of biological detail involved was substantially less than for mcgrath. by disbanding the mcgrath team, the bog lost the opportunity for future valuable input, including that from one resident of mcgrath who served on the team from the outset. for mcgrath, much of the groundwork was complete by 2003 as a result of the team’s efforts. nonetheless, the decision was made to proceed with wolf control despite the 2001 moose census that indicated nearly 4 times as many moose as estimated earlier. studies in progress at mcgrath on moose calf mortality, bear translocation, and moose population characteristics continued through 2004. similar studies are not in progress in any of the other areas, and the bog did not identify the need for such studies when it approved additional programs despite obvious the bog failed to recognize the imporcomponents of predator control programs including current, quantitative data on predator and prey numbers. this ignored the nrc guideline of evaluating the status of predator and prey populations prior to predator reduction. furthermore, the bog risked repeating the mistakes made in some control programs conducted in 1976-1983, as well as later at mcgrath, where prey numbers were greatly underestimated and wolf control was suspended when adequate censuses occurred. the bog’s approval of wolf control in unit 16b despite warnings from adfg that data alarming. the bog also retreated from the mcgrath model’s approach of requiring study plans that provided protocols for implementing, monitoring, and evaluating predator control actions and for conducting additional studies. peer review of the mcgrath plan in 2001 by biologists outside adfg with no stake in the plan’s outcome resulted in several adfg revisions to the study plan. similar reviews of plans for other areas, if they had been required, would undoubtedly have resulted in improved designs. most previous predator control programs in alces vol. 42, 2006 van ballenberghe predator control in alaska 9 alaska and canada had unclear outcomes, in part because the programs were primarily management actions based on particular assumptions about predator-prey dynamics. these programs were not designed to test those assumptions. “as a result, less has been have been possible had they been better deby continuing to implement similar managerecent bog actions will result in more unclear outcomes and continued inability to improve the design of future programs. a consistent and often repeated concern in the nrc review pertained to ungulate habitat quality and carrying capacity issues. obviously, predator reductions will not result in increased ungulate numbers if the necessary habitat to support more animals is lacking. in theory, all predators could be removed with no response in ungulate numbers if habitat linking ungulate nutrition, body condition, growth rates, pregnancy rates, and survival to habitat quality (klein 1981). furthermore, winter severity can lower carrying capacity as snow buries forage and increases the energy costs of movement (parker and robbins 1984). the nrc review recognized these important to predator control programs, and provided suggested guidelines for incorporating them in management actions. the bog’s approach in approving recent control programs was to accept crude, qualitative information and broad generalizations on habitat quality and carrying capacity rather than requiring quantitative data. this is a serious breach of recommended standards. in general, the bog’s recent approval of programs to reduce wolf and bear numbers, in an attempt to increase ungulates, represents a retreat from the sound science standard in place in alaska the previous decade. arguably, most of the important biological standards and guidelines recommended by the nrc (1997) have not been followed. the nrc strongly recommended that predator control should be done as adaptive management, that management actions should be planned so that outcomes are clear, and that programs with a low probability of success should be avoided. contrary to nrc recommendations, the bog has begun a process where there is less attenof results and more reliance on anecdotal and qualitative information to justify control programs. this approach jeopardizes progress made during the past 2 decades in applying science-based management to the controversial practice of predator control in alaska. a fundamental, underlying problem in applying recommended biological standards and guidelines to predator control in alaska is the state’s intensive management statute. this 1994 law requires a political standard aimed at restoring depleted ungulate populations to previously attained levels including historical highs. in many instances such highs resulted from irruptions linked to large-scale predator control in the 1950s and 1960s. peak populations were clearly unsustainable and restoring them now is likely unattainable. furthermore, estimates of the magnitude of peak populations, even those reached as recently as the 1980s, are often little more than guesses and despite these problems, the bog, guided by the intensive management statute, has consistently set ungulate population and harvest objectives at high levels, or, as was the case in mcgrath, raised previous objectives in the absence of data on habitat quality and carrying capacity. the net result of this is to commit the bog to approving perpetual predator control programs that chase unattainable objectives. such an approach may repeat the historical pattern of wolf and bear control that triggered ungulate irruptions and subsequent habitat damage and sharp ungulate declines. predator control in alaska – van ballenberghe alces vol. 42, 2006 10 intensive management and its accompanying widespread predator control will likely place established conservation principles at risk in those areas where predator control programs are implemented. fortunately, predator control will not occur on some federal land including national parks. conservation principles at risk include sustainability of certain carrying capacity as wolf and bear populations are suppressed. past predator control programs in alaska, including the federal poisoning — ungulate irruptions were triggered followed by habitat damage and ungulate declines. but the tanana flats moose that increased from 2,800 in 1975 to about 11,000 by the early 1990s following wolf control from 1976 to 1983. now, the population is estimated at 16,000 and shows density-dependent signs of including reduced twinning rates, poor body condition, reduced growth of calves, female reproductive pauses, and increased age of intensity of winter forage plants is very high. in response, the bog recently increased the moose population objective by about 10 %, a debatable management strategy. alaska’s record of managing high-density ungulate populations demonstrates a consistent carrying capacity or quickly responding once problems are apparent. clearly, the irruptions of the 1950s and 1960s were unmanaged and the resulting sharp declines were, in some rently, the tanana flats moose population is at high density as a result of past wolf control, but despite recognizing the problem, managers were unable to respond quickly; antlerless harvests did not begin until 2003. public opposition to harvesting cow moose has complicated matters. in the nelchina basin, moose increased during the 1980s as wolves were heavily harvested. moose declined in response to severe winters thereafter. managers failed to anticipate the decline, having overestimated carrying capacity. now, despite moderate moose densities, predator control aims to again increase the nelchina basin moose population and repeat past patterns of increases and declines in response to winter weather. the board of game’s recent approval of programs to reduce bear and wolf numbers in an attempt to increase ungulates represents a retreat from earlier programs that incorporated most of the nrc’s major biological standards and guidelines. arguably, most of those standards are not implemented currently with monitoring, and more reliance on anecdotal and qualitative information. this approach and wasted effort with failure of ungulate numbers to increase at worst, if undetected factors rather than predation are limiting. alaska’s intensive management statute is a major barrier to implementation of the nrc’s recommendations. efforts to chase unattainable population and harvest objectives with poorly designed predator control programs risk long-term sustainability of ungulates, protection of habitat integrity, and predator population viability. references (adfg) alaska department of fish and game. 1983. wolf management programs in alaska 1975-1983. department of fish and game, juneau, alaska, usa. rebuild the moose population in gmu 19d. prepared by the adaptive wildlife management team. department of fish and game, juneau, alaska, usa. (bog) board of game, state of alaska. 2003a. findings of the board of game and guidelines for a unit 19d east predation control program. department of fish and alces vol. 42, 2006 van ballenberghe predator control in alaska 11 game, juneau, alaska, usa. _______. 2003b. findings of the alaska board of game authorizing wolf control in portions of unit 13. department of fish and game, juneau, alaska, usa. _______. 2004a. findings of the alaska board of game authorizing predator control in the western cook inlet area in unit 16b with airborne or same day shooting. department of fish and game, juneau, alaska, usa. _______. 2004b. findings of the alaska board of game authorizing wolf predation control in the unit 19a portion of the central kuskokwim wolf predation control area with airborne or same day airborne shooting. department of fish and game, juneau, alaska, usa _______. 2004c. findings of the alaska board of game authorizing wolf and bear predation control in portions of the updepartment of fish and game, juneau, alaska, usa. franzmann, a. w. 1993. biopolitics of wolf management in alaska. alces 29:9-26. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose implications for conservation. wildlife monographs 120. _____, r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. harbo, s. j., and f. c. dean. 1983. historical and current perspectives on wolf management in alaska. pages 51-65 in l. n. carbyn, editor. wolves in canada and alaska. report series 45, canadian wildlife service, ottawa, ontario, canada. klein, d. r. 1981. the problems of overpopulation of deer in north america. pages 119-127 in p. a. jewell and s. holt, editors. problems in management of locally abundant wild mammals. academic (nrc) national research council. 1997. wolves, bears and their prey in alaska. national academy press, washington, d.c., usa. parker, k. l., and c. t. robbins. 1984. thermoregulation in mule deer (odocoileus hemionus hemionus) and elk (cervus elaphus nelsoni). canadian journal of zoology 62:1409-1422. state of alaska. 2004a. game management unit 13, wolf predation control area. 5aac 92.125. wolf predation control implementation plan. register 170, state regulations. department of fish and game, juneau, alaska, usa. _____. 2004b. game management unit 16b, wolf predation control area. 5aac 92.125. wolf predation control implementation plan. register 170, state regulations. department of fish and game, juneau, alaska, usa. _____. 2004c. game management unit 19a and 19b, wolf predation control area. 5aac 92.125. wolf predation control implementation plan. register 170, state regulations. department of fish and game, juneau, alaska, usa. van ballenberghe, v. 1985. wolf predation on caribou: the nelchina herd case history. journal of wildlife management 49:711-720. young, d. 2004. status of the gmu 20a moose population. report to the alaska board of game. department of fish and game, juneau, alaska, usa. alces37(1)_35.pdf f:\alces\vol_38\pagema~1\3808.pdf alces vol. 38, 2002 crichton horseshoe posture 109 the horseshoe posture in moose a reaction to perceived threats vince crichton manitoba conservation, box 24, 200 saulteaux crescent, winnipeg, mb, canada r3j 3w5 abstract: a behaviour pattern of bull moose was noted when confronted with aircraft during surveys and occasionally on the ground when threatened. the pattern is called the horseshoe posture and involves shifting of the back hooves toward the front hooves with the legs now forming a triangle with the abdomen. the pattern was noted in yearling, teen, prime, and senior bulls, but not cows. no head tilting was noted as described for caribou during the rut, but there was rubbing together of the tarsal glands. bulls did not exhibit the horseshoe posture when confronted with an artificial head and antlers. alces vol. 38: 109-111 (2002) key words: bull moose, horseshoe posture, identification feature, response to threat some postures related to specific activities of moose have not been well described. the posture described by bubenik (1975) as the horseshoe posture has been reported infrequently. he described a tilted head display frequently done by bull caribou (rangifer tarandus) during aggression displays in the rutting period. he further suggested that the tilted head is common in many ungulates and is universal in the 'horned' ungulates. in all these species the head is postured frontally and the attitude is used when a foreign object is seen. he explained the tilted head posture in reindeer as typical for an offensive-defensive threat and that it could be used as a single element with the neck in line with the body or turned to the side. it is also used as a reinforcing or a complementary element in other searchimages and resembles a ‘horseshoe’ posture. this display is common among odocoileinae. it shows slight specific variation in white-tailed deer (odocoileus virginianus) where it is performed in both sexes. in the moose (alces alces) it is limited to the male, and only when incited. in general, the hind legs are close to the forelegs, the back is arched, penis erect or hanging down, and tarsal glands together and (not always) moistened by urine (bubenik 1975). he suggests it is performed fully by alpha or solitary males. bubenik (1975) also suggests that in caribou the ‘white-of-eyes’ and husky sounds can be mixed in the display of the tilted head. bubenik (1975) describes encounters with bull caribou interspersed with females in which some bulls were responding with the horseshoe posture, an erect penis and moistening of the tarsal glands. crichton (1987) suggested that this behaviour is a reliable technique for the identification of bull moose observed during aerial surveys. timmermann and buss (1998) suggested that when aroused from beds by searching aircraft, adult bulls frequently move their hind legs forward in a urination-like posture and that only rarely do rising cows exhibit the same behaviour. beginning in the winter of 1972/73, while flying moose surveys in manitoba i noticed some animals when aroused from their beds would shift the back feet forward, positioning them just posterior to the front feet. in this position, the ventral portion of the abdomen along with the front and rear legs when horseshoe posture crichton alces vol. 38, 2002 110 viewed from the side resembled a triangle. further, it became apparent that the only moose performing this behaviour were males either with or without antlers and that it was being done by all ages including calves-ofthe-year. although the total number of occasions this behaviour was seen was not recorded, it is estimated that it was approximately 750 times. bulls would maintain this position for 10-20 seconds and then move off either running or walking. the first opportunity to observe this behaviour on the ground was in riding mountain national park in manitoba in early january when the snow depth was about 20 cm. on this occasion, while travelling on a park road, i noted 6 moose in a small meadow. as i approached on foot with only a camera, all ran off with the exception of 3 bulls which, based on antler architecture, i classified as yearlings and 2-year olds. the bulls watched intently as i slowly approached to within about 75 m at which time all 3 displayed the horseshoe posture. with heads erect and pointed forward, they maintained the posture for 10-20 seconds and then bolted. this behaviour was recorded on a video camera. i examined each location where they had been standing and there was evidence that they had urinated on their hind legs, as there was urine on the snow at each site. the second occasion for viewing this behaviour close at hand was during a september moose hunt. while sitting in a canoe on the bloodvein river in eastern manitoba a yearling bull appeared on the riverbank nonchalantly feeding and oblivious to our presence. we watched the animal for a few minutes until he noticed us at which time we were about 100 m away. he immediately stopped feeding and his body length was parallel to the river. he assumed the horseshoe posture with the head erect and pointing forward, held the position for 10-20 seconds and subsequently ran off. the site where he was standing was examined but i was unable to detect evidence of urination. the third, and perhaps best opportunity for viewing and recording this behaviour from the ground, occurred in july in the chapleau crown game preserve in northern ontario. while travelling north on a road within the preserve, i noted a mature bull with velveted antlers feeding in a small pond adjacent to the road. he moved off when the vehicle stopped. i was able to enter the bush without being seen and followed above and slightly behind him on a ridge as he moved slowly along stripping leaves from aspen. he stopped and laid down after about 300 m. i approached unobserved to within about 15 m of his resting location and remained here for about one half hour taking video and noting resting behaviours. eventually, i emitted the sound of a cow and after repeated calls he rose from his bed, looked in my direction and slowly moved off toward the road feeding as he went. as he approached the road, a moving vehicle alerted him and he stopped and remained motionless for a few minutes, turned around and moved back to where i was now standing in plain view on the ridge. as he approached, he stepped over a fallen log, stopped and assumed the horseshoe posture with an uplifted forward pointing head, shuffled the hind feet toward the front (fig. 1) and rubbed the tarsal glands together. he held this position for about 15 seconds and then bolted past me. at no time was there any vocalization. most of his activities, from the time he was aroused from his bed until he disappeared, including assuming the horseshoe posture, were recorded on video. i examined the site as well as the video but was unable to ascertain if urination had occurred. the fourth occasion to view this behaviour occurred while attending the 37th north american moose conference in northern alces vol. 38, 2002 crichton horseshoe posture 111 maine. while watching a bull moose early one morning at a mineral lick, the bull became rather agitated as the sound of traffic increased on the highway about 150 m away. when a large truck went by, the bull ran for a few steps, suddenly stopped, assumed the horseshoe posture and urinated profusely on his hocks. his level of agitation remained and about 2 minutes after assuming the posture the first time, he repeated the behaviour again, with some urination on the tarsal gland, before running off into the adjacent bush. the behaviour was again recorded with a video camera. in all situations, by assuming the horseshoe posture, the bulls, whether they be yearlings, teens, primes, or seniors, appear to be responding to perceived threats. nothing resembling the tilted head posture as described by bubenik (1975) for caribou was noted. neither the tilted head phenomenon nor the horseshoe posture have been observed in bull moose on the ground when i have approached them while wearing an artificial head and antlers during the rutting period. in each observation, whether it be on the ground or from an aircraft, the head was held in an upright and forward pointing position. during aerial surveys, animals which are observed to perform this behaviour can be classed as bulls. in all situations where the behaviour was observed, whether it be on the ground or from an aircraft, the situation could, from the moose’s perspective, be viewed as an anthropogenic threat, with the gesture by some being to assume the horseshoe posture. acknowledgements i am grateful for the assistance of jim johnson, pilot and owner of northway aviation, particularly for his friendship, flying, and observational skills and his keen interest in doing moose and other wildlife surveys for many years. references bubenik, a. b. 1975. significance of antlers in the social life of barren-ground caribou. university of alaska, special report 1:436-461. crichton, v.f.j. 1987. procedure for standardized moose sex and age surveys in manitoba. manitoba department of natural resources. winnipeg, manitoba, canada. timmermann, h. r., and m. e. buss. 1998. population and harvest management. pages 559-615 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. fig. 1. bull moose displaying horseshoe posture. note the forward pointing, uplifted head, position of the back hooves in relation to the front hooves and the closeness of the back legs compared to the wider stance of the front legs. from the side, the front and back legs and the abdomen resemble a triangle. 159 moose and deer population trends in northwestern ontario: a case history bruce ranta1 and murray lankester2 1311 austin lake road, kenora, ontario, canada p9n 4n2; retired; 2101-2001 blue jay place, courtenay, british columbia, canada v9n 4a8; retired. abstract: many interrelated factors contribute to the rise and fall of white-tailed deer (odocoileus virginianus) and moose (alces alces) populations in the mixed boreal forests of eastern north america where these species often cohabit. a question not satisfactorily answered is why do moose populations periodically decline in a pronounced and prolonged way while deer populations continue to do well during times when habitat conditions appear good for both? long-term historical data from the kenora district of northwestern ontario, canada provided an opportunity to better understand temporal relationships between trends in deer and moose numbers and landscape-level habitat disturbances, ensuing forest succession, climate, predators, and disease. over the past 100 years, moose and deer have fluctuated through 2 high-low population cycles. deer numbers were high and moose numbers were low in the 1940s and 50s following a spruce budworm (choristoneura fumiferana) outbreak. by the early 1960s, deer trended downwards and remained low during an extended period with frequent deep-snow winters; as deer declined, moose recovery was evident. moose increased through the 1980s and 1990s as did deer, apparently in response to considerable habitat disturbance, including another spruce budworm outbreak and easier winters. however, despite conditions that were favourable for both species, moose declined markedly beginning in the late 1990s, and by 2012 were at very low levels district-wide while deer numbers remained high. despite the moose decline being coincident with a short-lived winter tick (dermacentor albipictus) epizootic in the early 2000s and increasing numbers of wolves (canis lupus), we argue that the meningeal worm (parelaphostrongylus tenuis) likely played a major role in this moose decline. alces vol. 53: 159–179 (2017) key words: landscape disturbance, fire, wind, spruce budworm, forest succession, balsam fir, snow depth, white-tailed deer, moose declines, population fluctuations, meningeal worm, parelaphostrongylus tenuis several factors constrained white-tailed deer (odocoileus virginianus) densities and distribution in the mixed-forest ecotone and regions of the eastern boreal forest until about 200 years ago (seton 1909, voigt et al. 2000). expansion was made possible by forest rejuvenation resulting from human settlement and attendant land clearing, logging, and agricultural practices, as well as increased frequency of forest fires (mcshea et al. 1997). as well, mech et al. (1971) documented how widespread reduction or eradication of predators, primarily wolves (canis lupus), aided and abetted the expansion of deer northward. karns (1980) also argued that the density of deer in northern mixed forests was constrained mostly by the high frequency of cold, deep-snow winters rather than food limitations. notwithstanding a lack of agreement on the relative importance of these limiting factors, periodic increases in the abundance of deer in the northern forests of eastern north america have had consequences for caribou (rangifer moose and deer population trends – ranta and lankester alces vol. 53, 2017 160 tarandus) (racey and armstrong 2000) and moose (alces alces) (anderson 1972, lankester and samuel 2007). in the past century, deer at the northern limits of their range in ontario have reached sustained high densities at least twice; in the 1940s and 1950s and again in the 1990s and 2000s (thompson 2000b). moose declined noticeably in the kenora district in northwestern ontario (kd) in each of these deer growth periods. these events in ontario mirrored recent prominent deer eruptions concurrent with pronounced moose declines in the eastern forests of mainland nova scotia, the northern mixed forests of minnesota, and in adjacent northeastern north dakota (parker 2003, beazley et al. 2006, murray et al. 2006, maskey 2008, lankester 2010, lenarz et al. 2010). although not universally accepted (lenarz 2009), the concurrence of sustained high deer populations and falling moose numbers is supported by numerous anecdotal accounts (early authors reviewed by anderson 1972, lankester and samuel 2007) and by empirical data (whitlaw and lankester 1994a, b, maskey 2008). within the present day boundaries of the kd, changes in the presence and abundance of a variety of cervids have been particularly dynamic. this area includes the aulneau peninsula where, beginning in about 1997, moose declined from more than 1/km2 to almost none in <15 years. we review long-term records from kd to better understand the importance of landscape-level forest disturbances, climate, predators, and pathogens including the meningeal worm (parelaphostrongylus tenuis) in determining historical trends in deer and moose populations. study area the kd of the ontario ministry of natural resources and forestry (mnrf, formerly the ontario ministry of natural resources) is located in northwestern ontario (fig. 1) and is bounded by the province of manitoba to the west and the ontario districts of red lake, dryden, and fort frances to the north, east, and south, respectively. the size of the kd changed minimally in 1961, was reduced in total area from 31,5302 to 14,189 km2 in 1972, and was increased to 19,744 km2 in 1992. administratively, the kd consists largely of 3 wildlife management units (wmus 6, 7a, and 7b; fig. 1). wmu 6 is the most northerly covering ~4,700 km2 and has had recent and extensive forestry activity, wildfires, and blowdowns (mnrf unpublished). wmu 7a, the aulneau peninsula, is about 832 km2 and located south of the city of kenora in the middle of lake of the woods. it has a recent history of limited forest management and infrequent wildfire, and contains no allweather roads. wmu 7b lies immediately south of wmu 6 and is >9,000 km2 with limited agricultural activity near kenora and a recent history of extensive forest management and wildfire. moose aquatic feeding areas are abundant among numerous lakes, rivers, and beaver ponds in all 3 wmus. the forest of the more southerly portion of kd is representative of the great lakes – st. lawrence forest region and the more northerly part is classified as mixed-wood boreal forest (rowe 1972). the surficial geology is an area of bedrock with little to no topsoil because of repeated glaciation (zoltai 1961). rich, glacial-lacustrine deposits of varved clays occur, particularly in low-lying valleys. hills are often rugged but most rise less than a few hundred meters from valley floors. the climate is characteristically continental, with a slight moderating effect from the great lakes marine climate (omnr 1974); temperatures range alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 161 fig. 1. the kenora district including its 3 wildlife management units (wmus 6, 7a, 7b) in northwestern ontario, canada. " " " " " kenora forest whiskey jack forest shoal lake umfreville lake wabigoon lake pakwash lake rainy lake (lac à la pluie) eagle lake kakagi lake lake of the woods lac seul lake of the woods (lac des bois) sioux narrows redditt vermillion bay kenora dryden k e n o r a d i s t r i c t fort frances district red lake district dryden district wmu 3 wmu 9b wmu 8 wmu 5 wmu 2 wmu 9a wmu 4 wmu 7a wmu 7b wmu 6 produced by the ministry of natural resources and forestry © queen's printer for ontario, 2017. 0 10 20 30 40 505 kilometers ¯ m a n i t o b a u . s . a . moose and deer population trends – ranta and lankester alces vol. 53, 2017 162 from a january mean of -17 °c to a july mean maximum of 24.5 °c. methods the term kd refers hereafter to the actual geographical extent of the mnrf kd and applies collectively to wmus 6, 7a, and 7b, and as appropriate, to 2 forest management units (mu), the kenora mu and the whiskey jack mu. the boundary of the 2 mus combined is not identical to that of the mnrf kd; some area extends outside, but in total, the combined area is roughly equivalent in area (fig. 1). landscape-scale disturbance events prior to european settlement and pre industrial forest conditions are described in broad terms from available internal historical reports and from survey notes on forest cover recorded circa 1880 to 1930 (see elkie et al. 2009). more recent impacts including spatial and temporal aspects of fires, blowdowns, insect damage, and logging are reported at the district level and supported by empirical data from mnrf files. fire data for the period 1920 to 2010 were reviewed and expressed as numbers of ha burned annually from 1963 to 2007. large fires (>4000 ha) occurring from 1975 to 2010 were mapped, as were large blowdowns occurring since 1980. insect infestation data was mostly limited to outbreaks of the eastern spruce budworm (choristoneura fumiferana) and the jack pine budworm (c. pinus pinus). an index of winter severity has been measured in the kd since 1952. early data (passmore 1953) were converted to a cumulative, over-winter, snow depth index (sdi) (warren et al. 1998). two snow stations have existed in the kd since the onset of the program; a third was added in 1960. one was in wmu 6 near the town of minaki (mk); the other 2 were in wmu 7b near the towns of kenora (kr) and sioux narrows (sk). snow stations were located in open hardwood stands and snow depth (cm) was measured at 10 sites, 10 m apart, and averaged weekly. the weekly averages were summed from the first to last snow of the season. winter severity was equated to sdi values using the following classification: <590 = mild; 591 to 760 = moderate; >760 = severe (omnr 1997, warren et al. 1998, and with permission of mnrf snow network for ontario wildlife). the sdi values from each station were averaged to provide a district wide sdi ranking. mean differences between time periods for total rainfall, snow depth index, and length of growing season were examined using student’s t-test (twosample, unequal variance) and accepted as different when p < 0.05. historical weather data including total annual rainfall and the length of the frostfree season were obtained for the kr in the period 1960 to 2013 from the environment canada website (http://climate.weather. gc.ca/historical_data/search_historic_ data_e.html). the length of the frost-free season was determined as the difference between the first day of the first 5 consecutive days in spring with minimum temperature > 0 oc and the day before the first 5 consecutive days in autumn with minimum temperature < 0 oc. several data sources were used to estimate past trends in deer populations, with other information subjective in nature and formed by expert opinion. data included hunter numbers and deer harvest information collected at check stations and from voluntary questionnaires. district-wide data from 1955 to 1960 were limited to % hunter success, with total deer harvest and % hunter success available thereafter. two time periods were compared using information pertaining to the kd: 1961 to 1978 and 1999 to 2012. in the intervening time period, 1981 to http://climate.weather.gc.ca/historical_data/search_historic_data_e.html http://climate.weather.gc.ca/historical_data/search_historic_data_e.html http://climate.weather.gc.ca/historical_data/search_historic_data_e.html alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 163 1997, only data from wmu 7b were examined in detail. during the time period of 1963 to 1982, pellet group surveys in certain years provided additional deer density estimates in specific wintering areas and the larger landscape. pellet group surveys in wmus 6 and 7 from 1976 to 1978 followed king (undated), and a 1982 survey in wmus 7a and 7b followed jones (1981). the number of deer observed during moose aerial inventories (mai) in the kd were recorded as the average number of deer per plot in 2 periods: 1994 to 1999 and 2000 to 2012. since 1957, moose numbers and population trends in ontario (including the kd) have been estimated from mid-winter mais based largely on caughley (1977a, b). after 1972, mais were done at the wmu level and their frequency declined after 1992. mais were conducted using 16 mi2 plots until 1975 when standardized surveys for wmus were adopted using 25 km2 plots (mclaren 2006). surveys were random or random-stratified depending on a variety of factors, particularly prior knowledge of relative moose abundance and distribution patterns. they were conducted using both fixed-wing and rotary aircraft, and searches followed the methodology outlined by oswald (1997). generally, mais in wmus 6 and 7b were flown with the objective of achieving a 90% confidence level (± 20%). however, mais in wmu 7a were often done with 50% coverage which tended to provide higher confidence levels. voluntary provincial hunter questionnaires and mail surveys were also used as a proxy to provide estimates of moose populations and to aid moose management. black bear (ursus americana) harvests from 1987 to 2010 in wmu 7b were estimated using returns from voluntary provincial mail surveys (resident hunters) and information from the returns of mandatory validation certificates (non-resident hunters). wolf (canis lupus) sightings in wmu 7b were estimated in 2000 to 2010 using information from provincial mail surveys sent to resident and non-resident deer hunters. office files and the published literature were searched for evidence of the presence of meningeal worm in deer and moose, as well as cases of moose sickness attributed to meningeal worm infection in the kd and surrounding region. anecdotal information on the occurrence of giant liver fluke (fascioloides magna) and winter tick (dermacentor albipictus) were recorded. results logging and land clearing — the pre-industrial forest of ecoregion 4s that includes the kd is believed to have been rich in conifer species (elkie et al. 2009). compared to present-day forests, there were more pure stands but similar amounts of young disturbed forest. in general, the pre-industrial forests were believed to have been less fragmented with larger disturbance patches from larger fires. about 20% of ecoregion 4s is believed to have been in the mixed-wood condition (coniferous and deciduous trees), and only about 1% of the forested area was comprised of pure balsam fir (abies balsamae) stands; presently, ~55% of the forested area is mixed-wood of which 7% is pure balsam fir (elkie et al. 2009). recently approved forest management plans for the whiskey jack and kenora mus document that logging began in the kd sometime in the early 1800s and has been more or less continuous since the 1880s. initially, most harvesting was from the nearshore areas of lake of the woods and other large lakes in the vicinity, with the harvest rate increasing substantially after 1890. a paper mill was built in kenora in 1922 further increasing the area logged annually, and a large timberstrand plant opened in 2002. moose and deer population trends – ranta and lankester alces vol. 53, 2017 164 although the paper mill in the city of kenora closed in 2005, a number of local sawmills continue to operate. the area logged annually has varied, but generally, a few 1000 ha of forest are cut annually (< 1% of the total area of kd). harvest data for the kenora and whiskey jack mus were available starting in 1990 (table 1), and mnrf forestry staff report that the greatest amount of harvesting occurred during the 1990s. clearings associated with early european settlement created an area of a few 1000 ha of field and pasture near the present city of kenora. clearcut logging, the silvicultural practice most commonly used in the kd, produces an abundance of summer forage, although the interior of very large clearcuts (e.g., cover-to-cover distance >400 m) may be used little by deer (thomas et al. 1979, roseberry and woolf 1998) or moose (hamilton et al. 1980, thompson and vukelich 1981, allen et al. 1987). however, owing to terrain and other factors, clearcuts in the kd have tended to be relatively small. fire — the amount of area burned each year in the kd has varied from <100 to >100,000 ha. large areas burned in the 1920s and 1930s, with fires much less frequent in the 1940s and 1950s (table 2). more recently, in the mid-1970s to the late 1980s, large areas were again burned, mainly by big fires in 1976, 1980, 1983, and 1988 (fig. 2 and 3). the area burned annually from 1989 to 2007 was relatively small ( generally < 100 ha/year) and has remained so; of note is the absence of fires since 1933 on the aulneau peninsula (wmu 7a). blow-down — in some years, blowdown affects large swaths of living forest in the kd; forestry staff report that, in general, small blowdown events occur annually. a large blowdown in 1991 covered > 63,600 ha, much of it in wmu 6, and a number of blowdown events in 2005 totaled ~93,000 ha (table 2, fig. 4). insect damage — landscape-scale insect damage is attributed to spruce budworm, jack pine budworm, and forest tent caterpillar (malacosoma disstria). two infestations of spruce budworm in the past century caused substantial mortality of balsam fir, and to a lesser extent, white spruce (picea glauca). jack pine budworm outbreaks tend to be smaller and infrequent, although a large outbreak resulted in extensive mortality of jack pine in certain areas in 2007-2008. forest tent table 1. area of forest harvest (ha), 1990-2014, in the kenora district forests, ontario, canada. decade kenora mu whiskey jack mu total 1990 10,040 57,584 67,624 2000 13,263 32,425 45,688 2010 (4 years, 40% of decade) 5797 2368 8165 table 2. area of landscape-scale disturbances (ha) in the kenora district, ontario, canada. fire blowdowns spruce budworm (1980-98) year area (ha) year area (ha) defoliation area (ha) 1920s 108,942 1991 50,935 moderate to severe 26,175 1930s 77,028 2005 67,942 1940s 8,373 high tree mortality 2,301,341 1950s 1,758 alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 165 caterpillar outbreaks are cyclical (~10 years), but they prefer aspen (populus spp.) and trees tend to recover quickly from defoliation. in 1934, a spruce budworm outbreak reached ‘epidemic proportions’ and by the end of the outbreak in 1947, 5.3 million ha of ontario had been infested, including a sizeable portion of the kd. the second spruce budworm epidemic occurred from about 1980 to 1998, with >8.3 million ha of ontario forests infested. substantial tracts of forest in the kd were categorized as having “moderate to severe defoliation” and >2 million ha had “high tree mortality” (table 2, fig. 5). fig. 2. number of hectares burned from 1963 to 2007, kenora district, ontario, canada. fig. 3. forest fires >4,000 ha from 1975 to 2010, kenora district, ontario, canada. fig. 4. large blow-downs in forests from 1980 to 2010, kenora district, ontario, canada. moose and deer population trends – ranta and lankester alces vol. 53, 2017 166 before the second spruce budworm infestation, kenora omnr management staff estimated that balsam fir composition was ≥  40%  in  mixed-wood  forests.  in  the  later  stage of the epidemic, tree lichen (usnea spp.), was very abundant where balsam fir mortality was high (fig. 5). its availability peaked in the early 2000s, but by 2010, one author (ranta) observed that most dead balsams once laden with lichens had fallen, and most lichens had been (presumably) consumed. the aulneau peninsula — of particular note is wmu 7a, the aulneau peninsula which had no large fires or blowdown since 1933 (omnr 2003), although the effects from spruce budworm were widespread (fig. 5). beginning in 1964, however, ~15,000 m3 were logged annually, generally as small (<100 ha) dispersed cuts that district staff believe greatly improved moose habitat conditions; logging ceased in 1986 and has not resumed. a substantial portion of the aulneau and some parts of wmu 7b had a pronounced loss of coniferous canopy cover as a result of the jack pine budworm outbreak in 2007-2008. although these infestations result in minimal growth of arboreal lichen on dead and dying jack pine trees, removal of the over-story likely stimulated growth in the understory. winter snow depth — winter severity rankings for the 3 snow stations (sn, kr, and mk) ranged from very mild to severe (fig. 6). mean (± sd) annual sdis were greater (p = 0.03) in the period from 1960 to 1980 (801 ± 231) than in the subsequent period from 1981 to 2014 (632 ± 241). the most southerly station (sn) had lower (p = 0.03) mean sdi in the period from 1960 to 2014 (616 ± 235) than the two more northerly stations that were similar (kr = 732 ± 266; mk = 719 ± 244). severe winters with an sdi > 760 were most frequent in the 20-year period from 1960 to 1980 when 11 winters were rated as severe and only 3 as mild (<590); over the next 34 years (1981 to 2014) only 7 winters were rated as severe with 16 as mild (fig. 6). the average maximum snow depths from 1952 to 2014 at the 3 snow stations were 55.4 cm (kr), 54.8 cm (mk), and 50.8 cm (sn). weekly readings exceeded 80 cm on only a few occasions and those depths were generally of short duration; depths >90 cm were recorded in only 2 winters (1955-56 and 2013-14). in the mk depths >80 cm were recorded for 3 consecutive weeks in 1954-55, and for 7 consecutive weeks the following year. in the winter of 1965-66, all 3 snow stations had 1 weekly recording >80 cm; a single weekly reading was 81 cm at kr in 1977-78. during the recent severe winter of 2013-14, snow depth >80 cm occurred at both kr and mk; the maximum depth was 72 cm at sn. annual rainfall and length of the frostfree season — the amount of rainfall and fig. 5. forests with significant spruce budworm damage in 1998, kenora district, ontario, canada. alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 167 the length of the frost-free season are climatically important in the external survival and transmission of parasites such as d. albipictus and p. tenuis. the 20-year period from 1970 to 1990 that had several large fires also received less (p = 0.01) rain (474 ± 115 mm) than in the following 20-year period from 1991 to 2012 (609 ± 120 mm). the mean length of the frost-free period between these time periods was not different (185 ± 16 vs. 188 ± 18 days), ranging from 156 to 214 days and 156 to 223 days, respectively. historical cervid populations — in the late 1800s, caribou and moose occurred in what is presently kd (cumming 1972, darby et al. 1989). seton (1909) believed that deer were largely absent until the late 1800s, but some elk (cervus elaphus) were extant. caribou range began to recede northwards concurrent with the increase in deer numbers (racey and armstrong 2000), with elk disappearing also; moose remain extant to the present. trends in deer numbers — by the 1930s, deer were numerous in the kd and stayed high during the 1950s and early 1960s (cumming 1972). by the late 1960s, deer numbers began to decline, increased somewhat, again declined, then remained relatively low until the mid-1980s (fig. 7). thereafter, deer numbers steadily increased, peaking about 2007. in 2014 a severe winter resulted in high deer mortality and likely substantial recruitment failure. declining hunter success and field observations suggested that deer in the northern portions of kd and deer away from settlements were most affected. records of the number of hunters and deer harvest for wmus 6, 7a, and 7b showed a similar trend from 1974 to 2012 (fig. 8). the annual deer harvest fluctuated, but was relatively low through the 1970s. by the late 1990s, hunter interest and success rates had begun to increase and remained fig. 6. snow depth index (sdi) averaged for 3 snow stations (sioux narrows, kenora, and minaki) from 1952 to 2014, kenora district, ontario, canada. 0 200 400 600 800 1000 1200 1400 sd i year mild < 590 moderate 591-760 severe > 760 moose and deer population trends – ranta and lankester alces vol. 53, 2017 168 high except in 2008 and 2011. in 2002, deer hunting regulations were relaxed and hunters could purchase additional antlerless deer tags in wmus 6, 7a, and 7b. data from this additional deer kill is only available starting in 2008, hence, total deer kill in 2002-2007 is under reported. the number of deer observed per moose plot clearly increased in each of the 3 wmus by the late 1990s, continuing through 2006 (fig. 9). spring pellet group surveys provided a few disjunct estimates of winter deer populations in portions of wmu 7 in 1976 and 1977, and the entire wmu 6 in 1978 (ranta and shaw 1982). density estimates were: wmu 7 1976, 4/km2 (14,557 ± 54.4%); wmu 7 1977, 4/km2 (15,515 ± 28.9%); wmu 6 1978, 1/km2 (3,362 ± 36.93%). recalculation of the wmu 7 1977 survey fig. 7. changes in deer and moose numbers in relation to landscape scale habitat disturbances from 1955 to 2014, kenora district, ontario, canada. 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.2 0.4 0.6 0.8 1 1.2 d ee r h ar ve st s u cc es s (# d ee r p er h u n te r) ) mk qs / #( ytis ne d es o o m deta mitse moose pop'n density deer harvest success drier period we�er period fires (4) bud worm 18 yrs blowdowns less snowdeeper snow fig. 8. deer hunters and harvest for wmus 6, 7a, and 7b, 1974 to 2012, in the kenora district, ontario, canada. 0 1000 2000 3000 4000 5000 6000 7000 19 74 19 75 19 76 19 77 19 78 19 99 20 00 20 01 20 02 20 03 20 04 20 05 20 06 20 07 20 08 20 09 20 10 20 11 20 12 n u m b er year harvest no hunters fig. 9. number of deer observed per moose aerial survey plot in wmus 7a, 7b, and 6, kenora district, ontario, canada. 0 2 4 6 8 10 12 14 19 94 19 96 19 98 20 00 20 02 20 04 20 06 20 08 20 10 20 12 a ve . n o . d ee r se en /m o o se p lo t year wmu 6 wmu 7a wmu 7b alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 169 data led to a higher estimate of 31,000 deer. in 1979, in response to a suspected drastic population decline after a severe winter, supplementary pellet group surveys were performed and indicated that the deer population in wmu 6 had declined 75% from the previous year, and in wmu 7, 55% from 2 years previous. the mid-winter deer population estimate from pellet group surveys in 1982 was only 47 ± 76.8% in wmu 7a and 10,231 ± 41.2% in wmu 7b (ranta and shaw 1982). deer numbers were considered relatively low throughout the period of 1976 to 1982. trends in moose numbers — moose were described as fairly common in the lake of the woods area by europeans as early as 1731 (cumming 1972, darby et al. 1989). the population declined in the 1800s with the growing population of settlers, survey crews, and road builders relying largely on market meat. in response to perceived low populations, the moose hunting season was closed across the entire province from 1888 to 1895; thereafter, moose numbers apparently increased. there are few estimates of moose numbers in the kd in the early years of the 20th century. cumming (1972) reported that the royal ontario museum (from questionnaires) believed that the provincial moose population declined prior to wwii, increased during the war years, but was considered low in 1949 when the hunting season was again closed. it was re-opened in 1951 when populations across the province appeared to have increased, although actual population estimates only began in the late 1950s. in 1957, mai data indicated that moose were at fairly constant and moderate density of ~0.2/km2 (fig. 7). as deer numbers declined in the 1960s, moose numbers increased slowly, continuing into the late 1980s and 1990s when they peaked ~ 0.4/km2; beginning about 1995, moose began to decline reaching very low numbers by 2012 (fig. 7). moose hunter survey information in wmu 7b (fig. 10) was used to corroborate the mai data. increased harvest began in the late 1980s until about 2001, after which success rates began to decline to present day lows. concurrently, deer numbers increased until about 2007, remaining high to 2012 (fig. 8). after the severe winter of 2014, deer numbers declined throughout the kd and adjacent districts (unpublished mnrf data). when the mai data for the 3 wmus are examined separately for the years 1980 to 2010, it appears that the timing of the moose decline differed slightly in each (fig. 11). a decline from high density was first evident in the most southerly unit (wmu 7a) after 1995, a distinct decline occurred in wmu 7b after 2001, and decline occurred after 2004 in the most northerly wmu 6; numbers remain low in all. in 1972 on the aulneau peninsula (wmu 7a), the moose population was estimated at only ~80 animals (about 0.1/km2), but by 1994 had peaked at ~1000 animals (about 1.0/km2) with numbers still relatively high in 2000; however, rapid decline occurred thereafter, and an aerial survey estimated only ~30 animals in 2011 (< 0.04/km2). fig. 10. hunter harvest of moose and bear, and number of wolves observed by hunters in wmu 7b, 1984-2010, kenora district, ontario, canada. wmu 7b moose, bear harvest; wolves seen 0 100 200 300 400 500 600 19 84 19 87 19 90 19 93 19 96 19 99 20 02 20 05 20 08 n u m b er resident moose harvest wolves seen (10%) bear harvest (all) moose and deer population trends – ranta and lankester alces vol. 53, 2017 170 predators — both black bears and timber wolves are common and ubiquitous in all 3 wmus. data from early years are largely limited to anecdotal information, but no concerns about either animal being ‘rare’ or in decline are on record. bear harvest from 1988 to 2009 rapidly increased, peaking around 1996 in all wmus (fig. 10). thereafter, harvest declined sharply with the lowest combined harvests in 2006 and 2007 across the wmus; harvest declined least on the aulneau (wmu 7a) and most in wmu 7b. records of wolf sightings by deer hunters in wmu 7b began in 2000 and indicated an initial wolf decline for 3 years, followed by a large increase to 2010 (fig. 10). evidence of meningeal worm and other pathogens — a number of surveys have documented the continued presence of p. tenuis larvae in deer feces and moose deaths attributed to meningeal worm in northwestern ontario and adjoining regions. the prevalence of first-stage larvae in deer pellets ranged from 57-85% in the kd (saunders 1973, whitlaw and lankester 1994b, mcintosh 2003) and 86% in the adjacent fort frances district (mcintosh 2003). also in the fort frances district, 3 cases of moose sickness caused by p. tenuis were diagnosed by anderson (1965) and 14 cases were reported to the district office during the 12-year period 1980 to 1992 (whitlaw and lankester 1994b). in the early 2000s, one of the authors (ranta) examined a number of sick, dying, and dead moose from wmus 6 and 7b that displayed classical symptoms of meningeal worm infection. the disease has been documented in adjacent southeastern manitoba where lankester (1974) recorded 13 cases within a 12-month period in 1972-73. giant american liver fluke (fascioloides magna) is known to occur in deer of the kd but no data exist about its relative abundance. we know of no reports from hunters of noticeably infected moose for at least the last 3 decades. winter ticks are regularly seen on moose in early spring, and reports from outfitters, trappers, and moose hunters suggest that a substantial die-off of moose occurred in wmus 7a and 7b in 2000 and 2001 when moose density was high (fig. 7 and 11). anecdotal evidence on the aulneau peninsula included a number of moose skeletons located the following spring, summer, and fall. discussion data presented here indicate that deer and moose populations in kd have experienced significant population swings over the past 100 years, and disturbances at the landscape scale have impacted both species. logging and land clearing are likely responsible for the initial invasion and subsequent maintenance of deer in the district, despite periodic die-offs associated with severe winters. both logging and fire are also believed responsible for an increase in moose in british columbia and northern ontario (thompson and stewart 1998). in ontario, relatively high moose populations are typically found in forested areas with a mosaic of vegetation types providing a high interspersion of cover and forage (rempel et al. 1997). fig. 11. moose estimated from aerial inventory in wmus 6, 7a, and 7b in 1975 to 2012, kenora district, ontario, canada. 0 0.2 0.4 0.6 0.8 1 1.2 1.4 m o o se d en si ty ( m o o se /k m 2 year wmu 6 wmu 7a wmu 7b alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 171 formalized moose habitat guidelines used in ontario and the kd since 1988 provide a detailed summary of the benefits of good forest management practices (omnr 2010). logging, combined with forest fire suppression, leads to shifts in forest composition and structure (carleton 2000). balsam fir is one species that becomes more prominent in managed forests (thompson 2000a), making the forest more susceptible to spruce budworm infestation and blowdown. infestations cause stand mortality after 5 consecutive years of defoliation (fleming et al. 2002) followed by peeling bark, growth of draped arboreal lichen (usnea spp.), and top breakage culminating in wind-throw 5 to 8 years later. peaks of deer abundance in the mid-1900s and early 2000s appear to be strongly associated with fir mortality and associated abundance of lichen. while balsam fir is generally not considered preferred deer browse (e.g., ullrey et al. 1968, mautz et al. 1976), the arboreal lichen associated with dead and dying balsams is heavily used by deer during winter (hodgman and bowyer 1985). usnea spp. compares favourably with respect to crude protein, available energy, and relatively high digestibility of typical hardwood winter browse (hodgman and bowyer 1985, gray and servello 1995). no evidence of impacts to deer or moose were evident from forest tent caterpillar outbreaks, although both species would presumably have access to improved quality and quantity of understory forage in the immediate aftermath of an outbreak. similarly, the effects of the jack pine budworm and associated loss of coniferous canopy should seasonally benefit both deer and moose. the last peak in the kd moose population is attributed primarily to the large fire events of the 1980s; considerably less area has burned since. mai found high concentrations of animals in and immediately adjacent to the large burns of the 1980s, but more recent surveys indicate few moose in these same areas. the association of moose with early seral stages of post-fire habitat has long been recognized (peek 1997). kelsall et al. (1977) concluded that the optimal successional stage for moose in the boreal forest occurred 11 to 30 years postburn, and moose in alaska respond positively to fires as early as 5 years post-burn (schwartz and franzmann 1989). although deer have an abundance of food in the early aftermath of fire, the loss of conifer cover in winter yarding areas can seriously jeopardize winter survival (hanley and rose 1987, broadfoot and voigt 1996). fires can eliminate balsam fir from stands (thompson 2000a), and little balsam fir was left in the kd burns. because these large burned areas lacked winter conifer cover and associated lichen as winter forage (usnea spp. do not thrive on fire-killed balsam fir), these burns presumably become unsuitable for deer in deep snow. a severe winter can dramatically lower deer density on northern ranges and limit range occupancy (potvin et al. 1981). high mortality can occur when deep snows of long duration are coupled with extreme cold (severinghaus 1947, verme 1968, verme and ozoga 1971), conditions that affect fawns in particular (karns 1980). the combination of severe winter conditions and predation by wolves produces higher deer mortality than either factor acting alone (mech et al. 1971). winter severity indices are helpful to identify winters when substantial deer losses likely occur, but the typical values measured most years in the kd are not believed high enough to negatively impact moose. peek (1997) found moose tolerant of snow depths up to 80 cm, and coady (1974) identified 90 cm as a critical depth when adults have restricted movement and access to forage. moose and deer population trends – ranta and lankester alces vol. 53, 2017 172 winters with snow depths >90 cm are rare in the kd, but depths of >40 cm that restrict deer movement occur regularly (kelsall and prescott 1971). weekly sdi values indicate that snow depths >75 cm occur occasionally. at these depths, deer are in a severe energy deficit due to restricted and energy-costly movement, even when browse is abundant (potvin and huot 1983). both moose and deer populations in the kd increased during the 1980s and 1990s in response to increased forage created by a variety of large landscape scale disturbances, and low snow depths that specifically favour deer survival. moose experienced a pronounced decline beginning about 1995, reaching very low levels by 2012 as deer numbers remained high. similar moose declines occurred concurrently in eastern north america and in jurisdictions neighbouring the kd. for example, populations began to decline in the early 1990s and were reduced to low numbers by 2003 in nova scotia (beazley et al. 2006), moose declined in the late 1980s with few occurring by the early 2000s in northwestern minnesota (murray et al. 2006), and numbers peaked about 1995 but moose had virtually disappeared in northeastern north dakota by 2006 (maskey 2008). a similar increase in deer and decline in moose also occurred in southeastern manitoba during this time frame (v. crichton, manitoba fish and wildlife, retired, pers. comm.). these declines followed periods of shorter, less severe winters that sustained high density populations of deer with meningeal worm (lankester 2018); longer and wetter growing seasons were also associated with some of these declines (maskey et al. 2015). typically, declines continued for 15-20 years reaching very low levels that persist to the present. it has been argued that the meningeal worm was a principal cause of these declines (maskey 2008, lankester 2010, 2018), and our observations parallel those in other regions. moose with winter tick-associated hair loss were commonly observed during mai surveys in the kenora mu and unusually high overwinter mortality was reported following the winters of 2000 and 2001 when moose densities were still relatively high. carcasses and skeletal remains found in a fashion inconsistent with mortality from predation were likely due to disease or parasitism, but the exact cause of these winter mortalities was never identified. winter tick numbers are not influenced by the presence or absence of deer and they have their greatest impact when moose densities are high. these ticks typically cause late winter mortality for a few successive years and then subside in abundance at lower moose density or environmental conditions that reduce larval survival and/or the questing period. winter ticks alone are not thought to be capable of driving moose populations to low levels in a short time frame (lankester 2010). the giant liver fluke is not prominent in the kd and cannot be considered a major contributor to the moose decline, as this parasite has not been proven to cause large scale moose mortality. heavy fluke infections were interpreted as being significant in a declining moose population in northeastern minnesota (murray et al. 2006), yet flukes were equally common when that same moose population was increasing 20 years earlier (karns 1972, lankester 2010). flukes were not considered important in the moose decline in adjacent northeastern north dakota (maskey 2011) and do not occur in nova scotia (pybus 2001) where pronounced moose declines have occurred twice. records of wolf sightings by deer hunters became increasingly common in the kd from about 2000 to 2012, the same period in alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 173 which deer reached peak numbers and moose declined to low levels. this was also the period in which the effects of meningeal worm on moose were expected to be greatest making it difficult to separate the relative roles of parasites and wolves in the decline. classically, wolves increase in response to increased deer numbers and may depress productivity of co-habiting moose by preying disproportionately on calves. wolves are also likely to find moose handicapped by p. tenuis infection particularly easy prey. yet, in most instances, wolves are not expected to reduce their prey to extremely low levels (mech 1970, mech and karns 1977). as well, several studies have shown that where deer and moose co-exist, wolves tend to concentrate on deer whether deer numbers are increasing or declining (pimlott et al. 1969, mech et al 1971, potvin et al. 1988). a prominent role for wolves in declines elsewhere is even less likely as wolves do not occur in mainland nova scotia, and the resident eastern coyote (canis latrans) is not considered a significant predator of moose (parker 2003) or to have played a substantial role in mainland nova scotia moose being declared “endangered” after the recent decline. nor were wolves considered a main factor in moose declines in northwestern (lenarz et al. 2009) or northeastern minnesota (murray et al. 2006), or in neighbouring northeastern north dakota (maskey 2008). however, mech and fieberg (2014) argued for caution in accepting the conclusion of lenarz et al. (2009) that elevated winter temperatures caused the impending decline of moose populations in northeastern minnesota, and instead suggested a stronger role for wolves. current research has identified that p. tenuis and wolf predation are principal mortality factors in minnesota moose (m. carstensen, minnesota department of natural resources, pers. comm.). hunting can reduce deer and moose numbers and significant declines may result, especially when stricter regulations are not implemented quickly enough in response to natural stochastic population changes (fryxell et al. 2010). however, there is little evidence that inordinately high hunter harvest (fig. 10) caused the abrupt and prolonged decline of moose in the kd. deer invariably decline following severe winters, and hunter harvest played a minimal role in the 1970s decline (ranta 1982). further, deer began to increase in the 1980s and continued to increase until about 2007 despite increasing hunting pressure. several climatic factors known to enhance transmission of p. tenuis circumstantially support a major role for this parasite in the kd moose decline. shorter winters with less snow and lower sdis during the 1990s and 2000s allowed increased deer densities, and in particular, increased survival of fawns. fawns are the biggest producers of the parasite’s larval stages and an early spring increases larval survival (lankester 2018, in press). also, the length of frost-free seasons during this period increased, albeit marginally, but growing seasons were much wetter than in the previous 20 years. precipitation is an important driver of terrestrial gastropod populations and determines the extent to which they move on the forest floor to become infected and ingested by cervids (wasel et al. 2003). caribou are much more susceptible than other cervids to neurological disease caused by meningeal worm infection (anderson 1972). records of range recession of caribou in northwestern ontario indicate that caribou disappeared from most of the present day kd during the first era of high deer densities (darby et al. 1989), and are now found only on the northern fringe of the kd (ranta 2001). moose and deer population trends – ranta and lankester alces vol. 53, 2017 174 conclusion landscape level factors, working in synergy, have been primary population-level drivers behind widely fluctuating populations of deer and moose in the kd in much of the past century. however, habitat availability, winter conditions, and predation cannot adequately explain the moose decline in the kd. much evidence suggests that pronounced and prolonged declines in moose populations result when specific conditions occur concurrently: 1) when the distribution of moose and infected (p. tenuis) deer are sympatric, 2) when winter conditions are generally favourable for survival, growth, and expansion of deer populations for many consecutive years (e.g., > 10 years), and 3) when environmental conditions are favourable for the survival and mobility of terrestrial gastropods required for transmission of the meningeal worm as illustrated in fig. 7. we suggest that the meningeal worm played a major role in the recent moose decline in the kd and is likely to have done so repeatedly in several locations in eastern north america within the past century (lankester 2018). acknowledgements the authors thank the many staff of the ontario ministry of natural resources and forestry who helped in the collection and analyses of data over many years. in particular, we thank l. anderson and c. martin for their efforts in tabulating data and assisting in the production of the figures used in the manuscript. special thanks are extended to the kenora district fire and information management staff and those associated with the snow network for ontario wildlife who graciously provided the data on landscape level disturbances and winter severity. mr. g. gordon assisted with some graphics and dr. d. euler kindly read the manuscript and offered helpful suggestions. references allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. united states fish & wildlife service biological report 82 (10.155). u.s. department of the interior, fish and wildlife service research and development, washington, d. c., usa. anderson, r. c. 1965. an examination of wild moose exhibiting neurologic signs, in ontario. canadian journal of zoology 43: 635-639. _____. 1972. the ecological relationships of meningeal worm and native cervids in north america. journal of wildlife diseases 8: 304-310. beazley, k., m. ball, l. isaacman. s. mcburney, p. wilson, and t. nette. 2006. complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada. alces 42: 89-109. broadfoot, j. d., and d. r. voigt. 1996. white-tailed deer migration behaviour: a resource management perspective. sters technical report no. 5. ontario ministry of natural resources, toronto, ontario, canada. carleton, t. j. 2000. vegetation responses to the managed forest landscape of central and northern ontario. pages 178-197 in a. h. perera, d. l. euler, and i. d. thompson, editors. ecology of a managed terrestrial landscape. university of british columbia press, vancouver, british columbia, canada. caughley, g. 1977a. analysis of vertebrate populations. john wiley and sons, london, england. _____. 1977b. sampling in aerial survey. journal of wildlife management 41: 605-615. coady, j. w. 1974. influence of snow on behavior of moose. naturaliste canadien (quebec) 101: 417-436. alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 175 cumming, h. g. 1972. the moose in ontario. ontario ministry of natural resources, toronto, ontario, canada. darby, w. r., h. r. timmerman, j. b. snider, k. f. abraham, r. a. stefanski, and c. a. johnson. 1989. woodland caribou in ontario: background to a policy. ontario ministry of natural resources, toronto, ontario, canada. elkie, p., m. gluck, j. boos, j. bowman, c. daniel, j. elliott, d. etheridge, d. heaman, g. hooper, r. kushneriuk, g. lucking, s. mills, b. naylor, f. pinto, b. pond, r. rempel, k. ride, a. smiegielski, g. watt, and m. woods. 2009. science and information in support of the forest management guide for landscapes: package “a” simulations, rationale and inputs. ontario ministry of natural resources, forest policy section, toronto, ontario, canada. fleming, r. a, j. candau, and r. s. mcalpine. 2002. landscape-scale analysis of interactions between insect defoliation and forest fire in central canada. climatic change 55: 251-272. fryxell, j. m., c. packer, k. mccann, e. j. solberg, and b. e. sæther. 2010. resource management cycles and the sustainability of harvested wildlife populations. science 328: 903-906. gray, p. b., and f. a. servello. 1995. energy intake relationships for whitetailed deer on winter browse diets. journal of wildlife management 59: 147-152. hamilton, g. d., p. d. drysdale, and d. l. euler. 1980. moose winter browsing patterns on clearcuttings in northern ontario. canadian journal of zoology 58: 1412-1416. hanley, t. a., and c. l. rose. 1987. influence of overstory on snow depth and density in hemlock spruce stands: implications for deer management in southeastern alaska. resources note pnw-rn-459. usda forest service, pacific northwest research station, portland, oregon, usa. hodgman, t. p., and r. t. bowyer. 1985. winter use of arboreal lichens, ascomycetes, by white-tailed deer, odocoileus virginianus, in maine. canadian field-naturalist 99: 313-316. jones, s. l. 1981. how to conduct deer fecal counts or what you always wanted to know about doing a spring deer survey, but were afraid to ask. unpublished report. ontario ministry of natural resources, toronto, ontario, canada. karns, p. d. 1972. minnesota’s 1971 moose hunt: a preliminary report on the biological collections. proceeding of the north american moose conference and workshop. 8: 115-123. _____. 1980. winter – the grim reaper. pages 47-53 in r. l. hine and s. nehls, editors. white-tailed deer population management in the north central states. proceedings of 1979 symposium, 10 december 1979. 41st midwest fish and wildlife conference, urbana, illinois, usa. kelsall, j. p., and w. prescott. 1971. moose and deer behaviour in snow in fundy national park, new brunswick. canadian wildlife report series 15. _____, e. s. telfer, and t. d. wright. 1977. the effects of fire on the ecology of the boreal forest, with particular reference to the canadian north: a review and selected bibliography. canadian wildlife service occasional papers number 32. king, d. r. undated. instructions for conducting stratified random pellet group and dead deer surveys. unpublished report. ontario ministry of natural resources, toronto, ontario, canada. lankester, m. w. 1974. parelaphostrongylus tenuis (nematoda) and fascioloides magna (trematoda) in moose of southeastern manitoba. canadian journal of zoology 52: 235-239. _____. 2010. understanding the impact of meningeal worm, parelaphostrongylus moose and deer population trends – ranta and lankester alces vol. 53, 2017 176 tenuis, on moose populations. alces 46: 53-70. _____. 2018. weather-enhanced transmission of meningeal worm, parelaphostrongylus tenuis, in whitetailed deer and implications for moose. alces 54: in press. ______, and w. m. samuel. 2007. pests, parasites and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. lenarz, m. s. 2009. a review of the ecology of parelaphostrongylus tenuis in relation to deer and moose in north america. pages 70-75 in m. w. doncarlos, r. o. kimmel, j. s. lawrence, and m. s. lenarz, editors. summaries of wildlife research findings. minnesota department of natural resources, st. paul, minnesota, usa. _____, m. e. nelson, m. w. schrage, and a .j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. _____, j. fieberg, m. w. schrage, and a j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. the journal of wildlife management 74: 1013-1023. maskey, j. j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph. d. thesis, university of north dakota, grand forks, north dakota, usa. _____. 2011. giant liver fluke in north dakota moose. alces 47: 1-7. ______, r. a. sweitzer, and b. j. goodwin. 2015. climate and habitat influence prevalence of meningeal worm infection in north dakota, usa. journal of wildlife diseases 51: 670-679. mautz, w. w., h. silver, j. b. holter, h. h. hayes, and w.e. urban jr. 1976. digestibility and related nutritional data for seven northern deer browse species. journal of wildlife management 40: 630-638. mcintosh, t. e. 2003. movements, survival and habitat use by elk (cervus elaphus) reintroduced to northwestern ontario. m.sc. thesis, lakehead university, thunder bay, ontario, canada. mclaren, m. 2006. standards and guidelines for moose population inventories in ontario. technical report number ssi-121. southern science and information, ontario ministry of natural resources, north bay, ontario, canada. mcshea, w. j., h. b. underwood, and j. h. rappole. 1997. deer management and the concept of overabundance. pages 1-7 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance: deer ecology and population management. smithsonian institute press, washington d. c., usa. mech, l. d. 1970. wolf: the ecology and behaviour of an endangered species. university of minnesota press, st. paul, minnesota, usa. _____, l. d. frenzel, and p. d. karns. 1971. the effect of snow conditions on the vulnerability of white-tailed deer to wolf predation. pages 51-59 in l. d. mech and l. d. frenzel jr., editors. ecological studies of the timber wolf in northeastern minnesota. resource paper #nc-52. united states department of agriculture, forest service, north central forest experimental station, st. paul, minnesota, usa. _____, and j. frieberg. 2014. re-evaluating the northeastern minnesota moose decline and the role of wolves. the journal of wildlife management 78: 1143-1150. ______, and p. d. karns. 1977. role of the wolf in a deer decline in the superior national forest. resource paper nc-148. united states department of alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 177 agriculture, forest service, north central forest experimental station, st. paul, minnesota, usa. murray, d. j., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 116: 1-30. omnr (ontario ministry of natural resources). 1974. lake of the woods planning area: information package. ontario ministry of natural resources, queen's printer for ontario, toronto, ontario, canada. _____. 1997. the snow network for ontario wildlife. the why, when, what and how of winter severity assessment in ontario. ontario ministry of natural resources, queen's printer for ontario, toronto, ontario, canada. _____. 2003. aulneau peninsula enhanced management area wildlife plan. ontario ministry of natural resources. ontario ministry of natural resources, queen's printer for ontario, toronto, ontario, canada. _____. 2010. forest management guide for conserving biodiversity at the stand and site scales background and rationale for direction. ontario ministry of natural resources, queen's printer for ontario, toronto, ontario, canada. oswald, k. 1997. moose aerial observation manual. north east science and information technical manual tm-008. ontario ministry of natural resources, queen’s printer for ontario, toronto, ontario, canada. parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. nova scotia department of natural resources, kentville, nova scotia, canada. passmore, r. c. 1953. snow conditions in relation to big game in ontario during the winter of 1952-53. report 2. ontario department lands and forests, wildlife resources, toronto, ontario, canada. peek, j. m. 1997. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. the ecology and management of the north american moose. wildlife management institute, washington, d. c., usa. pimlott, d. h., j. a. shannon, and g. b. kolenosky. 1969. the ecology of the timber wolf in algonquin provincial park. research branch research report (wildlife) no. 87. ontario department of lands and forests, toronto, ontario, canada. potvin, f., and j. huot. 1983. estimating carrying capacity of a white-tailed deer wintering area in quebec. journal of wildlife management 47: 463-475. _____, _____, and f. duchesneau. 1981. deer mortality in the pohénégamook wintering area, quebec. canadian fieldnaturalist 95: 80-84. _____, h. jolicoeur, and j. huot. 1988. wolf diet and prey selectivity during two periods for deer in quebec: decline versus expansion. canadian journal of zoology 66: 1274-1279. pybus, m. 2001. liver flukes. pages 121-149 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. second edition. iowa state university press, ames, iowa, usa. racey, g. d., and t. armstrong. 2000. woodland caribou range occupancy in northwestern ontario: past and present. rangifer, special issue no. 12:173-183. ranta, w. b. 1982. the status of whitetailed deer in the kenora district of ontario. report series. kenora fish and wildlife district, ontario ministry of natural resources, kenora, ontario, canada. _____. 2001. report on woodland caribou and their use of habitats in the kenora management unit and southern portions moose and deer population trends – ranta and lankester alces vol. 53, 2017 178 of woodland caribou provincial par. unpublished report. kenora fish and wildlife district, ontario ministry of natural resources, kenora, ontario, canada. ______, and s. e. shaw. 1982. white-tailed deer pellet group and habitat inventory survey; wildlife management units 7a and 7b spring/82. report series. kenora fish and wildlife district, ontario ministry of natural resources, kenora, ontario, canada. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61: 517-524. roseberry, j. l., and a. woolf. 1998. habitat-population density relationships for white-tailed deer in illinois. wildlife society bulletin 26: 252-258. rowe, j. b. 1972. forest regions of canada. publication. no. 1300. department of fisheries and the environment, canadian forest service, ottawa, ontario, canada. saunders, b. p. 1973. meningeal worm in white-tailed deer in northwestern ontario and moose population densities. journal of wildlife management 37: 327-330. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose, and forest succession, some management considerations on the kenai peninsula, alaska. alces 25: 1-10. seton, e. t. 1909. life histories of northern animals. volume 1, grass-eaters. charles scribner’s sons, new york, new york, usa. severinghaus, c. w. 1947. relationship of weather to winter mortality and population levels among deer in the adirondack region of new york. north american wildlife conference transactions 12: 212-223. thomas, j. w., h. black, r. j. scherzinger, and r. j. pedersen. 1979. deer and elk. pages 104-127 in j. w. thomas, editor. wildlife habitats in managed forests. united states department of agriculture. agricultural handbook. no. 553. u.s. department of agriculture, forest service, portland, oregon. thompson, i. d. 2000a. forest vegetation of ontario. pages 30-53 in a. h. perera, d. l. euler, and i. d. thompson, editors. ecology of a managed terrestrial landscape. university of british columbia press, vancouver, british columbia, canada. _____. 2000b. forest vertebrates in ontario: patterns of distribution. pages 54-73 in a. h. perera, d. l. euler, and i. d. thompson, editors. ecology of a managed terrestrial landscape. university of british columbia press, vancouver, british columbia, canada. _____, and r. w. stewart. 1998. management of moose habitat. pages 377-401 in a. h. perera, d. l. euler, and i. d. thompson, editors. ecology of a managed terrestrial landscape. university of british columbia press, vancouver, british columbia, canada. _____, and m. f. vukelich. 1981. use of logged habitats in winter by moose cows with calves in northeastern ontario. canadian journal of zoology 59: 2103-2114. ullrey, d. e., w. g. youat, h. e. johnson, l. d. fay, b. e. brent, and k. e. kemp. 1968. digestibility of cedar and balsam fir browse for the white-tailed deer. journal of wildlife management 32: 162-171. verme, l. j. 1968. an index of winter weather severity for northern deer. journal of wildlife management 32: 566-574. _____, and j. l. ozoga. 1971. influence of winter weather on white-tailed deer in upper michigan. michigan department of natural resources and development report no. 237. michigan department of natural resources, lansing, michigan, usa. alces vol. 53, 2017 ranta and lankester. – moose and deer population trends 179 voigt, d. r., j. a. baker, r. s. rempel, and i. d. thompson. 2000. pages 198-233 in a. h. perera, d. l. euler, and i. d. thompson, editors. ecology of a managed terrestrial landscape. university of british columbia press, vancouver, british columbia, canada. warren, r., a. r. bisset, b. pond, and d. voigt. 1998. the snow network for ontario wildlife. ontario ministry of natural resources, peterborough, ontario, canada. wasel, s. m., w. m. samuel, and v. crichton. 2003. distribution and ecology of meningeal worm, parelaphostrongylus tenuis (nematoda), in northcentral north america. journal of wildlife diseases 39: 338-346. whitlaw, h. a., and m. w. lankester. 1994a. a retrospective evaluation of the effects of parelaphostrongylosis on moose populations. canadian journal of zoology 72: 1-7. _____, and _____. 1994b. the cooccurrence of moose, white-tailed deer and parelaphostrongylus tenuis in ontario. canadian journal of zoology 72: 819-825. zoltai, s. c. 1961. glacial history of part of northwestern ontario. geological association of canada 33: 61-83. alces37(1)_79.pdf 4011.p65 alces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 45 effects of overabundant moose on the newfoundland landscape brian e. mclaren1,2, bruce a. roberts3, nathalie djan-chékar4, and keith p. lewis5 1 government of newfoundland and labrador, department of natural resources, p.o box 2222, gander, nl, canada a1v 2n9; 3 natural resources canada, canadian forest service, p.o. box 960, corner brook, nl, canada a2h 6j3; 4 provincial museum of newfoundland and labrador, natural history unit, p.o. box 8700, st. john’s, nl, canada a1b 4j6; 5 programme in cognitive and behavioural ecology, departments of biology and psychology, memorial university, st. john’s, nl, canada a1c 5s1 abstract: the long-term effects of introduced and overabundant herbivores on community development must be monitored and managed in an ecosystem-based forest management approach. this paper builds on previously published ecological descriptions and hypotheses offered on the effects of moose overabundance in newfoundland. the island, in the absence of wolves, provides a setting for study of local irruptions in moose populations, which now affect an increasing area of the forest. moose effects occur most often after natural disturbances and logging, involving unique forest succession patterns. we describe some of these changes, along with anticipated and realised changes in associated forest biodiversity. we offer suggestions to improve or refine monitoring of moose populations, especially at local scales, to detect cases of overabundance. finally, we offer recommendations for the management of overabundant moose populations. alces vol. 40: 45-59 (2004) key words: alces alces, biodiversity, carrying capacity, conservation biology, disturbance, moose, newfoundland, overabundance, population dynamics the long-term effects of introduced and overabundant herbivores on forest community development must be monitored and managed in an ecosystem-based approach to forestry and wildlife interests. adapting forest management to shifting baselines created by the effects of overabundant herbivores, especially in eastern north america, increasingly defeats the interest of biodiversity protection (lindenmayer and franklin 2003). logging in the boreal forest and other forest management tailored to the spatial scale and frequency of wildfires or insect-related tree mortality are recommended to protect and/or restore ecological integrity (hunter 1993, niemelä 1999), as a minimum way of recognizing the adaptation of diverse organisms to the forests they occupy, by considering their disturbance regimes. disturbance is usually defined as a rapid release or reallocation of resources in a forest community (white and jentsch 2001), thereby possibly ignoring more gradual changes to forest community development caused by irruptive population phases or overpopulations of herbivorous mammals, especially deer (cervidae). such changes can be considerable, unpredictable, and relatively irreversible (davidson 1993, côté et al. 2004). in this paper, we offer a case study of introduced moose (alces alces andersoni), 2 present address: faculty of forestry and the forest environment, lakehead university, 955 oliver road, thunder bay, on, canada p7b 5e1 overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 46 its overabundance, and its enduring effect on biodiversity in specific areas of boreal forest on the island of newfoundland. we show that moose are significantly influencing several aspects of some ecosystems, including forest succession and composition, soils, and other wildlife. previously it has been shown that moose are capable of producing negative economic effects on the forests of newfoundland (pimlott 1963, thompson 1988), but such studies are often limited just to a portion of a forest rotation (thompson and curran 1993, mclaren et al. 2000a). this paper builds on published ecological descriptions and hypotheses offered for longer-term effects of moose overabundance in newfoundland (e.g., bergerud and manuel 1968, thompson and mallik 1989, thompson et al. 1992). we make recommendations toward monitoring, conservation planning, and management of overabundant moose from the perspectives of past examples and anticipated future challenges in newfoundland. moose in newfoundland: a background the island of newfoundland, canada, is a landmass of 112,000 km2 in the northwest atlantic, of which about two-thirds is forested and/or qualifies as excellent moose habitat. moose were introduced to central newfoundland in 1878 with the release of a male and female from nearby nova scotia (pimlott 1953). a second release of two males and two females from new brunswick, into western newfoundland, followed in 1904. moose rapidly colonized newfoundland (fig. 1), as they exploited new habitat and as wolves (canis lupus) were e x t i r p a t e d ( p i m l o t t 1 9 5 9 ) . c a r i b o u (rangifer tarandus terra-novae) are the only other ungulate in newfoundland, primarily occupying non-forested habitat and existing prior to the arrival of europeans. where the two cervid species now co-exist, there has been no recorded direct competition between them. today, moose occupy all ecoregions on the island, at densities in primarily forested habitat in many instances exceeding 4 moose / km2 (> 1,000 kg / km2). population reconstruction and aerial survey estimates for a moose management unit normally do not show this finer-scale spatial variation in density, but overabundance is suggested by our temporal series (fig. 1), considering that approximately 75,000 km2 of habitat supported on average about 2 moose / km2 during two periods, in the late 1950s and late 1980s (mercer and mclaren 2002). this average is met by considerable variation and any densities > 2 moose / km2 are considered above management targets (newfoundland and labrador inland fish and wildlife division, unpublished). the national parks in newfoundland, where hunting is prohibited, form special cases of overabundance. the second and likely only successful moose introduction to newfoundland was ca. 20 km from what is now the boundary of gros morne national park, in the western part of the island. local hunting kept the population relatively low until this activity ceased with the establishment of the park in 1974. within the park area of 1,805 km2, of which only ca. 30% is suitable habitat, moose have increased steadily from some 1,000 animals to > 7,000, with local densities as high as 7 moose / km2 by 1995 (mclaren et al. 2000b). management of moose hunting in newfoundland, which began with the first season in 1945 (mercer 1995), has achieved a legal kill of about half a million, the majority of which is by resident hunting (mclaren 2004). annual license issue since 1990 has been between 20,000 and 25,000 resident and non-resident tags combined, with a roughly equal number of either-sex and male-only licences issued (mercer and mclaren 2002). annual kill estimates, incorporating poaching, crippling, and highalces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 47 way losses, have been 20,000 30,000 moose for several years. first nations are currently included in the resident licensing system in the island portion of newfoundland and labrador. dating at least to 1934, the provincial government has recognized that sport hunting is a major attraction for visitors and the success of the moose introduction to the island has often been applauded as a source of tourist revenue. non-residents obtain about 10% of moose tags issued, and owing to the higher success offered by outfitters and guides, are responsible for > 10% of the annual moose kill. newfoundland is also home to a disproportionately large part of the north american moose population. the island population, at 125,000 moose, represents > 10% of the total continental number of moose (1.05 million), while the total island area, including areas unsuited to moose, is < 2% of the estimated continental moose range (6.44 million km2). throughout north america, density of moose and other deer species varies according to four main factors: the availability of habitat, the availability of alternate foods created by agriculture, management of hunting, and the presence of limiting factors like aridity or natural predators (crête and daigle 1999). overpopulation usually occurs following introduction into unexploited habitat and persists in situations without natural predators (mcshea et al. 1997), as is the case in newfoundland. in these situations, moose appear to be limited primarily by the productivity of the boreal forest, as described for québec by crête and courtois (1997). as productivity varies, so does the effect of moose, as expected in an unregulated trophic system. thus, we have two arguments for moose overpopulation in newfoundland: (1) as presented by crête and daigle (1999), newfoundland hosts an anomalous deer biomass compared with the rest of the continent, presumably because of the absence of wolves; and (2) as described by mercer and mclaren (2002), a stable equilibrium between the population and food resources does not appear to have occurred for newfig. 1. moose population trends in insular newfoundland since introduction, using mclaren’s (2004) assessment of the successful point of introduction, pimlott’s (1959) estimate of increase rate during 1904-44 (solid squares), hunter success during 1945-65 (open circles), and hunter reports of moose seen during 1966-99 (solid circles). these estimates, scaled to consolidated aerial survey estimates in the 1980s and in the 1990s (newfoundland and labrador inland fish and wildlife division, unpublished), occur in a wide range of habitats and densities. overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 48 foundland, particularly problematic where moose densities are higher, in more productive areas, and/or in areas less accessible to hunting. newfoundland thus allows us to illustrate specific cases where ecosystem management experiences new challenges as a result of moose overabundance. effects of moose on the forest ecosyst e m forest succession and composition — most forests in newfoundland consist of a c o m b i n a t i o n o f b a l s a m f i r ( a b i e s balsamea) and spruce (picea spp.), with some pioneer and shade tolerant hardwoods. fir is dominant in older forests, while insect outbreaks, fires, and logging have been frequent forest disturbances creating a generally young-forest landscape. since the arrival of moose, their consumption of balsam fir and hardwoods has affected forest regeneration following disturbance, particularly along edges and roadsides (bergerud and manuel 1968). in these affected forests, spruce and larch (larix laricina) grow normally, since they are species not normally found in moose diets, but balsam fir, a heavily-consumed species, can be prevented from reaching heights > 1 m. the resulting open ecotype has been described but not attributed specifically to moose in any general literature on forest management in newfoundland; the forest succession leading to the ecotype has been termed “old-field spruce succession” (damman 1964). in updating the forest site classification (ecological land classification) for newfoundland, roberts and bajzak (1984) used the term “ungulate induced” to describe the succession specific to richer sites, in which a shift occurs from closed canopy balsam fir and white spruce (picea glauca) to open-grown white spruce following disturbances (fig. 2). roberts (1989a, 1989b) attributed this change to occupation of young forests by overabundant moose during the 1960s. roberts’ (1989b) concern was for white spruce associations and other rare forest types. similar examples of vulnerable forest types are yellow birch (betula alleghaniensis) associations on the avalon peninsula and red maple (acer rubrum) associations in central newfoundland. these types, occurring with balsam fir, are frequently subject to logging and wind disturbance and occur in management areas where moose are kept near target densities to offer steady hunting opportunities. yellow birch and red maple are preferred species in a moose diet dominated by fir. in time, additional changes to natural forest succession caused by overabundant moose may become apparent. since the 1980s, the total area disturbed by logging, as well as the secondary road network, has increased as forestry has become more extensive in newfoundland. more common ecological associations may be threatened by the coexistence of regenerating commercial forest and overabundant moose, especially with declining interest in hunting. less accessible areas may be especially prone to a combination of natural disturbance like insect outbreak and moose overabundance. areas of both natural and logging disturbance that fail to regenerate into closed canopy forest are already at a scale readily visible on aerial photographs (fig. 2). our photographs show white spruce associations. thompson et al. (1992) and thompson and mallik (1989) extended their concerns to black spruce (picea mariana) associations. observations of forest composition change attributed to moose overabundance have been recorded in the national parks. terra nova national park, 344 km2, has been unaffected by logging since the 1950s, but is a special case of forest disturbance, in which small (< 1 km2) landscape patches were disturbed by insect outbreaks in the late 1970s, just before a peak in moose alces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 49 fig. 2. ungulate induced changes to the newfoundland landscape. the aerial photographs show cases where forest succession is interrupted by overabundant moose for (a) an area near blue hill in terra nova national park that was disturbed by an insect outbreak in 1978 and has failed to regenerate in 25 years (photo date 1996), (b) a partially forested watershed in 1988, logged in the 1950s, where balsam fir has failed to regenerate in a white spruce association, following disturbances that also included insect outbreaks, and (c) an area near halfway cliff in gros morne national park, where moose have created semi-open cover because they prevent a fir-dominated alpine forest understory from regenerating. the partial canopy in the three photographs (at arrows) is created by open-grown spruce. the scale bars in each photograph measure ca. 1 km. density during the 1980s. moose density is low today, ca. 0.7 / km2, but very low recruitment observed in all of the last 5 midwinter surveys suggests a declining population, exceeding carrying capacity (gosse et al. 2002). carrying capacity in the park has itself declined over time as forests have matured and moose have nearly completely removed hardwoods, such as red maple and mountain ash (sorbus americana) from mature-forest understories. in the foragel i m i t e d , d i s t u r b e d p a t c h e s , p i o n e e r hardwoods, such as white birch (betula spp.), aspen (populus tremuloides), and pin cherry (prunus pensylvanica), have also been affected. to illustrate the effect moose continue to have on limiting growth and survival of understory trees, park managers constructed several fenced areas or “exclosures” (terra nova national park, unpublished). after only 3 growing seasons, red maple density inside exclosures was up to 3 times higher than outside. for white birch, similar densities occur inside and outside, because stump and root sprouting is common. however, growth of white birch has been affected. for example, up to 90% of white birch stems are > 0.5 m inside the exclosures (this includes 14% > 1 m after 3 growing seasons), but outside, only 5% are > 0.5 m. outside the exclosures, reductions in stem density of 2–17% for balsam fir also occurred between 1999 and 2003. measured annually, this effect was highest when snow cover was low and moose were able to uproot young stems. in c b a overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 50 gros morne national park, canada yew (taxus canadensis), once common in coastal forest associations (robertson and roberts 1982), is nearly completely removed from the forest understory, presumably by moose. in addition, areas with once fairly closed forest now have frequent understory openings and associated changes in forest structure (connor et al. 2000). experimental introduction of moose to offshore islands in eastern newfoundland has provided other examples of their capacity to change forest composition very soon after arrival. on brunette island, off the south coast, several moose were introduced for the first time in 1974 and plant measurements using exclosures to assist comparisons were undertaken during the 10 following years (butler 1986). in this study, the ratio in annual production of balsam fir to faster growing hardwoods declined during the first 5 years from 39:1 to 15:1, measured in stems / ha (wildlife division, unpublished data). less common species, like mountain ash and wild raisin (viburnum cassinoides), declined faster than balsam fir as a ratio to the fastest growing hardwoods in the same period. these measurements were corrected for the “natural” succession changes observed inside the exclosures. the removal of forbs, grasses, and alder (alnus spp.), not normally observed in retrospective studies, was significant in observations of a tame moose, which consumed these plants in > 80% of observed bites in a summer period (butler 1986). recognizing the earliest changes created by an introduced species, often difficult to quantify, is important to biodiversity management. soils — a review of the effects of herbivores on multi-trophic interactions including soil effects has recently been provided by bardgett and wardle (2003). pastor and naiman (1992) discussed the subject for moose. we provide one unique newfoundland example. in many mountainous, serpentine plateaus in gros morne national park, toxic levels of magnesium and heavy metals occur in soils (roberts 1980, 1992). before 1980, moose were rarely encountered in these areas of the park, but their use of the serpentine plateaus is now prevalent (b. a. roberts, unpublished observations). in plants adapted to the specialized plateau ecosystems, such as dwarf birch (betula glandulosa) and alder (alnus crispa), nickel concentrations range from 48–77 ppm (roberts 1992, roberts and proctor 1992). while not normally including these species in their diet, moose are now consuming these toxic plants, presumably because of the lack of preferred forage species, a consequence of overabundance. this interaction may ultimately be detrimental both to the animals and to the long-term persistence of some plant species. our concern for landscape effects is that even the sparse cover of plants acts as a soil stabilizer. as moose consume and trample slowly growing plants, the rate of soil erosion increases, as once described by leopold (1949), in “thinking like a mountain.” other forest wildlife — a review of deer overabundance in many parts of the world and its cascading (indirect) effects on other plant and animal species is provided by côté et al. (2004). changes to forests caused by overabundant moose in newfoundland are also very likely to affect many more forest-dwellers than the plants directly affected by browsing. for example, forest songbirds dependent on hardwood and balsam fir trees (setterington et al. 2000) and epiphytic tree lichens with specific habitat requirements (yetman 1999) may actually be eliminated from portions of the landscape with overabundant moose. alternatively, ungulate-induced modification of forest structure and composition after natural disturbance or logging may result in changes to habitat selection among alces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 51 forest-dependent species. our examples below further illustrate these hypotheses. lichen diversity is generally related to the availability of different microhabitats (gustafsson et al. 1992, kuusinen 1995, rosentreter 1995, neitlich and mccune 1997). a mixture of deciduous and coniferous tree species provides an array of bark acidity and texture, offers diverse trunk structures, and creates a mosaic of moisture and light conditions for lichen growth. consequently, altered forest composition and structure caused by moose, such as the elimination of hardwoods, may have an indirect effect on epiphytic lichen community composition. this effect was illustrated through principal component analysis (pca) of lichen cover in various forest types of terra nova national park (yetman 1999). red maple in particular, in stands mixed with balsam fir, supports a unique lichen community according to the pca, in which the principle components correspond to site richness and bark acidity. as discussed above, maple is one of the trees being limited in both density and height growth as a result of moose in the park. an end result may be the loss of the lichen community this tree supports. in the avalon peninsula, yellow birch is known to be a specific host for the rare cyanolichen, degelia plumbia. as yellow birch is selectively browsed in this area, its subsequent failure to regenerate may limit opportunities for this hostspecific epiphyte. b o r e a l f e l t l i c h e n , e r i o d e r m a pedicellatum, a globally rare species, is now restricted to coastal nova scotia and the island of newfoundland. it was recently listed as vulnerable under the newfoundland and labrador endangered species act. a species of oceanic affinity, it is found in moist, cool forests where it grows predominantly on balsam fir and forms part of a characteristic cyanolichen community (ahti 1983, maass 1983). the known newfoundland population of boreal felt lichen is concentrated in the avalon peninsula and bay d’espoir areas. the avalon peninsula has had a long history of land development, logging, insect outbreaks, wind disturbance, and fires; it is also an area frequently used by moose in winter. the percentage of juvenile lichens is much lower in this area than in the second area, where forest cover is more complete and moose are less abundant (n. djan-chékar, unpublished data). we speculate that in the bay d’espoir area more opportunity for lichen colonization exists as a result of fewer moose and better balsam fir regeneration. to assess this hypothesis, biologists now monitor both the boreal felt lichen by recording its occurrence and abundance on the landscape, and also the use by moose of its critical habitat by annually counting pellet groups, an ecosystem approach to biodiversity monitoring. changes to vertebrate communities as a result of forest succession are probably best understood for birds (helle and niemi 1996). plant structure and diversity influence avian assemblages (e.g., macarthur and macarthur 1961, macarthur et al. 1962, james and rathbun 1981, cody 1985, willson and comet 1996), foraging behaviour (parrish 1995), and nest site selection (martin 1992). in addition, predation of nests can increase in areas with low foliage density (martin 1993). however, the indirect effects that herbivores exert on avian assemblages through modifying vegetation are not well known (rotenberry et al. 1995, mcshea and rappole 1997) and only a few researchers have explicitly studied the effects of herbivores on birds (degraaf et al. 1991, popotnik and giuliano 2000). indirect effects of moose overabundance on songbirds in the forest–heathland ecotone in central newfoundland were considered in a model of avian richness and abundance in black spruce-feathermoss forest and kalmia angustifolia heath. fire overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 52 suppression, logging, and plant consumption by moose all contribute to a conversion from black spruce-feathermoss forest to transitional black spruce-kalmia forests and, in cases of very poor regeneration, kalmia heath (cf. thompson and mallik 1989, thompson et al. 1992). a forest inventory describing the extent of this conversion has not been completed but the problem has demanded attention by silviculturists in the province (english and hackett 1994). we observed that songbird abundance and species richness is significantly lower in kalmia heath compared to black spruce-kalmia forest (lewis 2004). most birds in kalmia heath tended to be habitat generalists and were also common in the forests. species abundance increased with increasing vertical structure in the kalmia heath, indicating the importance of fire skips and snag retention, current elements of forest management. for example, common yellow throat (geothlypis trichas) and lincoln’s sparrow (melospiza lincolni) were associated with the fire skips. we were able to conclude that the continued suppression of many plant species by overabundant moose, as well as the invasiveness of kalmia with logging and disruption of a natural fire regime, has variable but significant effects. initial conversion of forest to black sprucekalmia transition types will result in increased songbird abundance and species richness. however, if severe restriction of regeneration in black spruce forest continues and heath increases in area, moose will have contributed to an increasingly impoverished avifauna. monitoring moose overabundance given that herbivores can influence forested ecosystems in a variety of ways, it is important to have programs to monitor their abundance and effects. increasingly, moose management plans must ask for local knowledge and advice on appropriate mitigative measures in cases of overabundance. we offer some suggestions to improve or refine monitoring of moose populations, especially at local scales, in newfoundland and in other jurisdictions where overabundance may be a concern. improved use of aerial surveys — surveys are not generally useful in assessing moose overabundance, because they are prioritized to areas of low, not high abundance, providing information to justify changes to licence quotas, the largest concern of hunters. in addition, survey areas are often much larger than areas of local overabundance. however, data from past surveys are often readily available and their innovative use, such as for identification of very high density areas or for location of census blocks of high individual counts, can be made to assess local overabundance and to verify interpretations from land capability indices. land inventories — management for forest sustainability must recognize land capability as the critical long-term factor determining productivity of any biological species. land inventories can assist in ecosystem management at various scales. most readily available maps are based on timber inventory, i.e., predicting economic value of the trees, and have limited applicability to predicting wildlife habitat (e.g., proulx and joyal 1981, potvin et al. 1999, mclaren and mahoney 2001). however, the soil-based canada land inventory (cli) is a more comprehensive classification that includes specific references to the habitat requirements of deer and other wildlife. it offers ecological comparisons of areas with varying forest capability and consequently shows relative capability to support ungulate populations. the inventory can become a reference for predicting “ungulate-induced” changes to forest succession on the landscape. for example, dryopteris–lycopodium-balsam fir and hylocomium–balalces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 53 sam fir, classed uniquely in the cli, are the dominant types where stand conversion to white spruce has taken place as a result of disturbance and overabundant moose (roberts 1989a, b). pellet-group counts — neff (1968) first described pellet-group (deer defecation) counts as a means of assessing local deer density. jordan et al. (1993) advise that pellet-group counts for moose are an effective means to make relative density comparisons over short periods and that long-term averages also compare well to information obtained from aerial surveys. the fact that moose can migrate seasonally, sometimes over long distances (mclaren et al. 2000b), may make pellet groups a more relevant index of local overabundance than an aerial survey conducted in one season. important factors in designing a monitoring program using pellet-group counts are to replicate counts in each area of interest and to conduct counts at the same time in successive years (jordan et al. 1993). pellet-group counts are subjective, because group definition is variable between observers and the season in which defecation occurs in a pellet form is variable between years and individuals. pellet groups also preserve longer in dry relative to humid soil conditions that vary with topography. simple attempts at calibration to correct for such errors are likely to be unsuccessful. browse surveys — often, plant-based protocols for monitoring herbivore effects are too labour-intensive or they include inaccurate assessments like browsing “severity indices” determined by visual inspection. indices designed for one plant architecture may not apply well to another. telfer (1967, 1972) advises on more accurate measurement of forage yield and browsing effects based on twig counts. literature on optimal foraging theory (e.g., gross et al. 1993) offers additional advice. dendrochronology — roberts (1989a) used dendrochronology to describe “oldfield spruce succession” by means of sampling balsam fir trees from several areas within one forest type to measure stem growth and ring width. from this work, a specific period of moose overabundance was determined. dendrochronology has since been applied to several studies of the long-term effects of mammals on ecosystems (e.g., sinclair et al. 1993, mclaren and peterson 1994). the technique can provide an accurate description of the cumulative effects of moose consumption on tree or shrub biomass (mclaren and peterson 1996). experimental exclosures — several ecologists have constructed exclosures (fences to keep mammalian herbivores out of an experimental area, e.g., mcinnes et al. 1992). exclosures may not produce immediate changes in forest succession but they may reveal other ecological effects. changes in the trophic pyramid may be most apparent at levels directly above and below the plants, i.e., in soil or in the abundance of herbivores that are not excluded by the fence. for example, mcinnes et al. (1992) gained insight into effects of moose on the boreal forest by measuring not only browsing of trees and shrubs, but also changes to leaf litter. there are several cautions in designing an exclosure system: 1. the fence itself introduces ecological effects. attracted by the forage regeneration inside the exclosure, herbivores may circle the fence and cause more extreme damage to plants adjacent to it. consequently, the “control” (unfenced) monitoring area should be well outside the fence. as vegetation in the area outside the fence continues to be browsed, it may retain an open or semi-open canopy depending on conditions at the time of fence construction. plants just inside the fence thus benefit overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 54 from a higher light supply than in the absence of herbivores and their measurement would constitute a bias. the “experimental” (fenced) monitoring area should therefore be located far enough inside the fence to eliminate this bias. finally, the fence material should be chosen so that it produces minimum soil change; e.g., galvanized material adds toxic zinc to soils. cribbing works well on fence posts in shallow soil. 2. sufficient replication to control for differences in site history and productivity is often difficult, both because of the difficulty in placing exclosure sites randomly and because of the expense of fence construction and maintenance. an obvious compromise is to construct larger exclosures; we suggest that 35 m on a side is minimum construction because it allows at least a 500-m2 unbiased sampling area. larger exclosures may be required to incorporate topographic variation. numerous plots inside exclosures may be unnecessary and they do not provide real landscape-scale replication. 3. monitoring protocols are often too ambitious or flawed. good advice is to begin fence construction only after very specific research hypotheses have been outlined. 4. exclosures will not mimic the forest succession that would occur as if the herbivore had never been present in the ecosystem. while this statement is a truism, it is often ignored in discussions of the results of exclosure monitoring. for example, trees regenerating inside an exclosure often grow vigorously from rhizomes, stump sprouts, or layering. such vegetative reproduction is obviously enhanced when herbivores dominate an ecosystem for many years allowing plants to allocate biomass below ground or to lateral branch growth. meanwhile, seed fall into an exclosure declines when plants outside the exclosure are not replaced because herbivores continue to suppress apical or floral growth, and seeds with longer persistence will have a competitive advantage inside the exclosure area. local ecological knowledge — forest managers rely on conservation officers, field technicians, and the general public, especially hunters and naturalists, to collect new field observations. access to wilderness areas and our ability to change them through industry, sport, and deliberate or accidental species introductions increase in tandem. thus, we have two related reasons to improve the reporting and analysis of local ecological knowledge. in many instances, descriptions of changing forest structures or landscapes through natural succession can only be made following direct field observations repeated over many years, usually by naturalists and field technicians (e.g., robertson and roberts 1982, crête et al. 2001). good examples of the systematic collection of this information already exist (e.g., ecological monitoring and assessment network, www.emanrese.ca, bird map canada, www.bsceoc.org/birdmap_e.htm, etc.). these can serve as baselines for new or improved monitoring systems. similarly, diligent reporting by hunters has allowed efforts to summarize overabundance at a continental scale (e.g., crête and daigle 1999), or at local scales (e.g., mercer and mclaren 2002). local ecological knowledge is increasingly incorporated into research and management plans (e.g., ferguson and messier 1997). moose in newfoundland: some management recommendations côté et al. (2004) challenged ecologists and wildlife managers to reduce deer numbers before and not after long-term impacts become difficult to reverse. angelstam et al. (2000) specifically reviewed management issues involving high moose densities. mercer (1995: 92) correctly took the posialces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 55 tion for newfoundland that more emphasis be placed on relationships between hunter moose-kill and moose density to stabilize populations. in areas of similar habitat, hunter kill density should be similar and proportional to moose density. in one area of newfoundland, local moose overabundance was effectively managed by temporarily directing resident hunting into the area (mclaren et al. 2000a). in other areas, where resident hunter demand is not high, or in less accessible areas, non-resident hunters might be encouraged to use permanent or temporary hunting camps. other innovations that may be part of future management of overabundant moose in newfoundland include implementation of a commercial hunt for the restaurant and luxury export trade and targeted areas of reduction, currently being considered for the national parks. conclusion considerable interest was generated by the ecological literature discussing when and where a species becomes a “keystone” for its effect on structuring ecosystems (paine 1995, power et al. 1996). we suggest that this discussion also be applied to introduced herbivores to determine where the effects of their introduction on the structure and function of native communities accumulate to the extent that they can be described as “wrecking balls”. understanding and predicting the continuous effects of herbivores on forest ecosystems following disturbance and managing these effects must be part of sustainable forest management. acknowledgements the material in this paper was first presented at the fourth international workshop on disturbance dynamics in boreal forests, university of northern british columbia, prince george, b.c., august 9–14, 2002. the authors wish to thank the canadian forest service for support in preparing this paper, joe brazil of the endangered species program and john maunder of the provincial museum, government of newfoundland and labrador, for their early discussions and review, and gordon eason, richard ward, and an anonymous reviewer for later critical reviews. references ahti, t. 1983. lichens. pages 319–360 in g. r. south, editor. biodiversity and ecology of the island of newfoundland. junk publishers, the hague. angelstam, p., p.-e. wikberg, p. danilov, w. e. faber, and k. nygrén. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland, and russian karelia. alces 36:133–145. bardgett, r. d., and d. a. wardle. 2003. herbivore-mediated linkages between aboveground and belowground communities. ecology 84:2258–2268. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. journal of wildlife management 32:729– 746. butler, c. e. 1986. summer food utilization and observations of a tame moose, alces alces. canadian field-naturalist 100:85–86. cody, m. l. 1985. habitat selection in birds. academic press, san diego, california, usa. connor, k. j., w. b. ballard, t. dilworth, s. mahoney, and d. anions. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 36:111–132. côté, s. d., t. p. rooney, j.-p. tremblay, c. dussault, and d. m. waller. 2004. ecological impacts of deer overabundance. annual review of ecology, overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 56 evolution and systematics 35:113–147. crête, m., and r. courtois. 1997. limiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal of the zoological society of london 242:765–781. ______, and c. daigle. 1999. management of indigenous north american deer at the end of the 20th century in relation to large predators and primary production. acta veterinaria hungarica 47:1–16. ______, j.-p. ouellet, and l. lesage. 2001. comparative effects on plants of caribou/reindeer, moose and whitetailed deer herbivory. arctic 54:407– 417. damman, a. h. w. 1964. some forest types of central newfoundland and their relationship to environmental factors. forest science monograph 8. davidson, d. w. 1993. the effects of herbivory and granivory on terrestrial plant succession. oikos 68:23–35. degraaf, r. m., w. m. healy, and r. t. brooks. 1991. effects of thinning and deer browsing on breeding birds in new england oak woodlands. forest ecology and management 41:179–191. english, b., and r. hackett. 1994. the impact of kalmia on plantation performance in central newfoundland. silviculture notebook no. 2, government of newfoundland and labrador, st. john's, newfoundland, canada. ferguson, m. a. d., and f. messier. 1997. collection and analysis of traditional ecological knowledge about a population of arctic tundra caribou. arctic 50:17–28. gosse, j., b. mclaren, and e. eberhardt. 2002. comparison of fixed-wing and helicopter searches for moose in a midwinter habitat-based survey. alces 38:47-53. gross, j. e., l. a. shipley, n. t. hobbs, d. e. spalinger, and b. a. wunder. 1993. functional response of herbivores in food-concentrated patches: tests of a mechanistic model. ecology 74:778– 791. gustafsson, l., a. fiskesjö, t. ingelög, b. pettersson, and g. thor. 1992. factors of importance to some lichen species of deciduous broad-leaved woods in southern sweden. lichenologist 24:255–266. helle, p., and g. j. niemi. 1996. bird community dynamics in boreal forests. pages 209–234 in r. m. degraaf and r. i. miller, editors. conservation of faunal diversity in forested landscapes. chapman and hall, norwall, massachusetts, usa. hunter, m. l., jr. 1993. natural fire regimes as spatial models for managing boreal forests. biological conservation 65:115–120. james, f. c., and s. rathbun. 1981. rarefaction, relative abundance, and diversity of avian communities. auk 98: 785–800. jordan, p., r. o. peterson, p. campbell, and b. mclaren. 1993. comparison of pellet counts and aerial counts for estimating density of moose at isle royale, a progress report. alces 29:267–278. kuusinen, m. 1995. epiphytic lichen diversity on salix caprea and populus tremula in old-growth forests of finland. mitteilungen der eidgenössischen forschungsanstalt für wald, schnee und landschaft 70:125–132. leopold, a. 1949. a sand county almanac and sketches here and there. oxford university press, london, england. lewis, k. p. 2004. processes underlying nest predation by introduced red squirrels (tamiasciurus hudsonicus) in the boreal forest of newfoundland. ph.d. alces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 57 thesis, cognitive and behavioural ecology programme, memorial university of newfoundland, st. john’s, newfoundland and labrador, canada. lindenmayer, d., and j. franklin. 2003. towards forest sustainability. island press, washington, d.c., usa. maass, w. 1983. new observations on erioderma in north america. nordic journal of botany 3:567–576. macarthur, r. h., and j. w. macarthur. 1961. on bird species diversity. ecology 42:594–598. _____, _____, and j. preer, jr. 1962. on bird diversity. ii. prediction of bird census from habitat measurements. american naturalist 46:167–174. martin, t. e. 1992. breeding productivity considerations: what are the appropriate habitat features for management? pages 455–473 in j. m. hagan and d. w. johnston, editors. ecology and conservation of neotropical migrant landbirds. smithsonian institution press, washington, d.c, usa. _____. 1993. nest predation and nest sites, new perspectives on old patterns. bioscience 43:523–532. mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan. ecology 73:2059–2075. mclaren, b. e. 2004. social and natural history of moose introduced to newfoundland in a. j. gaston, t. e. golumbia, j.-l. martin, and s. t. sharpe, editors. lessons from the islands: introduced species and what they tell us about how ecosystems work. canadian wildlife service occasional paper. in press. _____, and s. p. mahoney. 2001. comparison of forestry-based remote sensing methodologies to evaluate caribou habitat in non-forested areas of new foundland. the forestry chronicle 77:866–873. _____, _____, t. s. porter, and s. m. oosenbrug. 2000a. spatial and temporal patterns of use by moose of precommercially thinned, naturally-regenerating stands of balsam fir in central newfoundland. forest ecology and management 133:179-196. _____, c. mccarthy, and s. p. mahoney. 2000b. extreme moose demographics in gros morne national park, newfoundland. alces 36:217–232. _____, and r. o. peterson. 1994. wolves, moose, and tree rings on isle royale. science 266:1555–1558. ______, and _____. 1996. seeing the f o r e s t w i t h t h e t r e e s : u s i n g dendrochronology to investigate mooseinduced changes to a forest understory. alces 31:77–86. mcshea, w. j., and j. h. rappole. 1997. herbivores and the ecology of forest understory birds. pages 298–309 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance, deer ecology and population management. smithsonian institution press, washington, d.c., usa. _____, h. b. underwood, and j. h. rappole (editors). 1997. the science of overabundance, deer ecology and population management. smithsonian institution press, washington, d.c., usa. mercer, w. e. 1995. moose management plan for newfoundland. report on file with inland fish and wildlife division, newfoundland and labrador, corner brook, newfoundland and labrador, canada. _____, and b. e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose. alces 38:123– 141. neff, d. j. 1968. the pellet-group count technique for big game trend, census, overabundant moose in newfoundland – mclaren et al. alces vol. 40, 2004 58 and distribution, a review. journal of wildlife management 32:541–545. neitlich, p. n., and b. mccune. 1997. hotspots of epiphytic lichen diversity in two young managed forests. conservation biology 11:172–182. niemelä, j. 1999. management in relation to disturbance in the boreal forest. forest ecology and management 115:127– 134. paine, r. t. 1995. a conversation on refining the concept of keystone species. conservation biology 9:962–964. parrish, j. d. 1995. effects of needle architecture on warbler habitat selection in a coastal spruce forest. ecology 76:1813–1820. pastor, j., and r. j. naiman. 1992. selective foraging and ecosystem processes in boreal forests. american naturalist 139:690–705. pimlott, d. h. 1953. newfoundland moose. transactions of the north american wildlife conference 18:563–581. _____. 1959. reproduction and productivity of newfoundland moose. journal of wildlife management 23:381–401. _____. 1963. influence of deer and moose on boreal forest vegetation in two areas of eastern canada. transactions of the 6th international union of game biologists congress, bournemouth, england. popotnik, g. j., and w. m. giuliano. 2000. response of birds to grazing of riparian zones. journal of wildlife management 64:976–982. potvin, f., l. bélanger, and k. lowell. 1999. validité de la carte forestière pour décrire les habitats fauniques à l’échelle locale: une étude de cas en abitibi-témiscamingue. the forestry chronicle 75:851–859. power, m. e., d. tilman, j. a. estes, b. a. menge, w. j. bond, l. s. mills, g. daily, j. c. castilla, j. lubchenco, and r. t. paine. 1996. challenges in the quest for keystones. bioscience 46:609–620. proulx, g., and r. joyal. 1981. forestry maps as an information source for description of moose winter yards. canadian journal of zoology 59:73–80. roberts, b. a. 1980. some chemical and physical properties of serpentine soils from western newfoundland. canadian journal of soil science 60:231– 240. _____. 1989a. the growth and development of white spruce (picea glauca) in newfoundland. page 167 in the silvics and ecology of boreal spruces. proceedings of the 11th iufro northern forest silviculture and management working party s 1.05–12 symposium, st. john's, newfoundland, canada. _____. 1989b. the origin and development of white spruce, picea glauca, forest stands in central newfoundland. american journal of botany 76:138. _____. 1992. the serpentinized areas of newfoundland, canada. a brief review of their soils and vegetation. pages 53– 66 in a. j. m. proctor and r. d. reeves, editors. the vegetation of ultramafic (serpentine) soils. intercept scientific publication, france. _____, and d. bajzak. 1984. forest site classification for the boreal forest of central newfoundland, canada (b.28a) using a bio-physical soils approach. joint meeting of the working parties no. 1.02–06 and no. 1.02–10 of iufro on q u a l i t a t i v e a n d q u a n t i t a t i v e asseessment of forest sites with special reference to soil. birmensdorf, switzerland. _____, and j. proctor, editors. 1992. the ecology of areas with serpentinized rocks, a world view. kluwer academic publishers, netherlands. robertson, a., and b. a. roberts. 1982. checklist of the alpine flora of western alces vol. 40, 2004 mclaren et al. overabundant moose in newfoundland 59 brook pond and deer pond areas, gros morne national park. rhodora 84:101– 115. rosentreter, r. 1995. lichen diversity in managed forests of the pacific northw e s t , u s a . m i t t e i l u n g e n d e r eidgenössischen forschungsanstalt für wald, schnee und landschaft 70:103– 124. rotenberry, j. t., r. j. cooper, j. m. wunderle, and k. g. smith. 1995. when and how are populations limited? the roles of insect outbreaks, fire, and other natural perturbations. pages 55– 84 in t. e. martin and d. m. finch, editors. ecology and management of neotropical migratory birds. oxford university press, new york, usa. setterington, m. a., i. d. thompson, and w. a. montevecchi. 2000. woodpecker abundance and habitat use in mature balsam fir forests in newfoundland. journal of wildlife management 64:335–345. sinclair, a. r. e., j. m. gosline, g. holdsworth, c. j. krebs, s. boutin, j. n. m. smith, r. boonstra, and m. dale. 1993. can the solar cycle and climate synchronize the snowshoe hare cycle in canada? evidence from tree rings and ice cores. american naturalist 141:173–198. telfer, e. s. 1967. comparison of a deeryard and a mooseyard in nova scotia. canadian journal of zoology 45:485–490. _____. 1972. forage yield and browse utilization on two logged areas in new brunswick. canadian journal of forest research 2:346–350. thompson, i. d. 1988. moose damage to pre-commercially thinned balsam fir stands in newfoundland. alces 24:56– 61. _____, and w. j. curran. 1993. a reexamination of moose damage to balsam fir – white birch forests of central newfoundland: 27 years later. canad i a n j o u r n a l o f f o r e s t r e s e a r c h 23:1388–1395. _____, _____, j. a. hancock, and c. e. butler. 1992. influence of moose browsing on successional forest growth on black spruce sites in newfoundland. forest ecology and management 47:29– 37. _____, and a. u. mallik. 1989. moose browsing and allelopathic effects of kalmia angustifolia on balsam fir regeneration in central newfoundland. canadian journal of forest research 19:524–526. white, p. s., and a. jentsch. 2001. the search for generality in studies of disturbance and ecosystem dynamics. progress in botany 62:399–450. willson, m. f., and t. a. comet. 1996. bird communities of northern forests, ecological correlates of diversity and abundance in the understory. condor 98:350–362. yetman, d. 1999. epiphytic lichen diversity and abundance based on forest stand type in terra nova national park: implications for lichen conservation and forest management, b.sc. thesis, department of biology, memorial university of newfoundland, newfoundland and labrador, canada. f:\alces\vol_39\p65\3908.pdf alces vol. 39, 2003 wam and hjeljord wolf predation, hunter data 263 wolf predation on moose a case study using hunter observations hilde karine wam and olav hjeljord department of ecology and natural resource management, norwegian university of life sciences, p. o. box 5003, n-1432 ås, norway abstract: we studied predation by colonizing wolves on a high density and highly productive moose (alces alces) population in south-eastern norway (about 1.5 moose and 0.01 wolves per km2 in winter). as indices to population changes, we used hunter observations. over the summer, the wolf pack utilized about one tenth of their total territory (530 km2), with the den area as the centre of activity. of the main prey taken (moose, roe deer, and beaver), moose calves contributed 61% of the biomass ingested by wolves in summer. hunting statistics and hunters’ observations of moose showed no changes for the territory as a whole after wolves settled there in 1998. however, in the den areas (60 80 km2) the number of calves per cow and the total number of moose seen per hunter-day significantly decreased during the year of wolf reproduction. the following year, though, both indices increased again. we speculate that some of the lack of overall effects might be due to increased fecundity in cows that lost their calf. as the wolves changed their den from year to year, den areas were spatially spread over time. the pressure from wolf predation will differ between cohorts in the same area, and landowners should adjust their hunting quotas accordingly. alces vol. 39: 263-272 (2003) key words: alces alces, canis lupus, compensation, hunter observations, predation, territory use the return of wolves (canis lupus) to southern scandinavia introduces several problems to wildlife management. one is a predicted reduction in harvest of moose (alces alces) due to wolf predation. to ease the resistance among norwegian moose hunters, the directorate for nature management is evaluating the possibility of reimbursing landowners yearly losses of moose to wolves. due to the highly dynamic nature of the predator prey relationship between wolves and moose (messier 1994, ballard and van ballenberghe 1998, hayes and harestad 2000), an eventual reimbursement plan requires that the effects of predation are estimated locally. in this study we investigate the influence of wolf predation on a high-density, productive moose population in south-eastern norway. we expect wolf predation to have a relatively small effect on this population, compared to less dense moose populations with lower recruitment rates (andrèn et al. 1999). however, within a wolf territory, we also expect predation to vary locally. since the den with rendezvous sites is the centre of activity throughout the summer (mech 1970), we expect predation losses to be higher among moose living close to the den. hence, landowners in the neighbourhood of the den may suffer a higher loss of moose to wolves than will landowners in other parts of the territory. how landowners should adjust their hunting quotas to mitigate the effects of predation depends not only on the number of moose taken, but also on which sexand age-group is preyed upon. as there are few old individuals among scandinavian moose, we expect the wolves to prey particularly on calves (fritts and mech 1981, boyd et al. 1994, olsson et al. 1997). however, when wolf predation, hunter data wam and hjeljord alces vol. 39, 2003 264 table 1. development of a re-establishing wolf pack in south-central norway 1998-2002 (based on snow tracking and sightings at rendezvous sites). territory size was 530 km2, and there were no bordering packs. (a = adult wolves, j = juvenile wolves, and p = pups.) in february in august 1998 no wolves 1a 1999 2a 2a, 5p 2000 2a, 3j 3a, 4 or 5p 2001 3a, 2j 3a, 8p 2002 1,2a1, 5 or 6j no denning 98-02 0.009/km2 0.017/km2 1 no alpha male. nursing calves are killed by brown bears (ursus arctos), there appears to be an increase in the fecundity of the cow the next year (swenson et al. 2001). a compensation effect may also apply to wolf predation on moose. in norway, moose populations are monitored routinely by a system in which hunters report on moose seen during the hunting season. because of large confidence limits, the hunter observation indices are not suitable for predicting absolute values of population size and recruitment rate. however, they appear well suited to predict directional changes (fryxell et al. 1988, solberg and sæther 1999). in this study we use hunter observation indices to look for changes in the moose population at two different scales: (1) within the wolf territory as a whole; and (2) within the wolves’ den areas. study area the study area is located in southeastern norway (59°33´n, 11°02´e), about 30 km east of the oslo fjord. most of the area is forested, and part of the boreonemoral zone, with spruce (picea abies) and pine (pinus silvestris) being the dominant tree species. lakes cover < 0.5% of the area, and bogs are infrequent. mature forest is harvested by clearcutting, and birch (betula pubescens and b. pendula) and rowan (sorbus aucuparia) dominate on clearcuts soon after logging. clearcuts are small and usually < 10 ha. elevations range from 40 to 260 m and the topography is broken by small creek valleys. the ground is usually snow-covered from december to april, with an average snow depth of 20 cm. the study area lies within the most high-yielding populations of both moose and roe deer (capreolus capreolus) in norway (bjar and selås 1987, hjeljord and histøl 1999). there has been no census of the ungulate density, but yearly harvest may be used as an indication. during 1995-2000 with an apparently stable population of the two species, an average of 0.6 moose and 1.6 roe deer were shot per km2 each year. assuming a yearly finite rate of increase of 1.35 for moose (hjeljord, unpublished data) and 1.4 for roe deer (strandegaard 1974), the density may be estimated at 1.5 moose/ km2 and 4.0 roe deer/km2. it is likely that the density of roe deer is actually higher as not all roe deer shot are reported by hunters. wolves had been absent from this area for 150 years when a female wolf settled there in 1998. a male wolf joined her shortly after, and a pack was formed. thereafter the wolf density varied with an average of 0.009 wolves per km2 in the winter (for further details see table 1). methods wolf use of the area was investigated by following the radio-collared alpha male from may 1999 to november 2001. the male was located every 30 minutes over a continuous 10-day period. ten such 10-day periods of intensive triangulation were spread over the year to get a picture of territory use throughout the different seaalces vol. 39, 2003 wam and hjeljord wolf predation, hunter data 265 sons. wolf summer diet was investigated by analysing faeces collected at the den site in the second week of august, 2000. scats were analysed for prey remains like claws, teeth, bones, and hair using standard methods as described in ballard et al. (1987). samples of hair that we macroscopically judged to belong to different species were later identified microscopically. in some samples it was necessary to study a gelatine casting of the cuticula (teerink 1991). blind tests were conducted to check the reliability of the method. we calculated the prey proportions of wolf diet both by occurrence and by biomass (as defined in floyd et al. 1978). for the latter, we used the equation of weaver (1993), y = 0.439 + 0.008x, where y is kg biomass consumed per scat of a particular prey of live biomass x kg. for estimates of biomass of the different prey species, we used the figures given in a scandinavian study by olsson et al. (1997). to investigate the effect of wolf predation on moose population size and reproduction, we used hunting statistics (central bureau of statistics 1995 2001), and hunter observations recorded mainly during the first week of the moose hunting season in early october. the hunt on each unit of land is done by a team of moose hunters, and the leader of each team completes the observation forms. as an indication of moose population size and recruitment rate/fecundity, we used the number of moose observed per hunterday (8h), and the number of calves per cow (hereafter the c.c. ratio), respectively. we also used the number of calves per cowwith-calf to verify our data, since wolf winter predation in this initial phase of recolonization may have affected the number of maiden cows (so far there have been more female than male moose > 1 year of age killed by wolves in scandinavia, sand et al. 2002). in the statistical analysis we used the 3year average 1995 1997 for each hunting unit as reference data against which we tested changes within the same hunting unit after the wolves settled. the year when the wolf arrived (1998) was not included in the analyses. the territory comprised a total of 19 hunting units, and the den areas 3 4 hunting units each. due to the small number of replicates in the den areas, we grouped all 3 den areas when looking for changes here. on average, there were 37 ± 5.7 se moose seen within each hunting unit, of which 17 ± 2.4 se were cows. our data were normally distributed, and we used paired t-tests for all comparisons. we also compared the moose population within the wolf territory with the moose population in a control area bordering the wolf territory. initial tests showed that prior to 1998, hunter statistics and observations within the wolf territory were correlated with that of the control area. results summer diet moose, roe deer, and beaver (castor fiber) (later called main prey species) dominated the prey remains in the collected scats (n = 151), contributing 94% by occurrence. mountain hare (lepus timidus), birds, domestic animals, and unidentified food items contributed the remaining 6%. moose dominated among the main prey species, contributing 44% by occurrence of main prey. moose calves appear to be an important part of wolf diet in the study area over the summer, as they made up 95% of all the moose (i.e., 42% by occurrence of main prey). roe deer contributed 36%, and beaver the remaining 20% by occurrence. for the proportions of biomass ingested see fig. 1. when we estimated the biomass of main prey species, we did not include 12% prey remains from scats where we could not distinguish juvenile roe deer from juvewolf predation, hunter data wam and hjeljord alces vol. 39, 2003 266 nile moose. summer territory use from april 2000 to november 2001 we recorded 2,961 positions of the radio-collared male, and estimated annual territory size to be 530 km2. while the animal regularly used the entire territory during fall/winter, the spring/summer use was more restricted and apparently depended on location of the den. during the summer of 2000 (may august) more than 90% of the locations we obtained during our 10-day triangulation periods lay within 4 km of the den. the area of primary occupancy extended over approximately 50 km2 (fig. 2), less than one tenth of the total territory. in 2001 the wolves moved their den site 17 km to the south, and our radio-locations indicated a similarly restricted range use during summer (approximately 70 km2). in 1999, the first year when wolves reproduced in the area, no animal was radio-collared. however, sightings of pups, and systematic searches for faeces and prey remains, made it possible to locate the den site. based on these data we outlined an area of hunting units close to each of the 3 dens where we expected the predation pressure on moose calves to have been most severe. this area, later called the den area, consisted of 4 units around the 1999 den (80 km2), 4 units around the 2000 den, (60 km2), and 3 units around the 2001 den (80 km2). the larger sizes of the 1999 and 2001 den areas are due to both the variable size of individual hunting units and our impression of the area used by the wolves in these two years. hunter observations in the territory and control area hunter-observations and harvest data of moose showed few changes in the territory as a whole following wolf colonization in 1998 (fig. 3). there was neither any moose calves 61 % moose older 8 % roe deer fawns 14 % roe deer older 5 % beaver 12 % . fig. 1. consumed biomass of main prey for wolves in south-central norway based on scat analysis (n = 151) collected at the den in august 2000. adjusted after weaver (1993). fig. 2. den site, summer range and total territory size as determined by triangulation of a radiocollared alpha male wolf in south-central norway, 2000. total wolf territory summer range use may-aug. 2000 wolf den n 0 5km alces vol. 39, 2003 wam and hjeljord wolf predation, hunter data 267 change in the number of moose seen per hunter-day (mean = 0.58 ± 0.08 se without wolves and 0.60 ± 0.06 with wolves) (t = -0.28, 18 df, p = 0.39), nor in the total number of moose shot (108 without wolves and 112 with wolves) (t = -0.62, 18 df, p = 0.22) (fig. 3). harvest of calves also remained stable (28 without wolves and 27 with wolves), while the number of yearlings harvested slightly decreased, although not significantly (46 versus 43) (t = 0.17, 18 df, p = 0.47). during the study period the hunting quotas were reduced by 7%, while the fulfilment of them increased from 87 to 97%. the total number of days (8 hours) hunted within the territory decreased from 1,649 (1995-1997) to 1,491 (1999-2001). hence, the hunters apparently had no problem getting all the moose on their quotas. for the territory as a whole there was no decrease in the c.c. ratio after wolves settled in the area (0.74 ± 0.03 without wolves and 0.74 ± 0.04 with wolves) (t = -0.01, 18 df, p = 0.50). if we look only at the cows with calves, there was a slight but insignificant increase (1.31 ± 0.04 without wolves and 1.39 ± 0.04 with wolves) (t = -1.23, 18 df, p = 0.11). in the control area, the same numbers of moose were hunted in the years 19951997 and 1999-2001 (131 vs. 129). there was a small increase in the number of moose seen per hunter-day (from 0.47 ± 0.03 to 0.57 ± 0.09) (t = -1.86, 5 df, p = 0.06), and there was no change either in the c.c. ratio (from 0.89 ± 0.01 to 0.88 ± 0.02) (t = 0.29, 5 df, p = 0.39) nor in the number of calves per cow-with-calves (1.32 ± 0.03 vs. 1.34 ± 0.00) (t = -0.19, 5 df, p = 0.43). hunter observations in the den areas within the den areas, there were more obvious changes in the hunter observations of moose than for the territory as a whole (fig. 4). in the years of wolf denning, there was an insignificant decrease in the number of moose seen per hunter-day from what it had been before wolves re-established (from 0.50 ± 0.05 without wolves to 0.45 ± 0.02 in the year of denning) (t = 0.72, 10 df, p = 0.24). the year following active denning, there were more moose seen within the den areas than before wolves re-established (0.66 ± 0.17), although due to high variance this was not significant (t = -1.51, 10 df, p = 0.08). compared to 1995-1997, the c.c. ratio fig. 3. number of moose seen and harvested within a wolf territory and a control area, before and after wolf settled in the area in 1998, south-central norway. average winter density: 0.009 wolves and 1.5 moose per km2. (n.s. = not significant.) wolf predation, hunter data wam and hjeljord alces vol. 39, 2003 268 in the den areas significantly decreased in the years of wolf reproduction (from 0.78 ± 0.01 to 0.54 ± 0.03) (t = 3.6, 10 df, p = 0.00). the year after active denning, though, the c.c. ratios were higher than the pre-wolf levels (0.86 ± 0.08), but again the variance was high and the results were not statistically significant (t = -1.0, 10 df, p = 0.17). if we look at only the cows with calves, fewer calves were seen per cow with calves in the den year than before wolf re-establishment (from 1.35 ± 0.02 before wolves down to 1.14 ± 0.03 in den year) (t = 3.7, 10 df, p = 0.00). as with the c.c. ratio, more calves were observed per cow with calves the year after denning compared to the period before wolves re-established (from 1.35 ± 0.02 to 1.60 ± 0.09) (t = -2.1, 10 df, p = 0.03). discussion moose calves in the wolf diet moose are the primary prey of wolves in the northern boreal forest (fuller and keith 1980, peterson et al. 1984, messier and crete 1985, ballard et al. 1991, gasaway et al. 1992, gade-jørgensen and stagegaard 2000). in our study, wolves clearly preferred the calf segment of the moose population during summer. a preference for calves was also shown by olsson et al. (1997) in south-central sweden, where 51% of 65 moose killed by wolves were calves. apparently low-density, colonizing wolves kill a higher proportion of calves than do established wolf populations (fritts and mech 1981, boyd et al. 1994). smaller ungulates such as white tailed deer (murie 1944, carbyn 1983), red deer (murie 1944, carbyn 1983), caribou (hollermann and stephenson 1981, dale et al. 1995), and roe deer (olsson et al. 1997) seem to be the preferred prey where they occur together with moose. our data suggest that roe deer were killed at about the same rate as moose during summer (36% of occurrence for roe deer, and 44% for moose). since the density of roe deer probably was at least 3 times that of moose, this indicates a preference not for roe deer, but for moose. scats collected on forest roads (may-november) in the study area and 2 other wolf territories in the same region, also indicate a preference for moose over roe deer during summer (østreng 2000). in sweden, olsson et al. (1997) concluded from their scat analysis, that wolves killed roe deer about twice as often as moose (52% of occurrence for roe deer, and 25% for moose). with a moose density in their study area at about 3 times that of roe deer, their conclusion was the opposite fig. 4. number of moose seen and calf recruitment rates in areas adjacent to the den in a wolf territory in south-central norway. data from 3 consecutive dens (1999-2001) are grouped. there were no wolves in the area prior to 1998. (n.s. = not significant.) 0,50 0,45 0,66 0,78 0,54 0,86 1,35 1,14 1,60 0,00 0,20 0,40 0,60 0,80 1,00 1,20 1,40 1,60 95-97 den year year after 95-97 den year year after 95-97 den year year after stolpediagram 1 linjediagram 2 linjediagram 3 moose seen per hunter's day (8h) calves seen per cow calves seen per cow with calves n.s n.s ** ** * n.s * * n.s alces vol. 39, 2003 wam and hjeljord wolf predation, hunter data 269 of ours: that the wolves preferred roe deer over moose. in their study, however, scats were collected throughout the year. the relative vulnerability of roe deer to moose probably depends on seasonal factors such as snow depth. furthermore, when the study in sweden was started, moose had already been exposed to wolves for 7 8 years. this might have made the animals less naive as prey (berger et al. 2001) compared to our study area where scats were collected shortly after wolves had settled in the area. expected effects of predation using data from the scat analysis, estimated moose density, yearly calf production, and wolf daily food requirement, we can estimate the effects of predation on this particular moose population: with 210 km2 of moose habitat within the wolf territory, 315 moose (of which 105 were calves) were potential prey for the wolves each summer. there is no data in the literature to calculate the food needed to raise a litter of wolf pups. however, using data from mech (1970) and data from dog breeders (wam, unpublished data) we have set the average food intake by pups (average summer weight 9 kg) from mid may to the end of september at 1.4 kg per pup per day, and the average food intake for adults at 3.7 kg per animal per day. a litter of 5 and 8 pups then needs 945 kg and 1,512 kg of meat, respectively, during the summer. in our study area, 61% of this should be derived from moose calves. applying an average meat yield of 35 kg per moose calf (olsson et al. 1997), and including the number of adult wolves (2 adults during the summer of 1999, 3 adults in 2000 and 2001) in our calculations, we estimate that 34, 40, and 50 moose calves were consumed over the summers 1999, 2000, and 2001, respectively. within the den area of 2000 (5-6 pups), there were about 28 calves, given the moose density of 1.5 km2, and in the den area of 1999 and 2001 (5 and 8 pups, respectively) there were 37, assuming that all of the area was moose habitat. therefore, if wolves killed calves mostly within their den areas, very few if any would be left there by the end of the summer. for the territory as a whole, the estimated calf losses are 32, 38, and 48%, respectively, for 1999-2001. likewise, we can estimate the total number of moose killed since wolves first denned in this territory in 1999. using the actual wolf pack size for each year, a daily meat intake of 3.7 kg per animal per day, and a similar proportion of calf and adult moose in the winter kill as in the summer kill, wolves should have killed an approximate total of 235 moose by the fall of 2001, or about 15% of the summer population per year. theoretical vs. observed loss from predation the observed overall losses of moose from predation were diminishingly small, and lower than expected from our theoretical calculations. possible reasons include: (1) there was an increased immigration of moose into the territory; (2) more calves were born after the wolves established; or (3) the moose population was increasing at the time wolves re-established. theoretically the wolf territory could act as a sink for dispersing young moose from the surrounding forests. however, because the territory is enclosed by highways, lakes, and broad rivers, we believe the migration of animals into the area is negligible. furthermore, there is little evidence showing a selective colonization by moose of areas where the density has been reduced from hunting, predation, or other causes (hjeljord 2001). an increase in calf production by cows losing nursing calves has been documented wolf predation, hunter data wam and hjeljord alces vol. 39, 2003 270 in scandinavian brown bears (swenson et al. 2001), and this may also apply when calves are preyed upon by wolves. we found no decrease in the c.c. ratio in the territory at large after the wolves arrived, even though we did find a significant decrease within the den areas in the years of active denning. we speculate that this is partly due to higher fecundity in cows that lost a calf to wolf predation the previous year, as our data did show an increase in recruitment rates. however, our sample size is too small to draw any firm conclusion. as calves in a den area are also preyed upon the year after active denning (albeit not that intense), increased fecundity in spring will be reduced by the time of census when using hunters’ observations. the best way to assess the fecundity would therefore be to do a survey of the number of calves per cow as soon after birth as possible, and then compare it with hunter observations in the autumn. for the control area, the hunter observations indicated an increase in the moose population from 1995 1997 to 1999 2001. the moose in the control area and the wolf territory have similar conditions, apart from wolf predation. probably the wolves halted a similar increase in the moose population inside their territory and thereby masked some of the effects of predation. the discrepancy between calculated and observed losses was enhanced by our assumption that the proportion of roe deer in the wolves’ diet is similar during summer and winter. this assumption certainly is invalid for most winters, and this may be part of the reason why there has been a smaller impact of wolf predation than suggested by our estimates. this does not, however, affect what we observed. conclusion the establishment of a wolf pack in south-eastern norway caused little change to the highly productive moose population in the area. yearly harvest and population size as indicated by hunter observations remained stable. we believe, however, that the moose population was increasing when the wolves settled, and this may have masked some of the effects of predation from wolves. if so, we will see greater predation effects in coming years as the increase is halted, and possibly reversed. our data also indicate that the moose might have compensated for the loss of calves during summer. it is, however, premature to conclude from our small sample size. more data is needed on this topic. in our study area the wolves changed their den location from year to year. if this is something they ordinarily do, an area that is heavily preyed upon one year will get a chance to recover the following year, thereby dividing predation loss among land owners. however, over the long term and if hunting quotas are not adjusted, cohorts can be reduced below what is needed to replace harvested adult moose. we have shown that hunter observation indices can be a valuable tool in future management of moose populations preyed upon by wolves in scandinavia. they may be used to adjust hunting quotas for specific areas and years. to get an accurate account of the situation, though, it is important that the hunting units used as replicates are not too small. rather than having many replicates, we believe it might be a better option to group small replicates into bigger units (within the appropriate geographical scale). furthermore, data gathered by the hunters should be used to promote a good dialog between managers and the people who actually harvest the moose. acknowledgements the work was supported by grants from alces vol. 39, 2003 wam and hjeljord wolf predation, hunter data 271 the directorate for nature management, the norwegian research council, and the regional wildlife administration of østfold. references andrén, h., o. liberg, and h. sand. 1999. de stora rovdjurens inverkan på de vilda bytesstammarna i sverige. sou 1999: 146, miljödepartementet, stockholm, sweden. (in swedish). ballard, w. b., and v. van ballenberghe. 1998. pedator/prey relationships. pages 247-273 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , j. s. whitman, and c. l. gardner. 1987. ecology of an exploited wolf population in south-central alaska. wildlife monographs 98. , , and d. j. reed. 1991. population dynamics of moose in southcentral alaska. wildlife monographs 114. berger, j., j. e. swenson, and i.-l. persson. 2001. recolonizing carnivores and naïve p r e y : c o n s e r v a t i o n l e s s o n s f r o m pleistocene extinctions. science 291:1036-1039. bjar, g., and v. selås. 1987. sosial regulering og habitat bruk hos rådyr capreolus capreolus om sommeren. m.sc. thesis, agricultural university of norway, ås, norway. (in norwegian). boyd, d. k., r. r. ream, d. h. pletcher, and m. w. fairchild. 1994. prey taken by colonizing wolves and hunters in the glacier national park area. journal of wildlife management 58:289-295. carbyn, l. 1983. wolf predation on elk in riding mountain national park, manitoba. journal of wildlife management 47:963-976. dale, b. w., l. g. adams, and r. t. bowyer. 1995. winter wolf predation in a multiple ungulate system, gates of the arctic national park, alaska. pages 223-230 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, edmonton, alberta, canada. floyd, t.j., l. d. mech, and p. a. jordan. 1978. relating wolf scat content to prey consumed. journal of wildlife management 42:528-532. fritts, s. h., and l. d. mech. 1981. dynamics, movements, and feeding ecology of a newly protected wolf population in northwestern minnesota. wildlife monographs 80. fryxell, j. m., w. w. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52:14-21. fuller, t. k., and l. b. keith. 1980. wolf population dynamics and prey relationships in northeastern alberta. journal of wildlife management 44:583-602. gade-jørgensen, i., and r. stagegaard. 2000. diet composition of wolves canis lupus in east-central finland. acta theriologica 45:537-547. gasaway, w. c., r. d. boertje, d. v. grandgard, k. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. hayes, r. d., and a. s. harestad. 2000. wolf functional response and regulation of moose in the yukon. canadian journal of zoology 78:60-66. hjeljord, o. 2001. dispersal and migration in northern forest deer are there unifying concepts? alces 37:353-370. , and t. histøl. 1999. range-body mass interactions of a northern unguwolf predation, hunter data wam and hjeljord alces vol. 39, 2003 272 late a test of hypothesis. oecologia 119:326-339. hollerman, d. f., and r. o. stephenson. 1981. prey selection and consumption by alaskan wolves. journal of wildlife management 45:620-628. mech, l. d. 1970. the wolf: the ecology and behavior of an endangered species. natural history press, garden city, new york, usa. messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75:478-488. , and m. crête. 1985. moose-wolf dynamics and the natural regulation of moose populations. oecologia 65:4450. murie, a. 1944. the wolves of mount mckinley. u.s. national park service, fauna series no. 5. olsson, o., j. wirtberg, m. andersson, and i. wirtberg. 1997. wolf canis lupus predation on moose alces alces and roe deer capreolus capreolus in south-central scandinavia. wildlife biology 3:13-25. østreng, o.-c. 2000. ulv canis lupus i akershus og østfold, sommerdiett og byttedyrselektivitet. m.sc. thesis, agricultural university of norway, ås, norway. (in norwegian). peterson, r. o., j. d. woolington, and n. t. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88. sand, h., p. wabakken, c. wikenros, o. liberg, and h.-c. pedersen. 2002. p a t t e r n s o f p r e y s e l e c t i o n b y scandinavian wolves. abstract, 5th international moose symposium, hafjell, norway. solberg, e. j., and b. e. sæther. 1999. hunter observations of moose as a management tool. wildlife biology 5:107-117. strandegaard, h. 1974. the roe deer (capreolus capreolus) population at kalø and the factors regulating its size. danish review of game biology 7. no. 1, 205 pp. swenson, j. e., b. dahle, and f. sandegrän. 2001. brown bear predation on moose in scandinavia. nina fagrapport 048. teerink, b. j. 1991. hair of west-european mammals. atlas and identification key. cambridge university press, cambridge, uk. weaver, j. l. 1993. refining the equation for interpreting prey occurrence in gray wolf scats. journal of wildlife management 57:534-538. 99 winter habitat use of moose in cape breton, nova scotia jason i. airst and jason w. b. power nova scotia department of lands and forestry, wildlife division, kentville, nova scotia, canada abstract: aerial survey data collected between 2001 and 2020 were used to assess winter habitat use by moose (alces alces) in the greater highland ecosystem of cape breton, nova scotia. these data were analyzed using generalized additive mixed models that explored the influence of habitat variables. we compared abundance estimates developed directly from the surveys to those estimated from habitat use. moose generally occupied the same general area throughout the study despite a marked population decline. moose favoured areas comprised of greater proportions of coniferous forest showing preference for younger forest, and moose meadows, areas of predominantly coniferous forest but with abnormal or retarded regeneration due to high moose herbivory. moose occupied areas farther away from roads inferring that moose preferred areas with younger plant forage and lower human access. the use of long-term survey data coupled with related habitat use relationships provided a useful approach to assess temporal tends in abundance and habitat use of moose in cape breton. alces vol. 57: 99–111 (2021) key words: aerial survey; alces alces; cape breton; gis; habitat use; roads. a fundamental objective of wildlife management is to maintain healthy sustainable wildlife populations (fryxell et al. 2014). monitoring population-wide demographics such as abundance and vital rates is one effective approach for managing populations (williams 2011, boyce et al. 2012, fryxell et al. 2014). another important strategy is to anticipate effects of management actions on wildlife populations by understanding wildlife-habitat relationships in the context of environmental change (krausman 1999, hebblewhite and merrill 2008). in this study, our primary objective was to assess moose (alces alces) habitat use in the greater highland ecosystem of cape breton, nova scotia using 20 years of aerial survey data. aerial surveys are commonly used to assess large mammal populations active during winter (gasaway et al. 1986, kantar and cumberland 2013). this approach has been used in many jurisdictions, including nova scotia, to assess changes in moose populations and to inform management decisions (snaith et al. 2002, van beest et al. 2012, andreozzi et al. 2016). moose, like other mammals, show seasonal patterns of habitat use (schwartz and franzmann 1998, manly et al. 2002, van beest et al. 2012). therefore, by noting spatial variation in the abundance of moose observed during aerial surveys, we can infer habitat preferences across the landscape (manly et al. 2002, van beest et al. 2012, andreozzi et al. 2016). however, one must also account for how easily animals are detected during surveys, or their sightability. if not corrected for, sightability can bias survey results (anderson and lindzey 1996). in cape breton, moose habitat selection is predominantly shaped by a history of winter habitat use in nova scotia – airst and power alces vol. 57, 2021 100 spruce budworm (choristoneura fumiferana) infestations with the last major outbreak in the late 1970s (bridgland et al. 2007, smith et al. 2015). as a result, cape breton has experienced significant change in habitat composition and forest age structure in recent decades, providing a unique opportunity to examine habitat use by moose in response to this forest heterogeneity. further, because road density varies greatly across the cape breton landscape, we also examined its potential influence on habitat use. we predicted that moose should be attracted to younger softwood forests and avoid areas nearer roads (schwartz and franzmann 1998, manly et al. 2002, van beest et al. 2012). study area in cape breton, moose were abundant prior to european arrival. however, subsequent change in land use (i.e., habitat) and hunting pressure caused population decline that continued until the early 1900s when moose were virtually extirpated from the region (pulsifer and nette 1995, davis and browne 1996). in 1947 and 1948, parks canada reintroduced moose to the cape breton highlands national park (cbhnp) by translocating 18 moose from elk island national park in alberta (davis and browne 1996). this reintroduction was successful and moose gradually spread across the region, and by the mid-1980s the population was approximately 4000 animals (pulsifer and nette 1995). in the mid-1970s, a major spruce budworm outbreak caused extensive tree mortality spurring an abundance of new tree growth (bridgland et al. 2007, smith et al. 2015). taking advantage of this abundant food source, the population rapidly grew in the late-1990s until 2004 when it peaked at just over 8000 animals (bridgland et al. 2007, smith et al. 2015). the population gradually declined to ~ 4500 in 2015 and roughly halved again to 2300 animals by 2020 (smith et al. 2015, nova scotia lands and forestry 2020a). moose on cape breton island are almost exclusively found in one region, the greater highland ecosystem (ghe), with limited migration, immigration, and emigration. the ghe makes up the northwestern third of cape breton island (fig. 1) and is bordered by the gulf of saint lawrence to the west where the land rises rapidly from the ocean to a height of 500 m. the land forms a large plateau that slopes eastward and northward toward the atlantic ocean. south of the cbnhp, there is an extensive road network built in the 1980s in conjunction with increased logging. the area has a maritime climate with average winter and summer fig. 1. map of nova scotia’s greater highland ecosystem (ghe). the map also shows the location of cape breton highlands national park (cbhnp) and paved and unpaved roads. alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 101 temperatures of −5 and 18°c, respectively. annual precipitation averages 1053 mm of rainfall and 337 cm of snowfall (environment canada 2020). the ghe has 3 major land types: boreal forest, acadian forest, and taiga (smith et al. 2015) which are characterized by a mixture of forest types and ages over recent decades (table 1). following disturbance events, boreal succession typically starts with rapid growth of trees in the understory favouring faster growing shade-intolerant species such as white birch (betula papyrifera). these shade intolerant species are eventually overtaken by slower growing shade tolerant conifer species (maclean and ostaff 1989, smith et al. 2010). however, herbivores can affect this successional pattern through their foraging activity, most typically by over-browsing preferred plants (mclaren et al. 2004, smith et al. 2010). moose in many areas of the ghe consume nearly all conifer trees before they can grow large enough to escape herbivory. the resultant forest is characterized by open savannahs dominated by remnant white birch, alder (alnus spp.), black spruce (picea mariana), and herbaceous growth (smith et al. 2010, 2015), hereafter termed “moose meadows” that are mostly located on the western side of the cbhnp. methods aerial survey data were collected as part of 10 aerial surveys conducted over a 20-year period (2001, 2002, 2004, 2006, 2008, 2011, 2013, 2015, 2019, 2020). these helicopter surveys were conducted by 4 personnel: 2 back-seat observers, 1 front-seat recorder, and the pilot. all surveys were completed in 1–2 days during the first week of march and conducted in conditions of adequate visibility and minimal precipitation. effort was made to maximize sightability of moose by flying at low speed (< 130 km/h) and altitude (< 100 m above ground), and only recording animals within 150 m of either side of the helicopter (gasaway et al. 1986). this work was completed as part of an effort by nova scotia lands and forestry to monitor moose in the area and done in partnership with parks canada, the unama’ki institute of natural resources, and the confederacy of mainland mi’kmaq. to survey the area, we divided the ghe into 893 equal-sized survey blocks that were 2 min of longitude (~ 2.5 km) × 1 min of latitude (~1.9 km) large, or ~ 4.7 km2. data were collected on an east-west transect flight flown over the midline of each block; transect lines were 1 min of latitude apart (~ 1.9 km). observations (sightings) of moose were made within 150 m of the helicopter covering 1/6 of the block area; the number of moose observed per transect was recorded. a total of 8674 transects were flown over the 20-year study period. table 1. habitat composition of the nova scotia’s greater highlands ecosystem in 1999, 2009, and 2020. habitat type % of land cover 1999 2009 2020 non-forest 37.3 28.1 28.1 <25 year conifer 12.0 8.7 2.8 25–40 year conifer 8.9 18.2 5.2 >40 year conifer 3.9 5.4 24.5 uneven year conifer 4.4 4.6 4.4 <25 year deciduous 0.3 0.1 0.2 25–40 year deciduous 0.8 0.8 0.1 >40 year deciduous 9.2 8.0 8.5 uneven year deciduous 0.5 1.3 1.3 <25 year mixed wood 2.8 1.2 0.6 25–40 year mixed wood 3.2 5.0 1.0 >40 year mixed wood 10.5 9.9 14.6 uneven year mixed wood 2.2 5.0 4.8 moose meadow 3.9 3.9 3.9 winter habitat use in nova scotia – airst and power alces vol. 57, 2021 102 to control for sightability, we created a sightability correction factor (scf) in each survey year. this required flying 6 equidistant, east-west transect lines over 28–44 survey blocks each year, counting all moose observed; lines were spaced 300 m apart. these same areas were re-flown using a more intense survey regime of 12 flights spaced 150 m apart with half the viewing distance. moose counts from the two surveys were compared to determine if more animals were observed in the more intense survey. these comparisons were averaged to create a scf in each survey year (table 2). these same scfs were used when historical moose abundances were first calculated in the area (bridgland et al. 2007, smith et al. 2015, nova scotia lands and forestry 2020a). habitat and surface features habitat type underlying survey transects was determined using nova scotia’s 1999, 2009, and 2020 forest inventories (nova scotia lands and forestry 1999, 2009, 2020b). the 1999 and 2009 forest inventories were based on forestry records and satellite imagery, and the 2020 inventory was based on forestry records and age progression of the 2009 inventory. the 2001–2005 surveys were assessed with the 1999 inventory, the 2006–2014 surveys with the 2009 inventory, and those in 2015–2020 with the 2020 inventory. habitat types were categorized according to forest type (deciduous, conifer, mixed) and stand age (< 25 years old, 25–40 years old, > 40 years old, and uneven aged) and included an additional forest category for moose meadows. all non-forested habitats were combined into a single category. the forest inventories were also used to calculate the average percent crown closure within each transect. additionally, provincial road (distinguishing paved and unpaved) and surface water maps were used to assess abiotic habitat characteristics. because data collection was restricted to 150 m on either side of transect lines, we created a 150 m buffer from the transect using arcgis (version 10.5.1, environmental systems research institute 2018). we then calculated the percent habitat cover for each habitat type in each transect using the arcgis add-in patch analyst version 5.2 (rempel et al. 2016). to determine the average distance to paved and unpaved roads and surface water features in each transect, we first calculated the euclidian distance of all these features on the landscape. we then averaged these values in each transect to produce a measure of the distance to these features. we repeated the entire process for each survey block to subsequently use this information to determine the likely abundance and distribution of moose across the landscape. statistical analysis to determine how site composition affects moose numbers in each transect, we used a generalized additive mixed model (gamm) with a poisson distribution using the r package “gamm” (version 0.2–6, r table 2. sightability correction factors (scf) for each year of aerial moose surveys in nova scotia. scf were calculated based on the average difference between two sets of surveys flown over the same areas using different survey intensities. year n scf variance 2001 36 1.21 0.02 2002 37 1.36 0.05 2004 30 1.12 0.01 2006 37 1.14 0.01 2008 32 1.14 0.01 2011 28 1.30 0.02 2013 36 1.25 0.02 2015 36 1.51 0.05 2019 44 1.48 0.07 2020 38 1.18 0.01 alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 103 core team 2019, wood and scheipl 2020). a poisson (p) distribution was chosen over a negative binomial (nb) distribution as it yielded a lower akaike information criteria (aic) value when the two were compared (aicp = 11,941.9; aicnb = 12,183.3, theta = 10). generalized additive mixed models allow for both linear and non-linear effects in the model. for variables with non-linear effects, we used spline smoothers and a non-parametric reverse iterative approach to create separate model estimates for sections of the regression line (wood 2017). the procedure involves assessing how much model fit is improved as more spline smoothers are used to explain the non-linear effect, while simultaneously penalizing the model for each smoother added. the result of this was that an optimal number of smoothers, or effective degrees of freedom (edf), was chosen for each non-linear effect. this method, however, did not yield a single model estimate for each non-linear effect as multiple smoothers were used when creating these model effects (wood 2017, wood and scheipl 2020). our objectives were to understand moose habitat use with respect to roads and forest type and age, while controlling for the underlying effect of forest cover; thus, we included crown closure as a non-linear fixed effect. because survey blocks are spatially autocorrelated, we also used the x and y coordinates of the transect lines as a non-linear effect (kneib et al. 2009, wood 2017). the remaining fixed linear effects were percent cover for each habitat type and the average distance to roads and surface water features. survey year was used as a random effect to account for population differences between surveys. for model selection, we fitted a global model and then used stepwise backward selection to remove non-important variables from the model. this involved sequentially removing the variable with the lowest beta/ se absolute value until the aic value stopped declining (pagano and arnold 2009). once all non-important variables were removed form the model, we compared models using their aic values and estimated model weights. these weights were then used to create model averaged estimates that were unlogged to generate incidence rate ratios (irrs). an irr indicates how the relative count of moose changes as you increase an independent variable by 1 unit. in the case of habitat type, one unit was the difference between a site without or devoid of a habitat type (0) or entirely composed of that type (1). for roads and water, this was measured as distance (km) from the feature. to assess if our model was able to accurately predict moose abundance in the ghe, we compared historic abundance estimates from this area (historical estimates) to abundance estimates based on our model and the habitat and surface feature data from the survey blocks (model estimates). the model estimate included the random intercept for each survey year. we also used the survey block data to determine the road density across the ghe. because moose abundance was calculated to account for the ghe and not the entire area surveyed with transects, we applied a multiplier of 6 to the model estimates to account for the entire area. we further needed to multiply our estimates by the scf for each year to account for missed animals. after this adjustment, we compared the observed and estimated abundances using a pearson’s correlation coefficient. we calculated the average abundance in each survey block using the 1999, 2009, and 2020 forest inventories to determine the likely moose distribution during 3 time periods (2001–2005, 2006–2014, and 2015– 2020), and visually assessed these data for temporal and geographic trends in abundance. winter habitat use in nova scotia – airst and power alces vol. 57, 2021 104 results our best supported model identified relationships between habitat use and several categories of forest type and age, as well as distance to water and roads (table 3). while aic values indicated an initial best model, (aic = 11,933.5, variables = 16; table 4), the removal of uneven aged coniferous forest (11,934.6) and < 25-year-old deciduous forest (11,934.4) separately, and together (11,935.3), increased the aic value by < 2 points. because of the small difference between these four models, we used model averaging based on model weights to generate the final estimates. we found that the probability of observation was highest in moose meadows followed consecutively by younger coniferous forest types. the incidence rate ratios (95% ci) were 10.32 (8.25, 12.90) in moose meadows, 3.67 (3.02, 4.46) in < 25-year-old conifer forests, 3.16 (2.58, 3.88) in 25–40-year old conifer forests, and 1.30 (0.99, 1.69) in > 40-year old conifer forests (table 3). lower selection was found for >25-year-old deciduous forest, 25–40 year-old mixed wood forest, and > 40-year-old mixed wood forest: incidence ratios were 2.87 (0.98, 8.40), 3.12 (2.17, 4.50), and 4.33 (3.39,5.53), respectively. there was minimal chance of observation in uneven aged deciduous forest: incidence ratio = 0.45 (0.33, 0.610) (table 5). areas with greater average distance to roads had higher observations of moose: paved = 1.05 (1.03, 1.07), unpaved = 1.27 (1.24, 1.30). the opposite was true for areas with a greater average distance to surface water (0.07 (0.05, 0.11) (table 5). road density (km of road/km2) was lowest in the cbhnp and highest in the area to the south: north of park = 0.29 (0.07, 0.51), park = 0.15 (0.03, 0.27), south of park = 0.71 (0.70, 0.73). table 3. removal of non-informative variable from the habitat use model generated from transects flights over the greater highland ecosystem of nova scotia (10 surveys, n = 8674). variables were sequentially desegrated (bolded) based on their beta/se absolute values until the simplified model’s aic value began to increase. all models were general additive models with poisson distributions and with year as a random factor. transect location and % crown closure were treated as non-linear factors. model # of variables aic δaic ce, cy, cm, cu, de, dy, dm, du, me, my, mm, mu, moo, nf, paved roads, unpaved, water 21 11,941.85 8.35 ce, cy, cm, cu, de, dy, du, me, my, mm, mu, moo, nf, paved roads, unpaved, water 20 11,939.90 6.40 ce, cy, cm, cu, de, dy, du, me, my, mm, moo, nf, paved roads, unpaved, water 19 11,937.90 4.40 ce, cy, cm, cu, de, dy, du, my, mm, moo, nf, paved roads, unpaved, water 18 11,935.96 2.46 ce, cy, cm, cu, de, dy, du, my, mm, moo, paved roads, unpaved, water 17 11,934.31 0.81 ce, cy, cm, cu, de, du, my, mm, moo, paved roads, unpaved, water 16 11,933.50 0.00 ce, cy, cm, de, du, my, mm, moo, paved roads, unpaved, water 15 11,934.50 1.00 habitat types: <25 year conifer (ce), 25–40 year conifer (cy), >40 year conifer (cm), uneven year conifer (cu), <25 year deciduous (de), 25–40 year deciduous (dy), >40 year deciduous (dm), uneven year deciduous (du), <25 year mixed (me), 25–40 year mixed (my), >40 year mixed (mm), uneven year mixed (mu), moose meadow (moo), non-forest (nf) alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 105 table 4. comparison of models generated from flown transects over the greater highland ecosystem of nova scotia using aic to determine which habitat type and surface features best predict moose counts (10 surveys, n = 8674). all models were generalized additive mixed models with poisson distributions and with year as a random factor. transect location and % crown closure were both treated as non-linear factors. model # of variables aic δaic weight ce, cy, cm, cu, de, du, my, mm, moo, paved roads, unpaved, water 16 11,933.5 0 0.38 ce, cy, cm, cu, du, my, mm, moo, paved roads, unpaved, water 15 11,934.4 0.9 0.25 ce, cy, cm, de, du, my, mm, moo, paved roads, unpaved, water 15 11,934.6 1.1 0.22 ce, cy, cm, du, my, mm, moo, paved roads, unpaved, water 14 11,935.3 1.8 0.15 null 4 13,077.7 1144.2 0 habitat types: <25 year conifer (ce), 25–40 year conifer (cy), >40 year conifer (cm), uneven year conifer (cu), <25 year deciduous (de), uneven year deciduous (du), 25–40 year mixed (my), >40 year mixed (mm), moose meadow (moo) table 5. model averaged incidence rate ratio ratios (irrs) of seeing a moose on a transect in nova scotia’s greater highland ecosystem based on its habitat type and average distance from surface features (10 surveys, n = 8674). results were generated from a generalized additive mixed model with a poisson distribution, a with year as a random effect. transect location and % crown closure were treated as non-linear effects. habitat type irrs are based on a 100% cover of that habitat type. edf shows the number of spline smoothers used in each non-linear effect. linear variables variables incidence rate ratio 95% ci (intercept) 0.22 (0.16, 0.31) <25 year conifer 3.67 (3.02, 4.46) 25–40 year conifer 3.16 (2.58, 3.88) >40 year conifer 2.01 (1.54, 2.63) uneven year conifer 1.30 (0.99, 1.69) <25 year deciduous 2.87 (0.98, 8.40) uneven year deciduous 0.14 (0.05, 0.42) 25–40 year mixed wood 3.12 (2.17,4.50) >40 year mixed wood 4.33 (3.39, 5.53) moose meadow 10.32 (8.25, 12.90) avg. distance paved road (per km) 1.05 (1.03, 1.07) avg. distance unpaved road (per km) 1.27 (1.24, 1.30) avg. distance surface water (per km) 0.07 (0.05, 0.11) non-linear spline smoothers variables edf chi2 location (x, y) 24.90 328.28 % crown closure 3.17 27.39 winter habitat use in nova scotia – airst and power alces vol. 57, 2021 106 because the gam analysis precluded calculation of model estimates for the non-linear variables (average percent crown closure and moose group location), we plotted these data to assess potential relationships. observations increased as average percent crown closure rose to ~17%, and then declined. this decline slowed around 55% but continued to 80%, the upper end of the range (fig. 2). moose groups were highly clustered across the landscape, creating distinct high and low abundance areas (fig. 3). both these patterns fit with our expectations based on the best model (edf values; crown closure = 3.17; moose location = 24.90) (table 5). when we estimated moose abundance for the entire ghe, the abundance estimates based on the model were highly correlated with the historical estimates (r2 = 0.94; fig. 4). when examining the average moose distribution across the ghe in the 3 time periods (2001–2005, 2006–2013, and 2015–2020), we found that moose occupied the same areas throughout, although distribution contracted especially south of cbhnp (fig. 3). the principal areas of consistent occupation were just north of cbhnp, the western side of the cbhnp, and to a lesser extent an area southeast of cbhnp, all protected areas with poor road access and limited forestry. discussion as we predicted, moose showed higher use and preference for younger softwood forest and moose meadows. young coniferous forests provide optimal moose forage, and fig 2. the relationship between average % crown closure of a transect line and the number of moose seen on that line during flights over the greater highland ecosystem of nova scotia (10 surveys, n = 8674). the regression line used was fitted using spline smoothers. a b c fig. 3. expected numbers of moose in each survey block in the greater highlands ecosystem between 2001–2005 (a, est. pop. = 6883), 2006–2014 (b, est. pop. = 3563), and 2015–2020 (c, est. pop. = 3202) based on the best transect model and forest inventories in each period. the map also shows the location of cape breton highlands national park (cbhnp). alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 107 moose meadows, although over-browsed, are dominated with early successional species (schwartz and franzmann 1998, andreozzi et al. 2016). our hypothesis that moose would avoid areas closer to roads was also supported. roads pose a direct mortality risk to moose due to vehicle collisions and indirectly through habitat loss and fragmentation (forman and alexander 1998, beazley et al. 2004, eldegard et al. 2012). they also increase hunter access, which in turn leads to higher moose mortality (forman and alexander 1998, beazley et al. 2004). one factor we could not fully account for was how moose population changes might have affected habitat use. moose numbers in the ghe changed dramatically (> 4-fold) during the study from a high of 8000 animals to a low of 1350 (bridgland et al. 2007, smith et al. 2015, nova scotia lands and forestry 2020a). we accounted for the annual variation in habitat use by using a random intercept to account for overall population differences among years, but other subtle and unmodelled differences may have influenced habitat use such as resource (forage) competition. animals adjust foraging behavior and consumption to maximize caloric intake while minimizing energetic costs (emlen 1966), which moose often accomplish by foraging on the most abundant vegetation rather than selectively feeding on the most nutritious plants (wam and hjeljord 2010, bjornerass et al. 2012). overall, this pattern is mediated by resource availability and competition (wam and hjeljord 2010). our decision to use a single sightability correction factor each year may have affected the relative accuracy of our surveys because sightability often varies widely among habitat types (anderson and lindzey 1996, quayle et al. 2001, mcintosh et al. 2007). by not accounting for this variation, we likely reduced the accuracy of certain abundance estimates (quayle et al. 2001). however, since this method was used to calculate the historical abundance estimates and we did not have sightability information for each habitat type, we adopted this approach to compare the two data sets and believe our estimates are reasonable. additionally, it is likely that the crown closure values used in our analysis were not entirely representative of the field conditions during the surveys. crown closure values came from forest inventories based on conditions during the growing season, whereas our surveys occurred during winter. thus, areas with higher proportions of deciduous forest likely had inflated crown closure values due to the lack of leaves during the winter surveys. while only 10% of our site was dominated by deciduous trees (table 1), the results in certain blocks were presumably biased to a degree. we found that moose abundance increased more rapidly in response to increased distance to unpaved rather than paved roads. this suggests that hunter access, not vehicle collision, was the more import fig. 4. comparing the historical moose population estimates for the greater highland ecosystem and estimates based on the habitat use model. model estimates were extrapolated to the entire survey area and both sets of estimates have had the same set of sightability correction factors applied to them. results for 2013 were not used to calculate the r2 due to there not being an historical abundance estimate for this year. winter habitat use in nova scotia – airst and power alces vol. 57, 2021 108 risk factor driving moose avoidance of roads. other studies have shown that moose avoid areas near roads year-round (forman and alexander 1998, dussault et al. 2007, van beest et al. 2012). however, on cape breton this pattern might simply reflect the relative location of roads, as unpaved roads are mostly inland where we expect most moose activity to occur, whereas paved roads are largely near the coast where fewer moose are expected (fig. 1). further, the timing of the winter survey relative to the autumn hunting season may have some influence on winter abundance and location of moose, as animals in easily accessible areas may be harvested at higher rate. because moose move little during winter (wattles and destefano 2013), their ability to disperse and recolonize areas is limited until the following spring. roads also have a cumulative effect on moose populations. beazley et al. (2004) suggested that areas with road density >0.6 km/km2 in mainland nova scotia were incapable of supporting significant moose populations. we found that the average road density was 0.71 km/km2 in the ghe south of cbhnp and most of this area supported low numbers of moose. the same pattern was observed in the moose management area directly south of the ghe, an area with even higher road density and fewer moose (nova scotia lands and forestry 2020a). if managers seek to increase the number of moose south of the cbhnp, one course of action may be to decommission certain roads in this area, an action proposed in other parts of nova scotia to help local moose populations (beazley et al. 2004). this action would be in accordance with nova scotia’s endangered species act (government of nova scotia 1998) as several endangered species live within the area. if full decommissioning of roads is not feasible, managers could regulate access into areas of the ghe by using gates and signage during the hunting season. this approach has been successful in maintaining and growing wildlife populations in other jurisdiction with high human access (cole et al. 1997, crichton et al. 2004). our results provide further support for strategic planning and placement of new roads on the landscape. our predictive maps in the ghe show that moose generally used the same areas throughout the study, although their distribution was expected to contract given the measurable population decline. this same pattern was observed during the corresponding aerial surveys (bridgland et al. 2007, smith et al. 2015, nova scotia lands and forestry 2020a), suggesting that factors beyond habitat covariates were associated with the decline in the southern population; for example, the relative abundance of white-tailed deer (odocoileus virginianus). when deer and moose are sympatric, their general abundance pattern is an inverse relationship in that as deer numbers increase, moose abundance declines (snaith et al. 2002), a relationship believed to explain, in part, moose decline elsewhere in nova scotia (pulsifer and nette 1995, snaith et al. 2002). this effect is unlikely to be direct competition between the species, rather, the increased spread and abundance of parelaphostrongylus tenuis, or meningeal worm associated with high deer populations (anderson 1972, lankester 2010). white-tailed deer are the principal host of meningeal worm and rarely display symptoms or negative effects, whereas this parasite can be mortal to moose (lankester 2010). unfortunately, the prevalence of meningeal worm is unknown on cape breton and warrants further study. to conclude, our aerial surveys provided a useful assessment of the annual and longterm winter abundance and habitat use of moose on cape breton. we found similarity between population estimates from historical aerial surveys and habitat use models alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 109 developed from aerial survey data. our approach is applicable in many jurisdictions where long-term aerial survey data are available, but more specific gps-telemetry data from marked individuals are lacking. we encourage further development of similar approaches and use of long-term data sets from various sources to assess wildlife populations and improve understanding of how change in environmental factors affects moose and other wildlife over time. acknowledgements we would like to acknowledge that this work took place in unama’ki, the ancestral and unceded territory of the cape breton mi’kmaq. data collection was a joint effort of many groups including the nova scotia department of lands and forestry, parks canada, the unama’ki institute of natural resources, and the confederacy of mainland mi’kmaq. we would like to thank all participants for their help collecting this data. we would especially like to thank j. bridgeland and m. lemieux of parks canada for their advice early on when developing methods for interpreting the survey data. references anderson, c. r., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24: 247–259. _____, r. c. 1972. the ecological relationships of meningeal worm and native cervids in north america. journal of wildlife diseases 8: 304–310. doi: 10. 7589/0090-3558-8.4.304 andreozzi, h. a., p. j. pekins, and l. e. kantar. 2016. using aerial survey observations to identify winter habitat use of moose in northern maine. alces 52: 41–53. beazley, k. f., t. v. snaith, f. mackinnon, and c. david. 2004. road density and the potential impact on wildlife species such as american moose in mainland nova scotia. proceedings of the nova scotian institute of science 42: 339–357. doi: 10.15273/pnsis.v42i2.3610 bjørneraas, k., i. herfindal, e. j. solberg, b. e. sǣther, b. van moorter, and c. m. rolandsen. 2012. habitat quality influences population distribution, individual space use and functional response in habitat selection by a large herbivore. oecologia 168: 231–243. doi: 10.1007/ s00442-011-2072-3 boyce, m. s., p. w. baxter, and h. p. possingham. 2012. managing moose harvests by the seat of your pants. theoretical population biology 82: 340– 347. doi: 10.1016/j.tpb.2012.03.002 bridgland, j., t. nette, c. dennis, and d. quann. 2007. moose on cape breton island, nova scotia: 20th century demographics and emerging issues in the 21st century. alces 43: 111–121. cole, e. k., m. d. pope, and r. g. anthony. 1997. effects of road management on movement and survival of roosevelt elk. journal of wildlife management 61: 1115–1126. doi: 10.2307/3802109 crichton, v., t. barker, and d. schindler. 2004. response of a wintering moose population to access management and no hunting – a manitoba experiment. alces 40: 87–94. davis, d., and s. browne. 1996. the natural history of nova scotia. nova scotia museum, halifax, nova scotia, canada. dussault, c., j. p. ouellet, c. laurian, r. courtois, m. poulin, and l. breton. 2007. moose movement rates along highways and crossing probability models. journal of wildlife management 71: 2338–2345. doi: 10.2193/2006-499 eldegard, k., j. t. lyngved, and o. hjeljord. 2012. coping in a human-dominated landscape: trade-off between foraging and keeping away from roads by moose (alces winter habitat use in nova scotia – airst and power alces vol. 57, 2021 110 alces). european journal of wildlife research 58: 969–979. doi: 10.1007/ s10344-012-0640-4 emlen, j. m. 1966. the role of time and energy in food preference. american naturalist 100: 611–617. doi: 10.1086/282455 environment canada. 2020. grand etang nova scotia historic weather data. government of canada. (accessed december 2020). environmental systems research institute. 2018. arcgis. version 10.5.1. redlands, california, usa. forman, r. t., and l. e. alexander. 1998. roads and their major ecological effects. annual review of ecology, evolution, and systematics 29: 207–231. doi: 10.1146/annurev.ecolsys.29.1.207 fryxell, j. m., a. r. sinclair, and g. caughley. 2014. wildlife ecology, conservation, and management. john wiley and sons, new york, new york, usa. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. institute of arctic biology, university of alaska, fairbanks, alaska, usa. government of nova scotia. 1998. endangered species act. ratified 3 december 1998. (accessed december 2020). hebblewhite, m., and e. merrill. 2008. modelling wildlife–human relationships for social species with mixed-effects resource selection models. journal of applied ecology 45: 834–844. doi: 10.1111/j.1365-2664.2008.01466.x kantar, l. e., and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose abundance in maine. alces 49: 29–37. kneib, t., t. hothorn, and g. tutz. 2009. variable selection and model choice in geoaddative regression models. biometrics 65: 626–634. doi: 10.1111/ j.1541-0420. 2008.01112.x krausman, p. r. 1999. some basic principles of habitat use. idaho forest, wildlife and range experimental station bulletin 70: 85–90. lankester, m. w. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53–70. maclean, d. a., and d. p. ostaff. 1989. patterns of balsam fir mortality caused by an uncontrolled spruce budworm outbreak. canadian journal of forest research 19: 1087–1095. doi: 10.1139/x89-165 manly, b. f. j., l. l. mcdonald, d. l. thomas, t. l. mcdonald, and w. p. erickson. 2002. resource selection by animals: statistical analysis and design for field studies. second edition. kluwer academic publishers, boston, massachusetts, usa. mcintosh, t. e., r. c. rosatte, j. hamr, and d. l. murray. 2007. development of a sightability model for low-density elk populations in ontario, canada. journal of wildlife management 73: 580–585. doi: 10.2193/2007-550 mclaren, b. e., b. a. roberts, n. djan-chékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 45–59. nova scotia lands and forestry. 1999. forest inventory – second round. forestry division, truro, nova scotia, canada. ______. 2009. forest inventory – third round. forestry division, truro, nova scotia, canada. ______. 2020a. winter 2020 cape breton moose survey. wildlife division, kentville, nova scotia, canada. ______. 2020b. forest inventory – third round aged forward. forestry division, truro, nova scotia, canada. pagano, a. m., and t. w. arnold. 2009. detection probabilities for ground-based breeding waterfowl surveys. journal of wildlife management 73: 392–339. doi: 10.2193/2007-411 https://climate.weather.gc.ca https://climate.weather.gc.ca https://nslegislature.ca/legc/bills/57th_1st/3rd_read/b065.htm https://nslegislature.ca/legc/bills/57th_1st/3rd_read/b065.htm https://nslegislature.ca/legc/bills/57th_1st/3rd_read/b065.htm alces vol. 57, 2021 airst and power. – winter habitat use in nova scotia 111 pulsifer, m. d., and t. l. nette. 1995. history status and present distribution of moose in nova scotia. alces 31: 209–219. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–55. r core team. 2019. r: a language and environment for statistical computing. version 3.6.1. r foundation for statistical computing, vienna, austria. rempel, r. s., d. kaukinen, and a. p. carr. 2016. patch analyst and patch grid 5.2. center for northern forest ecosystem research, ontario ministry of natural resources, thunder bay, ontario, canada. schwartz, c. c., and a. w. franzmann. 1998. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. smith, c., k. f. beazley, p. duinker, and k. a. harper. 2010. the impact of moose (alces alces andersoni) on forest regeneration following a severe spruce budworm outbreak in the cape breton highlands, nova scotia, canada. alces 46: 135–150. smith, r. m. smith, p. paul, and c. bellemore. 2015. hyperabundant moose management plan for north mountain, cape breton highlands national park. parks canada. snaith, t. v., k. f. beazley, f. mackinnon, and p. duinker. 2002. preliminary habitat suitability analysis for moose in mainland nova scotia, canada. alces 38: 73–88. van beest, f. m., b. van moorter, and j. m. milner. 2012. temperature-mediated habitat use and selection by a heat sensitive northern ungulate. animal behaviour 84: 723–735. doi: 10.1016/j. anbehav.2012.06.032 wam, h. k., and o. hjeljord. 2010. moose summer and winter diets along a large scale gradient of forage availability in southern norway. european journal of wildlife research 56: 745–755. doi: 10.1007/s10344-010-0370-4 wattles, d. w., and s. destefano. 2013. space use and movements of moose in massachusetts: implications for conservation of large mammals in a fragmented environment. alces 49: 65–81. williams, b. k. 2011. adaptive management of natural resources – framework and issues. journal of environmental management 92: 1346–1353. doi: 10. 1016/j.jenvman.2010.10.041 wood, s. 2017. generalized additive models: an introduction with r. second edition. chapman and hall/crc press, london, united kingdom. _____, and scheipl, f. 2020. generalized additive mixed models using “mgcv” and “lme4.” version 0.2–6. (accessed december 2020). https://cran.r-project.org/web/packages/gamm4/gamm4.pdf https://cran.r-project.org/web/packages/gamm4/gamm4.pdf https://cran.r-project.org/web/packages/gamm4/gamm4.pdf 4208(49-54).pdf alces vol. 42, 2006 keigley and fager habitat-based adaptive management 49 habitat-based adaptive management at mount haggin wildlife management area richard b. keigley1 and craig w. fager2,3 1united states geological survey, 632 coulee drive, bozeman, mt 59718, usa; 2montana fish wildlife and parks, 1820 meadowlark lane, butte, mt 59701, usa abstract: the 22,743-hectare mount haggin wildlife management area was purchased in 1976, in part for moose (alces alces) winter range. observed moose populations climbed from a low of 7 in 1976 to a high of 56 in 2000. a 4-step management program was initiated in 2000 consisting of being primarily limited by local environmental conditions. a survey of geyer willow (salix geyeriana) in critical moose habitat indicated that browse plants were 100% intensely browsed, suggesting that browsing could prevent willow height growth. beginning in 2000, willow trend was monitored annually at 4 sites using an index based on the height of the tallest live stem and the height of the tallest, dead intensely browsed stem (ld index). low ld index values indicated that browsing did prevent height growth. in 2000 moose harvest quotas were increased by 40%; in 2002 harvest quotas were increased an additional 7%. from 2000 to 2002, willow growth increased at all 4 locations. from 2002 to 2004, growth indicators changed relatively little at sullivan creek, deep creek, and french creek; at these sites willow condition in 2004 had improved compared to willow condition in 2000. from 2002 to 2004, growth indicators declined markedly at american creek; in 2004, growth indicators at american creek were lower compared to measurements made in 2000. the improvement of willow condition at 3 sites was likely due to a combination of reduced moose numbers (due to an increase in harvest) and increased dispersal (due to low snow-cover conditions). over the study period, the sporting public complained of reduced moose sightability; harvest quotas were lowered substantially in 2003. alces vol. 42: 49-54 (2006) key words: adaptive management, browse, geyer willow, habitat, monitoring, moose the 22,743 ha mt. haggin wildlife management area (wma) in southwestern montana was purchased in 1976 and is managed by montana fish, wildlife & parks (fwp) for wildlife, public recreation, as well as controlled livestock grazing (newell and ellis 1982, frisina 1992). prior to 1976 the area was privately owned and heavily exploited for timber and minerals and by season-long livestock grazing. east of the continental divide moose (alces alces) are the only yearround, resident ungulate species on the wma as deep snow forces other species to migrate to lower elevations. prior to public ownership the area supported limited numbers of moose due to livestock grazing impacts and systematic attempts to reduce or eliminate moose habitat (willows) to promote livestock forage. observed moose populations climbed from a low of 7 in 1976 to a high of 56 in 2000 during the winter, moose congregate in broad riparian areas in which grow a variety of willows including geyer willow (salix geye3present address: montana fish wildlife and parks, 730 north montana, dillon, mt 59725, usa habitat-based adaptive management – keigley and fager alces vol. 42, 2006 50 riana), booth willow (s. boothii), drummond willow (s. drumondiana), planeleaf willow (s. planifolia), and wolf willow ( ) (fig. 1). a study initiated in 1997 indicated that browsing pressure on willow had increased. short (young) heavily-browsed plants grew stems of the tall plants had not been browsed. the age-related difference in growth forms indicated that browsing pressure must have been lower in the past as the older shrubs grew through the browse zone. based on dendrochronologic evidence, browsing pressure was found to have increased at about 1985 from a light-to-moderate level to an intense level; over the period 1976 – 1985, moose increased from a count of 7 to a count of 23 (keigley et al. 2003). in 1985 the number of moose counted was approximately half the number counted in 2000 (56). at mt. haggin wma, moose depend on the availability of browse that has grown above snow-cover. intense browsing can inhibit height growth, thus reducing the amount of forage available in the winter (keigley and frisina 1998). the management of a viable winter range will require regulating the use of browse to a level that sustains moose while at the same time protects the habitat on which the moose depend. this paper describes a method for the habitat-based management of moose on the mt. haggin wma. methods the management process was partitioned has been attained; (3) if necessary, developing and (4) if necessary, implementing the management strategy. the process must be repeated at regular intervals, with the management strategy being adapted to attain or maintain management objective given that moose require available forage focuses on the availability of browse at diverse heights above snow-cover. many factors ingrow, including climate, recent weather, and browsing pressure. among these, browsing is the factor that can be regulated by the land manager. for that reason, adaptive managebrowsing will not prevent young plants from attaining their potential stature, their growth being primarily limited by local environmental conditions. the monitoring program was being attained. monitoring geyer willow was selected as an indicator species. the effect of browsing on geyer willow was determined in 2 steps. first, we determined if browsing was a potential factor to determine the number that were intensely browsed if at some point in the life of the shrub: (1) complete annual segment was dead (causing an annual segment to develop from a segment that elongated prior to the previous year); and (2) the dead segment was browsed. fig. 1. overview of willow community in sullivan creek. alces vol. 42, 2006 keigley and fager habitat-based adaptive management 51 browsed if all annual segments developed from segments that elongated the previous year. an architecture-based survey was conducted in 2000 to classify key wintering areas into 2 categories: (1) areas where all plants were intensely browsed; and (2) areas where some plants were light-to-moderately browsed (keigley et al. 2002a, b). the architectures are described in keigley and frisina (1998). a would indicate that browsing was a potential factor. if browsing was determined to be a potential factor, the second step in the monitoring process was to determine if plants were likely to attain their potential stature. the likelihood of growth to potential stature was assessed using an index based on the height of the tallest live stem measured to the base of current-year-growth (hbcyg) and the height of the tallest intensely-browsed dead stem (hd) (ld index = hbcyg – hd; fig. 2). an ld index value of about zero indicates that ungulates are browsing plants down to the zone of mechanical protection. ld index values much greater than zero indicate that stems are growing above the zone of mechanical protection (thus the plants are growing taller), while values much less than zero indicate that plants are dying back to ground level. in 2000, monitoring sites were established in sullivan creek, deep creek, french creek, and american creek. at each site, ld index measurements were taken on 20 plants annually. plants were selected for measurement by walking a set number of paces and selecting the nearest plant that met the height criterion; plants were not marked for re-measurement (keigley et al. 2001). at each site photographs were taken down a permanently marked transect each year. trends documented by ld index data were compared to trends indicated by paired photographs of given shrubs. management strategy plants from attaining potential stature would require a reduction in browsing pressure if the this case, browsing pressure was regulated by changing the moose harvest quota. results and discussion the survey of critical willow habitat in 2000 found that 100% of geyer willow was intensely browsed in all surveyed areas (keigley et al. 2003). the widespread intense browsing indicated that browsing could potentially prevent young plants from attaining full stature. mean ld index values at the 4 monitoring sites in 2000 were, respectively: -12 ± 5 cm, -38 ± 10 cm, -9 ± 4 cm, and 2 ± 6 cm (± se, n = 20). the 3 negative values and the single value near zero (2 cm) were interpreted to indicate that browsing was preventing young plants from attaining their potential stature. based on the initial browse survey, the moose harvest quota was increased 40% in 2000 in the 2 hunting districts encompassing the wma (hunting districts 319 and 325) (fig. 3). in 2002, harvest quotas were increased an additional 7% and hunting district 341 was created from a portion of hunting district 319 to distribute moose hunting pressure away from the readily accessible portion of the wma west of the continental divide. harvest quotas were lowered to slightly below pre-study levels following the 2002 hunting fig. 2. ld index is a measure of the response of a plant after intense browsing. habitat-based adaptive management – keigley and fager alces vol. 42, 2006 52 season. percent harvest success remained above the management threshold of 80% during the period 2000-2002 (mfwp 2001, 2002, 2003). in 2003, 2 of the 3 hunting districts fell below the 80% threshold (mfwp 2004). mild fall conditions combined with fewer moose in easily accessible areas were the likely cause of the decline in hunting success. from 2000 to 2002 the ld index increased at all sites, indicating that stems had grown above the height of stems previously killed by browsing (fig. 4). from 2002 to 2004 the ld index changed relatively little at sullivan creek, deep creek, and french creek; at these sites, willow condition in 2004 had improved compared to willow condition in 2000. from fig. 4. ld index values at 4 monitoring sites over the period 2000 2004. fig. 3. mount haggin moose permits 1988 – 2004. alces vol. 42, 2006 keigley and fager habitat-based adaptive management 53 2002 to 2004 the ld index declined markedly at american creek; in 2004, the index at american creek was lower compared to measurements made in 2000. we attribute the improvement of willow condition at 3 sites to a combination of reduced moose numbers (due to an increase in harvest) and increased dispersal (due to low snow-cover conditions). repeat photography corroborated the ld index data at all sites. at sullivan creek, geyer willow shrubs photographed in 2004 had extensive new growth that was not visible in the photograph taken in 2000 (fig. 5). fig. 5. paired photographs of the same shrubs at sullivan creek taken in 2000 and 2004. stems are markedly taller in 2004 compared to 2000. this growth corresponds with an increase in ld index over that period. fig. 6. paired photographs of the same shrub at american creek taken in 2000 and 2004. stems in 2004 are about the same height as in 2000. this lack of growth corresponds with a decline in ld index over that period. habitat-based adaptive management – keigley and fager alces vol. 42, 2006 54 similar results occurred in paired photographs of shrubs in deep creek and french creek. at american creek, new growth on shrubs photographed in 2004 was not visible (fig. 6). repeat photographs provide tangible visible evidence of trends, but do not provide quantitative data amenable to statistical analysis. ld index values can be statistically compared. for example, at deep creek the mean ld index measured in 2000 (-38.3 cm) statistically differed from the mean measured in 2004 (-3.3 cm) at p = 0.0019. public education on the need for moose population reductions was initiated through local newspapers, sportsman’s organizations, and watershed groups. however, during the period of population reductions some individuals sought to lower the harvest based on fewer moose observations. concern was expressed over possibly losing a lifetime opportunity to harvest a moose; montana imposes a 7-year wait on hunters who draw a moose license, regardless of success. in response to these concerns, the harvest quota was reduced following the 2002 hunting season. current willow monitoring suggests the moose population rebounded within 2 years after the harvest quotas were reduced in 2002. wildlife managers must now decide: (1) if the try and effect change in the willow community with a public that demands a visible, readily accessible moose resource. acknowledgements fish wildlife and parks and the united states geological survey. references frisina, m. r. 1992. elk habitat use within a rest-rotation grazing system. rangelands 14: 93-96. keigley, r. b., and m. r. frisina. 1998. browse evaluation by analysis of growth form. montana fish, wildlife, and parks, helena, montana, usa. _____, _____, and c. fager. 2001. assessing browse trend at the landscape level. pages 15-26 in s. knapp and m. frisina, editors. port number. 1. montana fish wildlife and parks, helena, montana, usa. _____, _____, and _____. 2002a. assessing browse trend at the landscape level. part 1: preliminary steps and field survey. rangelands 24: 28-33. _____, _____, and _____. 2002b. assessing browse trend at the landscape level. part 2: monitoring. rangelands 24: 34-38. _____, _____, and _____. 2003. a method for determining the onset year of intense browsing. journal of range management 56: 33-38. (mfwp) montana fish, wildlife and parks. 2001. special big game hunting and harvest report. license year 2000. montana fish, wildlife and parks, helena, montana, usa. _____. 2002. special big game hunting and harvest report. license year 2001. montana fish, wildlife and parks, helena, montana, usa. _____. 2003. special big game hunting and harvest report. license year 2002. montana fish, wildlife and parks, helena, montana, usa. _____. 2004. special big game hunting and harvest report. license year 2003. montana fish, wildlife and parks, helena, montana, usa. newell, a., and d. ellis. 1982. mount haggin: living history. montana outdoors. may/june 1982: 27-31. f:\alces\vol_38\pagemaker\3810. alces vol. 38, 2002 courtois and beaumont – habitat and moose 167 a preliminary assessment on the influence of habitat composition and structure on moose density in clearcuts of north-western québec réhaume courtois and aldée beaumont société de la faune et des parcs du québec, direction de la recherche sur la faune, 675 renélévesque est, 11e étage, boîte 92, québec, pq, canada g1r 5v7; rehaume.courtois@fapaq.gouv.qc.ca abstract: aerial survey data were used to describe moose density changes in relation to habitat composition and structure in clear-cut areas, and to infer the impact of these variables on limiting factors. we hypothesized that moose density would be lower in cut areas due to increased hunting and predation. four habitat types (food and cover stands, cover stands, cuts, and other habitats) and 7 fragmentation indices were used in our analyses. aerial surveys conducted in seven 35-112km2 blocks showed that moose density was related to the proportion of deciduous and mixed (food and cover) stands within each block and edge between food and cover and resinous stands (cover). density, productivity, and harvest rate were not significantly influenced by clear-cuts. our results suggest that habitat models should consider food and food-cover border over other habitat components. alces vol. 38: 167-176 (2002) key words: alces alces, cover, edge, food, habitat management, mortality, productivity résumé: des inventaires aériens ont été utilisés pour décrire les changements de densité de l’orignal (alces alces) en fonction de la composition et de la structure des habitats dans des sites comportant des coupes forestières, et pour inférer l’impact de ces variables sur les facteurs limitatifs. nous avons émis l’hypothèse que la densité serait plus basse dans les aires coupées à cause d’un accroissement de la chasse et de la prédation. quatre catégories d’habitat (nourriture-abri, abri, coupe, autres habitats) ont été retenues et 7 indices de fragmentation ont été utilisés dans les analyses. les inventaires aériens réalisés dans 7 blocks de 35 à 112 km2 ont montré que la densité était reliée à la proportion de peuplements décidus et mélangés (nourriture-abri) à l’intérieur de chaque bloc et à l’effet de bordure entre les peuplements de nourriture-abri et ceux d’abri. la densité, la productivité et le taux d’exploitation ne semblaient pas influencés par la coupe. nos résultats suggèrent que la nourriture et l’effet de bordure entre la nourriture et l’abri devraient être considérés de façon prioritaire dans les modèles d’habitat. alces vol. 38: 167-176 (2002) mots clés: abri, alces alces, bordure, couvert, gestion de l’habitat, mortalité, nourriture, productivité over the long term (15-40 years), forest harvesting has a positive impact on moose by rejuvenating the forest (crête 1977). however, the introduction of mechanized clear-cutting techniques in the early 1970s has provoked controversy and confrontations with wildlife users and outdoor recreationists. at that time, the only guideline aimed at protecting habitat was the maintenance of forest strips, for the preservation of water quality, along main watercourses. cutovers were juxtaposed, creating large areas that were unsuitable for moose (girard and joyal 1984). moreover, no particular techniques were used to protect advanced regeneration. often, cutovers habitat and moose – courtois and beaumont alces vol. 38, 2002 168 did not regenerate and had to be scarified and planted 5 to 10 years later, diminishing browse and cover quality for a longer period. previous studies have reported very low moose densities in these habitats (eason et al. 1981, girard and joyal 1984, eason 1989), a situation also frequently cited and criticized by moose hunters. in québec, more restrictive forest harvesting guidelines were implemented in the late 1980s. in order to evaluate short term advantages of these regulation changes, moose densities were estimated by aerial survey over a period of 5 years in 7 study blocks with different types of clear-cuts. because forest harvesting created new roads and removed cover which could facilitate access to hunters and predators (eason et al. 1981, girard and joyal 1984, eason 1989, seip and cichowski 1996), we hypothesized that moose density would be lower in cutovers due to increased hunting and predation. consequently, in cut areas, we predicted that moose would experience a higher hunting rate, lower productivity, and lower density due to increased predation on adults and calves, and increased hunting vulnerability. the impact of forest harvesting on other species and on moose hunting has been previously reported (courtois et al. 1998a, 2001; dussault et al. 1998; potvin 1998; courtois and beaumont 1999; potvin et al. 1999; turcotte et al. 2000). study area the study was conducted in a 2,183 km2 area located in north-western québec. the dominant tree species of the area are black spruce (picea mariana), jack pine (pinus b a n k s i a n a ) , p a p e r b i r c h ( b e t u l a papyrifera), and trembling aspen (populus tremuloides). the terrain is gently rolling with hills rarely exceeding 350 m above sea level. temperature, as measured at the belleterre meteorological station, is -16.2°c ± 0.6 (29) in january (mean ± se [n years]) and 17.3°c ± 0.2 (29) in july. total precipitation is moderate with 1,013 mm ± 23 (24) including 291 mm ± 12 (26) of snow. maximum snow depth (66 cm ± 7 [17]) occurs in february and did not exceed 90 cm during the study. wolf (canis lupus) and black bear (ursus americanus) are found in the study area at ∼0.01 and ~0.14 individuals / km2 respectively (lamontagne et al. 1999, larivière et al. 2000). the area was divided into 25 blocks delineated using identifiable landmarks, mostly streams, roads, and forest harvesting plans, so as to contain different proportions of clear-cuts (see courtois and beaumont 1999 for a map). aerial surveys were conducted in 7 of these blocks (fig. 1). lakes and streams represented between 8 and 17% of the area depending on the study block, whereas cuts occupied between 4 and 68%. two main types of cutting methods were employed: large clear cuts without protected regeneration (ct) that were 7-11 years old at the time of the study, and recent cuttings with protected regeneration (cpr) that were mostly made between 1992 and 1994. characteristics of these two types of cuts have been previously described (courtois et al. 1998b). at the end of the project, block 5 (112 km2) was still dominated by undisturbed mature conifer stands (4% cut). blocks 3 (96 km2; cut during the summer and fall of 1992), 11 (35 km2; cut in 1989), 13 and 20 (74 and 59 km2, respectively; cut in the winters 1992 to 1994) were covered by large recent cuts on 29%, 45%, 46%, and 43% of their areas, respectively, but residual forests were relatively abundant and well dispersed within the blocks. in blocks 16 and 19, dominated by 7to 11-year-old clear-cuts (68 and 62 km2; 50% and 69% cut, respectively), remnant forest was small and restricted to the fringe of the blocks. alces vol. 38, 2002 courtois and beaumont – habitat and moose 169 methods aerial surveys moose density and population structure (adult males, adult females, and calves) were estimated by aerial surveys 3 to 5 times in each of the 7 blocks between 1991 and 1995 using standard methods (crête and goudreault 1980, courtois 1991). surveys were conducted between late january and mid-february. adjacent blocks were surveyed simultaneously to avoid counting the same animals in 2 blocks. as visibility bias likely varied according to habitat type, moose densities were corrected using yearand block-specific visibility rates estimated with marked animals. to remove the influence of hunting on annual density, winter densities were converted to fall densities by adding harvest density recorded at mandatory registration stations during the fall that preceded each survey. harvest rates (percent) were computed as the fall harvest density × 100 / fall moose density. in spite of the relatively small size of our study blocks, trends are likely to be correctly estimated since each block was surveyed 25 times over a 5-year period, and moose are resident and have relatively small home ranges. habitat composition and structure habitat composition was obtained from digitized 1:20,000 forest maps produced by interpretation of 1:15,000 aerial photographs (mer 1984). this information was used to produce a 10 m × 10 m raster map imported into arcview 3.1 and managed with the spatial analyst extension (esri 1996). habitat types were grouped into 4 classes according to food and cover quality for moose (food&cover: deciduous, mixed, and spruce budworm outbreak stands; cover: fig. 1. location of the study site in north-western québec (78° 40’ w, 47° 50’ n). numbers refer to blocks delineated for aerial surveys. habitat and moose – courtois and beaumont alces vol. 38, 2002 170 conifer; cuts: cut with and without protection of advanced regeneration; other: unproductive stands, usually alnus rugosa, and water). based on vegetation surveys conducted in the study area (courtois et al. 1998b, dussault et al. 1998, turcotte et al. 2000) and in comparable sites (dussault 2001), food&cover stands roughly correspond to those comprising > 10,000 deciduous stems / ha with at least 1 twig of browse per tree between 0.5 and 3.0 m with conifer trees occupying < 3.5 m2 / ha. cover stands represent those with scarce understory deciduous species (< 2,000 stems / ha) and where the basal area of conifer trees occupy >10 m2 / ha. recently cut sites usually had < 5,000 understory deciduous stems / ha with conifer trees occupying < 0.2 m2 / ha. old cuts and those recently done in mixed stands of block 20 supported dense deciduous browse (11,000-18,000 stems / ha) but very few conifer trees (< 0.2 m2 / ha). edge, food, interspersion, and the size and form of cuts are thought to influence moose distribution and are intensively used to define moose habitat guidelines (omnr 1988, mer 1989, rempel et al. 1997). the habitat structure was estimated with 7 f r a g s t a t s l a n d s c a p e p a t t e r n i n d i c e s (mcgarigal and marks 1995) computed with the arcview extension patch analyst (elkie et al. 1999). we selected 2 indices to measure edge (the length of the interface) between habitat types (ed: edge density, perimeter of all habitat patches per unit area [m / ha]; cwed: contrast-weighted edge density, edge between food&cover and cover stands [m / ha; weights: food&cover versus cover = 1; other edges = 0]). one metric was selected to estimate the importance of food&cover stands (cad: number of food&cover stand core areas [inside part beyond 100 m] / 100 ha). two indices measured the diversity of the landscape (sdi: shannon’s diversity index [mcgarigal and marks 1995], the amount of information per habitat patch [without units]; iji: interspersion and juxtaposition index, the extent to which patch types are equally adjacent to each other [%]). two others were selected to quantify the shape and size of cutovers (msi: mean shape index, average perimeterto-area ratio of the cuts [without units]; tcai: total core area index, relative importance of the core areas [buffer: 100 m] within the cuts [%]). data analysis the influence of habitat composition and structure on moose density and population structure among blocks and year were assessed using stepwise multiple regression analysis (proc stepwise, sas 1989). year, habitat composition, and landscape indices were considered independent variables. each survey was considered independent due to annual habitat changes, the elapsed time (12 months) between two surveys, movements of animals (200-300 m / d), and the impossibility of using repeated measures in proc stepwise. to prevent multicollinearity problems, pearson’s correlation analyses were used to identify redundant variables. only those variables that were not significantly correlated at p = 0.05 were retained in the stepwise regression analyses. moose density and population structure before and after cutting operations were compared in blocks 3 (coniferous) and 20 (mixed), which were harvested during the study, using t-tests and chi-square analysis, respectively. in 1994 and 1995, the same analyses were used to compare moose density and population structure between the forested part and the cut areas of these blocks. results effect of habitat composition and structure on moose density many habitat variables in the survey blocks were correlated (p < 0.01). among alces vol. 38, 2002 courtois and beaumont – habitat and moose 171 t h e 4 h a b i t a t c o m p o s i t i o n i n d i c e s , food&cover vs. cover (r = -0.51), cover vs. cuts (r = -0.89), cover vs. other habitats (r = 0.77), and cuts vs. other habitats (r = -0.82) were correlated. similarly, among landscape pattern indices, ed, cwed, and cad (r > 0.59) as well as msi and iji (r = 0.63) were correlated. cwed was also weakly correlated to tcai (r = 0.44, p = 0 . 0 2 7 ) . c o n s e q u e n t l y , o n l y b l o c k , food&cover, cuts, cwed, tcai, and iji were used in the stepwise regression analysis. among them, only food&cover and cwed were retained in the model (fall density = -0.496 + 0.100 food&cover + 0.066 cwed, r2 = 0.89, f = 90.63, p < 0.001) and both variables were highly significant (p < 0.007, r2 > 0.85). simple regression models illustrate the relationship between moose density and cwed, food&cover, and cover, and the absence of relationship between moose density and cuts (fig. 2). effects of clear-cuts on moose productivity and mortality the stepwise multiple regression analysis did not reveal any relationship between habitat composition or landscape indices and the number of calves per 100 females. on average, among blocks and years, harvest rate was 18.9% ± 2.3 before and 23.6% ± 2.2 after cutting (t = 1.23, p = 0.16). also, population structure did not differ before and after cutting. during the 5-year study, we observed a total of 35 males and 41 calves / 100 females before cutting and 37 males and 49 calves / 100 females after cutting (n = 371 moose; adult sex ratio: χ2 = 0.003, p > 0.05; productivity: χ2 = 0.437, p > 0.05). we found 45 calves / 100 females in the cutovers and 64 calves / 100 females in the remnant forest, this difference being non-significant (n = 163; χ2 = 1.160, p > 0.05). moose density being low relative to habitat carrying capacity (crête 1989), calves / 100 females were not fig. 2. influence of habitat structure and habitat composition on moose density by block, estimated by aerial surveys (n = 26 surveys), northwestern québec, january 1991 to january 1995. y = 0.0147x + 0.0023 r2 = 0.8029 0.0 0.2 0.4 0.6 0.8 0 10 20 30 40 contrast w e ighted edge density m o o s e / k m ² y = 0.0161x 0.0458 r2 = 0.848 0.0 0.2 0.4 0.6 0.8 0 20 40 60 % of the survey blocks in food stands m o o s e / k m ² y = -0.0049x + 0.4473 r2 = 0.3413 0.0 0.2 0.4 0.6 0.8 0 20 40 60 80 % of the survey blocks in cover stands m o o s e / k m ² 0.0 0.2 0.4 0.6 0.8 0 20 40 60 80 % of the survey blocks in cuts m o o s e / k m ² r² = 0.048 habitat and moose – courtois and beaumont alces vol. 38, 2002 172 correlated to moose density changes (r = 0.21, p > 0.05). discussion influence of habitat on population demography as observed by rempel et al. (1997), some cutting patterns may result in hunters having a major influence on moose populations. due to the creation of new roads and the removal of cover that potentially increase accessibility for hunters and predators, and increase moose visibility (eason et al. 1981, girard and joyal 1984, eason 1989, seip and cichowski 1996), we expected that the parameters of moose demography would be lower in the presence of clear-cuts, but the 2 predictions associated with this hypothesis were rejected. we failed to detect an influence of clear-cuts on moose mortality from hunting and on population productivity. as observed by courtois and beaumont (1999), density and harvest rate could be slightly modified by recent clear-cuts but the changes were not important enough to be significant. moose densities were relatively low in our study site (annual mean: 0.13-0.61 moose / km2 in fall depending on the survey block), hunters were quite evenly distributed, and hunting pressure (2-11 hunting-days / km2) and hunting rate (19-24%) were high but not related to the importance of clear-cuts (courtois and beaumont 1999). the study site has been hunted for a long time and moose hunters often access their hunting sites by boats and all-terrain vehicles with little consideration of forest roads (courtois and beaumont 1999). the impact of clearcuts on density and harvest rate would probably have been more important in an area experiencing a low hunting pressure before cutting. moreover, telemetry data suggest that moose preferred the forested part of their home range (courtois et al. 1998a ). this behavioural adaptation may have helped give access to habitats that minimized mortality risks. although our work should be seen as a preliminary assessment due to the relatively small size of our study blocks (35-112 km2), the short-term influence of forest cutting on moose seemed marginal. variations in moose density were not related to the importance of clear-cuts. after cutting, densities decreased by 23-30% in blocks 3 and 20 but these changes were likely not only related to clear-cuts, since a greater interannual variability was observed in block 5 where no cutting was carried out during the study (courtois and beaumont 1999). relatively few moose (2-28) were present in each survey block. movements of a few animals in and out of a survey block would suffice to explain annual density changes. based on aerial surveys, a significant dependence on edge was noted, but this relationship was partly due to the correlation between edge indices and food&cover. food&cover and cover stands were generally small (food&cover: 11.6 ± 0.3 ha, [n = 7,837]; cover: 11.6 ± 0.4, [7,685]; cuts: 46.5 ± 8.0, [1,405]; other: 14.0 ± 4.4, [5,776]), and edge was automatically high in areas supporting an important proportion of food&cover or cover stands. dalton (1989) reported higher productivity in a site with 30% of its area cut, compared to another where cuts covered 50% of the area. girard and joyal (1984) also observed a lower productivity in cutovers. in our study, productivity varied from year to year and among study blocks, but as in eason (1989), without any relation to the proportion of cuts, for up to 69% of area cut. the productivity we measured was intermediate between estimates made immediately south of the study site (40-41 calves / 100 females) and within the hunting zone where our study was conducted (5862 calves / 100 females; paré and courtois 1990; paré 1991, 1996). wolf and black alces vol. 38, 2002 courtois and beaumont – habitat and moose 173 bear were probably responsible for the relatively low calf recruitment (crête and jolicoeur 1987) but we did not detect any effects of clear-cuts on that variable. management implications the juxtaposition of several clear-cuts in order to facilitate logging operations and minimize costs of road construction has produced 15-250-km2 landscapes dominated by clear-cuts (f. potvin and r. courtois, unpublished data). at a scale of about 100 km2, this has a minor influence on moose, who can continue to use parts of their home range that remain forested. however, moose do not intensively use the cut part of the landscape which leads to the observation of a decline in moose numbers by hunters. because they hunt in small hunting sites and they try to be isolated from each other, hunters would benefit from forest harvesting guidelines that favour the maintenance of moose in cut areas. two alternatives have been suggested to maintain moose in cut landscapes (courtois et al. 1998b): (1) maintain abundant deciduous browse (> 10,000-15,000 stems / ha) and adequate cover (shrub layer > 2-3-m high; lateral obstruction > 50% at 15 m); or (2) distribute 50-100 ha cuttings over the landscape while keeping about 50% of the area uncut. however, each wildlife species has its own requirements and each needs specific habitat management guidelines. dussault et al. (1998) proposed keeping 50% of the basal area of mixed and deciduous stands in order to protect ruffed grouse (bonasa umbellus). f o r s p r u c e g r o u s e ( f a l c i p e n n i s canadensis), turcotte et al. (2000) suggested two non-contiguous 25-ha cuts per 200-ha block and per 25-30 years. for marten (martes americana), potvin (1998) proposed cuts of 50-150 ha at 20-30-year intervals in blocks of 10 km2 while maintaining 50% of the landscape in residual forests > 7 m. multi-scale planning which takes several species and several land uses into account, including the needs of the forest industry, native people and outdoor recreationists (courtois and beaumont 1999, hénault et al. 1999, potvin et al.1999), is essential in order to transform forest harvesting into a sustainable development activity. potvin et al. (1999) suggested a 3scale approach. first, at a regional scale (> 10,000 km2), the objective should be the protection of biodiversity through the maintenance of interconnected protected areas (species and rare ecosystems approach). second, at a forest landscape level (1,0005,000 km2), the objective should be the maintenance of a dynamic mosaic of forest stands in terms of composition, age, and spatial configuration (ecosystem management) through adequate planning of forest harvest, appropriate repartition of cuts in time and space, and judicious use of diversified harvest techniques (bergeron et al. 1999). finally, a local scale (50-500 km2) would take into account the specific needs of the general public (integrated forest management). guidelines previously suggested for grouse, marten, and moose are examples of management at a local scale that would favour not only the needs of the given targeted species but also satisfy the public through the maintenance of wildlife harvesting. as suggest by dussault (2001), our results suggest that food and edge between food and cover could be the most important criteria to explain moose density changes between areas. in such a case, these variables should be considered a priority in moose habitat models. acknowledgements we would like to thank michel crête and françois potvin from société de la faune et des parcs du québec (fapaq), habitat and moose – courtois and beaumont alces vol. 38, 2002 174 jean-pierre ouellet and luc sirois from université du québec à rimouski (uqar), christian dussault from université laval, and two anonymous referees for providing helpful comments on earlier versions of this paper. alain caron of uqar assisted in the preparation of habitat maps. we also thank luc bélisle, claude brassard, nicole blanchette, nancy delahaye, andré gaudreau, jean-pierre hamel, louis jourdain, alain lachapelle, marcel paré, and mario poirier from fapaq for assistance in fieldwork. references bergeron, y, b. harvey, a. leduc, and s. gauthier. 1999. stratégie d’aménagement forestier qui s’inspire de la dynamique des perturbations naturelles: considérations à l’échelle du peuplement et de la forêt. forestry chronicle 75:5561. courtois, r. 1991. résultats du premier plan quinquennal d’inventaires aériens de l’orignal au québec, 1987-1991. ministère du loisir, de la chasse et de la pêche du québec, direction de la gestion des espèces et des habitats, report 1921. , and a. beaumont. 1999. the influence of accessibility on moose hunting in northwestern québec. alces 35:41-50. , , l. b r e t o n , a n d c . dussault. 1998a. réaction de l’orignal et des chasseurs d’orignaux face aux c o u p e s f o r e s t i è r e s . m i n i s t è r e d e l’environnement et de la faune du québec, service de la faune terrestre, report 3875. , j.-p. ouellet, and a. bugnet. 2001. moose hunters’ perceptions of forest harvesting. alces 37:19-33. , , and b. gagné. 1998b. characteristics of cutovers used by moose (alces alces) in early winter. alces 34:201-211. crête, m. 1977. importance de la coupe forestière sur l’habitat hivernal de l’orignal dans le sud-ouest du québec. canadian journal of forest research 7:241-257. . 1989. approximation of k carrying capacity for moose in eastern québec. canadian journal of zoology 67:373-380. , and f. goudreault. 1980. les bois, la tache vulvaire et la couleur du museau pour déterminer le sexe des orignaux (alces alces americana) en janvier dans le sud-ouest du québec. proceedings of the north american moose conference and workshop 16:275-288. , and h. jolicoeur. 1987. impact of wolf and black bear removing on cow: calf ratio and moose density in southwestern québec. alces 23:61-87. dalton, w.j. 1989. use by moose (alces alces) of clear-cut habitat where 100% or 50% of the production forest was logged. ontario ministry of natural resources, report cofdra 32001. du s s a u l t , c. 2001. influence des contraintes environnementales sur la sélection de l’habitat par l’orignal (alces alces). ph.d. thesis, université laval, québec, québec, canada. , r. courtois, and j. ferron. 1998. impact à court terme d’une coupe avec protection de la régénération sur la gélinotte huppée (bonasa umbellus) en forêt boréale. canadian journal of forest research 28:468-477. eason, g. 1989. moose response to hunting and 1-km2 block cutting. alces 25:6374. , e. thomas, r. jerrard, and k. oswald. 1981. moose hunting closure in a recently logged area. alces 17:111125. elkie, p.r., r.s. rempel, and a.p. carr. alces vol. 38, 2002 courtois and beaumont – habitat and moose 175 1999. patch analyst user’s manual. ontario ministry of natural resources, northwest science and technology, publication tm-002. (esri) environmental systems research i n s t i t u t e , i n c o r p o r a t e d . 1 9 9 6 . arcview gis. the geographic system for everyone. environmental systems r e s e a r c h i n s t i t u t e i n c o r p o r a t e d , redlands, california, usa. girard, f., and r. j oyal. 1984. l’impact des coupes à blanc mécanisées sur l’orignal dans le nord-ouest du québec. alces 20:40-53. hénault, m., l. bélanger, a.r. rodgers, g. redmond, k. morris., f. potvin, r. courtois, s. morel, and m. mongeon. 1999. moose and forest ecosystem management: the biggest beast but not the best. alces 53:213-225. lamontagne, g., h. jolicoeur, and r. lafond. 1999. plan de gestion de l’ours noir, 1998-2002. société de la faune et des parcs du québec, report 3960. larivière, s., h. j olicoeur, and m. crête. 2000. status and conservation of the gray wolf (canis lupus) in wildlife reserves of québec. biological conservation 94:143-151. mcgarigal, k., and b.j. marks. 1995. fragstats. spatial pattern analysis program for quantifying landscape structure. oregon state university, corvalis, oregon, usa. (mer) ministère de l’energie et des ressources. 1984. normes d’inventaire forestier. ministère de l’énergie et des ressources, service de l’inventaire forestier, québec, québec, canada. . 1989. modalités d’intervention en milieu forestier. ministère de l’énergie et des ressources, québec, québec, canada. (omnr) ontario ministry of natural resources. 1988. timber management guidelines for the provision of moose habitat. ontario ministry of natur a l r e s o u r c e s , t o r o n t o , o n t a r i o , canada. paré, m. 1991. inventaire aérien de l’orignal dans la zone de chasse 16 en janvier 1990. ministère de l’environnement et de la faune du québec, direction régionale de l’abitibi-témiscamingue, report 1928. . 1996. inventaire aérien de l’orignal dans la zone de chasse 13 à l’hiver 1994. pages 13-18 in s. st-onge, r. courtois, and d. banville, editors. rapport annuel des inventaires aériens de l’orignal à l’hiver 1994. ministère de l’environnement et de la faune, report 3291. , and r. courtois. 1990. inventaire aérien de l’orignal dans la zone de chasse 12 en janvier 1988 et dans la zone de chasse 13 en janvier 1989. ministère de l’environnement et de la faune, direction régionale de l’abitibitémiscamingue, and direction de la gestion des espèces et des habitats, report 1764. potvin, f. 1998. la martre d’amérique (martes americana) et la coupe à blanc e n f o r ê t b o r é a l e : u n e a p p r o c h e télémétrique et géomatique. ph.d. thesis, université laval, québec, québec, canada. , r. courtois, and l. bélanger. 1999. short-term response of wildlife to clear-cutting in québec boreal forest: multiscale effects and management implication. canadian journal of forest research 29:1120-1127. rempel, r.s., p.c. elkie, a.r. rodgers, and m.j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. (sas) sas institute inc. 1989. sas/ stat user’s guide. sas institute inhabitat and moose – courtois and beaumont alces vol. 38, 2002 176 corporated, cary, north carolina, usa. seip, d.r., and d.b. cichowski. 1996. population ecology of caribou in british columbia. rangifer special issue 9:7380. turcotte, f., r. courtois, r. couture, and j. ferron. 2000. impact à court terme de l’exploitation forestière sur le t é t r a s d u c a n a d a ( f a l c i p e n n i s canadensis). canadian journal of forest research 30:1-9. 3921.pdf alces vol. 43, 2007 gaillard are moose only a large deer? 1 are moose only a large deer?: some life history considerations jean-michel gaillard unité mixte de recherche n°5558 “biométrie et biologie evolutive”, université claude bernard lyon 1, bâtiment 711, 43 boulevard du 11 novembre 1918, 69622 villeurbanne cedex, france abstract: body mass generally accounts for a large part of variation in life history traits of ungulates. however, phylogeny and ecological features such as habitat or diet have been shown to cause differences in life history patterns among species of similar size. to assess the factors that shape life history traits of moose (alces alces with both traits expected from allometric equations and traits of similar-sized bovids. both kinds of analyses led to the same results. while moose calves grow as expected from the size of their mothers, they start life at only about half the expected size. moose populations have higher growth rates and shorter generation times as compared to similar-sized ungulates. females reproduce earlier and have larger litters relative to their body size. the resulting faster than expected life cycle for moose canpopulation dynamics characterized by a low and variable juvenile survival as opposed to a high and constant survival of prime-age females. high reproductive output accounts for the fast life cycle of moose populations compared to other similar-sized ungulates. i propose that the high reproductive output has evolved in response to the unpredictable environmental conditions of early successional habitats preferred by moose. the evolutionary strategy of moose appears more similar to that of a very large roe deer (capreolus capreolus) than that associated with larger deer in general. alces vol. 43: 1-11 (2007) key words: output, survival patterns, ungulates since the pioneering work by stearns (1976), the study of variation in life history traits has become a popular task among evolutionary ecologists. the analyses of variation in life history traits can be performed at two different scales. first, the variation at the literature by using comparative analyses (sensu harvey and pagel 1991). second, the existence of evolutionary trade-offs between history variation generated by differences in phenotypic quality are usually performed at level, the variation in life history traits of vertebrates is mostly accounted for by three major structuring factors. variation in body size generally accounts for more than half of the variation in most life history traits (peters 1983, calder 1984, brown and west 2000 for reviews). in mammals, for instance, it is wellestablished that large mammals live longer, reproduce later, and produce fewer offspring per year than small ones (stearns 1983, gaillard et al. 1989). however, for a given size, taxa often show marked differences in life history traits. for example, it is well known that bats outlive similar-sized rodents. thus, ecological correlates of life history traits also occur. differences in diet and differences in habitat quality have been shown to generate are moose only a large deer? gaillard alces vol. 43, 2007 2 differences in life history traits (sæther and gordon 1994 for ungulates, fisher et al. 2001 for marsupials). moose (alces alces) are the largest members of the cervidae family (from 200 to 825 kg, novak 1993). therefore, i expect that its large body size may have markedly shaped life history traits currently observed in moose populations. from comparative analyses of maternal care and demographic patterns reported in populations of moose and related ungulate species, i assessed whether moose life history can simply be accounted for by large size (i.e., moose are only large deer), or independent of their size relative to other deer (i.e., moose are different than a large deer). methods to assess whether moose are simply large deer, i performed two types of analyses on life history traits related to maternal care (birth mass and early growth rate) and population dynamics (population growth rate, generation i used allometric analyses in a three-step ships without including moose for studied life history traits among ungulate species for which published information was available. although species do not represent independent ting usual linear models without accounting for phylogenetic relationships among species. my approach was based upon: (1) the similar results obtained from analyses on raw data (as performed here) and analyses including corrections for phylogenetic dependence (such as independent contrasts, see garland et al. 1992) often reported (e.g., fisher and owens 2000); and (2) criticisms of the usefulness of phylogenetic methods such as independent contrasts (ricklefs and starck 1996, björklund 1997, price 1997), mainly based on the strong assumptions made by such methods on evolutionary changes of traits (harvey and rambaut 2001). then i used the allometric equation to obtain the predicted value of the traits for a cervid with the same size as moose. lastly, i compared predicted trait values with those reported in literature for moose populations. the second type of analyses consisted of comparing life history traits observed in moose with those observed in similar-sized bovids. to the life history traits (birth mass and early size (polytocous and monotocous species). indeed, individual offspring of polytocous ungulates that produce 2 offspring per breeding attempt might be lighter at birth than single offspring of monotocous ungulates (roff 1992). moreover, birth mass of singletons is often higher than birth mass of twins in polytocous species (e.g., moose, schwartz and hundertmark 1993). i found data for 38 (birth mass) and 22 (growth rate) monotocous ungulates and for 8 (birth mass) and 6 (growth rate) polytocous ungulates. to assess demographic patterns of moose as well as of other ungulate populations, i (i.e., the mean age of mothers at the time of birth, tb, in a given population, leslie 1966) and population growth rate, r (i.e., the malthusian parameter, fisher 1930) from demographic data collected from the literature. to do that, i considered the following female of ungulate populations: the juvenile survival from birth to 1 year of age, the yearling survival between 1 and 2 years of age, the annual survival of prime-age females between 2 and 7 years of age (or 10 depending on the size, gaillard et al. 2000), the annual survival of females from 7 (or 10) years of age onwards, alces vol. 43, 2007 gaillard are moose only a large deer? 3 into leslie matrix models and estimated both r and tb (see caswell 2000 for further details). for species in which i obtained data from sevpopulations belonging to 22 species including 6 moose populations (in south-central alaska, ballard et al. 1991; in south coast barrens of newfoundland, albright and keith 1987; in northwest territories (canada), stenhouse et al. 1995; and 3 populations in northern norway, stubsjoen et al. 2000). to assess whether observed survival patterns account for the relatively rapid life cycle observed in moose populations, i used published estimates of both adult and juvenile survival in ungulate species. from a previous literature review (gaillard et al. 2000), i found data on adult survival in 61 populations belonging to 25 species (including 9 populations of moose) and on juvenile survival in 53 populations belonging to 25 species (including 7 populations of moose). because very low between-year variation in survival could also contribute to the higher than expected population growth rate of moose (see tuljapurkar 1989 for a discussion of the changes in population growth generated by environmental variation), i also compared the magnitude of annual variation of both juvenile and adult survival of annual estimates) in moose populations with the variation reported in other ungulate species. to account for the expected increase in survival with increasing body size (see peters 1983, calder 1984 for reviews), i regressed both mean survival and cv of survival for juveniles and adult females (measured as the a given species) on adult body mass. to assess whether observed reproductive patterns account for the relatively rapid life cycle observed in moose populations, i collected data for ungulate species on two and litter size), as well as on body mass and generation time (see above). i found data for differences in reproductive traits according to adult body mass with 1-way anovas using reproductive traits as factors (i.e., three classes age; and two classes of litter size: 1 or 2) and the log-transformed adult body mass as the dependent variable. i then compared the adult body mass of moose with the mean mass expected from species with similar reproductive traits. in a second step, i performed the same kind of analysis by using the log-transformed generation time instead of adult body mass. i assumed that once variation in adult body mass is taken into account, differences in reproductive traits between moose and other ungulates account for the relatively faster life cycle of moose compared to other ungulates, and moose should reproduce earlier and more frequently relative to their size but perform as expected from their generation time. results and discussion patterns of maternal care in moose: birth mass and early growth rate as expected, a strong positive relationship occurred between birth mass (bw) and adult body mass (abw) in both monotocous (ln (bw) = -1.366 + 0.902 ln (abw); r = 0.977, p < 0.0001) and polytocous (ln (bw) = -3.274 + 1.059 ln (abw); r = 0.892, p = 0.0029) species. there was no difference between slopes according to litter size (f = 0.903; df = 1, 42; p = 0.347). however, for a given adult body mass, birth mass was larger in monotocous than in polytocous species (difference in intercepts of 0.295 (se = 0.107); f = 7.630; df = 1, 43; p = 0.008). similarly, early growth rate (gr) was allometrically related to adult body mass (abw) in both monotocous (ln (gr) = -2.689 + 0.733 ln (abw); r = 0.968, p < 0.0001) and polytocous (ln (gr) = -0.691 + 0.561 ln (abw); r = 0.744, p = 0.0090) species. however, litter are moose only a large deer? gaillard alces vol. 43, 2007 4 ship between early growth rate and adult body mass (differences in slope: f = 0.574; df = 1, 24; p = 0.456; differences in intercept: f = 2.770; df = 1, 25; p = 0.109). using such allometric relationships to estimate expected values for moose, i obtained birth mass of 30.12 kg and 34.99 kg and early growth rates of 777.11 g/d and 639.61 g/d from the equations of monotocous and polytocous species, respectively. observed birth mass was only about half the expected values: 16.2 kg for monotocous moose and 13.5 kg for polytocous moose (schwartz and hundertmark 1993). on the other hand, an observed early growth rate of 785 g/d (reese and robbins 1994) was very similar to the expected values from allometric equations. comparison of moose to similar-sized bovids led to the same conclusions. moose had a much lighter birth mass than similar-sized species (table 1). birth mass in moose was similar to the birth mass of wildebeest (connochaetes taurinus) whose adult body size is only half that of moose. on the other hand, early growth rates in moose were mid-range to those measured in similar-sized bovids. comparative analyses of maternal care show that moose produce small newborns in relation to their size (about half the newborn size expected from other cervid species and similar-sized bovids). on the other hand, relative to their size, newborn moose grow at the same rate as other cervids and similarsized bovids. i can thus also conclude that moose allocate energy to maternal care as a monotocous species during the gestation period but as a polytocous species during the lactation period. demographic patterns of moose populations as expected, a marked positive allometric relationship occurred between tb and adult body mass (abw) among the 21 ungulate species other than moose (ln (tb) = 0.967 + 0.247 ln (abw); r = 0.646, p = 0.0016; fig. 1). the allometric exponent was very close to that expected for a measure of biological time such as generation time (0.25; calder 1984), indicating that populations of large ungulate species have relatively slower life cycles than populations of small ungulate species. from such a relationship, tb of moose would be expected to be 11.76 years. from the 6 moose populations for which i found published information, the estimated tb was consistently shorter (from 4.57 to 10.66 years) than the expected value (table lht wildebeest (connachaetes taurinus) cattle (bos taurus) eland (taurotragus derbianus) moose (alces alces) buffalo (syncerus caffer) adult mass (kg) 165 309 363 340-450 536 birth mass (kg) 16.5 24 31.5 13-16 37.2 growth rate (kg/d) 0.29 0.64 1.11 0.79 1.47 litter size 1 1 1 1-2 1 table 1. comparison of life history traits (lht) among moose to similar-sized bovids as related to maternal care. 765432 1,0 1,5 2,0 2,5 3,0 ln (a dult body m ass) l n (g en er at io n t im e) fig. 1. allometric relationship between generation observed generation time of moose (as measured by the median value from 6 populations) corresponds to the open square. alces vol. 43, 2007 gaillard are moose only a large deer? 5 2), meaning that moose have a relatively short tb for their size. the median values observed for moose in the allometric relationship linking tb and adult body mass led to the largest negative residual. likewise, according to previous work on a large range of taxa (e.g., blueweiss et (e.g., sinclair 1996, 1997), a negative allometric relationship tended to occur between r and adult body mass (abw) among 17 ungulate species other than moose that showed a positive r (i.e., increasing populations: ln (r) = -1.402 0.303 ln (abw); r = 0.395, p = 0.117; fig. 2). the slope was close to the theoretical expectation of -0.25 (calder 1984), meaning that the product between r and tb is a dimensionless number (life history invariant sensu charnov 1993). from such a relationship, r would be expected to be 0.039 for increasing populations of moose. from the 5 increasing moose populations for which i found published information, the estimated r was consistently higher (from 0.077 to 0.344; table 2) than the expected value, meaning that moose populations have a high growth rate relative to female body size. the median value observed for moose on the allometric relationship between r and adult body mass led to one of the two largest positive residuals with a colonizing population of bison (bison bison) (van vuren and bray 1986). such allometric analyses suggest that overall demographic patterns of moose populations are more similar to those of smallor medium-sized ungulates than to those of similar-sized species. from expectations based on their body size alone, moose populations increase faster and the turnover of individuals is faster. such overall demographic features can have three explanations: (1) survival of juveniles and/or adult female moose is much lower than that of similar-sized ungulates; (2) reproductive output of moose is much higher than that of similar-sized ungulates; or (3) both lower survival and higher reproductive output occur simultaneously in moose populations relative to similar-sized ungulates. do observed survival patterns account for the relatively rapid life cycle observed in moose populations? contrary to expectation, the logit of female adult survival (las, which corresponds to the log-transformed adult life expectancy) did not increase with increasing adult body mass among ungulate species (las = 2.205 + 0.040 ln (abw); r = 0.054, p = 0.801; fig. 3). female adult survival was high irrespective of body mass (from 0.710 in topi (damaliscus lunatus) to 0.978 in pronghorn (antilocapra americana); mean of 0.903, se = 0.012). female survival varied r tb reference 0.10 8.89 ballard et al. 1991 0.26 6.14 stubsjoen et al. 2000 0.34 4.57 stubsjoen et al. 2000 0.27 6.02 stubsjoen et al. 2000 -0.03 10.66 albright and keith 1987 0.08 6.91 stenhouse et al. 1995 table 2. population growth rate (r) and generation time (tb) estimated for 6 moose populaliterature. 765432 -4 -3 -2 -1 ln (a dult body m ass) l n (p o p u la ti on g ro w th ) fig. 2. allometric relationship between population data collected from 17 ungulate species with population growth rate of moose (as measured by the median value from 5 increasing populations) corresponds to the open square. are moose only a large deer? gaillard alces vol. 43, 2007 6 from 0.780 to 0.976 and averaged 0.907 (± 0.022) among the 9 moose populations for which data were available. therefore, we can conclude that female adult survival of moose is similar to adult survival reported for other female ungulates. likewise, there was no effect of adult body mass on cv of adult survival in female ungulates (cv = 0.069 + 0.001 ln (abw); r = 0.021, p = 0.927; fig. 4). cv of female adult survival was low, irrespective of body mass (from 0.017 in reindeer (rangifer tarandus ovis gmelini); mean of 0.073, se = 0.007). cv of adult survival of females varied from 0.009 to 0.051 and averaged 0.035 (± 0.005) among the 7 populations of moose for which data were available. such between-year variation appears to be a little lower than that observed in other ungulates. as expected, the logit of juvenile survival (lsj) tended to increase with increasing adult body mass among ungulate species (lsj = -1.648 + 0.409 ln (abw); r = 0.314, p = 0.135; fig. 5). expected juvenile survival indeed increased from 0.40 for an ungulate weighing 20 kg to 0.69 for an ungulate weighing 400 kg. juvenile survival varied from 0.235 to 0.835 and averaged 0.640 (± 0.088) among the 7 moose populations from which i found data. therefore, i can conclude that juvenile survival of moose is similar to juvenile survival reported for similar-sized ungulates. likewise, there was a trend in the cv of juvenile survival to decrease with increasing adult body mass among ungulates (cv = 0.648 – 0.072 ln (abw); r = 0.353, p = 0.099; fig. 6). expected cv in juvenile survival decreased from 0.430 for an ungulate weighing 20 kg to 0.220 for an ungulate weighing 400 kg. cv of juvenile survival of moose varied from 0.126 to 0.710 (average of 0.332 (± 0.130), median of 0.245) among the 765432 0 1 2 3 4 l n (a dult body m ass) l o g it (f em al e ad u lt su rv iv a l) fig. 3. allometric relationship between adult survival of females (after logistic transformaobserved female adult survival of moose in 9 populations corresponds to the open squares. 765432 0,00 0,05 0,10 0,15 l n (adult body m ass) c v (f em al e ad u lt su rv iv al ) fig. 4. allometric relationship between temporal variation in adult survival of females (as meatemporal variation in female adult survival of moose in 7 populations corresponds to the open squares. 765432 -2 -1 0 1 2 3 4 l n (adult body m ass) l o g it (j u v en il e su rv iv al ) fig. 5. allometric relationship between juvenile survival (after logistic transformation) and adult survival of moose in 7 populations corresponds to the open squares. alces vol. 43, 2007 gaillard are moose only a large deer? 7 4 populations of moose with available data. such between-year variation was similar to that observed in other similar-sized ungulates. comparison of moose survival to survival in similar-sized bovids led to the same conclusions (table 3). both survival estimates and temporal variation in survival of juvenile and adult female moose were very close to the values for similar-sized bovids. eral survival pattern of ungulates, especially among smaller species, characterized by both a high and constant survival of adult females, irrespective of the species considered, and a low juvenile survival with high variability among years (see gaillard et al. 1998, 2000 for similar conclusions). because survival found in other ungulate species, survival patterns cannot be an explanation for the relatively rapid life cycle in moose. do observed reproductive patterns account for the relatively rapid life cycle observed in moose populations? as expected, the mean body mass of ungulates differed according to the observed f = 4.808; df = 2, 18; p = 0.021), increasing from species that give birth at 1 year of age (20 kg; n = 1) to species n = 10). ungulates that usually start to give birth at 2 years of age had an intermediate mean from that of ungulates starting to give birth at 3 years of age or older (fisher’s lsd test, p = 0.028). moose often give birth at 2 years of age when they are faced with favorable environmental conditions (schwartz 1992). with a female adult body mass usually between 350 and 450 kg, moose do not belong to the size distribution of ungulates that normally start to reproduce at 2 years of age but belong to the size distribution of ungulates that normally do not start to reproduce before 3 years of age. likewise, the mean body mass of ungulates differed according to the observed litter size (f = 13.57; df = 1, 19; p = 0.002), decreasing from species that give birth to 765432 0,0 0,2 0,4 0,6 0,8 1,0 l n (adult body m ass) c v (j u v en il e su rv iv al ) fig. 6. allometric relationship between temporal variation in juvenile survival (as measured by tion in juvenile survival of moose in 4 populations corresponds to the open squares. traits wildebeest kudu moose bison buffalo adult mass (kg) 165 170 340-450 450 536 js1 0.58 0.45 0.71 0.97 0.45 cv (js)2 0.37 0.52 0.25 0.04 0.24 as3 0.88 0.91 0.91 0.95 0.93 cv (as)4 0.12 0.05 0.03 0.05 0.04 table 3. comparison of traits related to survival patterns among moose and similar-sized bovids. 1juvenile survival. 2 3adult survival. 4 are moose only a large deer? gaillard alces vol. 43, 2007 8 single offspring (139; n = 10) to species that give birth to twins (37 kg; n = 6). moose often give birth to twins when they are faced with favorable environmental conditions (boer 1992). with a female adult body mass usually between 350 and 450 kg, moose do not belong to the size distribution of ungulates that are expected to produce twins but belong to the size distribution of ungulates that normally produce single offspring. looking now at the relationship between i found that as expected, generation time tion (f = 5.89; df = 2, 18; p = 0.011) from yearlings to 10 years for those giving birth for of age had an intermediate generation time (fisher’s lsd test, p = 0.013). the generation times observed in moose populations (from 4.57 to 10.66 years) match the distribution of generation times of ungulates that reproduce generation time decreased as expected with increasing litter size (f = 11.27; df = 1, 19; p = 0.003) from 9.3 years for ungulates that normally produce single offspring to 5.7 years for those that can produce twins. the generation times observed in moose populations (from 4.57 to 10.66 years) match the distribution of generation times of ungulates that can produce twins. comparison of moose reproductive patterns with those of similar-sized bovids led to the same conclusions (table 4). the age at of smaller wildebeest and kudu (tragelaphus strepsiceros) than that of larger bison and buffalo (syncerus caffer). moreover, both the proportion of 2 year-old females that give birth and the litter size were greater in moose than in smaller wildebeest and kudu. i can therefore conclude that female moose reproduce earlier (often giving birth at 2 years of age instead of 3 years of age for similarsized ungulates) and have larger litters (often producing twins instead of single offspring as in similar-sized ungulates) than expected from their size. high reproductive output accounts for the rapid life cycle of moose populations compared to populations of other, similarsized ungulates. indeed, the distribution of generation times reported in moose populaexpected for ungulate populations that give often produce twins. conclusions: are moose a large roe deer or a very large deer? although only a few comparative analyses have reported clear ecological correlates of life history strategy, there is general agreement among evolutionary ecologists that traits wildebeest kudu moose bison buffalo adult mass (kg) 165 170 340-450 450 536 age of primiparity 2 2 2 3 3 % 21 0.27 0.1 0.4 0 0 % m2 0.9 0.9 0.81 0.62 0.7 mean litter size 1 1 1.32 1 1 table 4. comparison of reproductive traits among moose to similar-sized bovids. 1proportion of 2 year-old females giving birth in a given year in a population. 2proportion of multiparous females giving birth in a given year in a population. alces vol. 43, 2007 gaillard are moose only a large deer? 9 among-species differences in habitat and diet should lead to differences in life history traits (stearns 1992), especially in mammals (saether and gordon 1994 for ungulates, geffor marsupials). i therefore may ask whether for the relatively high reproductive output of moose? contrary to other large cervids such as red deer (cervus elaphus), moose appear to select early successional vegetation stages as a preferred habitat and are concentrate selectors (browsers) rather than grazers or mixed-feeders (hofmann 1989). from these features, moose are much closer to roe deer (capreolus capreolus), with which they occur often in sympatry, than to larger deer. like all cervids, roe deer and moose both patterns of other ungulates. strong selective pressures might have been operating during the evolutionary history of ungulates in response to predation (see byers 1997). the canalization of adult survival (gaillard and yoccoz 2003) and the production of large and fast growing offspring within allometic constraints might ungulate evolution, leading these life history traits to vary little among ungulate species as ecological conditions vary. on the other hand, both moose and roe deer have a relatively high reproductive output, maybe in response to unpredictable environmental conditions in early successional habitats (as proposed by liberg and wahlström 1995). female moose (weighing about 400 kg) cannot produce offspring as fast and as often as roe deer (weighing about 25 kg) because of allometric constraints (peters 1983, calder 1984, brown and west 2000). thus, most female roe deer under a large range of environmental conditions, while only about half of female moose in the most productive populations (e.g., vega island in norway where most females produce twins, solberg, personal communication) do the same (schwartz and hundertmark 1993). litter size of roe deer can be 3 offspring in very good conditions, while moose litter size is commonly 2 in the same situation. lastly, because of their large size, moose cannot be a true polytocous species. to reach their high reproductive output, moose have to trade quality of offspring (moose offspring are half the size of other ungulates’ offspring) for a higher quantity of offspring (female moose produce twins as soon as environmental conditions allow). this comparative analysis of moose life history traits suggests that moose are large roe deer rather than simply large deer, and supports current theory on life history evolution that species occupying unpredictable habitats live at a faster rate than species living in more predictable habitats (yodzis 1989). acknowledgements this paper was prepared for the 5th international moose symposium held at lillehammer, norway, august 2002. i thank erling solberg and bernt-erik saether for inviting me to the symposium. references albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of newfoundland. the canadian field-naturalist 101:373-387. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. björklund, m. 1997. are ‘comparative methods’ always necessary? oikos 80:607-612. blueweiss, l., h. fox, v. kudzma, d. nakashima, r. peters, and s. sams. 1978. relationship between body size and some life history parameters. oecologia 37:257-272. boer, a. h. 1992. fecundity of north amerare moose only a large deer? gaillard alces vol. 43, 2007 10 ican moose (alces alces): a review. alces supplement 1:1-10. brown, j. h., and g. b. west. 2000. scaling in biology. oxford university press, new york, new york, usa. byers, j. a. 1997. american pronghorn: social adaptations and the ghosts of predators past. the university of chicago press, chicago, illinois, usa. calder, w. a. iii. 1984. size, function and life history. harvard university press, harvard, cambridge, massachusetts, usa. caswell, h. 2000. matrix population models. construction, analysis and interpretation. sinauer associates, sunderland, massachuetts, usa. charnov, e. l. 1993. life history invariants. some explorations of symmetry in evolutionary ecology. oxford university press, oxford, u.k. fisher, d. a., and i. f. p. owens. 2000. female home range size and the evolution of social organization in macropod marsupials. journal of animal ecology 69:1083-1098. _____, _____, and c. n. johnson. 2001. the ecological basis of life history variation in marsupials. ecology 82:3531-3540. fisher, r. a. 1930. the genetical theory of natural selection. oxford university press, oxford, u.k. gaillard, j.-m., m. festa-bianchet, and n. g. yoccoz. 1998. population dynamics of large herbivores: variable recruitment with constant adult survival. trends in ecology and evolution 13:58-63. _____, _____, _____, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual review of ecology and systematics 31:367-393. _____, d. pontier, d. allaine, j. d. lebreton, j. trouvilliez, and j. clobert. 1989. an analysis of demographic tactics in birds and mammals. oikos 56:59-76. _____, and n. g. yoccoz. 2003. temporal variation in survival of mammals: a case of environmental canalization?. ecology 84:3294–3306. garland, t., p. h. harvey, and a. r. ives. 1992. procedures for the analysis of comparative data using phylogenetically independent contrasts. systematic bio-biology 41:18-32. geffen, e., m. e. gompper, j. l. gittleman, h. k. kuh, d. macdonald, and r. k. wayne. 1996. size, life history traits, and social organization in the canidae: a reevaluation. american naturalist 147:140-160. harvey, p. h., and m. d. pagel. 1991. the comparative method in evolutionary biology. oxford university press, oxford, u.k. _____, and a. rambaut. 2001. comparative analyses for adaptive radiations. philosophical transactions of the royal society of london series b 355:1599-1605. hofmann, r. r. 1989. evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. oecologia 78:449-457. leslie, p. h. 1966. the intrinsic rate of increase and the overlap of successive generations in a population of guillemot (uria aalge pont). journal of animal ecology 35:291-301. liberg, o., and k. wahlstrom. 1995. habitat stability and litter size in the cervidae; a comparative analysis. pages 1-60 in k. wahlström. natal dispersal in roe deer. an evolutionary perspective. unpublished ph.d. thesis, university of stockholm, sweden. novak, r. m. 1993. walker’s mammals of the world. fifth edition, volume ii. the john hopkins university press, baltimore and london, u.k. peters, r. h. 1983. the ecological implication of body size. cambridge university alces vol. 43, 2007 gaillard are moose only a large deer? 11 press, cambridge, u.k. price, t. 1997. correlated evolution and independent contrasts. philosophical transactions of the royal society of london series b 352:519-529. reese, e. o., and c. t. robbins. 1994. characteristics of moose lactation and neonatal growth. canadian journal of zoology 72:953-957. ricklefs, r. e., and j. m. starck. 1996. applications of phylogenetically independent contrasts: a mixed progress report. oikos 77:167-172. roff, d. a. 1992. the evolution of life histories. chapman and hall, london, u.k. saether, b. e., and i. j. gordon. 1994. the adaptive significance of reproductive strategies in ungulates. proceedings of the royal society of london series b 256:263-268. schwartz, c. c. 1992. reproductive biology of north american moose. alces 28:165-173. _____, and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. journal of wildlife management 57:454-468. sinclair, a. r. e. 1996. mammal populations: fluctuation, regulation, life history theory and their implications for conservation. pages 127-154 in r. b. floyd, a. w. sheppard, and p. j. de barro, editors. frontiers of population ecology. csiro publishing, melbourne, australia. _____, 1997. fertility control of mammal pests and the conservation of endangered marsupials. reproduction, fertility and development 9:1-16. stearns, s. c. 1976. life-history tactics: a review of the ideas. quarterly review of biology 51:3-47. _____, 1983. the influence of size and phylogeny on patterns of covariation among life history traits in the mammals. oikos 41:173-187. _____, 1992. the evolution of life histories. oxford university press, oxford, u.k. stenhouse, g. b., p. b. latour, l. kutny, n. maclean, and g. glover. 1995. productivity, survival, and movements of female moose in a low-density population, northwest territories, canada. arctic 48:57-62. stubsjoen, t., b. e. saether, e. j. solberg, m. heim, and c. m. rolandsen. 2000. moose (alces alces) survival in three populations in northern norway. canadian journal of zoology 78:1822-1830. tuljapurkar, s. d. 1989. an uncertain life: demography in random environment. theoretical population biology 21:141165. van vuren, d., and m. p. bray. 1986. population dynamics of bison in the henry mountains, utah. journal of mammalogy 67:503-511. yodzis, p. 1989. introduction to theoretical ecology. harper & row, new york, new york, usa. 4004.p65 alces vol. 40, 2004 dettki et al. real-time moose tracking 13 real-time moose tracking: an internet based mapping application using gps/gsm-collars in sweden holger dettki1, göran ericsson2, and lars edenius2 1department of forest resource management and geomatics, swedish university of agricultural sciences (slu), se – 901 83 umeå, sweden; 2department of animal ecology, swedish university of agricultural sciences (slu), se – 901 83 umeå, sweden abstract: to date, moose (alces alces) tracking has relied on techniques either based on ‘very high frequency’ (vhf) / ‘ultra high frequency’ (uhf) radio collars, or global positioning system (gps) collars, often requiring significant effort in the field to collect data. here we present a technique that automatically tracks and reports moose in almost real time, and presents moose positions and movement paths with an interactive web-based map service. we equipped 25 female moose with gps/gsm collars in västerbotten county, northern sweden. the gps receivers acquired a position every 30 minutes and transmitted them after 3.5 hours as a standard short messaging service (sms) message using the global system for mobile communications (gsm) cell phone network. the positions were automatically extracted from the receiving local gsm-modem and stored in a database. during 18 days in march 2003, 18,638 gps positions were transmitted by 2,719 sms messages. of all positioning attempts 98.1% resulted in a valid position, whereof 99.7% were 3-dimensional positions. the real-time approach allows for many new research studies; e.g., smallscale migrational studies with adapted gps schedules for different phases of migration. further, public access to the moose data by a web-based map can be of fundamental importance for public acceptance when dealing with local concerns. alces vol. 40: 13-21 (2004) key words: alces alces, boreal forest, collar, gps, gsm, moose, sms, sweden, tracking, ungulate a large fraction of the moose (alces alces) population in northern sweden migrates between summer and winter home ranges (ball et al. 2001), creating locally concentrated problems for forestry. areas prone to extensive winter browsing by moose are often associated with extensive areas of economically valuable sapling stands of scots pine (pinus sylvestris). access to accurate and up-to-date information is an important requisite for effective co-management of moose and forest resources. in order to mitigate and reduce conflicts arising because of strong browsing pressure in such winter concentration areas, improved knowledge about moose movement and ranging behaviour at the regional and landscape scale in relation to available winter browse is important. such insights may greatly facilitate co-operation among hunters, foresters, and managers over management issues. to date, ungulate tracking used to, for example, study migration and movements of large ungulates has relied either on techniques based on traditional radio-telemetry using ‘very high frequency’ (vhf)or ‘ultra high frequency’ (uhf) radio coll a r s ( e . g . , h e e z e n a n d t e s t e r 1 9 6 7 , hundertmark 1998, ericsson et al. 2001) or in recent years more and more on global positioning system (gps) telemetry (e.g., rempel et al. 1995, moen et al. 1996, edenius 1997, hundertmark 1998, turner et al. 2000). gps is a us based satellite-aided navigation system that allows the calculation of real-time moose tracking – dettki et al. alces vol. 40, 2004 14 positions worldwide with about ± 10 m precision. using a small handheld receiver a position can be calculated by triangulation when signals from at least 3 different of 21 available gps-satellites are received. extensive studies have been conducted to evaluate the performance of gps collars under different environmental conditions (e.g., moen et al. 1996, edenius 1997, moen et al. 1997, struch et al. 2001). however, existing systems have the disadvantage that significant labour is required in the field for acquiring locations (vhf/uhf) or to extract on-board stored positional data from gps collars via a local radio-link. the alternative, to download stored data after retrieval of a collar from an animal, always faces the risk of total data loss due to both mechanical collar failures or loss of the animal. further, visualisation and spatial analysis of the data often require considerable investments in geographic information systems (gis) software and education. the aim of this paper is to describe a framework for tracking and displaying moose positions and movement paths in almost real-time with a web-based map service, which provides simple statistics on positions and movements. study area the study area was located near åsele in the county of västerbotten, sweden (64 °06’n 17°18’e; fig. 1). the 2,200 km2 area has a winter density of 1.3 moose/km2 with local concentrations up to >12.5 moose/ km2 (andersson 2002). the climate is predominantly continental with relatively short summers (1 june 10 sept.) and the length of the growing season is 150-160 days with an onset between 10 may and 20 may. average annual temperature is 23°c. the onset of winter is normally the first week of november and lasts to midapril. annual precipitation ranges between 600 and 700 mm. the ground is usually snow-covered from the first week of november to the last week of april. maximum snow depth peaks at 70-80 cm in late february. climatic data were averaged from the period 1961-1990 (raab and vedin 1995). fig. 1. the county of västerbotten in northern sweden shown with the study area (rectangle), the city of umeå, and the arctic circle indicated. alces vol. 40, 2004 dettki et al. real-time moose tracking 15 methods gps/gsm collars we equipped 25 female moose with gps collars (gps/gsm plus 4d) from vectronic aerospace gmbh (fielitz 2003) and uniquely numbered ear-tags during 1 5 march 2003. the collars weighed ca 1.1 kg and were designed for a battery lifetime of 1.5 years. we immobilised the moose from a helicopter using a dart gun injecting a mixture of anaesthetic and tranquillizer (ethorphine and xylazine; sandegren et al. 1987). moose were aged according to tooth wear (ericsson and wallin 2001), which is in agreement with the cementum annuli method <5 years of age (k. wallin and g. ericsson, unpublished data). twelve-channel gps receivers mounted on the collars acquired a position every 30 minutes and stored them internally for later download. each collar was equipped with a cell phone unit using the widely available global system for mobile communications (gsm) network in europe, the second generation digital cell phone technology. the unit consisted of a ‘dual band’ module, i.e., it could use the common gsm frequencies in europe of 900 mhz and 1,800 mhz. after 7 positioning attempts with a time interval of 0.5 hours, the gsm unit was programmed to send these 7 positions each 3.5 hours as a standard short messaging service (sms) text message to a gsmmodem located at the swedish university of agricultural sciences (slu) in umeå, sweden (63°49’n 20°16’e; fig. 1). despite the fact that the gsm technology does not require a free line of sight towards a gsm relay transmitter, the study area contains gaps with low or no coverage due to weak signals or partial signal blackouts in very remote or mountainous areas. thus, not every sms message could be transmitted immediately after the seventh position was acquired. therefore the collars were programmed to always send a sms message with the latest 7 positions first and then check for unsent positions stored in the collar when within an area with sufficient gsm coverage. if unsent positions were found, up to 10 sms messages with a total of 70 positions were sent, starting with the latest unsent position. after sending, the gsm module went into a ‘sleep’ mode to preserve battery capacity and tried to send the next 7 positions and eventually the remaining unsent positions during the next sending opportunity (i.e., 3.5 hours later). after the gsm-modem received a sms message (fig. 2), the positions were automatically extracted and the co-ordinates converted to the national grid of sweden and stored in a sql-server database. together with the positions, the exact date and time, altitude, dilution of position (dop), type of position (2-dimensional or 3-dimensional), battery voltage, and the internal temperature of the collar were transmitted and stored in the database. after one year, the collars will be retrieved manually by darting the moose. to locate a collar for retrieval in areas without gsm coverage, collars were additionally equipped with a permanent vhf transmitter. after retrieval, all data can be downloaded from the collars. due to sms message size limitations not all registered data (e.g., data from the built-in activity sensors for the animals xand y-axes) for each position are transmitted by sms, but stored in the collars for later analysis. data retrieval and presentation digital topographical landcover maps (original scale 1:100,000) obtained from the swedish national land survey were used as background maps. they were stored on a web server and accessed by the arcims 4.0 engine (esri 2002) and the internet information server 5.0 (iis 5.0; microsoft corporation 2000). a web application built with the active server page technology real-time moose tracking – dettki et al. alces vol. 40, 2004 16 fig. 2. schematic description of the information flow: positional information from the gps-satellites is stored in the collar, and transmitted by the gsm network using sms messages to a database server to be accessed as a map on the internet. (asp) and visual basic, both in the .net environment (richter 2002), extracted the positions for one or more moose and overlaid them on the background maps (fig. 3a) using the activex connector and arcims (esri 2002) as positions or moving paths. further, position co-ordinates for each recorded position (fig. 3b) or simple statistics on path length or numbers of positions were shown as tables. results and discussion gps positions and data transfer between 1 18 march 2003, 18,638 gps positions were registered for the 25 moose in the database. of all positioning attempts in the field, 98.1% were successful (table 1), resulting in 99.7% 3-dimensional positions. the dilution of precision (dop) value as a measure of positional precision (moen et al. 1997) was < 2.0 for 75.1% of all positions. based on these figures, the estimated number of valid positions will be ca 15,120 positions per moose and year. all stored position data were transmitted by 2,719 sms messages. of all sms messages 95.3% were successfully sent immediately after 7 positions were registered, while 4.7% were delayed because of the moose being outside an area with gsm coverage during the time of transmission (table 2). maximum delay time between two consecutive sms transmissions for a 2-d 3-d total fix-attempts — — 19,004 valid positions 64 18,574 18,638 dop < 2.0 21 13,974 13,995 table 1. number of valid gps-positions in the database during the first 18 days (1 18 march 2003). data are given for 2-dimensional (2-d) and 3-dimensional (3-d) positions as well as positions with a dilution of position (dop) <2.0. the number of attempts to obtain a position (fix) is also given for the same period. alces vol. 40, 2004 dettki et al. real-time moose tracking 17 fig. 3a. example of a movement path for moose ‘432’ on 18 march 2003 near åsele in the county of västerbotten, sweden on the interactive web-site represented (a) as a map and (b) as a table of positions within the movement path. on the map, the circle at one end of the movement path (dashed line) indicated the last transmitted position. in the table the moose name or id, the local date and time for each position, the northing, easting, and height above sea level (‘rn north’, ‘rn east’, and ‘rn height’; coordinates in the national grid of sweden, in meters), and the dilution of position (dop) value were given. further, it was noted in field ‘nav’ whether the position was 3-dimensional (3d) or 2-dimensional (2d), and a temperature measure (‘temp’, in degrees celsius) was given. name date time rn north rh east rn height dop nav temp 432 2003-03-18 23:30:56 7115322 1580400 311.6 1.2 3d 6 432 2003-03-18 23:00:24 7115313 1580406 323.3 1.4 3d 6 432 2003-03-18 22:30:06 7115320 1580406 335.9 2 3d 6 432 2003-03-18 22:00:26 7115320 1580401 322.5 3.6 3d 6 432 2003-03-18 21:30:49 7115315 1580404 322.5 1.4 3d 5 432 2003-03-18 21:00:20 7115307 1580405 319.6 1.4 3d 5 432 2003-03-18 20:30:07 7115319 1580411 318.5 2 3d 5 432 2003-03-18 20:00:49 7115310 1580411 308.6 3.8 3d 5 432 2003-03-18 19:30:20 7115348 1580418 323.5 1.6 3d 2 fig 3b. real-time moose tracking – dettki et al. alces vol. 40, 2004 18 time (t) #sms received fraction at (t) 2,719 100.0% directly (after 3.5 h) 2,590 95.3% after 7.0 h 74 2.7% after 10.5 h 13 0.5% after 14.0 h 18 0.7% after 17.5 h 16 0.6% after 21.0 h 3 0.1% > 24 h 5 0.2% table 2. number and percentage of sms messages received during the first 18 days (1 18 march 2003) for different time intervals between sms receptions. single moose lasted 137.5 hours. we expect the percentage of delayed messages to increase during the year, as many of the collared moose may eventually migrate towards more remote, mountainous areas with a more fragmented gsm coverage. on the other hand, the gps positioning success rate should remain quite constant, as the vegetation coverage, which is the main obstacle for the reception of the satellite signals for position calculation, is similar throughout the study area. data access and presentation all moose data can be accessed by using a web page either as a map with positions and movement paths (fig. 3a), or as a table of positions (fig. 3b). further, as the data are stored on a sql-database server, the data are available on virtually any computer connected to the internet. hence, more sophisticated data analysis can be conducted locally using standard gis packages. the data were also made accessible to the public on an interactive web page showing moose locations and movement paths on a map (http://www-moosetrack.slu.se). data access, however, was restricted to positions older than 2 weeks and no tables with positional data were given to avoid disturbance of the animals. the application is already heavily used by local residents, media, hunter associations, and forest companies for different reasons. interest of local residents has so far been focused on comparing local knowledge on moose movement to actual movement as displayed data on maps on the internet. furthermore, several schools in sweden have started to use the real-time information on the web site in education on wildlife ecology and management. we also anticipate an increased interest from the tourist industry, for example, by using the internet interface to explore the charismatic value of moose to attract visitors, as a wild moose in many countries is recognized as a symbol for undisturbed wilderness. moreover, we foresee that forest companies and hunter associations increasingly will use real-time information on moose activity to facilitate communication to help resolve conflict situations. lastly, we expect to see an increased public interest channelled into projects such as ‘adopting’ a moose. hence, public interest in real-time moose movement may deepen acceptance of moose populations in managed forest areas. pros and cons monitoring of moose movement in almost real-time enables a quick reaction to possible problems with either animals or collars. for example, one of the collared moose stopped moving a few hours after darting and outfitting with a collar, and the collar temperature started to decrease over a period of 12 hours. it was possible to alces vol. 40, 2004 dettki et al. real-time moose tracking 19 immediately revisit the animal to check on the health status of the moose. though the moose was fine, technical problems with the collar were detected and corrected quickly with remote re-configuration to minimize data loss. the major advantage of the technology used in this study is the ability to collect a large amount of data in almost real-time with minimal labour required. in a previous p r o j e c t c o n d u c t e d i n t h e c o u n t y o f västerbotten (dettki et al. 2003), 15 moose were equipped with gps collars over a period of 4 years, resulting in 17,036 valid positions. a varying gps schedule was applied, resulting in 1 position per moose each hour, and up to 1 position per moose a day. the data were available first after the collars had been retrieved manually. in the current study, after only 18 days, 18,638 positions were recorded for 25 moose, applying a gps schedule with 1 position per moose and 0.5 hours. while data handling in the previous study was done manually, in the current project it is necessary to use automated routines to handle the relatively large amount of data. only basic supervision of the servers is required to handle the incoming data from the download of a sms message from the gsm-modem, conversion of positions into local grid format, error check of both sms and positional data, storage in the sql-database and finally presentation on a map. this results in high efficiency in data retrieval and manipulation and very low costs per position, as the number of received positions is high and labour required is low. however, some disadvantages with the technology exist. first, the amount of data transmitted by sms is mainly constrained by the allowed size of each sms, which is internationally standardized to a maximum of 160 characters per message. this prohibits transmission of additional data by sms, such as activity and mortality data. furthermore, sms transmission drains battery power and is costly, so there are tradeoffs to be made. even though the costs per position transmitted are small (approximately us $0.02 per position with 7 positions in each sms message), during 1 year the costs for data transmission add up to approximately us $350.00 per collar with a position taken every 0.5 hours. battery capacity continues to be a bottleneck of gps/gsm-systems, especially for long-duration studies or small-sized animals. the same problems arise for transmitting data for differential post-processing, which would increase the amount of necessary sms messages to be sent at least 3-fold. however, with a precision of ± 10 m with uncorrected positions, most applications likely do not have the need of differential real-time corrections. the far greatest disadvantage is the limitation due to the fragmented or nonexisting coverage of the gsm network in remote areas in many parts of the world. the scandinavian countries differ in this respect from, for example, north america or russia. in sweden, about 75% of the land area is covered by at least one gsm network (teliasonera 2003). however, even though all bigger roads and settlements in northern sweden have at least some degree of gsm coverage, coverage is poor in more remote areas. we tried to reduce this problem by programming the collars to deliver all stored positions at least once by sms when the animal is within or re-enters into an area with sufficient gsm coverage. during the first 18 days, all positions stored on-board were also retrieved by sms, even though some moose collars were out of contact with the gsmnetwork for up to 137.5 hours. it is important to note that sms transmissions require less net coverage than is required for oral communication by cell phones or gsm data transfer. this means that the techreal-time moose tracking – dettki et al. alces vol. 40, 2004 20 nique is applicable also in areas where it is not possible to use a cell phone for talking, due to fragmented gsm coverage. implications the gps/gsm-technique gives a new possibility to ‘supervise’ animals in almost real-time. for research this means one can see when, e.g., the migration starts, and then via sms, change the gps schedule scheme to acquire positions more frequently. it further opens up the possibility to determine in real-time when migration is over, i.e., when moose have reached their final destination. wildlife managers then can use lethal or non-lethal methods to immediately break up concentrations of moose if they are in conflict with other forestry management goals in the area, for example, sapling stands of regenerating scots pine. for researchers, the improved gps/gsmtechnique will mean less ad-hoc testing and less guess work in terms of what periods are interesting for data collection to test behavioral hypotheses. there is a great potential for time (habitat budget), disturbance studies, detailed movement, and habitat studies. the notion that the gpm/gsm-technique is only possible to use in europe or other densely populated areas is not necessarily true. the gsm-system is expanding in canada and the us, and moose populations near some urban settlements are growing. we suggest that a few moose equipped with gps/gsm collars which report to an interactive web-page will have fundamental importance; e.g., public acceptance when re-introducing moose. it will furthermore be of great help to make research more acceptable and stimulate public interest in wildlife management. acknowledgements we want to thank dipl.-ing. robert schulte from vectronic aerospace gmbh for hardand software development and mats högström for assistance on arcims issues. k.g. poole and an anonymous reviewer improved the manuscript with constructive comments. the project is financed by the county board administration of västerbotten, the forestry owners association, and the swedish association for hunting and wildlife management. references andersson, e. 2002. älginventering från flyg i borgsjö 2002. länsstyrelsen i västerbottens län. http://www.ac.lst.se/ jakt/algjakt/alginventering/. ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7:39-47. dettki, h., r. löfstrand, and l. edenius. 2003. modeling habitat suitability for moose in coastal northern sweden: empirical vs. process-oriented approaches. ambio 32:549-556. edenius, l. 1997. field test of a gps location system for moose alces alces under scandinavian boreal conditions. wildlife biology 3:39-43. ericsson, g., k. wallin, j. p. ball, and m. broberg. 2001. age-related reproductive effort and senescence in freeranging moose alces alces. ecology 82:1613-1620. _____, and k. wallin. 2001. age-specific moose alces alces mortality in a predator free environment: evidence for senescence in females. ecoscience 8:157163. (esri) environmental systems research institute. 2002. using arcims. environmental systems research institute incorporated, redlands, california, usa. fielitz, u. 2003. remote gps-data transmission via mobile phone. http:// alces vol. 40, 2004 dettki et al. real-time moose tracking 21 www.environmental-studies.de/products/02/gps-gsm_collars.html. heezen, k. l., and j. r. tester. 1967. evaluation of radio-tracking by triangulation with special reference to deer movements. journal of wildlife management 31:124-141. hundertmark, k. j. 1998. home range, dispersal and migration. pages 303-336 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. microsoft corporation. 2000. microsoft internet information services 5.0 documentation. microsoft press, redmond, washington, usa. moen r., j. pastor, and y. cohen. 1997. accuracy of gps telemetry collar locations with differential correction. journal of wildlife management 61:530539. _____, _____, _____, and c. c. schwartz. 1996. effects of moose movement and habitat use on gps collar performance. journal of wildlife management 60:659668. raab, b., and h. vedin, editors. 1995. sveriges nationalatlas. klimat, sjöar och vattendrag. sna, stockholm, sweden. rempel r. s., a. r. rodgers, and k. f. abraham. 1995. performance of a gps animal location system under boreal forest canopy. journal of wildlife management 59:543-551. richter, j. 2002. applied microsoft .net framework programming. microsoft press, redmond, washington, usa. sandegren, f., l. pettersson, p. ahlqvist, and b. o. röken. 1987. immobilization of moose in sweden. swedish wildlife research supplement 1:785-791. struch, m., c. angst, and r. eyholzer. 2001. möglichkeiten und grenzen des globalen positionierungssystems. ein methodenvergleich der gpsund vhftelemetrie im gebirgswald. wildark & k o r a . v e r e i n e f ü r w i l d t i e r biologische forschung. bern, switzerland. teliasonera. 2003. annual report 2003. h t t p : / / v p 0 3 4 . a l e r t i r . c o m / teliasonera_annualreport_2003en/. turner, l. w., m. c. udal, b. t. larson, and s. a. shearer. 2000. monitoring cattle behavior and pasture use with gps and gis. canadian journal of animal science 80:405-413. alces 47, 2011 a journal devoted to the biology and management of moose edward m. addison ecolink science vince f. j. crichton manitoba conservation printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 brian e. mclaren lakehead university kristine m. rines new hampshire fish and game associate editors chief editor peter j. pekins university of new hampshire submissions editor gerald w. redmond maritime college of forest technology business editor arthur r. rodgers ontario ministry of natural resources murray w. lankester lakehead university (retired) f:\alces\vol_39\p65\3905.pdf alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 161 the role of mammals as ecosystem landscapers a. r. e. sinclair centre for biodiversity research, 6270 university boulevard, university of british columbia, vancouver, bc, canada v6t 1z4 abstract: the role of mammals in ecosystems is to modify vegetation structure, alter pathways of nutrients, and thereby change species composition. their large-scale structuring effects make large mammals ‘ecological landscapers’. through this they influence ecosystem function and biodiversity. landscaping effects occur when mammals are regulated by food, rather than by predators. this condition is constrained by four factors: when (1) body size is large enough to avoid predators; (2) populations adopt large scale migration behaviour because predators are unable to follow them; (3) in multispecies communities (savanna, grasslands) with a range of predator and prey sizes, only the largest species can avoid predation because they subsidize predators that regulate smaller prey species; and (4) in single predator-prey systems (tundra, desert, boreal, and temperate forests), ecological conditions determine whether or not predators regulate prey. the structuring role of mammals in maintaining species diversity is evident not just in vegetation, but also in birds, other mammals, and invertebrates. this role makes them prime candidates as ‘umbrella species’ for conservation. protection of large mammal species and their habitats also conserves a large part of the remaining community. it also means that such mammals become the ‘indicator species’ for the health of the ecosystem. alces vol. 39: 161-176 (2003) key words: biodiversity, ecological landscapers, ecosystem function, mammals in terrestrial ecosystems, plants form the basis of all communities, in terms of both structure and function. the abiotic environment sets the particular form of structure, with cold and dry climates having slow and intermittent flow of resources. the structure moves from single layer lichens in the antarctic through arctic tundra to complex rainforests of the tropics. thus, plants determine the niche possibilities of animals. to what extent do mammals alter the basic structure of communities set by plants? because all animals in terrestrial systems depend directly or indirectly on plants, they must to some degree alter plant structure, rates of flow, and species composition. mammals, even at their most abundant, are numerically insignificant in comparison to such groups as birds and reptiles, not to mention insects, protozoa, and protists. nevertheless, mammals impact plant structure and function to a greater extent, relative to their abundance, than any other animal group. mammals are “ecological landscapers”. because mammals change the physical and biotic landscape, they can affect ecosystem function (hurlbert 1997, paine 2000). this impact leads naturally to the conservation issues of umbrella species and indicator species (landres et al. 1988). can mammals act as umbrella species purely because of their large scale structuring and functional effects so that in protecting them they also protect most other species? can they also act as surrogates for other groups so that easily detectable trends in mammals reflect similar trends in other less obvious groups? mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 162 mammals as ecological landscapers ‘ecological engineers’ are species that change the physical state of the biotic or abiotic environment in which they live, and thereby, alter the flow of resources to other species (jones et al. 1994, 1997). thus, in marine environments corals create their own local environment as well as largescale habitats for other species. perhaps, termites create similar households on a local scale in terrestrial systems. however, such ‘engineering’ operates on relatively local scales. there are also processes that take place on much larger scales (landscapes, watersheds, biomes) and determine not only physical structure, but also function and species composition of whole ecosystems. an example of one such process is fire, which in savanna biomes of africa, australia, and south america changes plant succession from a ‘fireless climax’ to a ‘fire disclimax’. in savanna, fire typically impedes the succession of trees and promotes grassland and fire tolerant herbs (frost 1985a). fire operates as a probability function on a biome scale, while individual fire events occur at least at landscape scales. mammals can have analogous effects to fire in savanna systems (hobbs 1996, sala et al. 1996). one could even describe a ‘mammal disclimax’ where plant succession is held in a different state as a result of the restructuring imposed by mammals. thus, mammals act as ecological landscapers. such impacts are evident in most terrestrial biomes where mammals are abundant. boreal and taiga forests the boreal forests of canada are dominated by a few species of conifer trees, in particular the white spruce (picea glauca). the dominant mammal there is the snowshoe hare (lepus americanus) that depends largely on woody shrubs such as willow (salix spp.) and birch (betula spp.) during winter. every 10 years hares reach high numbers and at those times they eat all of the terminal shoots of small white spruce within their reach (usually up to 120 cm) (fig. 1). the slow growing spruce then take about 10 years to recover from this browsfig. 1. the frequency of snowshoe hare browse heights on terminal shoots of small white spruce, yukon, canada. hares can prevent trees from growing above about 1 m for several decades. alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 163 ing, develop a new shoot, and add another 10 cm in height, only to have this browsed off at the next peak. thus, a tree, 50 cm high, may take 50-80 years to reach the escape height, and some of them never do so and die. in contrast, trees protected from browsing grow at an accelerating rate and can escape within 10 years. the boreal forest is subject to fires from lightening strikes and so a mosaic is formed of patches at different ages since a fire occurred. the effect of hares is to keep these patches in an open state for a century or more, creating a landscape suitable for other plants that like open areas (many herbs and shrubs) and indirectly promoting their own food supply (krebs et al. 2001.). furthermore, experiments have shown that hare numbers decline because of a combination of lack of food and an increase in predators. experimental increase in rate of food supply plus removal of predators kept hare numbers high for an extended period. the inference is that under these conditions hares could prevent forest regeneration altogether. there are similar browsing effects by moose (alces alces) on aspen and other species at isle royale, u.s.a. (risenhoover and maass 1987, pastor et al. 1988, mcinnes et al. 1992, pastor and naiman 1992), and on birch in scandinavia (danell et al. 1985, danell and bergstrom 1989, danell et al. 1994). moose abundance influences the density of trees and hence composition of the habitats, but these effects also depend on the abiotic conditions and composition of other vegetation. in general, a forest with many moose is different structurally from one without them. in both banff national park, canada, and yellowstone national park, usa, elk (cervus elaphus) browsed juvenile trees of aspen and willow (salix spp.) so intensively that they changed the landscape from a dense conifer and aspen woodland to an open parkland (grassland with scattered mature trees), and maintained it thus for 40 years (houston 1982,white 2001). temperate woodland the deciduous hardwood forests of north america are the home of white-tailed deer (odocoileus virginianus) (mcshea et al. 1997). the preferred habitat of these deer is young forest, regenerating after fire (or logging). in these conditions deer can reach numbers that prevent forests from regenerating, holding them in an open shrub state (schmitz and sinclair 1997). deer can maintain this state while inhibiting hemlock (tsuga canadensis) regeneration and extirpating rare herbaceous plants (alverson and waller 1997). without deer, forests proceed to dense hardwood forests with abundant infrastructure in the shrub layer that forms the nesting habitat for rare birds such as the kentucky warbler (oporornis formosus) (mcshea and rappole 1997). in essence, there are two woodland states and which state prevails is determined by the abundance of deer. tropical forest tropical forests, at least in the holocene, have been far less subject to major structuring forces of mammals. in both africa and asia, elephant species inhabit the forests but their large-scale impact (as opposed to local feeding effects) remains unknown. this is a subject that needs research for conservation reasons. i will refer later to the paleohistorical effects of megaherbivores in forests. mammals, however, do influence the distribution of trees in tropical forests and, thereby, the species diversity of tropical trees. in turn, the dispersion of trees influences the extraordinary diversity of insects that live in these trees (janzen 1970). mammals such as bush pig (potamochoerus p o r c u s ) a n d d u i k e r ( c e p h a l o p h u s , sylvicapra) species in african lowland formammals as ecosystem landscapers – sinclair alces vol. 39, 2003 164 est, and peccaries (pecari, tayassu), deer, pacas(agouti spp.), agoutis (dasyprocta spp.), and rats in new world tropical forests concentrate on fruit under parent trees, removing most of them, and transporting a few to other areas. we can see the effect of mammals by examining tree distributions in areas where mammals have been removed. thus, in tropical forests of costa rica disturbed by agriculture, mammals are at low density from hunting. the undispersed seeds of the tree, cassia grandi, are at high density and suffer unusually high mortality from bruchid beetles (janzen 1971). similarly, tropical islands often lack the large vertebrates found on the mainland, and tree population structure differs markedly. in puerto rican forests, trees such as trophis (moraceae) have dense stands of young trees under the canopy not seen in the costa rican mainland where seeds are removed by native rodents and birds (janzen 1970). savanna savannas are grasslands with scattered trees in the tropics and subtropics, typical of africa and australia, but with some representation in southern asia and south america. they intergrade with broad-leaved evergreen dry woodlands, particularly the ‘mopane’ woodland of southern africa and eucalypt woodland of australia. large mammals in africa impose substantial structural impacts in these biomes. in southern africa, heavy grazing of grasses by ungulates (monocot feeders) alters the balance of water relations between the tree component and the herb layer so that trees and shrubs become dominant. in turn browsing ungulates (dicot feeders) such as impala (aepyceros melampus ), greater kudu (tragelaphus strepsiceros), and giraffe (giraffa cemlopardalis) benefit and their numbers increase (walker 1985, owensmith 1988). fire is required to change the tree dominated vegetation state back to a grassland state. this was demonstrated in the acacia savanna of the serengeti-mara ecosystem, east africa (norton-griffiths 1979). between 1890 and 1950 acacia and related trees dominated the vegetation. a 20-year period (1950s, 1960s) of severe burning, where 80 % of the system was burnt each year, resulted in virtually no tree seedlings escaping fire. eventually senescence of mature trees, expedited only to a minor degree by african elephants (loxodonta africana), resulted in a grassland state. it is in this grassland state that elephants played their major structuring role (dublin et al. 1990, dublin 1995). elephants, by systematically browsing tree seedlings, were able to prevent regeneration of the trees and so hold the vegetation in a grassland state. later, in the 1980s, elephants were removed by poachers and trees regenerated in abundance forming dense thickets (fire having been reduced through grazing by wildebeest, connochaetes taurinus; sinclair 1995). thus, it was the combination of browsing by elephants and grazing by wildebeest that determined which of the two vegetation states, savanna or grassland, persisted (sinclair and krebs 2002, fig. 2). the return of savanna has resulted in an increase in impala. elephants as a structuring force were to be seen in another area of east africa, the tsavo national park, kenya. in the period 1850 – 1900, the ivory trade in east africa decimated elephant populations and none were to be found in the tsavo of the 1890s (patterson 1907). furthermore, the african tribe that lived around tsavo, the wakamba, were traditional elephant hunters and kept numbers low in their area during 1900 – 1949. the vegetation without elephants was dense shrubland with scattered trees. it was sufficiently dense that hunters had to crawl along tunnels made by alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 165 black rhino (diceros bicornis) (patterson 1907). in 1949, tsavo national park was formed and it contained few elephants and dense vegetation. elephants, now released from hunting, increased rapidly with the abundant food supply. they reduced the shrub and tree densities in inverse proportion to the distance from water. the population increase ceased with an initial die-off due to starvation (corfield 1973) and has since stabilized. in the presence of elephants the landscape is a mosaic of open grassland, shrubland, and savanna. this change in vegetation structure led to declines of browsers such as gerenuck ( l i t o c r a n i u s w a l l e r i ) , l e s s e r k u d u (tragelaphus imberbis), and giraffe and an increase in grazers such as zebra (equus burchelli) and buffalo (syncerus caffer) (parker 1983, owen-smith 1988). giraffe also structure savanna vegetation through two effects. one effect is by keeping small trees at a low height (1 – 2 m) through constant browsing (pellew 1981, 1983). this height makes small trees vulnerable to fire, whose effects can reach up to 3 m. when a browsed tree in this height range experiences a hot fire, all its above ground biomass is killed, and it must regrow from the rootstock (dublin 1986). thus, giraffe indirectly reduce the density of trees and maintain a more open vegetation structure. furthermore, abundant low level tree seedlings benefit other browsers such as greater kudu and impala (owen-smith 1988). the second effect of browsing is on mature trees that escape the fire window. these trees can be sculptured into a variety of shapes beginning with a ‘top’ shape, broad at the base and with a narrow central pole that giraffe cannot reach. this shape grows into an hourglass form as the central pole spreads out above the reach of giraffe. eventually, the tree forms the characteristic flat top or umbrella top. this phenotype is the direct result of giraffe browsing, for in their absence we see trees with branches that droop down to the ground, forming dense thickets. these differences in tree morphology determine their suitability as nest sites for birds. grasslands mammals play a major role in structuring grasslands, especially wet grasslands and swamps, in temperate and tropical regions. the treeless eastern serengeti plains are composed of short height grasses such a s a n d r o p o g o n g r e e n w a y i a n d sporobolus spicatus. they also support a large number of small herbaceous dicots. the structure and species composition of these plains is maintained by near continuous grazing from the large herds of migratory wildebeest and zebra (mcnaughton and sabuni 1988). when wildebeest numbers were reduced to 20 % of the present day population, as a result of rinderpest fig.2. savanna trees in serengeti can exist in two states with the same density of elephants. a high tree density in the 1950s (squares) was reduced by fires in the 1960s (triangles) through the inhibition of seedlings. the resulting low tree density was maintained by elephant browsing of seedlings (circles). only when elephants were removed by poaching (1980s) was the high tree density state restored (data from dublin et al. 1990, dublin 1995, personal observations. redrawn from sinclair and krebs 2002). mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 166 mortality that persisted for some 70 years (1890-1963), these eastern grasslands changed in structure, becoming taller (1 m). similar tall grasslands currently on the western serengeti plains show that most dicot species become overshadowed and drop out. the impact of wildebeest grazing is demonstrated from the measurement of grass consumption on the short grass plains compared to the long grass plains (fig. 3) (mcnaughton 1984, mcnaughton and sabuni 1988, augustine and mcnaughton 1998). furthermore, our studies reveal that the long grass structure provides the habitat for a wide range of grassland bird species including grass warblers (cisticola spp.) and larks (alaudidae). long-term grazing results in a short-grass structure. this in turn provides the habitat for a different bird community, including species such as the red-capped lark (calandrella cinerea), capped wheatear (oenanthe pileata), and desert cisticola (cisticola aridulus). grasshopper species also change with grass structure. thus, wildebeest grazing creates a niche for several different groups of species in the grassland community. similar changes in the dicot herbs occurred with grazing in the flooded pampa grasslands of argentina (sala 1988). in tundra biomes caribou or reindeer (rangifer tarandus) reduce lichen cover and promote crustose lichens and bryophytes (den herder et al. 2003). along the arctic shoreline of canada, geese rather than mammal herbivores are the major determinants of structure (jefferies et al. 1994). however, the low impact of mammals is a recent event. i address the pleistocene impacts below. swamps grazing by wildebeest, and their effect on grass structure, illustrates the process known as ‘facilitation’ whereby one species provides niches for other species. elephants can facilitate the coexistence of other ungulates in swamp grasslands by breaking down, trampling, and feeding on very tall (3-5 m high) woody grasses. the young regenerating shoots of these grasses combined with other species that can grow in the openings provide the niche for african buffalo, topi (damaliscus korrigum), and waterbuck (kobus defassa), a sequence that was classically named the ‘grazing succession’ (vesey-fitzgerald 1960). the kafue flood plains in zambia are annually flooded to depths varying from a few cm to several meters. kafue lechwe (kobus leche), a semi-aquatic antelope, occur at high density and impose considerable grazing pressure. the vegetation exhibits clear zonation determined by the degree of grazing, which in turn is determined by the degree of flooding and depth of water. in shallow zones, grazing is year round and grasses such as panicum repens have evolved a leaf structure that remains under water and protected from grazing. in deep water, out of reach of grazers for part of the year, grasses like vossia cuspidate fig. 3. serengeti migrant herbivores maintain the short grasslands by consuming most of the growth (triangles), measured from small exclosure plots. in contrast, herbivores have less impact on the long grass plains and savannas as seen from the lower consumption rates (circles). the taller the grass the less is consumed (data from mcnaughton and sabuni 1988). alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 167 have evolved a canopy that grows above the water surface (ellenbrock and werger 1988). it is likely that these different growth forms will support different animal communities. d e s e r t s in australia, burrowing bettongs (bettongia lesueur), a small macropod marsupial the size of rabbits, now extinct on the mainland, were thought to structure the vegetation over large areas of acacia shrubland (mulga) (noble 1999, noble et al. 2001). they formed large underground warrens from which they commuted several kilometres to feed on shrubs. under certain fire frequencies they were apparently able to prevent shrub regeneration and maintain an open herbaceous structure. the removal of bettongs through invasion of exotic red fox predators (short et al. 2000) changed the landscape to dense stands of acacia shrubs. incidentally, the pastoral value of these landscapes was probably much higher in the presence of bettongs than at present with the unutilized stands of shrubs. the chihuahuan deserts of mexico are presently characterized by dense stands of cactus such as opuntia. currently, cattle open up these stands but bison (bison bison), elk, and peccaries did so previously. openings increase diversity of herbs that are redistributed by mammals (janzen 1986). mammals and ecosystem processes mammals influence the rates of nutrient cycling in addition to altering physical structure. in boreal forests moose decrease nitrogen mineralization of the soil by decreasing the return of high quality litter: their browsing on deciduous trees reduces their leaf fall while promoting low quality white spruce inputs (pastor et al. 1988, 1993; pastor and danell 2003). in contrast, soil nitrogen cycling in yellowstone and other prairie areas of the u.s.a. is increased by large mammal grazers (hobbs 1996, frank and evans 1997). the woodland-savanna biomes of africa are dichotomous in terms of nutrients. soils formed from old granitic rocks are sandy, heavily leached, and low in nutrients (dystrophic), particularly in calcium and phosphorus. these areas tend to be in the miombo woodlands of southern africa. in contrast, soils formed from volcanic origins such as basalt are high in nutrients and are fine, clay types (eutrophic). these are more frequently found in east african acacia savanna (bell 1982, frost 1985b, naiman and rogers 1997). however, superimposed on this pattern is an effect of large ungulates: high soil nutrients lead to high ungulate densities, rapid grazing/browsing offtake, and high fecal deposition. nutrients in the feces are then returned rapidly to the soil to be used by forage plants. in essence, ungulates fertilize their own food, and thereby, create a positive feedback increasing their own density. these effects are observed in the medium rainfall range of 500-1000 mm per year (botkin et al. 1981). in arid areas there is high soil nutrient but insufficient rain to promote recycling, whereas in very wet (forest) areas there is too much leaching and insufficient herb layer to support high densities of ungulates. at a smaller scale within the serengeti system, mcnaughton et al. (1997) have found that concentrations of non-migratory ungulates occur at localities naturally high in sodium. these are similar to the ‘sodic’ sites found in kruger national park, south africa, and at yellowstone and prairie sites in north america (tracy and mcnaughton 1995). on the serengeti sites the concentration of ungulates produces higher levels of soil nutrients and hence higher nitrogen mineralization rates. such sites have been dubbed ‘hotspots’. mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 168 on the subarctic heathlands of finland reindeer reduce lichen biomass, which allows a higher mineralization by soil microbes. lichens are so efficient at removing nitrogen from rainwater that they reduce the amount reaching the soil (den herder et al. 2003). mammals and plant species composition black-tailed prairie dogs (cynomys l u d o v i c i a n u s ) a n d p o c k e t g o p h e r s (geomys bursarius), rodents that live in large colonies on the prairies of north america, provide one of the classic examples of mammals that structure the landscape, alter the species composition of the vegetation, and so facilitate other herbivores (huntly and inouye 1988, whicker and detling 1988a,b). miller et al. (1994) suggest that they act as ‘keystone species’ through their disproportionately large influence on vegetation composition. studies at wind cave national park show that prairie dogs graze grasses to a low level (a few cm) around their colonies. constant grazing changes grass species composition to low growing forms and many dicot species survive due to reduced competition from grass. american plains bison preferentially graze these short grasses and pronghorn antelope (antilocapra americana) feed on the dicots. originally prairie dogs affected at least 20 % of the prairies (coppock et al. 1983, whicker and detling 1988a,b). exclusion of prairie dogs and bison returns the vegetation to long grass prairie (cid et al. 1991). elk (cervus elephus) maintain grass patches in the understory of old growth conifer forests of olympic and yellowstone national parks, u.s.a. exclusion of elk results in grass species being replaced by mosses, ferns, and shrubs (schreiner et al. 1996, augustine and mcnaughton 1998). rabbits on the short grass chalk grasslands of sussex, england (the south downs), determine both their structure and plant composition. these effects were detected when marked vegetation changes took place after rabbits were removed through the epizootic myxomatosis in 1953 (ross 1982). short grasses and many dicots were replaced with tall tussock grasses, and there were subsequent changes in ants and lizards dependent on these plant forms. on tundra and subarctic heathland, selective grazing by reindeer keeps the community in an early succession stage. they reduce the preferred and competitively dominant cladina lichens, which allows other lichens, bryophytes, and dwarf shrubs to exist, but it also reduces regeneration of conifer seedlings (den herder et al. 2003). constraints on the role of mammals as ecosystem processors fundamentally mammals restructure landscapes when they are food limited. herbivore populations that are regulated by predators have only selective effects on vegetation, often in local (small-scale) predator-safe areas. what are the conditions, therefore, that determine when mammal herbivores are food limited? in essence there are 4 main conditions that affect the cause of regulation: body size very large herbivores are simply too large for predators. clearly elephants, rhinos, and hippos fall into this category despite a few newborn animals being killed occasionally. in addition, even animals the size of african buffalo and giraffe are large enough that predators have difficulty killing them. the consequence is that predation accounts for only 25 % of annual mortality in buffalo (sinclair 1977, 1979). it appears that wood bison (b. bison athabascae) also falls in this category, since a population alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 169 in the mackenzie bison sanctuary of canada continues to expand despite wolf predation of juveniles (larter et al. 2000). migration predators cannot follow, on a yearround basis, animals that migrate. this general rule is evident in all mammal migration systems including those by caribou (rangifer tarandus) in northern canada, white-eared kob (kobus kob), in sudan, gazelles in botswana, and wildebeest and gazelles in serengeti (sinclair 1983, fryxell and sinclair 1988). these species, therefore, escape from predator regulation. in addition, migration is an adaptation to access ephemeral, high-quality food resources not available to non-migrants. these two features of migration systems allow populations to become an order of magnitude greater in number compared to residents. thus, migrant wildebeest in serengeti occur at 50 animals/km2 compared to a sympatric resident population at 5 animals/ km2. such large populations have major structuring effects on the ecosystem. diversity of herbivore and carnivore guilds in some systems there is a high diversity of large mammal herbivores and carnivores. nearly all are associated with tropical savanna and grassland. whether a population of herbivores is limited by predators, and so has little structuring effect on vegetation, is determined by its place in the hierarchy of herbivores. in african savanna there are as many as 10 co-existing canid or felid carnivores feeding on ungulates, lagomorphs, and rodents. they vary in size from the 200 kg lion (panthera leo) to the 5 kg wild cat (felis sylvestris). the larger the carnivore, the greater is its range of prey sizes (fig. 4). thus, the diet of lions ranges from buffalo (450 kg) to dik dik (madoqua kirkii) (10 fig. 4. the range of mammal prey sizes for serengeti carnivores. the larger the carnivore the greater the prey range. thus, small ungulates can have as many as 9 different predator species whereas larger prey have only one or none (avenant and nel 1997, personal observations). kg), a small antelope, whereas that of caracal (felis caracal) ranges from duiker (15 kg) to 100 g rodents (avenant and nel 1997). the consequence of this is that in the serengeti system, for example, smaller ungulates have as many as 7 predator species, intermediate sized antelope such as topi (damaliscus korrigum) (120 kg) have 4 predators, and very large ones like eland (taurotragus oryx) have only one. thus, smaller ungulates must experience more predation and be predator regulated if they are not migrants, and so have less effect on vegetation. direct measures of mortality by predators are consistent with this prediction. we have found that in small antelope such as oribi (ourebia ourebi) (10 kg), nearly all mortality of adults is accounted for by predators; in zebra (equus burchelli) (250 kg) this is about 73 %, while in buffalo it is 23 % (sinclair 1977). therefore, in a multi-species mammal community, large herbivores will structure the landscape whereas smaller ones cannot. the effect of large herbivores on mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 170 vegetation structure is enhanced because they also feed in a broader range of habitats than do smaller species. thus, elephants can range from tropical forests to deserts whereas the smallest ungulates are confined to single habitats. the direction of regulation in many biomes of the temperate and arctic regions there is only one major predator and one or a few species of mammalian prey. landscape structure and composition in these areas are determined by whether there is top-down or bottom-up regulation. in some cases the direction of regulation is obligatory, in others, both top-down and bottom-up regulation can occur with the system switching from one state to another depending on disturbances from outside the system. obviously, bottom-up regulation occurs when predators are absent. in this case browsing mammals must have an impact on the rate of regeneration of juvenile woody plants simply from feeding. however, such effects would not necessarily prevent the formation of the eventual mature climax. the issue is whether mammals can hold the vegetation in a different state. in some cases they do. in europe, where predators have been removed, browsing by red deer (c. elaphus), wild boar (sus scrofa), and chamois (rupicapra spp.) changed forested areas into grasslands and held them there (miller et al. 1982); these browsers maintain a different state. examples of obligatory top-down regulation are less prevalent. wolves apparently regulate moose in parts of canada (messier 1994). when wolves were absent on isle royale, moose had major effects on vegetation structure. even when wolves arrived on the island their numbers appeared to track moose numbers and did not regulate that population (peterson and vucetich 2001). in other systems, either top-down or bottom-up regulation can occur with consequent differences in the landscape structure. for example, i have mentioned above the effect of elk in transforming aspen stands to open parklands in banff national park, canada. this occurred when wolves were removed in the 1930s. in 1985 wolves reappeared and since then they have both reduced elk numbers and confined them to a subset of the habitats. in elk-free areas, young trees are regenerating and more dense aspen stands should appear in the next decades (white 2001). the recent increase of wolf numbers in yellowstone should also change the aspen structure there. thus, which of the two landscapes occur (woodland or parkland) is determined by the presence or absence of both mammalian herbivores and carnivores. (note, either state is a normal or ‘natural’ configuration of indigenous species, neither should be regarded as aberrant in terms of conservation and management). megaherbivores and paleohistory present-day elephants, rhinos, and other megaherbivores clearly play a major role in shaping the landscapes in which they live. we must remember that megaherbivores were far more abundant in the pliocene and pleistocene, even 20,000 years ago they were common, and they died out a mere 10,000 years ago on all continents except africa (owen-smith 1988). before this, even larger ancient mammals such as brontotheres (30 m b.p.), indricotheres and chalicotheres (20 m b.p.), and litopterns (10 m b.p.) must have imposed important evolutionary pressures on the vegetation. mammoths, mastodons, and woolly rhinos roamed the tundra of the holarctic. on tundra only 13 % of annual production is currently consumed (c. j. krebs, personal communication), the remainder in the pleistocene most likely eaten by the mamalces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 171 moths. mastodons fed on trees and shrubs in both the boreal and tropical rain forests of the new world while giant ground sloths and glyptodonts fed in mexican deserts (janzen 1986). thus, the impacts of mammals in these biomes would have been more similar to those we currently see in african landscapes. as in africa, the biomes of eurasia and the americas would have evolved adaptations to tolerate or mitigate the impacts of megaherbivores, adaptations that are probably still present. janzen and martin (1982) have suggested that many of the seedpods in costa rican forests are designed for ingestion and dispersal by mastodons, and janzen (1986) proposes the fleshy fruits of opuntia cactus are designed for extinct large mammals. the lack of such dispersal agents today suggests new world forests now have a different structure and species dispersion pattern from the one when large mammals were common. similarly, the present-day boreal forests of north america and asia, characterized by homogeneous stands with few tree species, were once a mosaic of mature conifers and regenerating deciduous patches, and with a more diverse fauna associated with the mosaic. thus, the lack of large mammal impacts today could be an artifact of human induced extinctions in the recent past. conclusion the main theme of this paper is that the role of mammals in ecosystems is to modify vegetation structure, alter pathways of nutrients, and thereby change species composition. the large-scale structuring effects make large mammals ‘ecological landscapers’ that influence both ecosystem function and biodiversity. moose in boreal forests provide a classic example of such effects. landscaping effects occur when mammals are regulated by food, a bottom-up trophic process, rather than by predators. this condition occurs, or is constrained, by 4 factors: when (1) body size is large enough to avoid predators; (2) populations adopt large scale migration behaviour because predators are unable to follow them; (3) in multi-species communities with a range of predator and prey sizes (savanna, grassland)¸ only the largest species can avoid predation because they simultaneously subsidize predators that regulate smaller prey species; and (4) in single predator-prey systems (tundra, desert, boreal, and temperate forests) ecological conditions determine whether or not predators regulate prey. the landscaping role of mammals in maintaining species diversity is evident not just in vegetation, but also in birds, other mammals, and invertebrates. this role makes them prime candidates as ‘umbrella species’ for conservation. protection of large mammal species and their habitats also conserves a large part of the remaining community. such mammals become the ’indicator species’ for the health of the ecosystem. thus, if the wildebeest population were to go into a prolonged decline, for example, the serengeti ecosystem would disappear. acknowledgements i thank john pastor for insightful comments and anne sinclair for much help in the preparation of this paper. funding was provided by the canadian natural sciences & engineering research council. the norwegian department of nature management supported my travel to norway, and the csiro sustainable ecosystems division, australia, provided facilities during the writing. references alverson, w. s., and d. m. waller. 1997. deer populations and the widespread failure of hemlock regeneration in northmammals as ecosystem landscapers – sinclair alces vol. 39, 2003 172 ern forests. pages 280-297 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance. smithsonian institution press, washington, d.c., usa. augustine, d. j., and s. j. mcnaughton. 1998. ungulate effects on the functional species composition of plant communities: herbivore selectivity and plant tolerance. journal of wildlife management 62: 1165-1183. avenant, n. l., and j. a. j. nel. 1997. prey use by syntopic carnivores in a strandveld ecosystem. south african journal of wildlife research 27:86-93. bell, r. h. v. 1982. the effect of soil nutrient availability on community structure in african ecosystems. pages 359-404 in b. j. huntley and b. h. walker, editors. ecology of tropical savannas. ecological studies 42. springer-verlag, berlin, germany. botkin, d. b., j. m. mellilo, and l. s-y. wu. 1981. how ecosystem processes are linked to large mammal population dynamics. pages 373-387 in c. w. fowler and t. d. smith, editors. dynamics of large mammal populations. john wiley & sons, new york, new york, usa. cid, m. s., j. k detling, a. d.whicker, and m. a. brizuela. 1991. vegetational responses of a mixed-grass prairie site following exclusion of prairie dogs and bison. journal of range management 44:100-105. coppock, d. l., j. k. detling, j. e. ellis, and m. i. dyer. 1983. plant-herbivore interactions in a north american mixedgrass prairie. i. effects of black-tailed p r a i r i e d o g s o n i n t r a s e a s o n a l aboveground plant biomass and nutrient dynamics and plant species diversity. oecologia 56:1-9. corfield, t. f. 1973. elephant mortality in tsavo national park, kenya. east african wildlife journal 11:339-368. danell, k., and r. bergstrom. 1989. winter browsing by moose on two birch species: impact on food resources. oikos 54:11-18. , , and l. edenius. 1994. effects of large mammalian browsers on architecture, biomass, and nutrients of woody plants. journal of mammalogy 75:833-844. , k . h u s s -d a n e l l , a n d r . bergstrom. 1985. interactions between browsing moose and two species of birch in sweden. ecology 66:18671878. den herder, m., m-m. kytoiita, and p. niemala. 2003. growth of reindeer lichens and effects of reindeer grazing on ground cover vegetation in a scots pine forest and a subarctic heathland in finnish lapland. ecography 26: 3-12. dublin, h. t. 1986. decline of the mara woodlands: the role of fire and elephants. ph.d. thesis, university of british columbia, vancouver, british columbia, canada. . 1995. vegetation dynamics in the serengeti-mara ecosystem: the role of elephants, fire and other factors. pages71-90 in a. r. e. sinclair and p. arcese, editors. serengeti ii: dynamics, management, and conservation of an ecosystem. university of chicago press, chicago, illinois, usa. , a. r. e. sinclair, and j. mcglade. 1990. elephants and fire as causes of multiple stable states in the serengetimara woodlands. journal of animal ecology 59:1147-1164. ellenbrook, g. a., and m. j. a. werger. 1988. grazing, canopy structure and production of floodplain grasslands at kafue flats, zambia. pages 331-337 in m. j. a.werger, h. j. during, and j. t. a. verhoeven, editors. plant form and vegetation structure. spb academic alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 173 publishing, the hague, netherlands. frank, d. a., and r. d. evans. 1997. effects of native grazers on grassland n cycling in yellowstone national park. ecology 78:2238-2248. frost, p. g. h. 1985a. the responses of savanna organisms to fire. pages 232237 in j. c. tothill and j. j. mott, editors. ecology and management of the world’s savannas. australian academy of science, canberra, australia. . 1985b. organic matter and nutrient dynamics in a broadleaved savanna. pages 200-206 in j. c. tothill and j. j. mott, editors. ecology and management of the world’s savannas. australian academy of science, canberra, australia. fryxell, j. m., and a. r. e. sinclair. 1988. causes and consequences of migration by large herbivores. trends in ecology and evolution 3:237-241. hobbs, t. n. 1996. modification of ecosystems by ungulates. journal of wildlife management 60:695-713. houston, d. b. 1982. the northern yellowstone elk. macmillan, new york, new york, usa. huntly, n., and r. inouye. 1988. pocket gophers in ecosystems: patterns and mechanisms. bioscience 38:786-793. hurlbert, s. h. 1997. functional importance vs keystoneness: reformulating s o m e q u e s t i o n s i n t h e o r e t i c a l biocenology. journal of ecology 22:369382. janzen, d. h. 1970. herbivores and the number of tree species in tropical forests. american naturalist 104:501-528. . 1971. escape of cassia grandis l. beans from predators in time and space. ecology 52:964-979. . 1 9 8 6 . c h i h u a h u a n d e s e r t nopaleras: defaunated big mammal vegetation. annual reviews of ecology and systematics 17:595-636. , a n d p . s . m a r t i n . 1 9 8 2 . neotropical anachronisms: the fruits the gomphotheres ate. science 215:19-27. jefferies, r. l., d. r. klein, and g. r. shaver. 1994. vertebrate herbivores and northern plant communities: reciprocal influences and responses. oikos 71:193-206. jones, c. g., j. h. lawton, and m. shachak. 1994. organisms as ecosystem engineers. oikos 69:373-386. , , and . 1997. positive and negative effects of organisms as physical ecosystem engineers. ecology 78:1946-1957. krebs, c. j., s. boutin, and r. boonstra, editors. 2001. ecosystem dynamics of the boreal forest. oxford university press, oxford, england. landres, p. b., j. verner, and j. w. thomas. 1988. ecological uses of vertebrate indicator species: a critique. conservation biology 2:316-328. larter, n. c., a. r. e. sinclair, t. ellsworth, j. nishi, and c. c. gates. 2000. dynamics of reintroduction in an indigenous large ungulate: the wood bison of northern canada. animal conservation 3:299-309. mcinnes, p., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73:2059-2075. mcnaughton, s. j. 1984. grazing lawns: animals in herds, plant form, and coevoloution. american naturalist 124:863-886. , f. f. ba n y k w a , and m. m. mcnaughton. 1997. promotion of the cycling of diet-enhancing nutrients by african grazers. science 278:17981800. , and g. a. sabuni. 1988. large african mammals as regulators of vegetation structure. pages 339-354 in m. mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 174 j. a.werger, p. j. m. van der aart, h. j. during, and j. t. a. verhoeven, editors. plant form and vegetation structure. spb academic publishing, the hague, netherlands. mcshea, w. j., and j. h. rappole. 1997. herbivores and the ecology of forest understory birds. pages 298-309 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance. smithsonian institution press, washington, d.c., usa. , h. b. underwood, and j. h. rappole, editors. 1997. the science of overabundance. smithsonian institution press, washington, d.c., usa. messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478-488. miller, b., g. ceballos, and r. reading. 1994. the prairie dog and biotic diversity. conservation biology 8:677-681. miller, g. r., j. w. kinnaird, and p. cummins. 1982. liability of saplings to browsing on a red deer range in the scottish highlands. journal of applied ecology 19:941-951. naiman, r. j., and k. h. rogers. 1997. large animals and system-level characteristics in river corridors: implications for river management. bioscience 47:521-529. noble, j. c. 1999. fossil features of mulga acacia aneura landscapes: possible imprinting by extinct pleistocene fauna. australian zoology 31:396-402. , j. gillen, g. jacobson, w. a. low, c. miller, and the mutitjulu community. 2001. the potential for degradation of landscape function and cultural values following the extinction of mitika (bettongia lesueur) in central australia. pages 71-89 in a. j. conacher, editor. land degredation. kluwer academic publishing, netherlands. norton-griifiths, m. 1979. the influence of grazing, browsing, and fire on the vegetation dynamics of the serengeti. pages 310-352 in a. r. e. sinclair and m. norton-griffiths, editors. serengeti. dynamics of an ecosystem. university of chicago press, chicago, illinois, usa. owen-smith, n. 1988. megaherbivores. the influence of very large body size on ecology. cambridge university press, cambridge, uk. paine, r. t. 2000. phycology for the mammalogist: marine rocky shores and mammal-dominated communities how different are the structuring processes? journal of mammalogy 81:637-648. parker, i. s. c. 1983. the tsavo story: an ecological case history. pages 37-49 in r. n. owen-smith, editor. management of large mammals in african conservation areas. haum, pretoria, south africa. pastor, j., and k. danell. 2003. moosevegetation-soil interactions: a dynamic system. alces 39:177-192. , b. dewey, r. j. naiman, p. f. mcinnes, and y. cohen. 1993. moose browsing and soil fertility in the boreal forests of isle royale national park. ecology 74:467-480. , and r. j. naiman. 1992. selective foraging and ecosystem processes in boreal forests. american naturalist 139:690-705. , , b . d e w e y , a n d p . mcinnes. 1988. moose, microbes, and the boreal forest. bioscience 38:70-77. patterson, j. h. 1907. the maneaters of tsavo. macmillan, london, uk. pellew, r. a. p. 1981. the giraffe (giraffa cemlopardalis tippelskirchi matschie) and its acacia food resource in the serengeti national park. ph.d. thesis, university of london, uk. alces vol. 39, 2003 sinclair – mammals as ecosystem landscapers 175 . 1983. the impacts of elephant, giraffe, and fire upon the acacia tortilis woodlands of the serengeti. african journal of ecology 21:41-74. peterson, r. o., and j. a. vucetich. 2001. ecological studies of wolves on isle royale. annual report 2000-2001. michigan technological university, houghton, michigan, usa. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17:357-364. ross, j. 1982. myxomatosis: the natural evolution of the disease. pages 77-95 in m. a. edwards and u. mcdonnell, editors. animal disease in relation to animal conservation. academic press, london, uk. sala, o. e. 1988. the effect of herbivory on vegetation structure. pages 317-330 in m. j. a.werger, p. j. m. van der aart, h. j. during, and j. t. a. verhoeven, editors. plant form and vegetation structure. spb academic publishing, the hague, netherlands. , w . k . l a u e n r o t h , s . j . mcnaughton, g. rusch, and x. zhang. 1996. biodiversity and ecosystem functioning in grasslands. pages 130-149 in h. a. mooney, j. h. cushman, e. medina, o. e. sala, and e.-d. schulz, editors. functional roles of biodiversity: a global perspective. john wiley and sons, new york, new york, usa. schmitz o. j., and a. r. e. sinclair. 1997. rethinking the role of deer in forest ecosystem dynamics. pages 201-223 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance. smithsonian institution press, washington, d.c., usa. schreiner, e. g., k. a. krueger, p. j. happe, and d. b. houston. 1996. understory patch dynamics and ungulate herbivory in old-growth forests of olympic national park, washington. canadian journal of forest research 26:255-265. short, j., j. e. kinnear, and a. robley. 2000. surplus killing by introduced predators in australia – evidence for ineffective anti-predator adaptations in native prey species? biological conservation 103:283-301. sinclair, a. r. e. 1977. the african buffalo. a study of resource limitation of populations. university of chicago press, chicago, illinois, usa. . 1979. the eruption of the ruminants. pages 82-103 in a. r. e. sinclair and m. norton-griffiths, editors. serengeti. dynamics of an ecosystem. university of chicago press, chicago, illinois, usa. . 1983. the function of distance movements in vertebrates. pages 240258 in i. r. swingland and p. j. greenwood, editors. the ecology of animal movement. oxford university press, oxford, uk. . 1995. equilibria in plant-herbivore interactions. pages 91-113 in a. r. e. sinclair and p. arcese, editors. serengeti ii: dynamics, management, and conservation of an ecosystem. university of chicago press, chicago, illinois, usa. , and c. j. krebs. 2002. complex numerical responses to top-down and bottom-up processes in vertebrate populations. philosophical transactions of the royal society of london, series b 357:1221-1231. tracy, b. e., and s. j. mcnaughton. 1995. elemental analysis of mineral lick soils from the serengeti national park, the konza prairie and yellowstone national park. ecography 18:91-94. vesey-fitzgerald, d. f. 1960. grazing succession among east african game mammals as ecosystem landscapers – sinclair alces vol. 39, 2003 176 animals. journal of mammalogy 41:161172. walker, b. 1985. structure and function of savannas: an overview. pages 83-91 in j. c. tothill and j. j. mott, editors. ecology and management of the world’s savannas. australian academy of science, canberra, australia. whicker, a. d., and j. k. detling. 1988a. modification of vegetation structure and ecosystem processes by north american grassland mammals. pages 301316 in m. j. a. werger, h. j. during, and j. t. a. verhoeven, editors. plant form and vegetation structure. spb academic publishing, the hague, netherlands. , and . 1988b. ecological consequences of prairie dog disturbances. bioscience 38:778-793. white, c. a. 2001. aspen, elk, and fire in the canadian rocky mountains. ph.d. thesis. university of british columbia, vancouver, british columbia, canada. alces vol. 45, 2009 kaitala et al. – deer ked and moose in finland 85 deer ked, an ectoparasite of moose in finland: a brief review of its biology and invasion arja kaitala1, raine kortet1,4, sauli härkönen2, sauli laaksonen3, laura härkönen1, sirpa kaunisto4, and hannu ylönen5 1university of oulu, department of biology, p.o. box 3000, fi-90014 oulu, finland; 2finnish forest research institute, joensuu research unit, p.o. box 68, fi-80101 joensuu, finland; 3finnish food safety authority (evira), fish and wildlife health research unit, p.o. box 517, fi-90101 oulu, finland; 4university of joensuu, faculty of biosciences, p.o. box 111, fi-80101 joensuu, finland; 5university of jyväskylä, konnevesi research station, p.o. box 35, fi-40014 university of jyväskylä, finland abstract: the deer ked (lipoptena cervi) is an important ectoparasite of moose (alces alces) that has rapidly invaded finland during the last 50 years, and is currently found in southern parts of finnish lapland. we have studied the invasion, behavior, and ecology of this parasitic fly, and in this paper briefly review the effect of climate on the distribution of deer keds and our recent findings from host-choice experiments. the rapid increase of the deer ked is correlated with high moose densities in finland. we propose that the availability of suitable hosts, not climate, is the primary factor affecting its northward range expansion. our host-choice experiments indicated that deer keds are attracted by movement and large, dark objects. our results suggest that this parasite may continue to spread northwards in the near future, and that its potential impact on cervids and human health warrants attention. alces vol. 45: 85-88 (2009) key words: alces alces, climate, color preference, deer ked, hippoboscids, host choice, host search, lipoptena cervi, parasite. the deer ked (lipoptena cervi) is a bloodsucking ectoparasite of cervids especially moose (alces alces). the deer ked has a fairly extensive area of distribution including europe, some parts of siberia, northern china, and algeria in northern africa, and it has been introduced into north america (see maa 1969, dehio et al. 2004). it is assumed that the deer ked spread to finland from russia in the 1960s, and since then, has rapidly spread northward from southeastern finland (fig. 1) (hackman et al. 1983, reunala et al. 2008). in central europe and parts of scandinavia the deer ked parasitizes red deer (cervus elaphus), roe deer (capreolus capreolus) and, with a low prevalence, white-tailed deer (odocoileus virginianus) (haarløv 1964). the deer ked is also found in low numbers on wild forest reindeer (rangifer tarandus fennicus) and occasionally on semi-domestic reindeer (rangifer tarandus tarandus) (kettle and utsi 1955, kaunisto et al. 2009). in finland, the rapid invasion of the deer ked has been associated with the increase of moose density during the last 40 years. moose were regarded as almost extinct in finland during the 1930s and until the second world war. they were still threatened until the 1950s despite a slow recovery of their population. due to hunting regulations and changes in finnish forestry management, the population harvest during the 1970s increased to 50,000 moose, with a maximum annual harvest of 69,000 in the mid 1980s and 85,000 in the early 2000s (lavsund et al. 2003). currently, the moose population in finland has stabilized at about 90,000 (pusenius et al. 2008). we think that this may be the strongest single factor that enabled the rapid spread and increase of the deer ked. despite high deer ked and moose in finland – kaitala et al. alces vol. 45, 2009 86 local abundances of the deer ked in finland and high relative abundances in central and eastern europe, our current knowledge of its biology is limited. in the boreal areas of finland, adult deer keds emerge synchronously and seek hosts from late summer to the end of autumn (hackman 1977). almost immediately after finding a suitable host, usually moose, both males and females drop their wings and begin to suck blood from the host (hackman et al. 1983). the adults spend their entire lives on the same host. deer ked larvae develop inside the blood-sucking female that produces one pupa at a time; thus, the female gives birth to a single pupa. at birth, the whitish pupa turns dark during the rapid chitinization and drops to the ground vegetation layer or snow where over-wintering and development takes place (see review in haarløv 1964). our research group has experimentally investigated whether the distribution of the deer ked is affected by climate and if it could potentially spread northward in finnish lapland from its current range in the most southerly areas of reindeer husbandry (fig. 1). our results show that, although pupal hatching success decreases strongly with latitude, the deer ked has the potential to spread further north and reach the northern areas of scandinavia up to the arctic circle. even in the high arctic regions, some deer ked pupae could develop successfully (härkönen et al. 2009). in spring the pupae may survive at temperatures of -15o c and higher, but do not survive short periods of -20o c or below. even during extremely severe winter conditions, pupae may survive under the snow where temperatures are greatly moderated. thus, winter climate may not be the critical factor that regulates distribution of deer keds. we suggest that the northward spread of the deer ked will depend on the availability of suitable hosts for reproduction. even without new potential host species such as reindeer, the deer ked may be capable of expanding its range further north by depending only on moose, its major finnish host. this assumption is based on the fact that moose are notably numerous in the southern part of finnish lapland (pusenius et al. 2008). dispersal of the deer ked has been rapid in finland and its distribution has expanded, and related problems have increased for humans and domestic animals (e.g., horses and semi-domesticated reindeer). controlled fig. 1. the current distribution of the deer ked (lipoptena cervi) in finland according to the distribution records courtesy of the zoological museum of the university of oulu. sporadic observations occur 50-100 km above the current distribution area. in 50 years the deer ked expanded its range almost 1,000 km northward. the earliest records were from the southeastern corner of finland in 1960, suggesting spread from the russian side (hackman 1977). the distribution in 1970 is based on a report by hackman (1977), whereas the distribution in 1980 is based on hackman et al. (1983). alces vol. 45, 2009 kaitala et al. – deer ked and moose in finland 87 experimental infections demonstrated that deer keds can lower the physical condition of semidomesticated reindeer and cause short-term histological, physiological, and behavioral changes (s. kynkäänniemi et al., finnish food safety authority, unpubl. data). the effects of deer keds on moose are unknown, but deer keds can be extremely numerous on host moose and during the flying period. for example, up to 17,000 deer keds have been reported on a single moose (t. paakkonen et al., university of joensuu, unpubl. data). deer keds also cause major inconvenience, nuisance, and possible health threats for humans. although it has not been reported to reproduce after feeding on humans, their bites can cause serious health problems and symptoms, including chronic deer ked dermatitis (rantanen et al. 1982, reunala et al. 2008) and occupational allergic rhinoconjunctivitis (laukkanen et al. 2005). it may also act as a vector for other diseases (ivanov 1974, rantanen et al. 1982, dehio et al. 2004). our host-choice experiments (kortet et al. 2009) have revealed that the deer ked may use relatively simple cues for host choice (alekseev 1985). they are attracted by movement and to large sized, dark colored objects. studying host choice in the natural environment is challenging, but we have managed to explore this topic using volunteer human subjects. because this parasite also causes harm to humans (rantanen et al. 1982, reunala et al. 2008), more studies are needed to determine how attacks can be avoided, for example, by type and color of clothing. based on our recent studies, the colder and shorter growing season in northern finland may not constrain deer ked invasion. in northern finland, the densities of suitable (i.e., moose) and potential host species (i.e., semi-domesticated reindeer) are relatively high. as a result, the deer ked will likely continue spreading its range into new areas. continued research of the deer ked is warranted because of its potential negative impacts on wild and semi-domesticated cervids and human health. acknowledgements we would like to thank all the volunteers for their invaluable help during experiments. our studies were funded by the university of oulu (rk) and the finnish ministry of agriculture and forestry (ak, rk, sh, sl, and hy). references alekseev, e. a. 1985. initial experience with individual human protection from attack by the deer louse fly lipoptena cervi. medicinskaa parazitologia (mosk.) 6: 56-57. dehio, c., u. sauder, and r. hiestand. 2004. isolation of bartonella schoenbuchensis from lipoptena cervi, a blood-sucking arthropod causing deer ked dermatitis. journal of clinical microbiology 42: 5320-5323. haarløv, n. 1964. life cycle and distribution pattern of lipoptena cervi (l.) (dipt., hippobosc.) on danish deer. oikos 15: 93-129. hackman, w. 1977. hirven täikärpänen ja sen levittäytyminen suomeen. (a para-(a parasitic lousefly of moose and its spread in finland). luonnon tutkija 81: 75-77 (in finnish). _____, t. rantanen, and p. vuojolahti. 1983. immigration of lipoptena cervi (diptera, hippoboscidae) in finland, with notes on its biology and medical significance. notulae entomologicae 63: 53-59. härkönen, l., s. härkönen, a. kaitala, s. kaunisto, r. kortet, s. laaksonen, and h. ylönen. 2009. predicting range expansion of an ectoparasite – the effect of summer temperatures on deer ked (lipoptena cervi, diptera: hippoboscidae) performance along a latitudinal gradient. ecography, in press. ivanov, v. i. 1975. anthropophilia of deer deer ked and moose in finland – kaitala et al. alces vol. 45, 2009 88 blood sucker lipoptena cervi l. (diptera, hibboboscidae). medicinskaa parazitologia (mosk.) 44: 491-495. kaunisto, s., r. kortet, l. härkönen, s. härkönen, h. ylönen, and s. laaksonen. 2009. new bedding site examination-based method to analyse deer ked (lipoptena cervi) infection in cervids. parasitology research 104: 919-925. kettle, d. s., and m. n. p. utsi. 1955. hypoderma diana (diptera, oestridae) and lipoptena cervi (diptera, hippoboscidae) as parasites of reindeer (rangifer tarandus) in scotland with notes on the second stage larva of hypoderma diana. parasitology 45: 116-120. kortet, r., l. härkönen, p. hokkanen, s. härkönen, a. kaitala, s. kaunisto, s. laaksonen, j. kekäläinen, and h. ylönen. 2009. experiments on the ectoparasitic deer ked that often attacks humans; preferences for body parts, colour and temperature. bulletin of entomological research, in press. laukkanen, a., p. ruoppi, and s. mäkinenkiljunen. 2005. �eer ked-induced oc-2005. �eer ked-induced occupational allergic rhinoconjunctivitis. annals of allergy, asthma and immunology 94: 604-608. lavsund, s., t. nygrén, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. maa, t. c. 1969. a revised checklist and concise host index of hippoboscidae (diptera). pacific insects monographs 20: 261-299. pusenius, j., m. pesonen, r. tykkyläinen, m. wallén, and a. huittinen. 2008. hirvikannan koko ja vasatuotto 2006. (moose population size and calf production in 2006). pages 7-14 in m. wikman, editor. riistakannat 2007: riistaseurantojen tulokset (game populations 2007: results of game monitoring). riistaja kalatalouden tutkimuslaitoksen selvityksiä 5/2008. (in finnish). rantanen, t., t. reunala, p. vuojolahti, and w. hackman. 1982. persistent pruritic papules from deer ked bites. acta dermatol venereologica 62: 307-311. reunala, t., m. laine, m. vornanen, and s. härkönen. 2008. hirvikärpäsihottuma maanlaajuinen riesa. (deer ked dermatitis – a country wide nuisance). duodecim 124: 1607-1613. (in finnish). 113 estimating sustained yields for moose in central british columbia using a predator-prey model ian w. hatter nature wise consulting, 49-640 upper lakeview road, invermere, british columbia v0a 1k3, canada abstract: one of the fundamental principles of wildlife harvesting is that it must result in a sustained yield (sy), a harvest that can be taken year after year without jeopardizing future harvests. predator-prey models are rarely incorporated into estimates of sys for moose, despite predation of moose by wolf (canis lupus), grizzly bear (ursus arctos), and black bear (u. americanus) throughout much of western north america. a simple predator-prey model was parameterized from a stable moose-wolf-bear system in central british columbia during 1987–1998. modelled moose, wolf, and harvest parameters compared favourably with observed parameters when the annual rate of wolf removal (human-caused wolf mortality) was 31%. sy curves were modelled by incrementally increasing wolf removal rates from 0 to 40% while maintaining selective moose harvests of 16% bulls, 2% cows and 9% calves. sys displayed an s-shape curve with wolf removal rates, a hook-shape curve with wolf densities, and were linearly related to moose density. optimal harvests included a moderate harvest of bulls (16–21%), a nil-to-very low harvest of cows (0–0.2%), and moderate-to-high harvests of calves (15–43%) when wolf removal rates were ≥ 20%. higher cow harvest rates (2%) could be accommodated without substantially lowering sys if calf harvest rates were reduced. optimal harvest rates did not improve yields over bull-only hunting when wolf removal rates were 0–10% and management constraints were placed on adult sex ratios. this study supports previous findings that the optimal harvest strategy for moose should primarily target bulls and calves, whereas cows should be harvested minimally. however, for low-density, predator limited moose populations, bull-only harvests may provide equivalent yields while maintaining higher moose and wolf densities. alces vol. 57: 113–129 (2021) key words: alces alces, bears, harvest rate, moose, optimal harvests, predation rate, predator-prey model, sustained yield, wolves, yield-density curve one of the fundamental principles of wildlife harvesting is that it must result in a sustained yield (sy), a harvest that can be taken year after year without jeopardizing future harvests (sinclair et al. 2006). caughley (1976) proposed a general model for harvesting ungulates that produces a yield-density curve with a bell-shape, slightly skewed to the right, and with the maximum sustained yield (msy) occurring at ~ 70% of carrying capacity (k). crête (1987) expanded on caughley’s model and proposed two yield-density curves for moose (alces alces), one with wolves (canis lupus) and bears (ursus spp.) present (predator limited k = 400 moose/1000 km2), and the other without predators (food-limited k = 1000–2000 moose/1000 km2). the predator-limited curve implied that lowering moose density to 200–300 moose/1000 km2 through hunting will increase sys about 4-fold. that is, by intensively harvesting moose, wolf numbers and predation rates should decline, and moose growth rates modelling sustained yields for moose – hatter. alces vol. 57, 2021 114 and harvest should increase. gasaway et al. (1992) rejected the concept of a bell-shaped yield-density curve for moose in the presence of wolves and bears in alaska and yukon, and presented an approximate sy curve for moose where sustainable harvests increased more gradually with moose density. their empirical data suggested that wolves and bears remained effective predators on calves at very low densities, and that sys were lower than predicted by crête’s (1987) predator-limited, yield-density curve. gasaway et al. (1992) concluded that predation by lightly harvested wolf, grizzly bear (u. arctos), and black bear (u. americanus) populations in alaska and yukon lowered and maintained moose populations within a low density dynamic equilibrium (ldde, ≤ 417 moose/1000 km2), and that intensive harvesting of predators, rather than moose, was required to elevate sys. hatter (1999) developed a preliminary yield-density curve for moose in north-central british columbia (bc) that was similar in shape to the yield-density curve for alaska and yukon (gasaway et al. 1992), suggesting that some moose populations in bc may also exist within a ldde. although moose are limited by wolf and bear predation throughout much of north america (bergerud et al. 1983, gasaway et al. 1983, 1992, crête 1987, messier 1994, van ballenberghe and ballard 1994), predator-prey models are rarely used in determining sys for moose. van ballenberghe and dart (1982) used a simple conceptual model to examine harvest yields subject to wolf and bear predation and found that bull-only hunts provided an equivalent numerical yield to either-sex hunts, but had a much higher margin of safety for management errors. more recently, attention has been given to optimizing sys, or obtaining the maximum yield by selectively harvesting bulls, cows, and calves at different rates. sæther et al. (2001) concluded that the optimal harvest strategy in northern norway with few large predators, but within a fluctuating environment, involved a high harvest of calves and bulls, and that cows should hardly be harvested. nilsen et al. (2005) considered moose populations in south-eastern norway where due to strict management control there was no numerical response by wolves, and found that in the presence or absence of predation, a high proportion of calves in the harvest gave the highest sys. xu and boyce (2010) considered moose populations in central alberta subject to predation and stochastic weather events, and also concluded that when optimizing total yield, bulls and calves should be subject to intense harvest, with a low harvest of females. while the effect of predators on sys was considered by nilsen et al. (2005) and xu and boyce (2010), neither study modelled predator-prey dynamics in detail. the purpose of this study was to develop a model that explicitly considered predator-prey interactions over a range of fall wolf densities, and to use the model to investigate how predation affects sys and optimal harvesting of moose in central bc. i considered the optimal harvest to be the highest sustained yield of moose from all sex/age classes, as opposed to the greatest carcass weight or maximizing hunting opportunity. study area the 19,000 km2 study area, hereafter referred to as the prince george study area (pgsa), was located around prince george (53° 54’n × 122° 41’w) in central bc (fig. 1). the terrain is flat to rolling and forests are mainly hybrid white-engelmann spruce (picea glauca x engelmannii) and subalpine fir (abies lasiocarpa) with extensive successional stands of lodgepole pine (pinus contorta). forestry is a prominent industrial activity and cutblocks are common throughout the area. moose were probably the alces vol. 57, 2021 modelling sustained yields for moose – hatter. 115 predominant ungulate prey for wolves, black bears, and grizzly bears because other ungulates were rare, but included mule deer (odocoileus hemionus), white-tailed deer (o. virginianus), elk (cervus canadensis), and caribou (rangifer tarandus) (heard et al. 1999). the moose density estimated in 1998 was 1320 moose/1000 km2 from a stratified random block survey when predator densities were ~12.5 wolves/1000 km2, 20 grizzly bears/1000 km2, and 211 black bears/1000 km2 (heard et al. 1999). since 1980, selective fall harvests have governed moose hunting through a combination of differential licencing for bulls and cows and open seasons on 2-point bulls and calves (child 1983). from 1991 to 1998, licenced hunters harvested an average of 16% of the bulls, 2% of cows, and 9% of calves for a harvest rate of 7% of the pre-hunt population (heard et al. 1999). the first nations harvest was unknown, but likely less than the licenced harvest (heard, pers. comm.). moose densities and composition, harvests, and hunter success rates appeared to be relatively stable from 1987 to 1998. there was no trend in the number of wolves shot by licensed hunters or conservation officers during this period implying that wolf densities were also stable. heard et al. (1999) suggested that the elevated moose density was possible because wolf removals (humancaused wolf mortality) were high due to the combined effects of hunting, trapping, and killing wolves to protect livestock. grizzly bears were lightly hunted and black bears appeared to be moderately exploited (heard et al. 1997). further details of the pgsa are provided by heard et al. (1997, 1999). methods model structure moose the population was stage-structured as bulls (≥ 1 year-old males), cows (≥ 1 yearold females), and calves (< 1year-old with 50:50 sex ratio). density dependence in reproduction and survival were modelled fig. 1. location of the prince george study area (pgsa) in central british columbia, canada. the pgsa includes all of wildlife management units (mu) 707–715, and the lower portions of 716 and 724. modelling sustained yields for moose – hatter. alces vol. 57, 2021 116 followed the approach outlined by mcnay and delong (1998) and ranged from 51 to 85 calves/100 cows at birth, 15–39% summer calf mortality, 10–12% winter calf mortality, 2–12% winter cow mortality, and 4–12% winter bull mortality. all density dependent responses started at ~ 65% of k and generally increased in a linear manner; k was set to 2000 moose/1000 km2 (crête 1987, heard et al. 1999). summer adult mortality (2%) was assumed to be density independent. the modelled population was censused 3 times per year including the post-hunt, post-calving, and pre-hunt periods. wounding loss was set to 15% of the licenced harvest (kuzyk et al. 2018). first nations harvest was assumed to be unselective and occur during winter. the maximum harvest rate by first nations was presumed to be 7% of the post-hunt moose population at k, declining linearly to 0% at 100 moose/1000 km2. where not specified, harvest rates were based on the pre-hunt moose numbers, while moose densities and sex-age composition apply to the post-hunt period. wolves changes in wolf density were modelled as a ratio-dependent numerical response: w w rw w amt t t t t = + −      − − − − 1 1 1 1 1 (eberhardt 1997, hatter 2019) where wt denotes fall wolf density at time t, r is the maximal rate of increase for wolves (r = 0.46, keith 1983), a is the equilibrium ratio of wolves:moose, and mt is moose density. eberhardt et al. (2003) estimated a = 0.049 by linear regression of the observed finite rate of increase of wolves (λ) against the wolf:moose ratio. i added 6 additional data points (appendix 1) and estimated the relationship as: y e1.534 x10.72= − where y is the observed λ for wolves, x is wolves/moose, and a = 0.040 (25 moose/ wolf, 95% ci: 20–29) when y = 1.0. spring wolf numbers were estimated by subtracting the annual wolf removal from the fall population. the wolf removal rate included all human-caused mortality. several models have been proposed for estimating the annual kill rate of moose by wolves. messier (1994) found that winter kill rates were reduced at low moose densities, and proposed a type ii functional response. hayes and harestad (2000) modified the parameter estimates for the type ii functional response based on kill rates from the yukon, while eberhardt (1997) suggested kill rates were density independent and averaged 2.1 kills/wolf/100 days in winter. lake et al. (2013) concluded that kill rates were also density independent in alaska and yukon. serrouya et al. (2015), however, found that messier’s type ii response provided the best fit among competing models for a wolf-moose-caribou system in south-eastern bc. i used messier’s (1994) functional response: y x x 3.36 0.46 = + where y is the number of moose killed per wolf per 100 days in winter, and x is the number of moose/km2, but capped the maximum killing rate at 2.1 as suggested by eberhardt (1997). following eberhardt (1997), i assumed that summer kill rates were 75% of the winter rate during 5 months of summer (june–october). i used fall wolf densities for winter predation and spring wolf densities for summer predation. wolf kills were apportioned among the bulls, cows, and calves by assuming the relative vulnerability of calves was 10-fold greater alces vol. 57, 2021 modelling sustained yields for moose – hatter. 117 than adults during summer and 2-fold greater during winter. the high calf vulnerability reflected the preference by wolves for calves (peterson et al. 1984, sand et al. 2008). calculations for relative vulnerability followed mcnay and delong (1998). bears ballard (1992) reported that calf mortality rates due to grizzly bear predation ranged from 3 to 52% and predation rates were independent of moose density; mortality rates from black bears predation ranged from 2 to 50%. the predation rates of grizzly and black bears on moose calves in the pgsa were unknown. rea et al. (2019) found that only 2% of bear scats collected in the pgsa during spring and summer (n = 1381) contained moose calf hair; however, they noted that even this low percentage could result in substantial calf mortality. i assumed a combined annual calf predation rate by grizzly and black bears of 35%, that predation was density independent, and that these kill rates were additive to other forms of mortality. adult grizzly bears have been reported to kill an average of 0.5–2.2 adult moose annually (dahle et al. 2013), while black bears rarely kill adult moose (ballad 1992). i assumed that the modelled 2% summer mortality rate for adults included kills by bears. i did not consider the effects of bear hunting on moose population dynamics as it would have made the model more complex and further increased uncertainty in the model projections. model evaluation i evaluated the model by assessing how well the modelled moose and wolf parameters compared to the observed parameters in the pgsa when 16% bulls, 2% cows, and 9% calves were harvested annually. model fitting was primarily achieved by adjusting the wolf removal rate until the modelled parameters closely matched the observed parameters. i ran each simulation for 100 years to remove transient predator-prey dynamics and to ensure that moose, wolf, and harvest densities were stable. i also compared modelled estimates of moose calf mortality rates and total annual mortality rates with those from field studies in alaska and yukon summarized by boertje et al. (2009). additionally, i tested for consistency between modelled estimates of wolf density with low removal rates (0–15%) and those generated from a prey biomass regression model, where wolf densities were estimated from an ungulate biomass index at the regional level in bc (kuzyk and hatter 2014). as calf predation rates by bears were unknown, but could be substantial, i evaluated the sensitivity of different predation rates on the modelled parameters. i considered bear predation rates ranging from 10 to 50% in increments of 5%. for each predation rate, i used the optimization tool solver in microsoft excel (microsoft, redmond, washington, usa) to minimize the sums of squared differences between the observed and modelled parameters (hilborn and mangel 1997) by iteratively changing the wolf removal rate. the adjusted parameters included moose density, wolf density, moose harvest density, and moose sex:age ratios. sustained yield curve sys for moose in the pgsa were simulated by increasing annual wolf removal rates from 0 to 40% in increments of 1% while maintaining constant moose harvest rates for bulls (16%), cows (2%), and calves (9%). the sy was the moose harvest in the 100th year of the simulation. sy curves were generated for harvest density and wolf removal rates, harvest density and wolf density, and harvest density and moose density. modelling sustained yields for moose – hatter. alces vol. 57, 2021 118 i evaluated the modelled yield-density curve by comparing it to the yield-density curve developed by hatter (1999) for stable moose populations in north-central bc. optimal harvests optimal harvests were determined by iteratively changing the harvest rate for bulls, cows, and calves with solver until the maximum total harvest was achieved, and moose and wolf populations were stable. a similar procedure was used for optimizing bull and calf harvests with cows harvested at 2%, and for bull harvests with cows and calves unharvested. crête et al. (1981) recommended that in order to optimize moose harvest in south-western quebec, at least 40% bulls should be retained among adults to ensure that sex ratio-dependent fertilization was not adversely affected. the provincial moose harvest management procedure for bc (flnro 2013) states that the lower range of the post-hunt adult sex ratio should not fall below 30 bulls:100 cows, or 50 bulls:100 cows in low density (≤ 200/1000 km2) moose populations. therefore, i added a sex ratio constraint when optimizing moose harvests with solver to ensure the adult sex ratio was maintained at either ≥ 30 bulls:100 cows or ≥ 50 bulls:100 cows depending on moose density. results the modelled moose population, in the absence of hunting or predation, stabilized at 2000 moose/1000 km2, with 100 bulls/100 cows and 32 calves/100 cows (table 1). the finite rate of increase below population levels where density dependent effects became operative was λ ~ 1.3. the stable moose density was 9% lower with bears only (wolves and hunting absent), and 72% lower with wolves only (bears and hunting absent). with wolves and bears (no hunting), the stabilizing density was almost 90% lower, and consisted of 87 bulls/100 cows and 35 calves/100 cows. the estimated parameters from the predatorprey model compared favourably with the observed parameters for the pgsa during the study period when the wolf removal rate was 31%. both the modelled and observed wolf densities were 12.5 wolves/1000 km2. the modelled moose density was 1310/1000 km2 (observed: 1320/1000 km2) and the harvest density was 100/1000 km2 (observed: 99/1000 km2), or 7% of the pre-hunt moose density = 99/ (99+1320). the modelled adult sex ratio (40 bulls:100 cows) was lower than the observed ratio (46 bulls:100 cows), while calf:cow ratios were similar (modelled = 38 calves:100 cows, observed = 41 calves:100 cows). the modelled and observed moose:wolf ratios were the same (105:1). modelled summer moose calf mortality rates from wolves and bears, as well as total mortality rates, fell within the range reported from studies in alaska and yukon (table 2a). the combined predation rates from wolves table 1. comparison of modelled moose density, bull:cow ratios and calf:cow ratios for the pgsa without hunting or predation, with bear predation only, wolf predation only, and wolf and bear predation, central british columbia, canada. treatment moose/1000 km2 bulls:100 cows calves:100 cows no hunting or predation 2000 100 32 bear predation only 1828 100 22 wolf predation only 567 90 58 wolf and bear predation 218 87 35 alces vol. 57, 2021 modelling sustained yields for moose – hatter. 119 and bears, however, were slightly lower. annual wolf and bear predation rates on the post-calving population, as well as the total mortality rate, also fell within the range from alaska and yukon (table 2b). modelled estimates of wolf density with wolf removal rates ≤ 15% (x̄ = 7 wolves/1000 km2, range = 6–7) were similar to those from the ungulate biomass regression (x̄ = 6 wolves/1000 km2, range = 5–8). the modelled parameters were sensitive to different bear predation rates on moose calves (table 3). moose densities, wolf densities, sex/age ratios, and moose harvests declined as bear predation rates increased. wolf removal rates and the moose:wolf ratio increased as bear predation rates increased. sys varied from 94 moose/1000 km2 with 50% bear predation to 119 moose/1000 km2 with 10% bear predation. sys for moose displayed an s-shape curve with increasing wolf removal rates (fig. 2a). sys gradually increased when wolf removal rates rose from 0 to 15%, while removal rates between 27 and 32% greatly increased sys; removal rates > 31% only slightly increased sys. wolves sustained removal rates up to 31%, declined rapidly when removals were 31–38%, and were eliminated with 40% annual removal (fig. 2b). sys for moose based on wolf densities displayed a hook-shape curve (fig. 2c). wolf densities were 6/1000 km2 when wolf removals were 0%. moose densities (154/1000 km2) and harvest densities (12/1000 km2) were also low at this wolf density. as wolf removal rates increased, moose numbers and harvest also increased. wolves responded to the moose increase due to the ratio dependent numerical response, and wolf densities rose until removal rates table 2. comparison of modelled moose mortality rates for the pgsa in central british columbia, canada with studies in alaska, usa and yukon, canada. harvest rates were 16% bulls, 2% cows, and 9% calves. a. summer mortality rates (%) on calves. area, treatment wolves bears predation other total pgsa, 30% wolf removal 3 35 38 15 53 pgsa, 20% wolf removal 5 35 40 15 55 pgsa, 10% wolf removal 6 35 41 15 56 alaska/yukon, min.1 2 25 45 2 47 alaska/yukon, max.1 25 67 72 15 80 1minimum (min.) and maximum (max.) estimates of collared moose calves killed among 8 radiotelemetry studies in alaska and yukon (from boertje et al. 2009). b. annual mortality rates (%) on total post-calving moose population.1 area, treatment wolves bears other hunting total pgsa, 30% wolf removal 5 13 10 9 37 pgsa, 20% wolf removal 7 13 10 7 37 pgsa, 10% wolf removal 8 13 10 7 37 alaska/yukon, min.2 8 9 1 2 27 alaska/yukon, max.2 15 27 6 6 47 1the annual mortality rate of the post-calving population = no. of moose deaths (including calves)/number of moose alive after all calves were born. 2minimum (min.) and maximum (max.) estimates of annual predation rates and mortality rates among 4 postcalving moose populations during radiotelemetry studies in alaska and yukon (from boertje et al. 2009). modelling sustained yields for moose – hatter. alces vol. 57, 2021 120 reached 31%. wolves were not able to compensate for removal rates above 31% and densities declined. moose densities and harvest continued to increase as wolf densities declined, but at a much slower rate due to density dependent declines in moose reproduction and survival. although wolf densities from 6 to 12/1000 km2 were table 3. sensitivity analysis of different bear predation rates on modelled wolf, moose and harvest parameters for the pgsa in central british columbia, canada. harvest rates were 16% bulls, 2% cows and 9% calves. the sums of squares fit is the minimum sum of squared differences between the observed and modelled moose density, wolf density, harvest density, and bull:cow and calf:cow ratios. bear predation rate (%) 10 15 20 25 30 35 40 45 50 wolf removal rate (%) 21 23 25 27 29 31 33 35 37 moose harvest rate (%) 8 8 8 7 7 7 7 7 7 harvest/1000 km2 119 117 116 115 103 101 99 97 94 wolves/1000 km2 31 28 24 21 16 12 9 6 2 moose/1000 km2 1400 1400 1400 1400 1319 1319 1319 1320 1320 moose/wolf 45 51 57 67 84 106 145 238 711 bulls/100 cows 49 48 47 46 41 40 38 36 34 calves/100 cows 60 56 53 49 41 38 34 31 28 sums of squares fit 7817 7526 7271 7061 156 136 171 269 441 fig. 2. relationship between modelled moose and wolf parameters and sys (moose harvest density) for the pgsa in central british columbia, canada with annual harvests of 16% bulls, 2% cows and 9% calves. each data point (open circle) was generated by modelling predator-prey dynamics for 100 years with wolf removal rates ranging from 0 to 40%. (a) relationship between sys and wolf removal rate; (b) relationship between wolf density and wolf removal rate; (c) relationship between sys and wolf density; and (d) sy (yield-density) curve for moose harvest density and moose density. alces vol. 57, 2021 modelling sustained yields for moose – hatter. 121 associated with both low and high sys, wolf removal rates and moose population dynamics varied greatly within this range. the sy or yield-density curve was linear over the range in moose density (fig. 2d), as harvest rates were constant. sys ranged from 12 kills/1000 km2 at 154 moose/1000 km2 to 121 kills/1000 km2 at 1571 moose/1000 km2. the yield-density curve for the pgsa was similar to the preliminary yield-density curve for moose in north-central bc (fig. 3). optimal harvests of bulls, cows, and calves (“fully optimized harvest rates,” table 4a), optimal bull and calf harvests with a cow harvest of 2% (“optimized bull and calf harvest rates,” table 4b), and optimal harvest of bulls in bull-only seasons (“optimized bull harvest rates,” table 4c) increased with increasing wolf removal rates. optimized bull and calf harvest rates produced only slightly lower yields than fully optimized harvest rates. fully optimized harvest rates and optimized bull harvest rates produced equivalent yields when wolf removal rates were low (0–10%). this was because fully optimized harvests reduced moose densities to < 200/1000 km2 which required ≥ 50 bulls/100 cows to comply with bc’s harvest management procedure, while optimized bull harvests maintained densities > 200/1000 km2 where only ≥ 30 bulls/100 cows were required by the procedure. in all cases, the greatest harvests were achieved when adult sex ratios were maintained at the minimum bull:cow ratio objective. calf:cow ratios were variable and reflected wolf removal rates, calf harvest rates, and density dependence. fully optimized harvest rates with low wolf removals (0–10%) were 6% for bulls, 0% for cows, and 28–30% for calves. with moderate to high wolf removals (20–35%), optimal harvest rates were 17–21% for bulls, 0–0.2% for cows, and 15–34% for calves. optimal harvest rates, when wolves were absent, were 16% for bulls, 0.2% for cows, and 43% for calves. moose harvest rates ranged from 8 to 15% and were highest when wolves were absent. y = 4e-05x2 + 0.049x r² = 0.86 0 10 20 30 40 50 60 70 80 0 100 200 300 400 500 600 700 800 h ar ve st /1 00 0 km 2 moose/1000 km2 fig. 3. yield-density curves for moose in central british columbia, canada. the solid line is the modelled yield-density curve for moose in the pgsa with a harvest of 16% bulls, 2% cows and 9% calves. the open circles are the observed moose harvest and density estimates from 13 game management zones in north-central british columbia with stable moose populations (from hatter 1999), and the dashed line is the fitted regression. modelling sustained yields for moose – hatter. alces vol. 57, 2021 122 discussion i used a predator-prey model to investigate how predation may affect sys and optimal harvest strategies for moose in central bc. i used a modified type ii functional response for wolves (messier 1994), a ratio dependent wolf numerical response (eberhardt 1997), and a density independent calf predation rate by bears. the modelled predator-prey parameters closely matched the observed parameters from the pgsa when the wolf removal rate was 31%. modelled moose calf and annual mortality rates were generally consistent with alaska and yukon studies table 4. optimal harvest rates based on moose and wolf population parameters using wolf removal rates from 0 to 40% for the pgsa in central british columbia, canada. a. optimized harvest rates for bulls, cows and calves. wolf removal rate (%) 0 10 20 25 30 35 40 wolves/1000 km2 5 5 8 10 15 5 0 moose/1000 km2 120 166 363 609 1400 1400 1400 harvest/1000 km2 10 14 31 48 130 206 245 bulls/100 cows 50 50 30 30 30 30 30 calves/100 cows 26 26 31 33 38 29 24 bull harvest rate (%) 6 6 17 18 21 18 16 cow harvest rate (%) 0 0 0 0 0.2 0.2 0.2 calf harvest rate (%) 30 28 19 15 15 34 43 moose harvest rate (%) 8 8 8 7 8 13 15 b. optimized bull and calf harvest rates with cow harvest rate = 2%. wolf removal rate (%) 0 10 20 25 30 35 40 wolves/1000 km2 4 5 8 10 15 5 0 moose/1000 km2 100 161 350 585 1400 1400 1400 harvest/1000 km2 10 13 30 46 123 200 239 bulls/100 cows 50 50 30 30 30 30 30 calves/100 cows 28 32 36 38 43 34 29 bull harvest rate (%) 9 10 21 22 25 22 20 cow harvest rate (%) 2 2 2 2 2 2 2 calf harvest rate (%) 27 18 9 6 6 26 35 moose harvest rate (%) 9 7 8 7 8 12 14 c. optimized bull harvest rates with bull-only hunting. wolf removal rate (%) 0 10 20 25 30 35 40 wolves/1000 km2 9 10 13 18 16 6 0 moose/1000 km2 229 319 563 1040 1490 1606 1662 harvest/1000 km2 10 14 26 50 79 81 81 bulls/100 cows 30 30 30 30 30 30 30 calves/100 cows 36 37 38 39 40 34 31 bull harvest rate (%) 19 20 20 21 22 21 20 cow harvest rate (%) 0 0 0 0 0 0 0 calf harvest rate (%) 0 0 0 0 0 0 0 moose harvest rate (%) 4 4 4 5 5 5 5 alces vol. 57, 2021 modelling sustained yields for moose – hatter. 123 (boertje et al. 2009), and modelled wolf densities were similar to those generated from a biomass regression model. the modelled maximum rate of increase for moose (λ ~ 1.3) was similar to 1.35 estimated by eberhardt (1997), and unhunted adult sex ratios (87–100 bulls/100 cows) were near parity as documented from published studies of naturally fluctuating moose populations (timmermann 1992). the yield-density curve for the pgsa was comparable to the yield-density curve for moose in north central bc (hatter 1999) suggesting that the sy curve for the pgsa may be broadly applicable to other moose populations in bc. optimal harvests included a moderate harvest of bulls (16–21%), a nil-to-very low harvest of cows (0–0.2%), and moderate-to-high harvests of calves (15–43%) when wolf removal rates were ≥ 20%. these results support previous findings from nilsen et al. (2005) and xu and boyce (2010) who also noted that optimal moose harvests in the presence of predators involved a high harvest of bulls and calves with a minimal cow harvest. xu and boyce (2010) found that harvest rates to optimize moose yields in alberta included 40–45% of bulls, 0.1–5% of cows, and 35–40% of calves. the lower optimal harvest rates for bulls in this study were likely due to the harvest management constraints placed on adult sex ratios (i.e., ≥ 30 bulls:100 cows for > 200 moose/1000 km2, and ≥ 50 bulls:100 cows for ≤ 200/1000 km2). optimal yields of bulls and calves, with a cow harvest of 2% were only slightly lower than fully optimized yields. some advantages of harvesting a greater proportion of cows (i.e., 2% vs. 0.2%) include an increased sample size for monitoring moose reproduction and nutritional status (heard et al. 1997, boertje et al. 2007), recovery of low bull:cow ratios (young and boertje 2008), and more options to intensify moose management at high density (young et al. 2006). bull-only hunting, however, may be preferable for low density moose populations limited by predation (van ballenberghe and dart 1982, environment yukon 2016). i found that bullonly harvests with low wolf removal rates (0–10%) provided equivalent yields to selective harvests while maintaining higher moose and wolf densities. wolf control studies in alaska and yukon indicated that wolf reductions within low-density, predator-limited moose populations led to elevated moose and harvest densities followed by elevated wolf densities that equalled or exceeded pre-control levels (gasaway et al. 1992, boertje et al. 1996). modelled moose densities and harvests similarly increased with intensified wolf removals due to the ratio-dependent numerical response which enabled wolf densities to increase. for example, moose densities were elevated from 207 to 1103 moose/1000 km2, harvests from 16 to 84 moose/1000 km2, and wolf densities from 7 to 12 wolves/1000 km2 when wolf removal rates were increased from 10 to 30%. the modelled wolf population was able to sustain removal rates up to 31% which was consistent with adams et al. (2008) who analysed information from 39 north american wolf populations and determined that populations were able to compensate for removal rates up to 29%. while the predator-prey model appeared to provide a reasonable portrait of wolfmoose-bear dynamics in the pgsa from 1987 to 1998, increasing model realism would have helped to validate the model. several studies have shown that winter kill rates by individual wolves are inversely related to pack size, and that wolf predation should be modelled by the number and size of packs (ballard et al. 1987, mcnay and delong 1998, hayes et al. 2000). adams et al. (2008) discussed how wolves adjust dispersal rates as a primary mechanism to modelling sustained yields for moose – hatter. alces vol. 57, 2021 124 compensate for human harvest which also could be modelled. most northern wolf-ungulate studies have identified stochastic weather events as a significant component in predator-prey systems (gasaway et al. 1983 gasaway et al. 1992, boertje et al.1996, 2009, ballard and van ballenberghe 1998, mcnay and delong 1998). i did not model stochastic predator-prey-weather interactions because many parameters in predator-prey models are influenced by weather which greatly increases model complexity, and because quantitative measures of imprecision (ses) were difficult to parameterize. however, xu and boyce (2010) cautioned that predation and stochastic weather events can drive moose populations to low levels and negatively influence sys. a growing number of studies suggest ratio-dependence may be common in wolf-ungulate systems (hebblewhite 2013). i used a ratio-dependent, wolf numerical response (eberhardt 1997) with an equilibrium ratio of 25 moose/wolf. studies of moose:wolf ratios during winter suggest moose densities may stabilize with 20–30 moose/wolf. gasaway et al. (1983) summarized these studies and identified 3 general categories of moose/wolf relationships: predation was sufficient to cause a decline in moose abundance at <20 moose/ wolf, appeared to control moose numbers at 20–30 moose/wolf, and was insufficient to limit growth at > 30 moose/wolf. person et al. (2001) and bowyer et al. (2013), however, cautioned against even a general interpretation of such ratios for interpreting impacts of wolf predation on moose. the predator-prey model was based on moose-wolf-bear relationships in the pgsa prior to the moose population decline in the early 2000s. the decline coincided with a mountain pine beetle (dendroctonus ponderosae) outbreak where habitat changes and increased salvage logging and road building may have resulted in greater vulnerability to moose from human harvest and predation, while elevating nutritional constraints and health/disease concerns (kuzyk and heard 2014). while disease was not considered a substantive cause of moose mortality in british columbia from 2012 to 2019 (kuzyk et al. 2019), it is reported as a significant mortality factor in certain north american moose populations (murray et al. 2006). sys appear to have been impacted from the beetle outbreak (kuzyk et al. 2018) and optimal harvest rates may now be considerably lower than projected by the predator-prey model. further work on anthropogenic and environmental factors affecting moose population dynamics could improve and make the model more suitable for contemporary conditions. management implications and recommendations sy curves assume that moose populations are stable, while most populations fluctuate and rarely, if ever, achieve a stable equilibrium (sæther et al. 2001). thus, the main value of estimating sys was to provide an expectation of harvest under longterm, stable predator-prey interactions in central bc. the principal value of the optimal harvest calculations was to identify how various sex/age classes should be harvested under different levels of predation. estimating sys from the predator-prey model was limited by the lack of data on first nations harvest, calf vulnerability to predation during summer, and uncertainty in the wolf functional and numerical responses. nonetheless, the model findings that moose harvests should consist primarily of bulls and calves with nil to very low cow harvests was consistent with other published studies, and suggest these findings are broadly applicable to moose-wolf-bear systems where alternate prey are rare. alces vol. 57, 2021 modelling sustained yields for moose – hatter. 125 managers who wish to elevate cow harvests in order to monitor moose reproduction and nutritional status, or to recover low bull:cow ratios, should reduce calf harvests to compensate for the increased harvest of cows. further restrictions such as bull-only harvests should be considered for low density, predator-limited moose populations. when moose populations are food-limited, harvests of bulls, cows, and calves should be elevated to maintain populations below k and enhance yields. finally, and most importantly, moose harvests should be set below optimal yields to account for stochastic variation in predator-prey interactions, changing environmental conditions, and management uncertainty. acknowledgements d. heard, m. anderson, and m. scheideman kindly reviewed an early draft of the manuscript and made numerous helpful and insightful suggestions. i am also grateful for the constructive reviews from e. addison and two anonymous reviewers which further improved this manuscript. references adams, l. g., r. o. stephenson, b. w. dale, r. t. ahgook, and d. j. demma. 2008. population dynamics and harvest characteristics of wolves in the central brooks range, alaska. wildlife monographs 170(1): 1–25. doi: 10.2193/2008-012 ballard, w. b. 1992. bear predation on moose: a review of recent north american studies and their management implications. alces supplement 1: 162–176. _____, and v. van ballenberghe. 1998. predator-prey relationships. pages 247– 273 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. _____, j. s. whitman, and c. l. gardner. 1987. ecology of an exploited wolf population in south-central alaska. wildlife monographs 98: 3–54. bergerud, a. t., w. wyett, and b. snider. 1983. the role of wolf predation in limiting a moose population. journal of wildlife management 47: 977–988. doi: 10.2307/3808156 boertje, r. d., m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314–327. doi: 10.2193/ 2007-591 _____, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, and b. w. dale, 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. doi: 10.2193/2006-159 _____, p. valkenburg, and m. e. mcnay. 1996. increases in moose, caribou and wolves following wolf control in alaska. journal of wildlife management 80: 474–489. doi: 10.2307/3802065 bowyer, r. t., j. g. kie, d. k. person, and k. l. monteith. 2013. metrics of predation: perils of predator-prey ratios. acta theriologica 58: 329–340. doi: 10.1007/ s13364-013-0133-1 caughley, g. 1976. wildlife management and the dynamics of ungulate populations. pages 183–246 in t. h. coaker, editor. applied biology, vol. 1. academic press, london, england. child, k. 1983. selective harvest of moose in the omineca: some preliminary results. alces 19: 162–177. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1: 553–563. _____, r. j. taylor, and p.a. jordon. 1981. optimization of moose harvest in southwestern quebec. journal of wildlife management 45: 598–611. doi: 10.2307/ 3808693 modelling sustained yields for moose – hatter. alces vol. 57, 2021 126 dahle, b., k. wallin, g. cederlund, i.-l. persson, l. s. selvaag, and j. e. swenson. 2013. predation on adult moose alces alces by european brown bears ursus arctos. wildlife biology 19: 165–169. doi: 10.2981/10-113 eberhardt, l. l. 1997. is wolf predation ratio-dependent? canadian journal of zoology 75: 1940–1944. doi: 10.1139/ z97-824 _____, r. a. garrott, d. w. smith, p. i. white, and r. o. peterson. 2003. assessing the impact of wolves on ungulate prey. ecological applications 13: 776–783. doi: 10.1890/1051-0761 (2003)013 [0776:atiowo] 2.0.co;2 _____, and r. o. peterson. 1999. predicting the wolf-prey equilibrium point. canadian journal of zoology 77: 494– 498. doi: 10.1139/z98-240 environment yukon. 2016. science-based guidelines for management of moose in yukon. yukon fish and wildlife branch report mr-16–02. whitehorse, yukon, canada flnro. provincial moose management team. 2013. draft provincial framework for moose management in british columbia. unpublished report. ministry of forests, lands and natural resource operations, victoria, british columbia, canada. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larson. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120: 1–59. _____, r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey and man in interior alaska. wildlife monographs 84: 1–50. hatter, i. w. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. _____. 2019. an assessment of eberhardt’s ratio-dependent wolf-ungulate model. journal of wildlife and biodiversity 3: 1–8. hayes, r. d., a. m. baer, u. wotschikowsky, and a. s. harestad. 2000. kill rate by wolves on moose in the yukon. canadian journal of zoology 78: 49–59. doi: 10.1139/z99-187 _____, and a. s. harestad. 2000. wolf functional response and regulation of moose in the yukon. canadian journal of zoology 78: 60–66. doi: 10.1139/ z99-188 heard, d. c., s. barry, g. watts, and k. child. 1997. fertility of female moose (alces alces) in relation to age and body composition. alces 33: 165–176. _____, k. l. zimmerman, g. s. watts, and s. p. barry. 1999. moose density and composition around prince george, british columbia, december 1998. final report for common land information base. project no. 99004. ministry of forests, lands and natural resource operations, victoria, british columbia, canada. hebblewhite, m. 2013. consequences of ratio dependent predation by wolves for elk population dynamics. population ecology 55: 511–522. doi: 10.1007/ s10144-013-0384-3 hilborn, r., and m. mangel. 1997. the ecological detective: confronting models with data. princeton university press, princeton, new jersey, usa. keith, l. b. 1983. population dynamics of wolves. pages 66–77 in l. n. carbyn, editor. wolves in canada and alaska: their status, biology, and management. canadian wildlife service report series 45. ottawa, ontario, canada. kuzyk, g., and i. hatter. 2014. using ungulate biomass to estimate abundance of wolves in british columbia. wildlife society bulletin 38: 878–883. doi: 10. 1002/ wsb.475 alces vol. 57, 2021 modelling sustained yields for moose – hatter. 127 _____, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. british columbia ministry of forests, lands and natural resource operations, victoria, british columbia, canada. _____, _____, s. marshall, c. procter, b. cadsand, d. lirette, h. schindler, m. bridger, p. stent, a. walker, and m. klaczek. 2018. moose population dynamics during 20 years of declining harvest in british columbia. alces 54: 101–119. _____, c. procter, s. marshall, h. schindler, h. schwantje, m. scheideman, and d. hodder. 2019. factors affecting moose population declines in british columbia. 2019 progress report: february 2012–may 2019. wildlife working report no. wr 127. british columbia ministry of forests, lands and natural resource operations, victoria, british columbia, canada. lake, b. c., m. r. bertram, n. guldager, j. r. caikoski, and r. o. stephenson. 2013. wolf kill rates across winter in a low-density moose system in alaska. journal of wildlife management 77: 1512–1522. doi: 10.1002/jwmg.603 mcnay, m. e., and r.a. delong. 1998. development and testing of a general predator-prey computer model for use in making management decisions. federal aid in wildlife restoration, research final report, grants w-24-1 and w-24-5, study 1.46. alaska department of fish and game, juneau, alaska, usa. messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478–488. doi: 10.2307/1939551 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. doi: 10.2193/ 0084-0173 (2006) 166[1:pndaci] 2.0.co;2 nilsen, e. b., t. pettersen, h. gundersen, j. m. milner, a. mysterud, e. j. solberg, h. p. andreassen, and n. c. stenseth. 2005. moose harvesting strategies in the presence of wolves. journal of applied ecology 42: 389–399. doi: 10.1111/ j.13652664.2005. 01018.x person, d. k., r. t. bowyer, and v. van ballenberge. 2001. density dependence of ungulates and functional responses of wolves: effects on predator-prey ratios. alces 37: 253–273. peterson, r. o., j. d. woolington, and t. n. bailey.1984. wolves of the kenai peninsula, alaska. wildlife monographs 88: 1–52. rea, r.v., l. ajala-batista, d. a. aitken, k. n. child, n. thompson, and d. p. hodder. 2019. scat analysis as a preliminary assessment of moose (alces alces andersoni) calf consumption by bears (ursus spp.) in north-central british columbia. animal biodiversity and conservation 42: 369–377. doi: 10.32800/ abc.2019.42.0369 sand, h., p. wabakken, b. zimmermann, o. johansson, h. c. pedersen, and o. liberg. 2008. summer kill rates and predation pattern in a wolf–moose system: can we rely on winter estimates? oecologia 156: 53–64. doi: 10.1007/ s00442-008-0969-2 sæther, b.-e., s. engen, and e. j. solberg. 2001. optimal harvest of age-structured populations of moose alces alces in a fluctuating environment. wildlife biology 7: 171–179. doi: 10.2981/wlb.2001.021 serrouya, r., b. n. mclellan, and s. boutin. 2015. testing predator-prey theory using broad-scale manipulations and independent validation. journal of animal ecology 84: 1600–1609. doi: 10. 1111/1365-2656.12413 modelling sustained yields for moose – hatter. alces vol. 57, 2021 128 sinclair, a. r. e., j. m. fryxell, and g. caughley. 2006. wildlife ecology, conservation, and management, second edition. blackwell publishing, malden, massachusetts, usa. timmermann, t. 1992. moose sociobiology and implications for harvest. alces 28: 59–77. van ballenberghe, v., and j. dart. 1982. harvest yields from moose populations subject to wolf and bear predation. alces 18: 258–275. _____, and w. b. ballard. 1994. limitation and regulation of moose populations: the role of predation. canadian journal of zoology 72: 2071–2077. doi: 10.1139/ z94-277 xu, c., and m. s. boyce. 2010. optimal harvesting of moose in alberta. alces 46: 15–35. young, d. d., jr., and r. d. boertje. 2008. recovery of low bull:cow ratios of moose in interior alaska. alces 44: 65–71. _____, _____, t. seaton, and k. a. kellie. 2006. intensive management of moose at high density: impediments, achievements, and recommendations. alces 42: 41–48. alces vol. 57, 2021 modelling sustained yields for moose – hatter. 129 appendix 1. relationship between wolf finite rate of increase (λ) and the wolves:moose ratio from 17 north american wolf populations. y = 1.534e-10.77x r² = 0.78 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0 0.02 0.04 0.06 0.08 0.1 w ol f fi ni te ra te o f i nc re as e, λ wolves/moose location wolves/moose1 wolf λ reference alaska 0.0083 1.340 keith 1983 alberta 0.0085 1.460 keith 1983 michigan 0.0135 1.390 keith 1983 minnesota 0.0148 1.310 keith 1983 denali park 0.0159 1.280 mcnay and delong 19982 kenai, alaska 0.0174 1.180 eberhardt and peterson 1999 interior alaska 0.0183 1.280 mcnay and delong 19982 denali park 0.0260 0.900 mcnay and delong 19982 alberta 0.0306 1.210 keith 1983 michigan 0.0310 1.150 keith 1983 nelchina, alaska 0.0313 1.280 eberhardt and peterson 1999 ontario 0.0336 1.200 keith 1983 nc minnesota 0.0373 1.070 eberhardt and peterson 1999 isle royale 0.0376 1.004 https://isleroyalewolf.org/2 isle royale 0.0382 0.820 mcnay and delong 19982 quebec 0.0400 1.080 eberhardt and peterson 1999 interior alaska 0.0870 0.590 mcnay and delong 19982 1wolves per moose biomass equivalent (keith 1983). 2additional data used in this study to estimate the wolves:moose equilibrium ratio. https://isleroyalewolf.org/ 91 movements and resource use by moose in traditional and nontraditional habitats in north dakota james j. maskey jr.1,2 and rick a. sweitzer1,3 1department of biology, 10 cornell street, stop 9019, university of north dakota, grand forks, north dakota 58202, usa; 2department of biology, university of mary, 7500 university drive, bismarck, north dakota 58504, usa; 3great basin institute,16750 mt. rose hwy, reno, nevada 89511, usa abstract: in the past several decades, moose (alces alces) have expanded their range in north dakota from primarily forested areas to the prairie/agriculture mosaic of the state. as a result, moose are now wellestablished in a large portion of north dakota, yet little is known about their ecology in the state. we examined the home ranges, habitat selection, and diets of moose in both traditional (forested) and nontraditional ranges (prairie/agricultural) and inferred whether range expansion is the result of agriculture-related landscape changes. from 2004 to 2006, we placed gps radio-collars on a total of 14 moose in two study areas: turtle mountains (forested) and lonetree (prairie/agricultural). total and seasonal home ranges were larger for lonetree moose, and moose in both study areas selected strongly for wooded habitat. in both study areas seasonal diets ranged from 65 to 99% woody browse, with forbs 15% of summer diets. in the lonetree area row crops made up the second highest consumed forage in fall (12%) and winter (29%) diets. larger home ranges in the lonetree area may reflect the low availability and scattered distribution of wooded habitat. further, the strong selection for planted woodlands and the high proportion of woody browse and row crops in the diet of lonetree moose suggests that conversion of the native prairie to agriculture has facilitated range expansion by moose in north dakota. alces vol. 55: 91–104 (2019) key words: alces alces, browse, cropland, diet, habitat selection, home ranges, north dakota, prairie, woodland introduction moose (alces alces) are native to north dakota with their traditional range encompassing the aspen (populus tremuloides) and bur oak (quercus macrocarpa) forests of the turtle mountains and pembina hills along the northern edge of the state (knue 1991). while moose were extirpated from north dakota by the late 1800s, they had begun to re-establish a population in the state by the 1960s. after re-colonizing their historic range, by the 1980s moose had expanded their range to include large expanses of former tall and mixed grass prairie that had been greatly modified by conversion to agriculture and widespread planting of tree rows to reduce wind erosion subsequent to the dust bowl years of the 1930s (knue 1991, licht 1997). the colonization and range expansion by moose in north dakota are likely the result of conversion of the native prairie landscape to an agricultural mosaic that provides suitable cover and forage otherwise absent in unaltered tall or mixed grass prairie habitats. although moose are known to persist in other landscapes modified by humans such as clear-cuts and agricultural areas within forested landscapes (leptich and gilbert 1989, rempel et al. 1997, schneider and wasel 2002), the agriculture-dominated landscape of the northern great plains movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 92 represents a unique habitat for the species that was not occupied prior to human-induced habitat change. while numerous prior efforts have provided insight into moose movements and resource use in traditional habitats (kearney and gilbert 1976, leptich and gilbert 1989, cederlund and sand 1994, maccracken et al. 1997, labonte et al. 1998), the ecology and behavior of moose in the prairie ecoregions of north america is relatively unknown. the purpose of this project was to investigate the ecology of moose in both traditional woodland habitats and the recently colonized prairie region of north dakota, including how this species may be taking advantage of landscape alterations to extend its range. the specific objectives were to 1) examine seasonal and annual movements and habitat use of moose in the prairie and woodland regions of north dakota, 2) investigate the diet of moose in prairie and woodland regions of north dakota, and 3) compare movements, habitat use, and diets of moose in these two regions. to meet these objectives, we selected study areas that were representative of traditional and more recently colonized habitats. first, the forested turtle mountains comprise a major portion of the historic range of moose in north dakota and was one of the areas first recolonized upon their return (knue 1991, seabloom et al. 2011). second, the lonetree wildlife management area (wma) is an agricultural mosaic characteristic of the habitats more recently colonized by moose. it is representative of most of the landscape of eastern and central north dakota, which now comprises much of the primary range of this species in the state. study area the turtle mountains (48° 57ʹ n, 99° 53ʹ 00” w; fig. 1) are located along the canadian border and are characterized by hilly wooded terrain and numerous small lakes and wetlands with interspersed agricultural fields, pastureland, and hay fields, especially near the southern edge of the area. the forest of the turtle mountains is comprised primarily of aspen and bur oak along with green ash (fraxinus pennsylvanica), paper birch (betula papyrifera), balsam poplar (populus balsamea) , and box elder (acer negundo), with an understory of chokecherry (prunus virginiana), hazel (corylus cornuta), and several species of willow (salix spp.). typical herbaceous species include sarsaparilla (aralia nudicaulis), alfalfa (medicago sativa), brome (bromus spp.), fescue (festuca spp.), wheatgrass (agropyron spp.), sedges (carex spp.), baneberry (actea spp.), false lily of the valley (maianthemum canadensis), wild vetch (vicia americana), and virginia anemone (anemone virginiana) (stevens 1966, bakke 1980, nd forest service 2003). the lonetree wma is large, encompassing 134 km2 in the central part of the state (47°30ʹ n, 100°15ʹ w; fig. 1). it consists of farmland initially purchased by the u.s. bureau of reclamation to be used as part of the missouri river garrison diversion project (garrison diversion project 2019). following the cancellation of that portion of the project, management of the land was turned over to the north dakota game and fish department. habitats include native mixed grass prairie, corn (zea mays) and sunflower (helianthus annuus) food plots (range = 6–31 ha), numerous seasonal and semi-permanent wetlands, small impoundments along the sheyenne river, and planted woodlands in the form of linear tree rows or larger block plantings (smith et al. 2007). the surrounding area is comprised primarily of pasture and hay land alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 93 as well as crop fields consisting mostly of small grains. planted tree rows and woodlots are present with some natural woodlands occurring in woody draws along the missouri escarpment that marks the border between the northern glaciated plains and missouri coteau ecoregions (usepa 1996). typical grassland plants found in the area include prairie junegrass (koeleria macrantha), indiangrass (sorghastrum nutans), needle and thread grass (hesperostipa comata), brome, wheatgrass, and alfalfa. common tree species in planted and/or native woodlands include green ash, box elder, american elm (ulnus americana), russian olive (elaeagnus angustafolia), plum (prunus spp.), apple (malus spp.), chokecherry, fireberry hawthorn (crataegus chrysocarpa), serviceberry (amelanchier alnifolia), and willow. methods study animals global positioning system (gps) radio collars (lotek wireless inc. newmarket, ontario, canada) were placed on 14 adult moose (5 cows, 1 bull in the lonetree wma; 4 cows, 4 bulls in the turtle mountains) in january 2004–2006. only moose in the lonetree wma study area were captured in 2004, with subsequent expansion to the turtle mountains in 2005 and 2006. moose were captured from helicopter with the use of a net gun. capture operations were performed by leading edge aviation (lewiston, idaho, usa), and all methods were approved by the university of north dakota (iacuc project #0506-3). collars were set to acquire a location every 4 h, and location data were stored on board. after ~ 52 weeks, radio-collars were recovered when moose were fig. 1. location of the lonetree wildlife management area and turtle mountains study areas in north dakota, usa. boundaries delineate north dakota counties. movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 94 recaptured or after they dropped off via timed-release mechanisms. moose were periodically monitored by aerial and groundbased vhf telemetry. home range estimation total and seasonal moose home range sizes (km2) were estimated with a 95% fixed-kernel estimator, with seasons defined as winter (1 january–30 april), summer (1 may–31 august), and fall (1 september–31 december) for all analyses. we carried out these calculations using all available locations for non-dispersing moose, with dispersal defined as locations for a moose that deviated from other grouped relocations for that animal (dodge et al. 2004). seasonal home range sizes were estimated for each moose for all seasons for which at least 30 locations were available, and total home ranges were estimated for moose with at least 30 locations in every season (seaman et al. 1999). moose were considered to exhibit seasonal migrations if < 25% of their seasonal home ranges overlapped (dodge et al. 2004). location data were input into arcmap 9.2 (esri inc., redlands, california, usa) and home range calculations were performed using the home range extension (rodgers and carr 1998). at the time of our study, least squares cross validation (lscv) was the most recommended technique to determine the optimal smoothing parameter for fixed-kernel home range estimation (worton 1995, seaman and powell 1996, seaman et al. 1999). however, we experienced similar problems with this method as reported by others (silverman 1986, hemson et al. 2005); lscv was unable to calculate a smoothing parameter for most sets of locations, and if it did, the multi-modal nature of the locations produced home ranges that were dramatically under-smoothed. to deal with these problems, we used biased cross validation (bcv) to calculate the smoothing parameters for all fixed-kernel home range estimations (wand and jones 1995, rodgers and carr 1998). although bcv has not been commonly applied to estimate the smoothing parameters for home range estimates, the statistical literature has demonstrated its utility in selecting kernel bandwidth, as well as its potential superiority to lscv (sain et al. 1994, wand and jones 1995, rodgers and carr 1998). we compared total home range sizes between study sites with a two-sample t-test. for all moose with home range estimates for all seasons, we also compared seasonal home range sizes among seasons and study sites with repeated-measures anova. when necessary, home range sizes were natural log transformed to meet the assumptions of parametric tests. all statistical comparisons were performed in the statistical package r 2.6 (r core development team 2007). habitat selection we first estimated the extent of the area available to moose in each study area by constructing 99% fixed-kernel home ranges for each moose, and then combined the home ranges for each study site into a single polygon. we then used land cover data from the united states geological survey’s gap analysis program (gap) compiled from 1992 to 1999 (strong et al. 2005) as well as national wetland inventory 1:24,000 digital quadrangles (usfws 2000) to determine habitat types available to moose. to make the comparison of habitat use between study sites possible, land cover data were collapsed into 4 habitat types using spatial analyst in arcmap 9.2 (esri inc., redlands, california, usa). these were defined as woodland (all planted and naturally occurring woodlands), wetland (temporary, seasonal, permanent, and semi-permanent wetlands), grasslands (planted non-native grasses, hay fields, old fields, and planted alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 95 or naturally occurring prairie), and crops (all planted row crops or grains). next, because preliminary analysis indicated that the coarse spatial resolution (30m) of the gap data were insufficient to detect small areas of habitat, we improved the resolution of terrestrial habitat layers by re-digitizing data based on 1-m resolution aerial photos of each study site (national agricultural imagery program, usda-fsa 2005). we did this by overlaying gap habitat layers onto the aerial photos in arcmap, then manually re-digitizing the gap layers to conform to the habitat boundaries indicated on the photos. we modified the wetland habitat layer by considering all temporary and seasonal wetlands to be part of the terrestrial habitat in which they were imbedded, as these wetlands are typically inundated only during the spring and do not provide a source of emergent or submergent aquatic vegetation (usfws 2000). we also adjusted wetland habitat availability to account for the presence of several larger lakes in the 2 study areas. because the deep-water areas in these lakes were unlikely available to moose, we created a 100 m buffer layer that extended from the shoreline into each lake. this distance was chosen as a conservative estimate of the extent of the littoral zone, where water depth was shallow and emergent and submergent plants would occur. the area of this buffer layer was considered the amount of lake habitat available to moose. we measured the area of each habitat type in each study area using the x-tools extension for arcmap 9.2 (esri inc., redlands, california, usa) and then determined the proportional availability of each habitat (table 1). all locations for individual moose at each study site were separated into seasons. we then calculated manly’s standardized selection ratios for each moose with ≥30 locations in a season (manly et al. 2002, osko et al. 2004). this method produces selection ratios that represent the probability of a moose using a particular habitat if all habitats were equally available, and the selection ratios for all habitats sum to 1. because there were four habitat types in this study, a selection ratio of 0.25 indicates non-selection (not different than by chance), while a selection ratio >0.25 represent positive selection for a habitat type. diet in 2005–2006 we collected 5 samples of fresh moose feces monthly from each study area by searching several locations distributed across each study site, with no more than 2 samples/month collected from a single location (e.g., from the same clear-cut). samples were combined to generate a series of 2-month composite fecal samples. samples were sent to the wildlife habitat and nutrition laboratory at washington state university (pullman, washington, usa) for microhistological determination of plant fragments and estimates of diets to the genus and species level (van vuren 1984). forage plants were classified into 5 categories: woody browse, grasses and sedges, forbs, crops, and other (fruits, nuts, aquatic vegetation). results of this diet analysis were grouped by season based on the same criteria used for home range and habitat use analyses. results home range and movement a high rate of collar failure in 2005 prevented the calculation of all seasonal and table 1. proportional availability of each of the four major habitat types in the lonetree and turtle mountains study areas in north dakota, usa. woodland wetland grass crop lonetree 0.025 0.053 0.400 0.522 turtle mountains 0.451 0.170 0.252 0.127 movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 96 annual home ranges. estimates of home range size (95% fixed kernel method) ranged from 59.2 to 262.6 km2 (n = 5) in the lonetree wma study area which were larger than in the turtle mountains (9.6–47.7 km2, n = 4; t5.3= 3.7, p = 0.01; mean number of locations = 2709). seasonal home range estimates were also larger in lonetree than the turtle mountains (f1,25 = 13.3, p = 0.0012, mean number of locations = 807), ranging from 18.8 to 292.8 km2 and 1.0 to 44.7 km2, respectively. home range size did not differ among seasons (f2, 25 = 0.1, p = 0.91). one moose was excluded from comparisons because its 30 locations occurred in a single month. none of the moose exhibited seasonal migrations. one dispersed from the lonetree wma in march 2004, with the 5 others remaining in the general vicinity. radiocollars were recovered successfully from 7 animals (the 8th failed) in the turtle mountains; all remained in the turtle mountains throughout the study. habitat use moose strongly selected for wooded habitat in all seasons in both study areas; conversely, no selection for cropland or grassland habitats was measured in either study area. moose in the turtle mountains also selected for wetland habitats during the summer (table 2). moose diets moose consumed mostly woody plants in both the lonetree (≥65%) and turtle mountains (≥83%) areas in all seasons of the year (fig. 2). consumption of woody browse was particularly high in the turtle mountains; for example, moose consumed 99% woody browse during winter, primarily aspen (36%) and willow (20%). willow and aspen were also important components of the diets in the turtle mountains during summer (15 and 12%) and fall (19 and 23%). bur oak stems and leaves were also common forage items in these seasons, representing 20% and 23% of summer and fall diets, respectively. in the lonetree area, russian olive was the most common woody browse consumed in all seasons and was 50% of the fall diet, followed by willow (10% in summer) and cottonwood (11% in winter). row crops (primarily corn) were also a major component of the diets in the lonetree area during fall (12%) and winter (29%; fig. 2). in contrast, row crops were absent from samples collected in the turtle mountains, although alfalfa was an important component in summer and fall diets (13%), representing 90% of forbs consumed in these seasons. grasses (≤3%) and fruits and nuts (≤1%) were minor components of the diet in both study areas, while emergent and submergent aquatic vegetation were ≤1% of the diet during the open water seasons of summer and fall. table 2. mean manly’s standardized selection ratios (se) for four habitat types based on data from 13 gps-collared moose in the lonetree and turtle mountains study areas in north dakota, usa. bold numbers indicate positive selection (> 0.25) for a habitat type. study site season n woodland wetland crop grassland lonetree winter 6 0.95(0.008) 0.013(0.005) 0.013(0.003) 0.024(0.004) summer 6 0.89(0.016) 0.048(0.018) 0.024(0.006) 0.034(0.005) fall 5 0.84(0.033) 0.067(0.017) 0.051(0.017) 0.038(0.009) turtle mountains winter 7 0.76(0.01) 0.16(0.021) 0.031(0.017) 0.048(0.012) summer 4 0.56(0.06) 0.30(0.039) 0.015(0.007) 0.13(0.068) fall 4 0.54(0.1) 0.21(0.052) 0.10(0.040) 0.15(0.085) alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 97 discussion sample size and radio-collar performance the low density of moose limited the number of animals that could be studied in the lonetree wma. observations from fixed-wing aircraft indicated that only 5 moose were in the vicinity in 2004, and all were successfully captured and radio collared that year; likewise, in subsequent years we successfully radio-collared nearly every known moose in the area. radiocollar failures limited our ability to make comparisons between moose in their traditional range and the prairie habitats. overall, 9 of 22 radio-collars failed prematurely, fig. 2. seasonal diet composition (%) of moose in the lonetree (a) and turtle mountains (b) study areas in north dakota, usa. movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 98 with 6 of 10 in the turtle mountains including 4 of 5 deployed in 2005. we do not believe that these failures generated any obvious sources of bias within our data. more importantly, we radio-collared a large proportion of the moose in the study areas and our results provide novel information about habitat and forage use in the north dakota landscape. migration and dispersal our results indicate that moose in north dakota are largely non-migratory. elsewhere, moose may migrate to avoid deep snows at high elevation or to seek conifer forests that provide thermal cover or reduced snow depth (pierce and peek 1984, ballard et al. 1991, hundertmark 1998, thompson and stewart 1998, poole and stuart-smith 2006). because moose range in north dakota lacks these characteristics, elevational differences do not lead to variability in snowfall or depth, and conifer forests are absent. additionally, habitat selection and diet composition of moose in both study areas indicated that moose selected for wooded habitats and consumed primarily woody browse in all seasons. other habitats such as croplands realized increased seasonal use, but these habitats were interspersed within the mosaic of woodland patches used yearround by moose. as a result, moose did not need to migrate to gain access to seasonally preferred forage. home range detailed comparisons of home range sizes across studies are difficult because of differences in the number of locations collected and the variety of estimators used. however, home range size is expected to be a function of the energetic requirements of an animal and the spatial distribution of necessary resources (mcnab 1963, elchuk and weibe 2003, mitchell and powell 2004). thus, where required resources are widely dispersed, home range size will be larger. mean total home range size (160.5 km2, se = 38.9) for lonetree moose was near the upper range of averages reported for non-migratory moose (174–290 km2; grauvogel 1984, ballard et al. 1991, stenhouse et al. 1995). ballard et al. (1991) attributed large home ranges to the high proportion of unavailable habitat within home ranges. similarly, 92% of the landscape in the lonetree wma consisted of grassland and cropland habitats that moose mostly avoided, while the wooded habitats that moose selected for comprised only 2.5% of the landscape. in contrast, in the turtle mountains with a high proportion of woodland habitat (45.1%), moose had smaller total home ranges (x̄ = 27.7 km2, se = 10.0) similar to those (2–43 km2) for other predominantly wooded areas in eastern north america (phillips et al. 1973, addison et al. 1980, leptich and gilbert 1989, garner and porter 1990, dodge et al. 2004). we did not observe significant differences in home range size among seasons, but seasonal home-range size varied considerably among moose. although energy constraints associated with moving through deep snow or predator avoidance may result in smaller home ranges during winter (phillips et al. 1973, thompson and vukelich 1981, dussault et al. 2005), snow depths considered limiting to moose are rare (> 70 cm; hundertmark 1998) and large predators are absent in north dakota. thus, seasonal home ranges were more likely determined by the distribution of seasonally-important forage resources (doerr 1983, lynch and morgantini 1984, leptich and gilbert 1989). as such, the differences in size of seasonal home ranges among moose likely reflected the spatial pattern of available seasonal resources where moose resided. alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 99 habitat selection and diet while the characterization of habitat types was general in order to facilitate comparisons between the study areas, the habitat selection analyses nevertheless provided important insight into how moose utilized available habitats in north dakota. numerous researchers have demonstrated the importance of a variety of types of woody habitats in providing forage and/or cover for moose (e.g., peek et al. 1976, peek 1998, courtois et al. 2002). therefore, it was not surprising that moose in north dakota exhibited a strong selection for woody habitats in all seasons in both study areas. in adjacent minnesota and many areas of canada, moose inhabited early successional forest created by periodic fire or insect outbreaks and that are now maintained largely by forest harvesting (peterson 1955, phillips et al. 1973, peek et al. 1976). the forests that cover nearly half of the turtle mountains study area represent this “typical” moose habitat, and the tree plantings on and around the lonetree wma, though more scattered across the landscape, also appear to provide important forest habitat for moose. woody browse dominated the diets of moose in both study areas, similar to prior research demonstrating the importance of woodlands in providing forage for moose (belovsky 1981, renecker and schwartz 1998). diets in the turtle mountains consisted in large part of browse species such as aspen, willow, birch, juneberry, and cherry that are typically considered common forage items of moose. in addition to many of these traditional browse species, the turtle mountains also had an abundance of bur oak which was a major component of summer and fall diets. in the lonetree area, the most common woody plants (except willow) are not typically found in traditional moose range, and this was reflected in the local diet. for example, russian olive is a common shrub in tree plantings and was the most abundant browse item (25% of overall diets), and green ash and box elder, also commonly planted, were ~11% of the winter and summer diets. woodlands was the only habitat moose selected for in all seasons, but seasonal changes in diet and the use of croplands and wetlands suggest that moose also selectively used seasonally available forage in other habitat types. thus, selection ratios may not entirely reflect the importance of habitats other than woodlands. for example, while moose avoided grassland habitats in both study areas, alfalfa was 13% of the summer and fall diets in the turtle mountains, indicating that this forb was an important supplemental seasonal forage. likewise, moose exhibited negative selection for croplands in all seasons, even though its use was greater in fall than in other seasons and corn was an important part of the fall and winter diets of lonetree moose. this apparent lack of selection may reflect the different composition of croplands in the two study areas. crops in the turtle mountains consisted almost entirely of small grains (wheat, barley) that were not expected to serve as moose forage, and in sheridan and wells counties where lonetree wma is located, ~26% of the total land area was planted in wheat and barley in 2005 with only 2% corn and 3% sunflowers (usda 2005). in contrast, if habitat selection analyses were confined to the boundary of the lonetree wma where the only croplands were corn and sunflower food plots, then moose would show an overall positive selection for cropland habitats (manly’s standardized selection ratio = 0.29). moose avoided wetlands in most seasons, possibly due to avoidance of open areas during warm daytime temperatures (olson et al. 2016). however, relative to winter, we movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 100 observed an increase in wetland use during summer and fall which was primarily driven by turtle mountain moose; use of wetlands was low year-round for lonetree moose (table 2). the increased use of wetlands was not reflected in the diets, as aquatic vegetation was a minor component (≤1%) of summer and fall diets in both study areas. however, these plants likely play an important role because it is believed moose consume aquatic plants for their critical minerals (de vos 1958, belovsky and jordan 1981) and high digestibility (maccracken et al. 1997). the latter may represent a potential limitation of this study because it may cause underrepresentation of aquatic plants in fecal samples. alternatively, the relative increase in wetland use in summer and fall may have been independent of forage requirements and triggered by thermoregulatory behavior or to avoid insects (de vos 1958, belovsky and jordan 1981). the combined results of home range, habitat use, and diet analyses provide insight into the factors influencing space use by moose in both traditional and prairie habitats in north dakota. while wooded habitats appear to be critical for moose throughout their range in north dakota, other seasonally available resources such as corn and alfalfa may provide supplemental food sources. further, the strong selection for planted woodlands and use of crops as a food source in the lonetree wma support the hypothesis that range expansion by moose is the direct result of landscape modifications occurring since european settlement. management implications moose are a prized big game species in north dakota, with >15,000 hunters applying annually for a once-in-a-lifetime license (north dakota game and fish department 2019). this study provides ecological information about the state’s moose population that will help managers make informed decisions to maintain and enhance this unique wildlife resource. while moose have expanded their range to include areas of north dakota that were historically prairie, the woodland habitats that they depend on constitute a very small proportion of the overall landscape in these areas, thereby requiring moose to have large home ranges to acquire sufficient resources. as a result, managers should recognize that prairie habitats are likely capable of supporting fewer moose than forested areas, and that the continued persistence of prairie populations of moose will be dependent on the maintenance of forest habitat. additionally, the planted woodlands and food plots of the lonetree wma may make this area particularly attractive to prairie moose. the continued management of this and other wmas to provide food and cover for wildlife should help support the state’s moose population in non-traditional range where availability of preferred habitats is limited. acknowledgments we thank north dakota epscor, the north dakota game and fish department, the u.s.d.i. bureau of reclamation, the u.s. fish and wildlife service, the north dakota chapter of the wildlife society, the university of north dakota biology department, and the wheeler scholarship for funding this project. we would also like to thank w. jensen, r. johnson, r. kreil, and s. peterson for permitting us to capture moose and conduct work on north dakota game and fish lands. additionally, we are grateful to r. newman, b. rundquist, and j. vaughan for their helpful input on this manuscript. finally, gratitude is expressed to j. smith, e. pulis, t. manuwal, j. faught, and j. rubbert for their assistance in the field. alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 101 references addison, r. b., j. c. williamson, b. p saunders, and d. fraser. 1980. radiotracking of moose in the boreal forest of northwestern ontario. canadian fieldnaturalist 94: 269–276. bakke, e. l. 1980. movements and habitat use of ruffed grouse in the turtle mountains, north dakota. m. s. thesis. university of north dakota, grand forks, north dakota, usa. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114: 1–49. belovsky, g. e. 1981. food plant selection by a generalist herbivore: the moose. ecology 62: 1020–1030. doi:10.2307/1937001 _____, and p. a. jordan. 1981. sodium dynamics and adaptations of a moose population. journal of mammalogy 62: 613–627. doi:10.2307/1380408 cederlund, g. h., and h. sand. 1994. home-range size in relation to age and sex in moose. journal of mammalogy 75: 1005–1012. doi:10.2307/1376618 courtois, r., c. dussault, f. potvin, and g. daigle. 2002. habitat selection by moose (alces alces) in clear-cut landscapes. alces 38: 117–192. de vos, a. 1958. summer observations on moose behavior in ontario. journal of mammalogy 39: 128–139. doi:10.2307/ 1376618 dodge, w. b., jr, s. r. winterstein, d. e. beyer, jr., and h. campa iii. 2004. survival, reproduction, and movements of moose in the western upper peninsula of michigan. alces 40: 71–86. doerr, j. g. 1983. home-range size, movements, and habitat use in two moose, alces alces, populations in southeastern alaska. canadian field-naturalist 97: 79–88. dussault, c., j. p. ouellet, r. courtois, j. huot, l. breton, and h. jolicier. 2005. linking moose habitat selection to limiting factors. ecography 28: 619–628. doi:10.1111/j.2005.0906-7590.04263.x elchuk, c. l., and k. l. wiebe. 2003. home range size of northern flickers (colaptes auratus) in relation to habitat and parental attributes. canadian journal of zoology 81: 954–961. doi:10.1139/z03-077 garner, d. l., and w. f. porter. 1990. movements and seasonal home ranges of bull moose in a pioneering adirondack population. alces 26: 80–85. garrison diversion project. 2019. history and federal legislation. (accessed september 2019). grauvogel, c. a. 1984. seward peninsula moose population identity study. federal aid in wildlife restoration final report. alaska department of fish and game, juneau, alsaka, usa. hemson, g., p. johnson, a. south, r. kenwards, r. ripley, and d. macdonald. 2005. are kernels the mustard? data from global positioning system collars suggests problems for kernel home-range analysis with least-squares cross validation. journal of animal ecology 74: 455–463. doi:10.1111/j.1365-2656.2005. 00944.x hundertmark, k. j. 1998. home range, dispersal, and migration. pages 303–335 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, dc, usa. kearney, s. r., and f. f. gilbert. 1976. habitat use by moose and white-tailed deer on sympatric range. journal of wildlife management 40: 645–657. doi:10.2307/3800559 knue, j. 1991. big game in north dakota: a short history. north dakota game and fish department, bismarck, north dakota, usa. labonte, j., j. p. quellet, and r. courtois. 1998. moose dispersal and its role in the maintenance of harvested populations. journal of wildlife management 62: 225–235. doi:10.2307/3802282 leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by http://www.garrisondiv.org/about/historyfederallegislation/ http://www.garrisondiv.org/about/historyfederallegislation/ http://www.garrisondiv.org/about/historyfederallegislation/ movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 102 moose in northern maine. journal of wildlife management 53: 880–885. doi:10.2307/3809581 licht, d. s. 1997. ecology and economics of the great plains. university of nebraska press, lincoln, nebraska, usa. lynch, g. m., and l. e. morgantini. 1984. sex and age differential in seasonal home-range size of moose in northcentral alberta. alces 20: 61–78. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. journal of wildlife management 61: (suppl. 136) 1–52. manly, b. f. j., l. l. mcdonald, d. l. thomas, t. l. mcdonald, and w. p. erickson. 2002. resource selection by animals: statistical design and analysis. 2nd edition. kluwer academic publishers, dordrecht, the netherlands. mcnab, b. k. 1963. bioenergetics and the determination of home range. american naturalist 97: 130–140. doi:10.1086/ 282264 mitchell, m. s., and r. a. powell. 2004. a mechanistic home range model for optimal use of spatially distributed resources. ecological modelling 177: 209–232. doi:10.1016/j.ecolmodel.2004.01.015 north dakota forest service. 2003. legendary forests: state forests guide. north dakota forest service, bottineau, nd, usa. north dakota game and fish department. 2019. moose hunting. (accessed september 2019). olson, b. t., s. k. windels, r. a. moen, and n. p. mccann. 2016. moose modify bed sites in response to high temperatures. alces 52: 153–160. osko, t. j., m. n. hiltz, r. j. hudson, and s. m. wasel. 2004. moose habitat preferences in response to changing availability. journal of wildlife management 68: 576–584. doi:10.2193/0022-541x (2004)068[0576:mhpirt] 2.0.co;2 peek, j. m. 1998. management of moose habitat. pages 377–401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, dc, usa. _____, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 1–65. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northeastern minnesota. journal of wildlife management 37: 266–278. doi:10.2307/3800117 pierce, j. d., and j. m. peek. 1984. moose habitat use and selection patterns in north-central idaho. journal of wildlife management 48: 1335–1343. doi:10.2307/ 3801794 poole, k. g., and k. stuart-smith. 2006. winter habitat selection by female moose in western interior montane forests. canadian journal of zoology 84: 1823–1832. doi:10.1139/z06-184 r core development team. 2007. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. (accessed september 2019). rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber management and natural disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61: 517–524. doi:10.2307/3802610 renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behavior. pages 403–439 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, dc, usa. https://gf.nd.gov/hunting/moose https://gf.nd.gov/hunting/moose http://www.r-project.org/ alces vol. 55, 2019 movements and resource use by moose – maskey and sweitzer 103 rodgers, a. r., and a. p. carr. 1998. home range extension (hre) for arcviewgis. ontario ministry of natural resources centre for northern forest ecosystem research, thunder bay, ontario, canada. sain, s. r., k. a. baggerty, and d. w. scott. 1994. cross-validation of multivariate densities. journal of the american statistical association 89: 807–817. doi: 10.1080/01621459.1994.10476814 schneider, r. r., and s. m. wasel. 2000. the effect of human settlement on the density of moose in northern alberta. journal of wildlife management 64: 513–520. doi:10.2307/3803249 seabloom, r. w., j. w. hoganson, w. f. jensen. 2011. the mammals of north dakota. institute for regional studies, north dakota state university, fargo, north dakota, usa. seaman, d. e., and r. a. powell. 1996. an evaluation of the accuracy of kernel density estimators for home range analysis. ecology 77: 2075–2085. doi:10.2307/ 2265701 _____, j. j. millspaugh, b. j. kernohan, g. c. brundige, k. j. raedeke, and r. a. gitzen. 1999. effects of sample size on kernel home range estimates. journal of wildlife management 63: 739–747. doi:10.2307/3802664 silverman, b. w. 1986. density estimation for statistics and data analysis. chapman and hall, london, united kingdom. smith, j. r., r. a. sweitzer, and w. j. jensen. 2007. diets, movements, and consequence of providing wildlife food plots for white-tailed deer in central north dakota. journal of wildlife management 71: 2719–2726. doi:10.2193/2006-379 stenhouse, g. b., p. b. latour, l. klutny, n. maclean, and g. glover. 1995. productivity, survival and movement of female moose in a low-density population, northwest territories, canada. arctic 48: 57–62. doi:10.14430/ arctic1224 stevens, o. a. 1966. plants of bottineau county, north dakota. north dakota school of forestry, bottineau, north dakota, usa. strong, laurence l., h. thomas sklebar, and kevin e. kermes. 2005. north dakota gap analysis project. northern prairie wildlife research center, jamestown, north dakota. online. (version 12jun2006; accessed september 2019). thompson, i. d., and r. w. stewart. 1998. management of moose habitat. pages 377–401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, dc, usa. _____, i. d., and m. f. vukelich. 1981. use of logged habitats in winter by moose calves in northeastern ontario. canadian journal of zoology 59: 2103–2114. doi:10.1139/z81-287 united states department of agriculture. 2005. national agricultural statistics service website. u. s. department of agriculture, national agricultural statistics service, north dakota field office, fargo, nd, usa. (accessed september 2019). _____-farm service agency. 2005. statewide naip photography 2005. usda-fsa aerial photography field office, salt lake city, utah, usa. (accessed september 2019). united states environmental protection agency. 1996. level iii ecoregions of the continental united states. national health and environmental effects research laboratory map m-1, various scales. united states environmental protection agency, corvallis, orgeon, usa. united states fish and wildlife service. 2000. national wetlands inventory http://www.npwrc.usgs.gov/projects/ndgap/ http://www.npwrc.usgs.gov/projects/ndgap/ http://www.nass.usda.gov/nd/ http://www.nass.usda.gov/nd/ http://www.apfo.usda.gov movements and resource use by moose – maskey and sweitzer alces vol. 55, 2019 104 website. u.s. department of the interior, fish and wildlife service, washington, dc, usa. (accessed september 2019). van vuren, d. h. 1984. summer diets of bison and cattle in southern utah. journal of range management 37: 260–261. doi:10.2307/3899151 wand, m. p., and m. c. jones. 1995. kernel smoothing. chapman and hall. london, united kingdom. worton, b. j. 1995. using monte carlo simulation to evaluate kernel-based home range estimators. journal of wildlife management 59: 794–800. doi:10.2307/ 3801959 http://www.fws.gov/nwi/ 138 distinguished moose biologist award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public's attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (>10 in journals including alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. 6. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual's interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3-8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special "distinguished moose biologist" presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http://bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca 97 assessing age of harvested moose prior to population declines in british columbia gerald w. kuzyk1, kaitlyn d. schurmann2, shelley m. marshall3, and chris procter4 1ministry of forests, lands and natural resource operations and rural development, 205 industrial road g, cranbrook, british columbia v1c 7g5, canada; 2ministry of forests, lands and natural resource operations and rural development, p.o. box 9391, victoria, british columbia v8w 9m8, canada; 3ministry of forests, lands and natural resource operations and rural development, 2000 s ospika boulevard, prince george, british columbia v2n 4w5, canada; 4ministry of forests, lands and natural resource operations and rural development, 1259 dalhousie drive, kamloops, british columbia v2c 5z5, canada. abstract: moose populations in parts of british columbia, canada have been declining since about the mid-2000s with the licensed harvest dropping by more than half from 1987 to 2014. a tooth reporting program for harvested moose from 1982 to 2003 enabled us to assess the relationship between age of harvested moose and 1) time (1982–2003), 2) level of licensed harvest of bulls and cows, and 3) estimated populations prior to declines with age data collected after decline in the province. we used age data determined from cementum annuli of teeth collected from hunter returns from 72,888 moose (n = 57,376 bulls and n = 15,512 cows). we found average age of harvested bulls and cows to be 3.32 ± 0.02 and 4.99 ± 0.06 years, respectively, similar to ranges reported elsewhere in western north america. age of bulls declined linearly by year, whereas age of cows declined in the latter half of the study period. the average age of cows harvested from 1983 to 2003 prior to the population decline (n = 2,016; mean = 3.84 years, sd = 3.03) was 7 years younger than that of a small sample of cows dying of multiple causes (harvest and natural) during the decline (n = 47; mean = 10.93 years, sd = 3.72). we acknowledge the logistical and financial constraints required to gather a representative sample of teeth from harvested moose, but recommend reimplementation of a tooth collection program to provide continuous information on the age structure of moose populations to help guide management decisions. alces vol. 56: 97–106 (2020) key words: age, alces, bulls, cows, harvest, population, survival, tooth determining age from teeth of harvested ungulates provides important demographic information about populations to help guide management strategies. understanding age and sex of ungulate populations assists with determining productivity, age-specific survival, and reproduction (gove et al. 2002), evaluating expected population dynamics (loison et al. 1999), and estimating maximum yield in harvested systems (sæther et al. 2001, clutton-brock et al. 2002, nilsen et al. 2005). survivorship may also be inferred from age distribution of ungulates, though it may be biased if age structure changes during the period of survival estimation (eberhardt 2002). using the age of harvested moose is common and useful in developing effective management strategies and understanding moose population dynamics (timmerman and buss 2007). it is used to develop standing age distributions or estimate age structure of moose populations, recognizing there is no standard population age structure because of moose harvest ages in british columbia – kuzyk et al. alces vol. 56, 2020 98 fluctuations in survival and reproduction (van ballenberghe and ballard 2007). the population age structure of moose and how it varies over time (peterson 1977) is useful to interpret population trends because variation in survival and reproduction may be related to age structure (solberg et al. 2000, ericsson et al. 2001, sæther et al. 2001). harvest age data can also serve as an index of harvest rate and provide for assessment of its effects on sex structure and the mean age of females and males in the population (langvatn and loison 1999, milner et al. 2007). generally, the mean age of moose populations declines as harvest rate increases and relationships between mean age and population density indicate that mean ages are lower in low density populations with few older individuals (bowyer et al. 1999). recent declines in moose populations in parts of north america (timmermann and rodgers 2017, jensen et al. 2018) have created challenges for maintaining sustainable hunting opportunities. populations in british columbia have declined in ~70% of moose range in the last decade (kuzyk et al. 2018) and licensed harvest has declined by approximately half since 1987 (kuzyk 2016). declines within central british columbia coincided with a mountain pine beetle (dendroctonus ponderosae) outbreak where habitat changes and increased salvage logging and road building were hypothesised to increase vulnerability to harvest and predation (kuzyk and heard 2014). although british columbia had a voluntary tooth return program for licenced resident and non-resident hunters from 1982 to 2003 to provide harvest age information, it was discontinued due to financial constraints prior to the onset of population declines in the mid 2000s. the purpose of this study was to determine and assess change in the age of harvested moose populations in british columbia long-term, both prior to (1982–2003) and post-population decline in the mid-2000s. this assessment used two regions in british columbia with sufficient samples of harvested cows. specific objectives were to: 1) estimate the mean age of moose populations and how it changed over a 21-year period, 2) assess relationships between harvest age of bulls and cows over time and levels of harvest, and 3) compare the mean age of moose populations prior to, during, and after population decline. study area we assessed age of bulls and licenced harvest levels of moose at a provincial scale in 186 wildlife management units (wmu) where hunting was authorized from 1982 to 2003. we also examined cow age and harvest, but this was limited to only two regions of the province, region 7a and region 3 (fig. 1) where sufficient samples of harvested cow moose were available. in these areas moose occupy habitats in northern boreal forests, dry interior forests, and mountainous habitats (eastman and ritcey 1987, meidinger and pojar 1991) and are sympatric with mule deer (odocoileus hemionus), white-tailed deer (o. virginianus), elk (cervus elaphus), bison (bison bison), and caribou (rangifer tarandus) (shackleton 1999). moose co-exist with 4 main predators: wolves (canis lupus) and black bears (u. americanus) throughout moose range, grizzly bears (ursus arctos) in all areas except parts of the south-central interior, and cougars (felis concolor) primarily in the central and southern areas (spalding and lesowski 1971). methods licenced hunters voluntarily submitted a front incisor from moose harvested between 1982 and 2003, except in 1999 when the collection program was suspended temporarily. from 1982 to 1988, teeth were alces vol. 56, 2020 moose harvest ages in british columbia – kuzyk et al. 99 collected voluntarily through the harvest card tooth return program or at hunter check stations (hatter 1993). from 1989 to 2003 (excluding 1999), the voluntary tooth return program was made more prominent and easier to participate in as hunters received a harvest data envelope when purchasing their hunting license or by mail if they received a limited entry hunting (leh) authorization (hatter 1993, child et al. 2010). hunters were requested to document date, location (management unit), and sex of kill on the harvest data envelope, include an incisor, and mail the pre-paid envelope to the provincial wildlife agency. ages of bulls and cows were determined from sectioned incisor teeth (sergeant and pimlott 1959) by two trained wildlife staff (child et al. 2010). there was sufficient harvest of cows in region 7a and region 3 to enable an assessment of cow age and harvest levels at a regional scale. we also compared ages of harvested cows (n = 2,016) in region 3 and region 7a from 1982 to 2003 (prior to the decline) to a small sample of cows (n = 47) dying of harvest and natural causes during the decline (2012–2018) as part of related research (kuzyk et al. 2019, sittler 2019). ages were determined in a professional laboratory (matson’s laboratory, montana, usa) fig. 1. distribution of moose with licensed hunting seasons in british columbia at the provincial scale (grey, includes regions 7a and 3), region 7a (green), and region 3 (yellow). regions 7a and 3 had licensed cow hunting seasons and included the research sites where mortality of radio-collared cow moose was monitored. moose harvest ages in british columbia – kuzyk et al. alces vol. 56, 2020 100 specializing in using cementum annuli to age sectioned teeth. because teeth were aged by multiple trained individuals, we compared a sample of 35 teeth aged by a government biologist and matson’s laboratory; 70% were aged within ±1 year of age (authors, unpublished data). the majority of difference occurred with moose ≥10 years of age because separation of annuli declines with age. licenced resident harvest was estimated annually in 1982 to 2003 from a provincial survey conducted with mail-out questionnaires sent to a random sample of resident moose hunters. these estimates (± 95% confidence interval) were produced from an annual average of 14,278 ± 2,732 questionnaires with an average response rate of 69.4 ± 2.7%. non-resident licenced harvest was obtained from mandatory reports completed by guideoutfitters immediately following their hunts. total harvest was a combination of bulls (>1 year of age), cows (>1 year of age), and calves (<1 year of age). trends in age of harvested moose from 1982 to 2003 were assessed by using linear, second-degree polynomial, and third-degree polynomial regression analyses. polynomial regression analysis was used to identify possible increases and decreases in annual harvest age over time. the best regression model was selected using akaike’s information criteria (aic) (burnham and anderson 2002). analysis of general harvest trends was limited by variability in annual harvest data. for simplicity, licensed harvest trends were assessed with two-sample t-tests to examine for difference between the estimated harvest in the first 5 years (1982–1986) and the last 5 years (1999–2003) of the study period. pearson’s correlation coefficient was used to test for relationships between bull and cow harvest age and licensed harvest levels; significance was set at p < 0.05. results the mean age of harvested bulls province-wide was 3.32 ± 0.02 years (n = 57,376) (fig. 2), 3.17 ± 0.04 years (n = 19,610) in region 7a (fig. 3), and 2.96 ± 0.09 years (n = 3,226) in region 3 (fig. 4). age of bulls declined linearly by year in all study areas (table 1, fig. 2–4). age of bulls and harvest level were negatively correlated in region 7a (r = −0.69, p ≤ 0.01; fig. 3), positively correlated in region 3 (r = 0.76, p ≤ 0.01; fig. 4), but not correlated province-wide (fig. 2). mean age of harvested cows (n = 15,512) was 4.99 ± 0.06 years province-wide (fig. 2), 4.53 ± 0.07 years (n = 8,819) in region 7a (fig. 3), and 5.34 ± 0.19 years (n = 1,358) in region 3 (fig. 4). age of harvested cows followed a second-order polynomial trend in all study areas, with age declining in the latter half of the study period (table 2). age and harvest level were positively correlated in region 3 (r = 0.44, p ≤ 0.05; fig. 4), but unrelated in region 7a (fig. 3) and provincewide (fig. 2). in regions 3 and 7a, the mean age of harvested cows prior to the population decline (3.84 years, sd = 3.03; n = 2,016) was 7 years younger (t = −12.98, p ≤ 0.01) than that of cows (10.93 years, sd = 3.72; n = 47) dying from a variety of causes between 2012 and 2018 during and after the decline. the mean annual harvest was 11,302 ± 736 moose province-wide (fig. 2), with 3,355 ± 202 moose in region 7a (fig. 3) and 550 ± 117 moose in region 3 (fig. 4). harvest trends were stable province-wide (p ≥ 0.05; fig. 2) and in region 7a (p ≥ 0.10; fig. 3), but declined in region 3 (p ≤ 0.05; fig. 4). the mean annual bull harvest was 8,975 ± 568 province-wide (fig. 2), 1,926 ± 195 in region 7a (fig. 3) and 429 ± 80 in region 3 (fig. 4). the trend in bull harvest was considered stable province-wide (p ≥ 0.10; fig. 2), increased in region 7a (p ≤ 0.01; fig. 3) alces vol. 56, 2020 moose harvest ages in british columbia – kuzyk et al. 101 fig. 2. moose harvest data in 1982–2003, british columbia, canada. (a) average age of bull (blue dots) and cow (red dots) moose and estimated provincial harvest (blue bar for bulls, red bar for cows, grey bar for calves) of moose error bars are 95% confidence intervals. (b) average age of bull (blue dots) and cow (red dots) moose and estimated harvest (blue bar for bulls, red bar for cows, grey bar for calves) of moose in region 7a in british columbia from 1982 to 2003. error bars are 95% confidence intervals. (c) average age of bull (blue dots) and cow (red dots) moose and estimated harvest (blue bar for bulls, red bar for cows, grey bar for calves) of moose in region 3 in british columbia from 1982 to 2003. error bars are 95% confidence intervals. a b c moose harvest ages in british columbia – kuzyk et al. alces vol. 56, 2020 102 and declined in region 3 (p ≤ 0.05; fig. 4). the mean annual cow harvest was 1,079 ± 210 province-wide (fig. 2), 597 ± 44 in region 7a (fig. 3), and 84 ± 31 in region 3 (fig. 4). the cow harvest trend was stable (p ≥ 0. 05) province-wide (fig. 2) and in regions 7a (fig. 3) and 3 (fig. 4) throughout the study period. discussion the average age of harvested moose by licensed hunters in british columbia from 1982 to 2003 was 3.3 and 5.0 years for bulls and cows, respectively. these data were collected prior to population declines that began in the mid-2000s and later in parts of the central interior (kuzyk et al. 2018), and similar to those reported elsewhere in western north america. for example, in alberta the average age of harvested bulls was 2.5–2.7 years in heavily hunted areas and 3.5 years where hunting pressure was lighter (lynch 2006). we also found that the majority (66%) of bull moose harvested in british columbia was ≤3 years old. in alaska, 84% of harvested bull moose on the kenai peninsula was ≤3 years old (schwartz et al. 1992), and on kalgin island bowyer et al. (1999) found a relationship between the mean age of moose and population density such that, at low density, the mean age of both bulls and cows was ~2 years, increasing to ~3 years at higher densities. we found a higher average age for harvested cows than bulls in regions 3 and 7a. the mean age of harvested cows (5 years) in british columbia was lower than that in alberta where the average age was 7.2 years in an area with minimal first nations harvest (lynch 2006). the different population age structure of cows in british columbia and alberta in those respective timeframes suggests that either cow mortality rates were higher in british columbia due to natural causes and/or higher harvest pressure (licensed and/or first nation), or that calf recruitment rates that affect age structure differed in the two areas. we found a higher average age of harvested cows than reported by bowyer et al. (1999) in alaska (i.e., 2–3 years) in a heavily harvested population where the objective was to reduce population density. the average age of harvested cows in region 3 and region 7a prior to the population decline was younger (3.84 years) than that of radio-collared cows (10.93 years) that died of all causes (i.e., predation, hunting, health, natural accident) in portions of both regions from 2012 to 2018 (kuzyk et al. 2019, sittler 2019). a more accurate comparison would use only table 1. comparison of aic values associated with regression analyses used to evaluate trends in age of harvested moose age in 1982–2003, british columbia, canada. grey indicates top selected model. polynomial = degree of polynomial; df = degrees of freedom; aic = akaike’s information criterion. study area sex polynomial df aic province bulls 1 3 −15.08 bulls 2 4 −13.08 bulls 3 5 −13.86 cows 1 3 33.55 cows 2 4 17.18 cows 3 5 18.71 region 7a bulls 1 3 8.26 bulls 2 4 8.97 bulls 3 5 7.06 cows 1 3 19.07 cows 2 4 14.32 cows 3 5 13.82 region 3 bulls 1 3 25.46 bulls 2 4 25.78 bulls 3 5 25.43 cows 1 3 41.88 cows 2 4 38.87 cows 3 5 40.83 alces vol. 56, 2020 moose harvest ages in british columbia – kuzyk et al. 103 harvested radio-collared cows, but sample size of this group was insufficient. in the latter study, most moose died at an older ager than cows harvested prior to the decline in regions 3 and 7a, suggesting that a preponderance of older individuals were radio-collared as most mortality was unrelated to age. for these radio-collared cows, the average age of death was 11 years by predation (11 years for wolves, 10 years for bears, and 14 years for cougar), 9 years for apparent starvation, and 9 years at harvest (kuzyk et al. 2019). these data suggest that the individuals captured and monitored for research purposes reflected the age structure of a moose population skewed to older individuals, assuming captures were random. calf survival and recruitment may be a primary factor explaining moose population declines, which provides additional support for an age structure skewed toward older individuals, especially if lower calf recruitment occurred over a period of several years (kuzyk et al. 2018). aerial survey data within the research areas indicate that calf recruitment rates were lower in recent years (kuzyk et al. 2019). an alternative explanation is that age at death is not reflective of the age structure of the standing moose population because older cows are more vulnerable to certain mortality causes, including predation as suggested elsewhere (peterson 1977, montgomery et al. 2014). further, the risk of mortality from all causes other than harvest increases with age and becomes most apparent after 10 years of age (ericsson and wallin 2001). harvest trends of bull moose were stable province-wide, increased in region 7a, and decreased in region 3. surveys in the southern part of 7a indicated that stable populations (hatter 1999) and improved hunter access may have contributed to the increasing harvest trend. the primary cause of lower bull harvests in region 3 was due to season changes that purposefully reduced the harvest level. in 1984, the 55-day general open season (gos) for any bull was reduced to 24 days which reduced harvest initially, and in 1993 the any bull gos was closed and replaced with a limited entry hunt (leh) for bulls and an antler-restricted gos for spike/fork bulls. ultimately, the trends in bull harvest in region 3 were unrelated to population trends during this timeframe. cow harvests were stable province-wide and in regions 7a and 3. we found a positive correlation between age and harvest level of bulls in region 3, a negative correlation in region 7a, and no relationship at the provincial level. the sharp decline in average age of harvested bulls in region 3 in the early 1990s reflected unsustainable harvest as evidenced by a declining bull ratio during that timeframe, and provided the rationale for the substantial season changes in 1993. thereafter, lower harvest age of bulls was mostly due to implementation of the gos focused on yearling bull moose (i.e., spike/fork gos). increasing trends in bull harvest age through the late 1990s and early 2000s were correlated with an increasing population indicated by aerial surveys during that period. the negative relationship between age and harvest level in region 7a is consistent with the idea that as harvest increases, population age structure declines. the average bull age in region 7a suggests that harvest pressure was lower in this region than in alberta (lynch 2006) and alaska (schwartz et al. 1992). no significant relationship was found between age and harvest level of cows province-wide or in region 7a. an unexpected positive correlation was found in region 3 that contrasted with stratified random block survey data (risc 2002) indicating that moose populations were moose harvest ages in british columbia – kuzyk et al. alces vol. 56, 2020 104 increasing; harvest of antlerless moose was purposely reduced for reasons other than trends in moose numbers. although licensed harvests of antlerless moose were stable to declining, first nation harvest levels were unknown and if increasing, would reduce the mean age of cows. alternatively, this trend may reflect the addition of younger animals to the moose population; here, moose populations increased as mean age began to drop and concurrent survey data indicated higher calf:cow ratios in many units during mid-winter surveys (i.e., 50–71 calves/100 cows). however, this is in contrast to when mean age increases with increasing density in a harvested population (bowyer et al. 1999) and our results with bull harvests. we acknowledge the logistical and financial constraints required to gather a representative sample of teeth from harvested moose, but recommend reinitiating a tooth collection program in british columbia. harvest age provides critical information about moose populations and insight regarding population trends when combined with age data from unlicensed harvests and non-hunted animals. population management decisions that consider harvest level, hunter effort, and hunt structure would benefit measurably from such data, especially in areas where survey data is lacking or difficult to obtain. acknowledgements we would like to thank the many hunters that submitted moose teeth for aging that enabled this analysis. d. heard, a. walker, m. carstensen, and 2 anonymous reviewers provided important reviews which improved this manuscript. references bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–89. burnham, k. p., and d. s. r. anderson. 2002. model selection and multimodel inference: a practical informationtheoretic approach, 2nd edition. springer-verlag, new york, new york, usa. child, k., d. a. aitken, and r. v. rea. 2010. morphometry of moose antlers in central british columbia. alces 46: 123–134. clutton-brock, t., t. n. coulson, e. j. milner-gulland, d. thomson, and a. m. armstrong. 2002. sex differences in emigration and mortality affect optimal management of deer populations. nature 415: 633–637. doi: 10.1038/415633a eastman, d., and r. ritcey. 1987. moose habitat relationships and management in british columbia. swedish wildlife research supplement 1: 101–117. eberhardt, l. l. 2002. a paradigm for population analysis of long-lived vertebrates. ecology 83: 2841–2854. doi: 10.1890/ 0012-9658(2002)083[2841: apfpao] 2.0.co;2 ericsson, g., and k. wallin. 2001. agespecific moose (alces alces) mortality in a predator-free environment: evidence for senescence in females. ecoscience 8: 157–163. doi: 10.1080/11956860.2001. 11682641 _____, _____, j. p. ball, and m. broberg. 2001. age-related reproductive effort and senescence in free-ranging moose, alces alces. ecology 82: 1613–1620. doi: 10.1890/0012-9658(2001)082 [1613: arreas]2.0.co;2 gove, n. e., j. r. skalski, p. zager, and r. l. townsend. 2002. statistical models for population reconstruction using age-at-harvest data. journal of wildlife management 66: 310–320. doi: 10.2307/ 3803163 hatter, i. w. 1993. yearling moose vulnerability to spike-fork antler regulation. alces vol. 56, 2020 moose harvest ages in british columbia – kuzyk et al. 105 british columbia environment wildlife branch memorandum, 26 january 1993. wildlife branch, british columbia environment, victoria, british columbia, canada. _____. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. jensen, w. f., j. r. smith, m. carstensen, c. e. penner, b. m. hosek, and j. j. maskey, jr. 2018. expanding gis analyses to monitor and assess north american moose distribution and density. alces 54: 45–54. kuzyk, g. w. 2016. provincial population and harvest estimates of moose in british columbia. alces 32: 1–11. _____, i. hatter, s. marshall, c. procter, b. cadsand, d. lirette, h. schindler, m. bridger, p. stent, a. walker, and m. klaczek. 2018. moose population dynamics during 20 years of declining harvest in british columbia. alces 54: 101–119. _____, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. british columbia ministry of forest, lands and natural resource operations, victoria, british columbia, canada. _____, c. procter, s. marshall, h. schindler, h. schwantje, m. scheideman, and d. hodder. 2019. factors affecting moose population declines in british columbia. wildlife working report. no. wr-127. british columbia ministry forest, lands and natural resource operations and rural development, victoria, british columbia, canada. langvatn, r., and a. loison. 1999. consequences of harvesting on age structure, sex ratio and population dynamics of red deer cervus elaphus in central norway. wildlife biology 5: 215–223. doi: 10.2981/wlb.1999.026 loison, a., m. festa-bianchet, j. gaillard, j. t. jorgenson, and j. jullien. 1999. age-specific survival in five populations of ungulates: evidence of senescence. ecology 80: 2539–2554. doi: 10.1890/0012-9658(1999)080 [2539:assifp]2.0.co;2 lynch, g. m. 2006. does first nation’s hunting impact moose productivity in alberta? alces 42: 25–31. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. special report series number 6. british columbia ministry of forests, victoria, british columbia, canada. milner, j. m., e. b. nilsen, and h. p. andreassen. 2007. demographic side effects of selective hunting in ungulates and carnivores. conservation biology 21: 36–47. doi: 10.1111/j.1523-1739. 2006.00591.x montgomery, r. a., j. a. vucetich, g. j. roloff, j. k. bump, and r. o. peterson. 2014. where wolves kill moose: the influence of prey life history dynamics on the landscape ecology of predation. plos one 9(3): e91414. doi: 10.1371/journal. pone.0091414 nilsen, e. b., t. pettersen, h. gundersen, j. m. milner, a. mysterud, e. j. solberg, h. p. andreassen, and n. c. stenseth. 2005. moose harvesting strategies in the presence of wolves. journal of applied ecology 42: 389–399. doi: 10.1111/j.1365-2664.2005.01018.x peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. national park service scientific monograph series, no. 11. united states government printing office, washington, d. c., usa. resources information standards committee (risc). 2002. aerial-based inventory methods for selected ungulates: bison, mountain goat, mountain sheep, moose, elk, deer and caribou. standards for components of british columbia’s biodiversity no. 32, version 2.0. british moose harvest ages in british columbia – kuzyk et al. alces vol. 56, 2020 106 columbia ministry of sustainable resource management, victoria, british columbia, canada. sæther, b.-e., s. engen, and e. j. solberg. 2001. optimal harvest of age-structured populations of moose alces alces in a fluctuating environment. wildlife biology 7: 171–179. doi: 10.2981/wlb.2001.021 schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula, alaska. alces 28: 1–13. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23: 315–321. doi: 10.2307/ 3796891 shackleton, d. 1999. hoofed mammals of british columbia. royal british columbia museum handbook. university of british columbia press, vancouver, canada. sittler, k. l. 2019. moose limiting factors investigation. wildlife infometrics inc. report no. 678. annual report 2018–19. wildlife infometrics inc., mackenzie, british columbia, canada. solberg, e. j., a. loison, b. saether, and o. strand. 2000. age-specific harvest mortality in a norwegian moose alces alces population. wildlife biology 6: 41–52. doi: 10.2981/wlb.2000.036 spalding, d. j., and j. lesowski. 1971. winter food of the cougar in south central british columbia. journal of wildlife management 35: 378–381. timmerman, h. r., and m. e. buss. 2007. population and harvest management. pages 559–615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. _____, and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. van ballenberghe, v., and w. b. ballard. 2007. population dynamics. pages 223–245 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. alces39_89.pdf alces vol. 39, 2003 nygrén – multiple fecundity in moose 89 the potential for multiple fecundity of moose in finland tuire nygrén finnish game and fisheries research institute, ilomantsi game research station, fi-82900 ilomantsi, finland abstract: multiple fecundity (i.e., >2 fetuses or calves per female) is a rare and poorly known phenomenon in moose (alces alces). in this paper i: (1) report the frequency of multiple fecundity of moose in finland; (2) study the frequencies of multiple fecundity in different years and areas; (3) discuss the viability of litters with different numbers of progeny; and (4) discuss the possible fecundity effects of selective harvest and the evolutionary aspects of multiple fecundity. the embryo numbers of harvested cows were counted during 1980-89 (n = 2,347) and the proportion of single, twin, and triplet calves were determined from the 1986-99 moose observation material recorded in the field by hunters during the hunting season (n = 585,149). the material includes 4 sets of quadruplet calves, 1 set of stillborn sextuplets, and a moose female with 5 sets of triplet calves; a total of 30 calves in 15 years. in finland, 60.38 % of pregnant moose cows had one, 39.37% two, 0.21% three, and 0.04% four embryos. in the observation material, 61.79% of the cows had one calf, 38.18% twin calves, and 0.03% triplet calves. the proportion of multiple cases decreased from south to north. the viability of single and twin calves was found to be very high, but only 15% of the sets of triplet calves seemed to survive up to the first fall. calf survival rate was clearly higher in 1980-99 than in 1963-66, possibly depending on the different age structures of the female populations. according to the literature, the frequency of multiple fecundity in moose appears to be lower in north american than european moose populations. alces vol. 39: 89-107 (2003) key words: alces alces, calf survival, fecundity, moose, multiple fecundity, quadruplet calves, reproduction, sextuplet calves, triplet calves, twinning in favorable conditions, the reproductive potential of a moose (alces alces) population can be very high (markgren 1969, stålfelt 1974, nygrén 1983, cederlund and sand 1991, danilkin and ulitin 1998, schwartz 1998). most females annually produce either single or twin calves. the non-reproducers are pubescent females (12 years old), very old cows (15-20 years), or females unable to give birth every year while living on the poor forage or harsh conditions of the north (peek 1962, blood 1974, markgren 1974, albright and keith 1987). the most productive females can give birth to as many as 3 (peterson 1955, heptner et al. 1966, danilkin 1999) or 4 calves (skuncke 1949, knorre 1959, ling 1974, martin 1989, vitakova and minajev 2000). in different moose populations the proportion of singles and twins varies (markgren 1969, mech et al. 1987, boer 1992, gasaway et al. 1992, danilkin and ulitin 1998, schwartz 1998) but in all moose populations multiple fecundity (i.e., >2 fetuses or calves per single moose female) is a very rare phenomenon (pimlott 1959, heptner et al. 1966, franzmann 1981, kozlo 1983, schwartz 1998). in addition, the number of reports of triplet or quadruplet sets is small (table 1) and analytical papers dealing with the maximal productivity of moose are practically nonexistent. ling (1974) and kozlo (1983) might be the only authors who report multiple fecundity in moose – nygrén alces vol. 39, 2003 90 t ab le 1 . r ep or ts o f m ul ti pl e fe cu nd it y ca se s in fe m al e m oo se ( a lc es a lc es ). r eg io n1 m et ho d2 sa m pl e n um be r o f % n um be r o f a ut ho r si ze 3 t ri pl et s t ri pl et s q ua dr up le ts n a 1 + se to n (1 92 7) c it ed b y pe te rs on (1 95 5) n a , a la sk a 1 1 h os le y an d g la se r ( 19 52 ) n a , o nt ar io 1 3 pe te rs on (1 95 5) n a , n ew fo un dl an d 3 1, 57 9 2 0. 13 pi m lo tt (1 95 9) n a , i sl e r oy al e 1 1 m ol l a nd m ol l ( 19 76 ) c it ed b y m ar ti n (1 98 9) n a , a la sk a 3 3, 31 4 2 0. 06 b ai le y an d b an gs (1 98 0) n a , a la sk a 1 1 fr an zm an n (1 98 1) n a , a la sk a 3 1 fr an zm an n an d sc hw ar tz (1 98 5) n a , i sl e r oy al e 1 1 1 m ar tin (1 98 9) ru 1 1 b ut ur li n (1 89 0) c it ed b y t im of ej ev a (1 97 4) r u , k om i 2 >2 1 k no rr e ( 19 59 ) r u , t at ar sk ij 2 3 z ar ip ov a nd z na m en sk ij i ( 19 64 ) c it ed b y h er uv im ov (1 96 9) r u , l en in gr ad 2 2, 13 9 4 0. 19 d em en te v (1 96 7) c it ed b y l in g (1 97 4) r u , k ir ov 2 1 pa vl ov a nd ja za n (1 96 7) c it ed b y t im of ej ev a (1 97 4) r u , v la di m ir 2 58 1 1. 72 sy so je v (1 96 7) c it ed b y l in g (1 97 4) r u m os co w 3 24 8 1 0. 40 m ak ar ov a (1 96 9) c it ed b y l in g (1 97 4) r u t am bo v 2a 14 1 2 1. 42 h er uv im ov (1 96 9) r u , l en in gr ad 2 1, 85 0 4 0. 22 n ov ik ov (1 97 0) c it ed b y l in g (1 97 4) r u , n ov os ib ir sk 2 35 8 1 0. 28 z in ov je v (1 97 1) c it ed b y d an il ki n (1 99 9) r u , v ol ga 2 54 1 1. 85 ja za n (1 97 2) r u , l en in gr ad 2 3, 36 4 4 0. 12 t im of ej ev a ( 19 74 ) r u , b ye lo ru ss ia 2a 19 0 3 1. 58 k oz lo (1 98 3) r u , m ur m an sk 2 1, 20 9 1 0. 08 m ak ar ov a (1 98 1) r u , a rk ha ng el 2 67 8 2 0. 29 fi lo no v (1 98 3) r u , v ol og da 2 2, 41 6 3 0. 12 fi lo no v (1 98 3) r u , t ve r 2 1, 72 6 1 0. 06 fi lo no v (1 98 3) r u , j ar os la vl 2 1, 31 2 2 0. 15 fi lo no v (1 98 3) r u , m os co w 2 2, 82 6 1 0. 04 fi lo no v (1 98 3) r u , k al ug a 2 62 7 2 0. 32 z ai ki n an d v or on in (1 98 6) c it ed b y d an il ki n (1 99 9) alces vol. 39, 2003 nygrén – multiple fecundity in moose 91 r u , k ir ov 2a 32 7 3 0. 92 g lu sh ko v (1 98 7) r u , k om i f ar m 4 23 0 1 0. 43 k oz hu ho v (1 98 9) r u , c he rn oz em je 2 2, 04 0 2 0. 10 pr os ta ko v (1 99 6) c it ed b y d an il ki n (1 99 9) r u , k os tr om a fa rm 4 31 5 9 2. 86 1 v it ak ov a an d m in aj ev (2 00 0) b c , e st on ia 3 6, 72 1 60 0. 89 16 l in g (1 97 4) b c , e st on ia 2a 67 1 k ir k an d t ön is so n (1 99 4) b c , e st on ia 1 >4 k ir k an d t ön is so n (2 00 0) b c , e st on ia 2a 11 4 2 1. 75 k irk (2 00 1) sc , s w ed en 2 + l ön nb er g (1 92 3) sc , s w ed en 1 9 1 sk un ck e (1 94 9) sc , f in la nd 2a 40 2 3 0. 75 k oi vi st o an d r aj ak os ki (1 96 6) sc , n or w ay 1 2 l in g (1 97 4) 1 n a = n or th a m er ic a, r u = r us si a, b c = b al ti c c ou nt ri es , s c = f en no sc an di a. 2 1 = ra nd om o bs er va ti on s, 2 = f et us es , d et er m in ed b y hu nt er s, 2 a = fe tu se s, d et er m in ed b y sc ie nt is ts , 3 = c ow -c al f ob se rv at io ns , 4 = c al ve s bo rn o n m oo se f ar m s. 3 p re gn an t f em al es o r co w -c al f ob se rv at io ns . and also analyze triplet or quadruplet observations or frequencies. the purpose of this paper is to: (1) report the frequency of multiple fecundity of moose in finland from 1980 to 1999; (2) study the spatial and temporal variation of the frequency of multiple fecundity and the effects of environmental and geographical factors; (3) study the viability of litters with different numbers of progeny; (4) discuss the possible fecundity effects of selective harvest and evolutionary aspects of multiple fecundity; and (5) discuss possible differences in multiple fecundity of european and north american moose populations. study area the study area was the whole of finland. finland mainly belongs to the boreal forest or taiga region, but is situated 5001,000 km further to the north than boreal forests elsewhere. in fennoscandia, the vegetation is boreal but the light conditions are arctic (solantie 2001). the typical vegetation types in finland are coniferous forests (70%) and mires (30%). approximately 50% of the forests are pine (pinus spp.) dominated, 30% spruce (picea spp.) dominated, and 7% birch (betula spp.) dominated. since the 1950s, artificial forest regeneration with pine plantations throughout finland has dramatically increased the amount of moose forage, but at the same time, it has restricted the distribution of other forage species. in finland, all boreal vegetation zones are represented (fig. 1). the bedrock of northern and eastern finland (lapland, oulu, kainuu, and pohjoiskarjala) is mainly composed of ancient, nutrient-poor rock. the bedrock is more fertile in the southwestern part of finland. the terrain of finland is quite flat and the total area of lakes constitutes approximately 10% of the land surface. due to the strong effect of the sea, the climate is moist. the winters are not as cold as in other boreal multiple fecundity in moose – nygrén alces vol. 39, 2003 92 northern boreal middle boreal southern boreal hemiboreal sweden norway russia estonia gulf o f finlan d gu lf o f b ot hn ia n oroarctic fig.1. vegetation zones and game management districts of the study area in finland. areas. there are 110 snow-days per year in the hemi-boreal area and 200 days around the border of the central and northern boreal zones. the deepest snow layer (long-term average in march > 60 cm, 30-year maximum 90-120 cm) and the mostly continental climate is typical for the districts of pohjoiskarjala, kainuu, oulu, and lapland where a small number of large predators exist and take their share of the moose population. in these areas (except lapland), moose are the only numerous ungulate prey for brown bear (ursus arctos) and wolf (canis lupus). in coastal finland, the seasons are less pronounced, the climate is more windy and humid, and the snow depth is lower (longterm average < 30 cm, 30-year maximum 60-80 cm). data were collected in 15 game management districts: etelä-häme (eh), eteläsavo (es), kainuu (ka), keski-suomi (ks), kymi (ky), lapland (la), oulu (ou), pohjanmaa (po), pohjois-häme (ph), pohjois-karjala (pk), pohjois-savo (ps), ruotsinkielinen pohjanmaa (rp), satakunta (sk), uusimaa (um), and varsinais-suomi (vs). the districts were grouped into 4 regions: coastal finland (eh, ky, um, vs, sk, and rp), inland finland (es, ks, ph, po, pk, and ps), oulu (ka and ou), and lapland (la) (figs. 1 and 2). the administrative divisions of the data were based on the established practices of moose management in finland. the division into regions was based on vegetation zones, the position/distance from the coastline, and comparative calculations and long-term observations of the spatial characteristics of the moose population (nygrén and pesonen, finnish game and fisheries research institute, unpublished data). in 1981-96, the estimated numbers in the winter moose population in the study area were between 67,000 and 93,000; the number of annual moose kills was between 26,000 and 69,000, and the average percentage of kill in the total population was 39% (t. nygrén and pesonen 1989; nygrén 1996; t. nygrén, finnish game and fisheries research institute, unpublished data). methods the finnish game and fisheries research institute (fgfri) collected data with the aid of finnish hunters as part of the finnish moose research program. the program has been conducted mainly for management purposes since 1972. two types of data and samples were used. r e p r o d u c t i v e o r g a n s o f p r e g n a n t females the reproductive organs of female moose were collected during the hunting seasons of 1980, 1984, 1985, and 1989. in 1980 and 1985, the collection area was the whole of finland (fig. 2); in 1984, the samples were collected from a 5,617-km2 area in the western part of the oulu management alces vol. 39, 2003 nygrén – multiple fecundity in moose 93 district and in 1989, from the lapland game management district (fig. 2). in 1980, 1985, and 1989, the moose kills in coastal and inland finland were between 15 october and 15 december, and in oulu and lapland regions between 1 october and 15 december. in 1984, the samples were collected during an exceptional winter hunt between 14 january and 31 january. specimens of genitalia and jaws were also collected. the hunting teams were advised to take random samples from adult (> 1 year old) females and to send them as quickly as possible to the fgfri ilomantsi game research station by mail. no preservative or freezing was used. in most cases, the material arrived in good condition within 1-4 days after the kill. the samples included ovaries, uteruses, vaginas, and, to determine age, jawbones or front teeth. information about the date and place of kill, sex, carcass weight (weighed, estimated, or both), as well as information about lactation and possible calves following or killed with the mother was included. the organ samples were studied and measured and the numbers and measurements of the embryos/ fetuses were determined. all triplet and quadruplet cases were photographed. the ages of the females were determined by matson’s lab in montana, usa, using cementum annuli. the viability of single, twin, and triplet embryos were determined by comparing the embryo numbers with the numbers of single, twin, and triplet calves in the moose observation data. moose observations since 1976, moose observations have formed the most important source of information for moose management in finland. each hunting day, the hunting teams classify observed moose as bulls, cows without calves, cows with 1 calf, cows with 2 calves, or as unidentified animals. the teams are advised to record the observed animals on a card once a day, even if they were seen many times or by many hunters during the day. usually, hunters give additional information if they have made an unusual observation of triplet calves or killed some or all of them. when the hunting season is over, the teams mark the estimated number of moose still living in the hunting area on the observation card. the completed observation card is sent to the ilomantsi fgfri by mail. in 1986-99, the average annual number of cards totaled about 4,300 and the average number of moose observations exceeded 201,000. the average coverage of the returned cards was 83% (the coverage % is the proportion of licenses allowed to the hunting teams). the frequency of cows with 1, 2, or 3 calves was determined annually and regionally from the observation data. in addition, the average weights of killed male and female triplet calves was determined and compared with the longterm average weights of killed moose calves 1980 and 1985 1980, 1984 and 1985 1980, 1985 and 1989 sampling years 70 66 o 60 o o fig.2. sampling areas and regions of finland used in the study. multiple fecundity in moose – nygrén alces vol. 39, 2003 94 table 2. proportions of pregnant female moose (alces alces) with 1-4 embryos in utero, collected from different areas of finland between 1980 and 1989. area/year number number of embryos in utero of uteri 1 2 3 4 n % n % n % n % coastal finland 922 561 60.85 360 39.05 1 0.11 inland finland 931 510 54.78 417 44.79 4 0.43 oulu 186 119 63.98 67 36.02 lapland 308 227 73.70 80 25.97 1 0.33 1980 749 461 61.55 286 38.18 2 0.27 1984 67 49 73.13 18 26.87 1985 1,248 698 55.93 547 43.83 3 0.24 1989 283 209 73.85 73 25.80 1 0.35 finland 1980-89 2,347 1,417 60.38 924 39.37 5 0.21 1 0.04 (nygrén and pesonen 1989). additional information about unusually large litters was also given by mail or by phone to the ilomantsi fgfri. the reliability of this information was carefully checked before the results were included in the data set. among them were reports about the triplet mother “elli”, 3 sets of quadruplets, and a set of stillborn sextuplets. results multiple pregnancies in finland in the data from 2,347 uteri of pregnant moose, 5 included triplets and 1 had quadruplet embryos (table 2). four out of the five triplet sets were found in inland finland and one in coastal finland (table 2, fig. 3). the only quadruplet set was found in lapland. two triplet sets were recorded in 1980 and three triplet sets were recorded in 1985 (table 2). no triplet or quadruplet sets were found in 1984 and 1989. case studies of triplets and quadruplets: 1. an 8.5-year-old cow (eh 2267/80) was killed on 2 november 1980 (fig. 3). the lengths of the embryos were 24 mm, 23 mm, and 24 mm, respectively. 2. an 8.5-year-old cow (ks 803/80) was killed on 16 november 1980 (fig. 3). the embryo lengths were 82, 78, and 75 mm, respectively. three primary yellow bodies existed in the right ovary. 3. a 9.5-year-old cow (es 4207/85) was killed on 15 october 1985 (fig. 3). the cow had milk in its udder. the estimated carcass weight was 230 kg. the lengths of embryos were 4 mm; 3 primary yellow bodies were found in the ovaries. 4. a 9.5-year-old cow (es 2945/85) was killed on 20 october 1985 (fig. 3). it did not have milk. the measured carcass weight was 175 kg. the embryo lengths were 16 mm, 6 mm, and 17 mm, respectively. one of the embryos had died before the cow was killed and its amnion was twisted around the axis. all 3 primary yellow bodies were in the right alces vol. 39, 2003 nygrén – multiple fecundity in moose 95 ovary. 5. a 12.5-year-old cow (ps 4961/85) was killed on 3 november 1985 (fig. 3). it had milk and the estimated carcass weight was 215 kg. the embryo lengths were 13 mm, 11 mm, and 13 mm, respectively. two embryos (13 and 11 mm) were in the left horn and one in the right horn (13 mm). the membranes in the left horn were tightly together. the membranes of the smaller embryo were twisted around the main axis, and the amount of fluid was small. the embryo itself seemed to be in good condition. only 2 primary yellow bodies existed, both in the left ovary. quite obviously, the embryos in the left horn were single-egg twins. 6. an un-aged cow (la 958/89) was killed on 3 november 1989 (fig. 3). the estimated dressed weight was 175 kg. the cow had milk and a male calf was following it (dressed weight 78 kg). the calf was killed just before its mother. the embryo lengths were 36 mm, 33 mm, 35 mm, and 34 mm, respectively. there were 2 primary yellow bodies in both ovaries. triplet observations in the field the 585,149 calf-cow observations collected in 1986-99 included 191 triplet cows (0.033 %). the annual number of triplets varied between 1 (0.003%) and 28 (0.047%) (table 3). the triplet degree correlated significantly with the twinning degree annually (r = 0.605, df = 12, p = 0.022; fig. 4) and an apparent 4-year cycle of twinning and triplet rate was found in 1988-99 when annual twin and triplet frequencies were compared (fig. 5). in the game management districts, the triplet frequencies varied between 0.014% (lapland) and 0.071% (varsinais-suomi) (table 4). the frequencies were highest in coastal finland and decreased gradually to the north. the degree of twins and triplets was indicative of spatial correlation (r = 0.462, df = 13, p = 0.083; fig. 6). an interesting case was the triplet mother “elli” (figs. 3 and 7). this moose cow, with a smaller than average head, was easy to identify. she was a yearling without triplet calves triplet embryos quadruplet embryos sextuplet calves quadruplet calves “elli”the moose cow with five triplets fig.3. locations of multiple fecundity cases in finland, 1980-2001. 34 35 36 37 38 39 40 41 42 0,00 0,01 0,02 0,03 0,04 0,05 % t r ip l e t s % twins fig.4. correlation of annual twin and triplet frequencies in finland, 1986-99 (r = 0.605, df = 12, p = 0.022). multiple fecundity in moose – nygrén alces vol. 39, 2003 96 table 3. annual proportions of single, twin, and triplet moose (alces alces) cow-calf observations in finland, 1986-99. cows with cows with cows with total cow-calf 1 calf 2 calves 3 calves observations year n % n % n % n 1986 19,546 62.248 11,847 37.729 7 0.022 31,400 1987 21,412 61.761 13,256 38.236 1 0.003 34,669 1988 25,328 63.151 14,771 36.829 8 0.020 40,107 1989 28,334 63.580 16,218 36.393 12 0.027 44,564 1990 26,953 61.927 16,556 38.039 15 0.034 43,524 1991 25,226 59.993 16,806 39.969 16 0.038 42,048 1992 26,757 62.449 16,080 37.530 9 0.021 42,846 1993 28,939 63.971 16,282 35.992 17 0.038 45,238 1994 27,722 59.687 18,703 40.268 21 0.045 46,446 1995 25,678 63.826 14,541 36.144 12 0.030 40,231 1996 22,756 65.619 11,916 34.361 7 0.020 34,679 1997 21,559 60.371 14,135 39.582 17 0.048 35,711 1998 25,474 58.515 18,039 41.437 21 0.048 43,534 1999 35,890 59.665 24,235 40.289 28 0.047 60,153 total 361,574 61.792 223,385 38.176 191 0.033 585,150 calves in 1989. during her life span “elli” gave birth to 5 sets of triplets and 30 calves (table 5). because of a protection decision by hunters, only 6 of its offspring were killed as calves. “elli's” family lived throughout the years in a territory of 1,200 ha. often the family foraged in local gardens. if disturbed, “elli” protected her calves with exceptional violence. almost all triplets were strong and similar-sized. the carcass weight of the single calf killed in 2000 was 114 kg significantly over the 83 kg average carcasss weight of finnish moose calves in the fall. in 1999, at the age of 11.5 years, “elli” gave birth to her last triplet set. one of the calves, especially well-protected by the mother, was much smaller than the others were. in 2002 “elli” gave birth to twins, was aging rapidly and limped with her hind leg. the last calf was born in 2003, probably prematurely, because the hunters never saw any calf that summer. “elli” herself was finished at the age of 15 years on 14th july 2003 after the joint decision of hunters and scientists. she was very weak, carcass weight was 111 kg. hardly any fat was found in her organs, but her teeth were in fig.5. the 4-year cycle in twin and triplet frequencies in finland, 1986-99. alces vol. 39, 2003 nygrén – multiple fecundity in moose 97 good condition. under the skin was found hundreds of lead pellets. one of the pellets had caused the chronic inflammation in the hip joint of the right leg. also parasitic pneumonia was diagnosed and very high concentrations of cadmium were found in her liver and kidneys, but lead concentrations were low. quadruplet observations four quadruplet observations have been documented in finland: 1 in 1957 (anon 1957) and 3 in 2000 (fig. 3). in 1957, only 1 calf lived more than 1 week. in 2000, the table 4. proportions of single, twin, and triplet moose (alces alces) cow-calf observations in different areas of finland, 1986-99. game management district/area cows with cows with cows with total 1 calf 2 calves 3 calves cow-calf observations n % n % n % n etelä-häme 12,837 62.276 7,768 37.685 8 0.039 20,613 kymi 19,401 62.083 11,830 37.856 19 0.061 31,250 ruotsink. pohjanmaa 17,249 58.680 12,139 41.296 7 0.024 29,395 satakunta 20,056 60.966 12,831 39.004 10 0.030 32,897 uusimaa 18,714 60.440 12,237 39.521 12 0.039 30,963 varsinais-suomi 17,729 57.063 13,318 42.866 22 0.071 31,069 coastal finland 105,986 60.155 70.123 39.800 78 0.044 176,187 etelä-savo 26,521 59.714 17,877 40.252 15 0.034 44,413 keski-suomi 31,225 60.838 20,074 39.112 26 0.051 51,325 pohjanmaa 26,507 57.311 19,725 42.648 19 0.041 46,251 pohjois-häme 13,937 63.046 8,162 36.922 7 0.032 22,106 pohjois-karjala 21,992 63.126 12,840 36.856 6 0.017 34,838 pohjois-savo 25,566 56.351 19,793 43.627 10 0.022 45,369 inland finland 145,748 59.659 98,471 40.307 83 0.034 244,302 kainuu 26,129 65.349 13,846 34.629 9 0.023 39,984 oulu 47,498 64.861 25,719 35.120 14 0.019 73,231 oulu 73,627 65.033 39,565 34.947 23 0.020 113,21 5 lapland 36,213 70.390 15,226 29.596 7 0.014 51,446 finland 361,574 61.792 223,385 38.176 191 0.033 585,150 28 30 32 34 36 38 40 42 44 0,01 0,02 0,03 0,04 0,05 0,06 0,07 0,08 % t r ip l e t s %twins fig. 6. correlation of twin and triplet frequencies in different areas of finland, 1986-99 (r = 0.462, df = 13, p = 0.083). multiple fecundity in moose – nygrén alces vol. 39, 2003 98 fig.7. moose female “elli” with her triplet calves (photo: vesa mustonen 1995-96). table 5. annual calving history of “elli”, a female moose (alces alces) from coastal finland. year “elli’s” age number of number of (years) calves born calves killed 1988 0 0 1989 1 0 1990 2 2 2 1991 3 2 2 1992 4 2 1 1993 5 3 0 1994 6 2 0 1995 7 3 0 1996 8 3 0 1997 9 3 0 1998 10 2 0 1999 11 3 0 2000 12 1 1 2001 13 1 0 2002 14 2 0 2003 151 12 0 total 30 6 1died 14th july. 2probably prematurely born, determined by inspection of “elli's” genitalia. viability of 2 litters could be followed through the summer and in both cases only 3 of the 4 calves survived until the hunting season. in the third case, the quadruplet calves were observed for the first time in september and no further observations were made during the hunting season. in three of the documented quadruplet cases, the size-difference of the calves was clear. in 1957, the total weights of the 3 dead quadruplets were 4.7, 7.0, and 9.5 kg. in 2000, one of the quadruplets was much smaller and weaker than the others were in 2 of the 3 observed cases of that year. only the quadruplets observed in september were all of similar size but they were much smaller than average moose calves during that time of the year. premature birth of sextuplets on 18 may 2001, a calving place of 6 premature moose calves was checked (fig. 3). one stillborn calf lay at one end of the calving ground and 5 at the other end. according to the condition of the carcasses, the birth had taken place about 1 week earlier. the weights of the calves were 4.0 alces vol. 39, 2003 nygrén – multiple fecundity in moose 99 kg (female), 3.8 kg (male), 3.5 kg (male), 3.3 kg (male), 2.7 kg (male), and 2.5 kg (female). the two smallest calves pk 1013/01 and pk 1014/01 (fig. 8) had died some weeks earlier in the uterus. vague signs of respiration were evident only in the male calf of 3.3 kg (pk 1012/01). survival of multiple fetuses and calves the proportion of early triplet embryo sets in the data was 0.21%, but the proportion of triplet calves observed in fall was only 0.03% (fig. 9). according to this, the viability of triplet embryos is quite low. only about 15% of triplet embryo sets seem to be complete up to the age of 6 months. the viability of quadruplets is probably even lower. compared with the survival rate of triplet and quadruplet sets, the survival rate of single and twin embryos to the age of 6 months seems to be very high (fig. 9). from the early embryonic period to the age of 6 months, the proportion of singles increased in the total data by 2.4%. the proportion of twin pairs decreased by 3.0%, while the proportions of triplet sets decreased by 85.5% and quadruplet sets by 100%. consecutively, from embryos in 1985 until calves in 1986, the figures were: singles +11.3%, twin pairs –13.9%, and triplet sets –90.8%. multiple calves obviously have an insignificant effect on calf productivity per fertile female (table 6). carcass weights of triplets the average carcass weight of triplets killed during the hunting period was 72.0 kg. the annual averages varied between 58.5 and 78.9 kg. for male triplets, the average weight was 74.9 kg (n = 62), and for female triplets 69.3 kg (n = 62). the carcass weights of female triplets varied between 37 and 100 kg and for male triplets between 41 and 106 kg. discussion kozlo (1983) expressed strong doubt about the viability of triplet embryos. no definite evidence was available at the time that the triplet or quadruplet sets were observed to ensure they were the offspring of one moose female only. later kozhuhov (1989) as well as vitakova and minajev (2000) reported that in farming conditions, 9 viable triplet sets and 1 quadruplet set were fig. 8. the stillborn sextuplets found on 18 may 2001 in nurmes, finland (photo: tuire nygrén). multiple fecundity in moose – nygrén alces vol. 39, 2003 100 ovaries of the cow. koivisto and rajakoski (1966), markgren (1969), as well as kirk and tõnisson (1999) have also reported cases of identical twins as a part of triplet embryo sets. obviously monozygotic twins are a normal, but rare event, in the european moose. in this study, 0.21% of the pregnant females had triplet embryos. about 20 years earlier, koivisto and rajakoski (1966) found 0.75% triplet embryos in the same area (n = 402). heruvimov (1969) reported 1.42% triplet embryos in the tambov region of russia (n = 141), kozlo (1983) 1.58% in byelorussia (n = 190), and kirk (2001) 1.75% in estonia (n = 114). the frequencies of triplets in the former soviet union are often much higher than those observed in finland. this is consistent with the finding that the frequency of triplets decreases towards the north (south and central russia 0.28-1.69%, north russia 0.04-0.22%) as reviewed by danilov (1987). in the surveys based on the information provided by russian hunters, the frequencies of triplet embryo sets varied from 0 to 1.72%. in reports with at least 1 triplet set observation, the average frequency of triplets was 0.14% (table 1). compared with the specialists´ results of embryo numbers, these lower frequencies were predictable. as, for example, timofejeva (1974) has pointed out, the skills of the hunters seldom are good enough to reliably count the embryo numbers in the uteri during early fall. fig.9. proportions of single, twin, triplet, and quadruplet litters of embryos and calves in finland. table 6. number of embryos per pregnant moose (alces alces) cow and calves per calf-cow in finland. area n embryos/pregnant cow n calves/calf-cow coastal finland 922 1.39 176,190 1.40 inland finland 931 1.46 244,301 1.40 oulu 186 1.36 113,215 1.35 lapland 308 1.27 51,446 1.30 born. most of the farmed triplets had, however, low birth weight and their death rate was high (kozhuhov 1989). in this study, i have shown that multiple births also occur in nature and some triplet sets even survive until the next autumn. “elli”, the 15 year-old cow moose, gave birth to 5 sets of triplets, 6 pairs of twins, and 3 single calves, and brought most of its 30 calves to adulthood. on the kostroma farm in russia, one of the most productive moose females, “alysha”, had given birth to triplets once at the age of 13 years (kozhuhov 1989). altogether, “alysha” had been pregnant 14 times and given birth to 26 calves. the present material includes 1 triplet embryo set with only 2 yellow bodies in the alces vol. 39, 2003 nygrén – multiple fecundity in moose 101 in most studied populations, the proportion of multiple fecundity cases has been zero but sometimes, in favorable conditions and/or in reports based on small sample sizes, the percentage of triplet embryos has been higher than 2%. in farming conditions, vitakova and minajev (2000) reported 2.20% triplets and 0.24% quadruplets (n = 409). most probably, if all studied natural populations were included, the proportion of triplet embryo sets of european moose would vary between 0 and 1%. the overall percentage of triplet calf observations in fall was 0.03% in my material (n = 585,149). twenty years earlier koivisto (1963) reported not a single triplet cow in finland (n = 39,818). makarova (1969 cited by ling 1974) reported 1 triplet set (0.40%) in the moscow region (n = 248). ling (1974) reported 60 cases of triplets (0.89%) in estonia (n = 6,721). as expected, the triplet calf frequencies are lower than the triplet embryo frequencies. in the present material, triplet calf frequencies were highest in southwest finland and lowest in lapland ( table 3). this is also consistent with the earlier result that the number of multiple fecundity cases is less frequent in the north than in the south (danilov 1987). the opportunities to compare the temporal change of multiple fecundity frequencies are very few. koivisto and rajakoski (1966) found 3 triplets (0.75%) in the uteri of 402 pregnant moose females. the twinning percentage was 49.8%. in this study, the proportion of triplet embryos was 0.21% and with twin embryos, it was 39.4%. however, the proportion of twin calves in fall was lower in 1963-65 (24.7%) (koivisto 1963) than in 1980-99 (38.2%). during 1963-65, all triplet sets and about 50% of twin pairs had been lost before the fall. in 1980-89, all of the quadruplet sets, 85% of the triplet sets, but only 3% of twin pairs, were lost from the population before the fall. the difference is considerable and hardly can be explained by the difference in methods. the most probable explanation is the differences of the age structure of female populations. in 1963-65, the average age of harvested females was 3.3 years (n = 1,075) and only 6.3% of the killed females were > 7 years old (koivisto, fgfri, unpublished). in the 1997-99 material (n = 2,584), the average age was 4.3 years and 19.5% of females were > 7 years old (nygrén et al. 1999; t. nygrén, fgfri, unpublished). according to glushkov (1987, 1991), the calves of older, experienced females survive better than the calves of young females. older females usually give birth earlier than younger ones and the earlier calves have a higher probability of surviving until or during the following winter (ericsson and wallin 1999, keech et al. 2000). ericsson and wallin (1999) and testa et al. (2000) found no significant difference in the survival rate of twin and single moose calves. my results, as well the results of glushkov (1987, 1991), are the opposite, but according to glushkov’s age-structured material (1987, 1991), the mortality rate of triplets, twins, and singletons in northeast russia was significantly higher than in finland. all triplet sets and 82% of twin pairs were eliminated from the population before the first fall and the proportion of singletons increased by 8%. in the younger age classes (< 6 years), the elimination frequencies were higher than in the older age classes. compared with glushkov’s results (1987, 1991) and, for example, osborne (1991) and ballard et al. (1991), the viability prognosis for twin pairs (3% decrease in frequency) and singletons (2% increase in frequency) in finland is high. a multitude of factors probably lie behind these differences; at least the larger populations of large predators in russia and north america as well the different age structure of multiple fecundity in moose – nygrén alces vol. 39, 2003 102 populations explain some of the difference in calf survival. the average carcass weight of the harvested triplets was 11 kg lower than the average overall carcass weight (83 kg) of moose calves in finland (nygrén and pesonen 1989). the difference was 12.7 kg for female and 9.1 kg for male calves. at the time of birth, the normal weight of finnish moose calves is about 8-13 kg (t. nygrén, fgfri, unpublished). calves with a weight less than 7-8 kg are seldom viable (heptner et al. 1966). this also seems to be the case in north america, where the maximum fetal mass of a cow moose is, according to geist (1974), nearly 23 kg. the upper limit of the total weight of the fetuses of european moose is unknown. however, if we assume that the total weight of normal twins is about 16-26 kg, we can calculate that a strong adult female could give birth to 3 viable 7 kg calves but for most females, 4 calves of this minimum size would be too much. at least this appeared to be the case with the observed quadruplet sets. none of these survived until the first fall. in 1957, only one survived; the dead calves weighed 4.7 kg, 7.0 kg, and 9.5 kg. according to the embryo numbers per pregnant female and the calf numbers per calf-rearing cow, the potential for reproduction is highest in inland finland and lowest in lapland. the differences in female age structure (nygrén 1999) can explain most of the spatial productivity differences existing; in the north, there are more young females without a calf than in the south, but in both areas, fertile females seem to be very capable to rear their calves to the first fall. in the north, no signs of higher calf losses were found. knowing the harsh climate of the north, the result remains unexpected. compared with the mortality of russian and north american moose, the total natural mortality of moose is low in finland. excluding hunting mortality, the average annual mortality of moose (both adults and calves) has been estimated at approximately 3% (t. nygrén, fgfri, unpublished). usually, only during severe winters with deep snow (> 1 m) and a dense moose population (4-8 moose/1,000 ha of land area) a few dead yearlings are found in southern finland in april-may. more regularly, dead calves and yearlings are found in northern finland, but even there the numbers are comparatively low. in sweden, with substantially higher moose densities than in finland but with almost similar natural and hunting conditions, cederlund and sand (1991) estimated that the natural mortality of calves was no higher than 1%. the low death rate of moose calves in finland also gained support from a helicopter survey conducted by heikkinen (1998). he found that calves comprised 50.7% of the total number of moose in february-march 1998 in a 1,600 km2 sample area where the observed proportion of calves had been 48.4% two to three months earlier at the end of the hunting season in 1997 (t. nygrén, fgfri, unpublished). obviously, the natural calf mortality of the present moose population of finland is very low: the population of large predators is small, the amount and quality of forage is high, and, as a result of selective hunting, the population structure is very productive and the proportion of experienced females high. the triplet frequencies in finland were highest in those areas with the highest calf production; the frequencies decreased towards east and north where the twinning degree has permanently been lower (nygrén et al. 2000). the spatial correlation between the twin and triplet frequencies was close to significant and the temporal correlation was significant. earlier, ling (1974) found a significant correlation between the frequencies of twins and triplets in estonia. later kirk (2001) also reported high frealces vol. 39, 2003 nygrén – multiple fecundity in moose 103 quencies of multiple pregnancies in estonia. ling’s survey methods (i.e., hunters’ observations) have been criticized (e.g., kozlo 1983) because the number of triplet and quadruplet observations was much higher than in all previous reports (table 1). the main point raised by the critics was the possibility that orphan calves join with strange calf-rearing cows and thus cause false observations of triplet or even quadruplet sets. however, moose research in finland from 1971 to 1999 has not recorded a single case of moose females accepting orphan calves. on the contrary, there are numerous examples of aggression rather than normal agonistic behavior between calf-cows and other moose (k. nygrén, fgfri, personal communication). therefore, i find it quite possible that ling’s result (1974) was sound. the existing correlation between twin and triplet frequencies is consistent with the assumption that multiple fecundity is an extreme case of twinning. when the hereditary potential for productivity is high, the age structure optimal, and the living conditions favorable for reproduction and survival, the frequencies of twinning and multiple fecundity cases tend to increase. the most important critical factors seem to be: (1) low population densities compared with the carrying capacity of the feeding grounds; (2) weak or nonexistent populations of large predators; (3) mild winter conditions and cool, humid summers; and (4) an age structure with a high frequency of females at their best reproductive age. all these factors existed in finland during the survey. according to a theory by geist (1974), cows conceiving twins or even triplets are favored in the rapidly expanding but slowly contracting moose habitats where forage is seasonally superabundant and the long daylight hours of the northern summer permit carbohydrate accumulation and a greater rate of milk production. when resources are marginal, single calves are favored. geist theorizes that there must be mandatory selection for high reproductive rates in expanding populations of moose, favoring the evolution of twinning. he also thinks that under conditions of favorable forage availability and quality it can be expected that natural selection will act against cow moose bearing single calves, which can grow too large and chance dystocia. i have no record of such delivery difficulties as recorded for caribou (bergerud, personal communication, according to geist 1974) and find the selection against cow moose bearing single calves not very probable – not at least in finnish circumstances. the living conditions are favorable. the heavily hunted populations are in a constant state of expansion (without any real increase in population numbers). the twin-rearing cows are better protected than the cows without calves and single-calf cows by legislation and selective hunting recommendations. as a result, the success of twin cows and twin and multiple calves to survive and reproduce remains high, and the frequency of genes for higher reproductive potential can slowly increase. the number of multiple fecundity reports is much larger in the european than in the north american literature. there are several possible reasons behind the difference. the first is the modern research methods, which seldom allow a sufficiently numerous sample size (schwartz 1998). this could explain the small amount of multiple fecundity cases reported in north america. the second possible reason is a lower probability to make calf-cow observations in north america, where hunting takes place closer to roads with smaller hunting teams. in finland, big moose hunting teams (koskela and nygrén 2002) use vast areas with a very dense road network. the third possibility is that the reported difference between multiple fecundity cases multiple fecundity in moose – nygrén alces vol. 39, 2003 104 in europe and north america is real, as has been speculated by geist (1998). perhaps the small number of multiple embryo and calf observations in north america is a result of a genetic difference in reproductive potential compared to european moose populations. no direct evidence is available on the matter, but it would be unjustified to exclude the possibility of hereditary differences in multiple fecundity between the area with 68 chromosomes (gustavson and sundt 1968) in europe and the area with 70 chromosomes (wurster and benirscke 1967) to the east of yenisei river in russia and in north america (boeskorov 1997). acknowledgements this study was made possible by the work of tens of thousands of finnish hunters who annually aided with the samples, observations, and reports of moose during 1980-99. in particular, the keen eyes and interest of håkan gustafsson, yrjö hiljanen, per lundström, vesa mustonen, jukka peltola, pentti piiroinen, eero salmi, and arvo sorsa were indispensable to this study. i also wish to thank heikki hyvärinen, jorma tahvanainen, and kaarlo nygrén for their valuable comments on the manuscript, maija wallén, riitta tykkyläinen, and mauri pesonen for their assistance with the data and leo bljudnik and robert kinghorn for the translations. references albright, g. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of newfoundland. canadian field-naturalist 101:373-387. anon. 1957. hirvellä neloset. metsästys ja kalastus 7-8:304-305. (in finnish). bailey, t. n., and e. e. bangs. 1980. moose calving areas and use on the kenai national moose range, alaska. alces 16:289-313. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. blood, d. a. 1974. variations in reproduction and productivity of an enclosed herd of moose (alces alces). transactions of the international congress of game biologists 11:59-66. boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1:1-10. boeskorov, g. g. 1997. chromosomal differences in moose (alces alces l., artiodactyla, mammalia). genetika 33:974-978. (in russian with english summary). cederlund, g. n., and h. k. g. sand. 1991. population dynamics and yield of a moose population without predators. alces 27:31-40. danilkin, a. a. 1999. olenji (cervidae). mlekopitajushshije rossii i sopredel´nyh regionov. geos, moskva, russia. (in russian). _____, and a. a. ulitin. 1998. a review of moose fertility in russia. alces 34:459-466. danilov, p. i. 1987. population dynamics of moose in ussr (literature survey, 1970-1983). swedish wildlife research supplement 1:503-523. ericsson, g., and k. wallin. 1999. evaluation of moose calf mortality in a hunted swedish population. pages 1-17 in g. ericsson, demographic and life history consequences of harvest in a swedish moose population. doctoral thesis. acta universitatis agriculturae sueciae. silvestria 97. filonov, k. p. 1983. losj. lesnaja promyshlennost, moskva, russia. (in russian). franzmann, a. w. 1981. alces alces. mammalian species 154:1-7. _____, and c. c. schwartz. 1985. moose alces vol. 39, 2003 nygrén – multiple fecundity in moose 105 twinning rates: a possible population condition assessment. journal of wildlife management 49:394-396. gasaway, w. c., r. d. boertje, d. v. grandgard, k. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs120. geist, v. 1974. on the evolution of reproductive potential in moose. naturaliste canadien 101:527-537. 1998. moose. pages 223-253 in v. geist, editor. deer of the world. their evolution, behavior and ecology. stackpole books, mechanicsburg, usa. glushkov, v. m. 1987. vosproizvodstvo i produktivnost losja i ih prognozirovanije. ekologia 6:31-39. (in russian). . 1991. productivity and dynamics of european moose. pages565-567 in b. bobek, k. perzanowski, and w. regelin, editors. global trends in wildlife management. transactions of the eighteenth iugb congress, krakow 1987. swiat press, krakow-warzawa, poland. gustavsson, i., and c. o. sundt. 1968. karyotypes in five species of deer (alces alces, capreolus capreolus, cervus elaphus, cervus nippon nippon, and dama dama). hereditas 60:233-248. heikkinen, s. 1998. pienhelikopterin käyttö hirvikannan runsauden ja rakenteen selvittämisessä itä-suomessa. oulun yliopisto. report. (in finnish). available from author of this paper. heptner, v. g., a. a. nasimoviè, and a. g. bannikov. 1966. die säugetiere der sovietunion. band i: paarhufer und unpaarhufer. veb gustav fischer verlag, jena, germany. (in german). heruvimov, v. d. 1969. sravnitel´noje isseldovanije na primere tambovskoi populjacii. centralno-chernozemnoje knizhnoje izdatel´stvo, voronezh, russia. (in russian). hosley, n. w., and f. s. glaser. 1952. triplet alaskan moose calves. journal of mammalogy 33:247. jazan, ju. p. 1972. ohotnichiji zveri pechorskoi taigi. biologia populjatchii, mehanizmy reguljatciji chislennosti. kirovskoje otdelenije volgo-vjatskogo knizhnogo izdatel´stva, kirov, russia. (in russian). keech, m. a., r. t. bowyer, j. m. verhoef, r.d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64:450-462. kirk, a. 2001. põdralehmad saavad varsti vasikatega maha. jahimees 4:156-158. (in estonian). , and j. tõnisson. 1994. on potential productivity of moose in estonia. (in estonian with english summary). mimeo. available from author of this paper. , and . 1999. reproduction of moose (alces alces l.) in estonia, 1993/94-1995/96. proceedings of the latvian academy of sciences b 53:9395. , and . 2000. potential productivity of moose (alces alces) population in estonia 1993/94-1998/99. folia theriologica estonia 5:57-62. knorre, je. p. 1959. ekologija losja. trudy pechoro-ilychskogo gosudarstvennogo zapovednika vypusk vii, syktyvkar, russia. (in russian). k o i v i s t o, i. 1963. hirvikantamme r a k e n t e e s t a , l i s ä ä n t y m i s e s t ä j a verotuksesta. (composition, productivity and kill of the finnish moose, alces alces, population). suomen riista 16:722. (in finnish with english summary). , a n d e . r a j a k o s k i . 1 9 6 6 . förekomsten av gula kroppar samt multiple fecundity in moose – nygrén alces vol. 39, 2003 106 äggcellernas fertilisation och migration hos älg i finland. proceedings of xth nordic veterinary congress, stockholm, sweden. (in swedish with english summary). ko s k e l a, t., and t. ny g r é n. 2002. hirvenmetsästysseurueet suomessa vuonna 1999. suomen riista 48:65-79. (in finnish with english summary). kozhuhov, m. v. 1989. reproductive potential of domesticated moose (alces alces). zoologischeskij zhurnal 5:150153. (in russian with english summary). k o z l o , p . g . 1 9 8 3 . e k o l o g o morfologicheskii analiz populjacii losja. nauka i tehnika, minsk, russia. (in russian). ling, h. j. 1974. multifetation and population productivity in elks. bjulleten moskovskogo obschestva ispytatelej prirody otdel biologitseskij 79:5-14. (in russian with english summary). lönnberg, e. 1923. älgen, alces alces lin. pages 11-36 in sveriges jaktbara djur. albert bonnier´s förlag, stockholm, sweden. (in swedish). makarova, o. a. 1981. losj v murmanskoi oblasti. pages 160-166 in v kn.: ekologia nazemnyp pozvonochnyh severo-zapada sssr. karelskii filial a n s s s r , i n s t i t u t e b i o l o g i i . petrozavodsk, russia. (in russian). markgren, g. 1969. reproduction of moose in sweden. viltrevy, swedish wildlife 6:1-299. . 1974. factors affecting the reproduction of moose (alces alces) in three different swedish areas. transactions of the international congress of game biologists 11:66-70. martin, c. j. 1989. observation of a female moose, alces alces, accompanied by possible quadruplet calves at isle royale national park, michigan. canadian field-naturalist 103:418-419. mech, l. d., r. e. mcroberts, r. o. peterson, and r. e. page. 1987. rel a t i o n s h i p s o f d e e r a n d m o o s e populations to previous winter’s snow. journal of animal ecology 56:615-627. nygrén, t. 1983. the relationship between reproduction rate and age structure, sex ratio and density in the finnish moose population. proceedings from xvi congress of the international union of game biologists. vysoké tatry, štrebské pleso, èssr. . 1996. hirvi. pages 103-108 in h. lindén, m. hario, and m. wikman, editors. riistan jäljille. finnish game and fisheries research institute, edita, helsinki, finland. (in finnish with english summary). , and m. pesonen. 1989. hirvisaaliit ja hirvenlihantuotanto suomessa vuosina 1964-87. suomen riista 35:128-153. (in finnish with english summary). , m. pesonen, r. tykkylainen, and m . w a l l e n . 1 9 9 9 . h i r v i k a n n a n ikäjakautumassa näkyvät verotuksen jäljet. riistantutkimuksen tiedote 158:115. (in finnish). , , , and ,. 2000. syksyn suurjahdin kohteena erittäin tuottava, nopeasti kasvanut kirvikanta. risstantutkimuksen tiedote 168: 1-16. (in finnish). osborne, t. o., t. f. paragi, j. l. bodkin, a. j. loranger, and w. n. johnson. 1991. extent, cause, and timing of moose calf mortality in western interior alaska. alces 27:24-30. peek, j. m. 1962. studies of moose in the gravelly and snowcrest mountains, montana. journal of wildlife management 26:360-365. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. pimlott, d. h. 1959. reproduction and productivity of newfoundland moose. alces vol. 39, 2003 nygrén – multiple fecundity in moose 107 journal of wildlife management 23:381401. schwartz, c. c. 1998. reproduction, natality and growth. pages 141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. skuncke, f. 1949. älgen. studier, jakt och vård. p. a. norstedt and söners, stockholm, sweden. (in swedish). solantie, r. 2001. suomen ilmaston erityispiirteitä. tieteellisten seurain valtuuskunta. tieteessä tapahtuu 2001/ 3 [online] pp. 28-31. [accessed 21.2.2002]. available from http:// www.tsv.fi/ttapaht/013/solantie.htm. stålfelt, f. 1974. älgpopulationerna i län med samordnad älgjakt. pages 5-23 in f. stålfelt, i. norling, c. jägnert, and b. lundahl, editors. rapporter angåcnde f ö r s ö k m e d s a m o r d n a d ä l g j a k t i k r o n o b e r g s , v ä s t m a n l a n d s o c h norrbottens län. statens naturvårdsverk, solna 1974, snv pm 485. (in swedish). testa, j. w., e. f. becker, and g. r. lee. 2000. temporal patterns in the survival of twin and single moose (alces alces) calves in southcentral alaska. journal of mammalogy 81:162-168. timofejeva, je. k. 1974. los´ (ekologia, r a s p r o s t r a n e n i j e , h o z j a i s t v e n n o j e znachenije). izdatjel´stvo leningradskogo yniversiteta, leningrad,russia. (in russian). vitakova, a. n., and a. n. minajev. 2000. fertility and life duration of kostroma farm moose (alces alces) females. nauchnyje issledovanija v zoologicheskih parkah 13:182-190. (in russian with english summary). wurster, d. h., and k. benirschke. 1967. the chromosomes of twenty-three species of the cervoidea and bovoidea. mammalian chromosome newsletter 8:226-229. multiple fecundity in moose – nygrén alces vol. 39, 2003 108 139 previous meeting sites of the north american moose conference and workshop and international moose symposia 1963 – st. paul, minnesota 1964 – st. paul, minnesota 1966 – winnipeg, manitoba 1967 – edmonton, alberta 1968 – kenai, alaska 1970 – kamloops, british columbia 1971 – saskatoon, saskatchewan 8th 1972 – thunder bay, ontario 9th 1973 – québec city, québec, in conjunction with the 1st international moose symposium 10th 1974 – duluth, minnesota 11th 1975 – winnipeg, manitoba 12th 1976 – st. john’s, newfoundland 13th 1977 – jasper, alberta 14th 1978 – halifax, nova scotia 15th 1979 – soldotna – kenai, alaska 16th 1980 – prince albert, saskatchewan 17th 1981 – thunder bay, ontario 18th 1982 – whitehorse, yukon territory 19th 1983 – prince george, british columbia 20th 1984 – québec city, québec 1984 – uppsala, sweden, 2nd international moose symposium 21st 1985 – jackson hole, wyoming 22nd 1986 – fredericton, new brunswick 23rd 1987 – duluth, minnesota 24th 1988 – winnipeg, manitoba 25th 1989 – st. john’s, newfoundland 26th 1990 – regina and ft. qu’apelle, saskatchewan 1990 – syktyvkar, russia, 3rd international moose symposium 27th 1991 – anchorage and denali national park, alaska 28th 1992 – algonquin park, ontario 29th 1993 – bretton woods, new hampshire 30th 1994 – idaho falls, idaho 31st 1995 – fundy national park, new brunswick 32nd 1996 – banff national park, alberta 33rd 1997 – fairbanks, alaska, in conjunction with the 4th international moose symposium 34th 1998 – québec city, québec 35th 1999 – grand portage, minnesota 36th 2000 – whitehorse, yukon territory 37th 2001 – carrabassett valley, maine 38th 2002 – hafjell, norway, in conjunction with the 5th international moose symposium 140 39th 2003 – jackson hole, wyoming 40th 2004 – corner brook, newfoundland and labrador 41st 2005 – whitefish, montana 42nd 2006 – baddeck, nova scotia 43rd 2007 – prince george, british columbia 2008 – yakutsk, russia, 6th international moose symposium 44th 2009 – pocatello, idaho 45th 2010 – international falls, minnesota 46th 2011 – jackson hole, wyoming 2012 – bialowieza, poland, 7th international moose symposium 47th 2013 – whitefield, new hampshire 48th 2014 – girdwood, alaska 49th 2015 – middle park, colorado 50th 2016 – brandon, manitoba, in conjunction with the 8th international moose symposium 51st 2017 – ingonish, nova scotia 52nd 2018 – spokane, washington 53rd 2019 – carrabasset valley, maine future meetings 2020 – finland, 9th international moose symposium (cancelled due to covid-19) 2021 – tbd alces37(1)_227.pdf alces37(1)_129.pdf alces37(2)_315.pdf alces39_215.pdf alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 215 temporal and spatial aspects of predator-prey dynamics rolf o. peterson1, john a. vucetich1, richard e. page2, and anne chouinard3 1school of forest resources and environmental science, michigan technological university, houghton, mi 49931, usa; 2page and associates, 3915 scolton road, victoria, bc, canada v8n 4e1; 3centre d’études nordiques, université laval, sainte-foy, pq, canada g1k 7p4 abstract: ungulates are both major consumers of vegetation and are themselves consumed by carnivores, so food web dynamics, both top-down (predation) and bottom-up (food and weather effects), are prominent in theoretical and applied research involving ungulates. the long generation time of ungulates induces long lags in population responses. over broad geographic regions, ungulates commonly achieve high density only when predation is relatively low (< 2 species of predator), suggesting that predation provides a pervasive limitation of large herbivores. ungulate stability is fundamentally a trophic-dynamics issue, usually a mix of top-down and bottom-up influences. the isle royale case history, spanning 4 decades, reveals a wolf-moose system fluctuating at 2-decade intervals with significant predation, food, and weather effects on ungulates. after a century, an equilibrium between moose and forest vegetation has not yet been reached, and a historical context seems necessary to understand trophic relationships. components of predation compared at large spatial scales reveal different predator-prey patterns than the single system at isle royale, and analyses involving substitution of space for time also run counter to studies of single systems. choice of spatial and temporal scales for field studies and meta-analyses appear to have a strong bearing on the results and their interpretation. thus temporal and spatial scales enter influentially in the actual dynamics of carnivore-ungulate interaction as well problematically in our analyses of them. alces vol. 39: 215-232 (2003) key words: bear, bottom-up, herbivory, moose, population, predation, scale, space, time, topdown, trophic, wolf interactions among trophic levels have been pervasive themes in animal ecology since its inception (elton 1927). ungulates are both major consumers of vegetation and are themselves consumed by carnivores, so studies of their population dynamics should reveal much about dominant trophic linkages in terrestrial systems (schmitz et al. 2000). ungulates are also large-bodied organisms; they may forage at local scales (risenhoover 1987, spalinger and hobbs 1992), but they make decisions about movements that may cover tens or even hundreds of square kilometers. their population dynamics and ecological influence may likewise reflect ecological phenomena that occur at different scales, e.g., vegetation patch dynamics and local nutrient fluxes may be tracked by foraging ungulates (white 1983, pastor et al. 1997, etzenhouser et al. 1998, shipley et al. 1999), but their population dynamics may reflect broad-scale forest successional patterns or pervasive loss to predators with very large home ranges (schwartz and franzmann 1991, gasaway et al. 1992). their large body size also introduces potential complexity to predatorprey relationships. ungulates may be larger than sympatric predators, thus difficult to kill as adults but more easily killed as slowgrowing juveniles (peterson 1977). ungulates are relatively long-lived, with individuspace and time in predation dynamics – peterson et al. alces vol. 39, 2003 216 als persisting for many annual cycles of seasonal changes (caughley and krebs 1983), so ecological lag effects relating to maternal effects and individual vigor can be anticipated (mech et al. 1991). here we will attempt to assess how predator-prey dynamics, and our understanding of such dynamics, depends critically on several spatio-temporal scale issues, as illustrated by the wolf-moose-fir system at isle royale. top-down and bottom up the seminal essay by hairston et al. (1960) provides a useful starting point for “modern” consideration of trophic linkages, by clearly laying out concepts that were spawned earlier by elton (1927) and lindemann (1942). paine (1980) attached the name “trophic cascade” to the conceptual world of hairston et al. (1960). models by rosenzweig (1968, 1969, 1971, 1973), oksanen et al. (1981), and oksanen (1983) furthered understanding of how length of food chains and variations in primary productivity might alter outcomes (fretwell 1987). opposing viewpoints were numerous, both then (murdoch 1966, ehrlich and birch 1967) and now (paine 2000, polis et al. 2000, schmitz et al. 2000). for the most part, research on ungulate population dynamics has been a separate but parallel parade, highlighting the respective roles of density-dependence and predation (krebs 1995). ungulate researchers seemed “data-challenged”, with findings limited by a paucity of experimental studies, the slow temporal scale of population fluctuations, and logistical challenges that have limited the spatial and temporal context of research. the search for general patterns in population ecology has not been spearheaded by large-mammal ecologists and, as a result, ungulate dynamics have usually been understood through a theoretical looking-glass that was built to view much smaller animals, even invertebrates (eberhardt 1997). or, more commonly, ungulate population dynamics were viewed in a game management context built around the dichotomy that populations were commonly limited by food or predators, with weather contributing to complexity (bergerud 1980, sinclair 1985, mech et al. 1987, fryxell and s i n c l a i r 1 9 8 8 , m e s s i e r 1 9 9 1 , vanballenberghe and ballard 1994, ballard and vanballenberghe 1998). the food or predation dichotomy surfaces in the recent ecological literature as bottom-up or topdown control of food webs (holt 2000, power 2000), respectively, and may be generalized and enlarged into recent debate about ratio-dependent predation (arditi and ginzburg 1989) and trophic cascades (paine 2000, schmitz et al. 2000). ungulates and their predators in the “real” world several species of wild ursids, felids, and canids together consitute the predator fauna for northern hemisphere ungulates. by virtue of their widespread geographic distribution, group-hunting nature, and yearround activity, we argue that the gray wolf (canis lupus) is the most significant predator of ungulates in the northern hemisphere. where additional large predators coexist with wolves, along with human hunters, it is probably reasonable to assume that predation by these different species is at least partially additive, thus enlarging the ecological influence of carnivores. it is also important to emphasize that humans continue to occupy an ecological niche as a top carnivore. as with other large predators, we have enormous potential to influence ungulate dynamics through additive mortality (crête 1987), and certainly there is plenty of historical evidence that demonstrates our prowess in excluding other carnivores (hampton 1997). although we have instituted regulations to control our alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 217 harvest, we have also increased in number ourselves and, through technology, we have greatly increased our ability to exploit and overexploit. two studies, on opposite sides of the north american continent, illustrate the significance of predation in an ecosystem context. gasaway et al. (1992) compared moose (alces alces) densities in 19 study areas in the yukon and alaska. the combination of wolf predation and bear (ursus spp.) predation was sufficient to reduce moose to a level far below “ecological carrying capacity,” where density-dependent responses to food shortage would be evident. this was true even in denali national park, where moose are not hunted by humans. crête and manseau (1996) contrasted predictions of prey-based trophic dynamics models (conforming to ideas of hairston et al. 1960) with models relying on ratio-dependent predation, along a latitudinal productivity gradient (south to north) in the québec-labrador peninsula. preybased models predict that a change at one trophic level will prompt changes of alternating sign in successively lower trophic levels, while models of ratio-dependent predation predict that consumers and their resources will increase in parallel as ecosystem productivity increases (arditi and ginsburg 1989). crête and manseau (1996) found that tundra areas with inherently low productivity, where caribou (rangifer tarandus) are the only ungulate present, supported a basic 2-link system (caribou and forage) and wolf predation was unimportant. forage production increased within the forested zone where caribou and moose supported wolf populations, but both herbivores and carnivores were relatively scarce. further south, just north of the saint lawrence river, forage production was high but predation by wolves and black bears (ursus americanus) was thought to limit moose density. where wolves were extirpated on the gaspé peninsula, moose were 7 times more abundant than in the wolf-inhabited area by the saint lawrence river, even though preferred-forage production was relatively low for that latitude. results were interpreted as support for the socalled trophic-cascade model (polis et al. 2000). gasaway et al. (1992) asserted that each additional large carnivore species resulted in a stepwise reduction in moose density. if one considers the wolf, brown bear, black bear, and human as the suite of potential predators of moose, we see a general pattern of reduced moose density with each additional predator, at least when large areas of contiguous habitat are considered (fig. 1). gasaway et al. (1992) compared moose density where predators were lightly exploited to those where predafig. 1. ungulate density is related to number of predator species in large areas of contiguous habitat where they are principal prey species for black bears, brown bears, gray wolves, and humans. moose, elk, and bison are indicated by unfilled circles, white-tailed deer are indicated by “x”, and elk in yellowstone’s northern range (prior to wolf introduction in 1995) are indicated by filled circle. data points provided in appendix 1. space and time in predation dynamics – peterson et al. alces vol. 39, 2003 218 tors were reduced by humans. the latter areas had an average moose density of 0.66 moose/km2, while the former had 0.15 moose/km2. likewise, in the absence of predators, caribou exist at densities 2 orders of magnitude higher than when coexisting with multiple species of carnivore (appendix 1). caribou employ a spacingout strategy to avoid wolf predation, and so exist at low density when coexisting with wolves (bergerud 1983a, b). where moose coexist with only one predator species, they can achieve high population density (fig. 1). moose reach densities exceeding 2/km2 in sweden where they face only human hunters, now carefully regulated. in a park with no hunting on the gaspé peninsula in québec, and only black bears as predators, moose density is also > 2/km2. moose exist at a comparably h i g h d e n s i t y a t i s l e r o y a l e , w h e r e unmanipulated wolves are the only predator. add a second predator species to any of these moose-dominated systems and one should expect a decline in moose density, a prediction borne out in many areas (fig. 1). put bears, wolves, and humans together in the same area, and only rarely would moose density exceed 1/km2. with exceptionally favorable, but transient, habitat, moose reached 0.8/km2 on alaska’s kenai peninsula, where a hunted moose population was also preyed on by wolves and black bears (bangs and bailey 1980, peterson et al. 1984). but across the northern hemisphere, where moose are typically hunted by bears, wolves, and humans, their density is usually on the order of 0.4/km2 or less. anything that increases the reproductive potential of a prey population will tend to reduce the impact of large carnivores ( s e i p 1 9 9 5 ) . d e e r ( o d o c o i l e u s virginianus), for example, should be able to coexist with wolves at a higher density than prey with lower rates of increase (e.g., caribou). potentially, habitat improvement could accomplish the same (orians et al. 1997). thus, deer exist at higher density in north-central minnesota, with intensive forest management, than in the old-growth forests of extreme northern minnesota (nelson and mech 1986, fuller 1989). additionally, the presence of buffer prey may modify (increase or decrease) predation on a single prey species. the mechanism by which predation limits ungulate populations is pervasive removal of pre-reproductive juveniles (pimlott 1967); where juveniles survive poorly, populations tend to decline. at isle royale moose density increased during periods when calf overwinter survival was ~80%, but declined when survival was ~50% (r. peterson, unpublished data). predator reduction experiments improved moose calf survival about 3-fold and increased finite rate of population increase from 1.0 to 2.3 (gasaway et al. 1992). over the past 25 years there have been numerous studies to determine the extent and nature of juvenile ungulate mortality (reviewed by ballard and larsen 1987, van ballenberghe 1987, orians et al. 1997). in 11 studies located in 9 areas of alaska, yukon, and british columbia where wolves and bears existed, radio-collars have been placed on 623 moose and 462 caribou soon after birth (orians et al. 1997). survival was monitored for variable periods, from 2 months to 1 year, but clearly there was an early period of greatest risk during the first 2 months of life. for both species, average (mean for all studies) survival rate of juveniles in their first year was only 40%. ballard and larsen (1987) and van ballenberghe (1987) cited several studies which indicated that predation on young moose may account for 79% of neonate deaths; survival in the first 8 weeks can be as low as 17%. where moose were hunted in alaska, losses to predation (31% of postcalving numbers) alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 219 greatly exceeded loss to hunting (1.5%) and other losses (6%) (gasaway et al. 1992). in an unhunted population free of bears and wolves (rochester, alberta), annual survival of moose calves was as high as 67%, and 41% of calf/cow groups in winter included twins (rolley and keith 1980). where 2 or more species of large carnivore were present (8 studies with 469 moose calves total), average survival to age of 6 months was 30%. in 3 study areas with 0-1 species of carnivore, average 6-months survival of moose was 67% (3 studies with 111 calves). most mortality occurred within the first 6 weeks of life – black bears being responsible for 2-50%, grizzlies 3-52%, and wolves 2-18%. if moose density was high, approaching k carrying capacity, predatorinduced mortality was considered more likely to be compensatory (orians et al. 1997). bear predation may be densityindependent, but nevertheless a significant mortality factor in the first 6 months of life. on the other hand, wolf predation has its greatest impact in winter, when calves make up 30-40% of observed kills (peterson 1977, ballard et al. 1987, mech et al. 1995). page’s (1989) analysis of cohort survival indicated that overwinter survival of moose calves ranged from only 30% to almost 100%, depending on wolf density and relative nutritional stress (caused by deep-snow winters and high population density). predation may likewise reduce adult survivorship (van ballenberghe and ballard 1998), but variance in adult survival is much less than for juveniles. adult survival rates depend heavily on the intensity of hunting. peterson (1977) estimated that equilibrium survival of an unhunted moose population was ~87% for yearlings and adults; gasaway et al. (1983) reported that the adult moose survival rate improved from 80% to 94% after wolf densities were reduced in the tanana flats, alaska. messier (1994) reviewed 27 studies in which moose were the primary prey, and from his analysis he concluded that wolf predation was density-dependent at the low range of moose density, therefore regulatory. he predicted that moose would stabilize at 2 moose/km2 in the absence of predators via density-dependent mechanisms, at 1.3 moose/km2 in simple wolf/moose systems, and 0.2-0.4 moose/km2 in a single stable-state (low density) equilibrium with multiple predators. thomas (1995) asserted that caribou have no intrinsic (e.g., social) population limitations. in his view the evolutionary pressure of wolf predation pervades caribou ecology. the suite of predators for caribou is dominated by wolves but also includes lynx (lynx canadensis), coyote (canis latrans), brown bear, black bear, cougar (felis concolor), and wolverine (gulo gulo). he stated that predators “can keep caribou populations depressed for long periods if alternate prey are abundant.” farnell and mcdonald (1988) reported that recruitment of caribou increased 113% and adult mortality declined 60% in a population where wolf numbers were reduced 80%. thomas (1995) concluded that when predators are present, average natural mortality of caribou is approximately 50% in calves in forest-tundra areas and 50-70% in forestalpine and forest-forest moose zones; annual mortality was 7-30% in adults (bergerud 1980, farnell and mcdonald 1988, seip 1992). in contrast, annual mortality approaches zero where wolves are absent or rare. from his review of the literature, ranging from kelsall (1968) to seip (1992), thomas (1995) concluded that (1) wolf predation is the major direct cause of natural mortality of calf and adult caribou and (2) dense caribou populations occur only in the absence of wolves. from studies of deer in northern minnesota, mech et al. (1987) showed that deer fawn abundance was correlated not with space and time in predation dynamics – peterson et al. alces vol. 39, 2003 220 wolf density but rather the cumulative severity of 3 previous winters. this does not indicate that wolves were unimportant in deer dynamics (cf. mech and karns 1977), but might instead mean that the dominant variance in the system was winter severity (boyce and anderson 1999). further south in minnesota, where timber harvest created optimum deer habitat, fuller (1989) found that hunting mortality was more important than wolf predation in deer dynamics. resource abundance may indeed modify the effect of predation through effects on reproduction and individual vigor. wolves and moose on isle royale: dynamics of a simple system a perspective on wolf-moose relations in isle royale national park (544-km2 island in lake superior) is offered, to provide an update on a very dynamic case history and to illustrate temporal and spatial issues involved in interpreting predator-prey interactions. as moose and wolf populations have changed over the past 40 years (fig. 2) there have been significant changes in how the system has been interpreted by observers, and valuable perspective has also been provided by studies elsewhere. annual winter counts of wolves began at isle royale about 10 years after the island was colonized by wolves in the late 1940s. efforts were begun to estimate moose numbers, with methods steadily evolving into a “gasaway-type” stratified plot count in which about 17% of the land area is intensively searched from aircraft in winter. an independent method to track historic change in moose numbers has been retrospective reconstruction, based on recoveries of approximately one-third of the moose after death (page 1989). temporal chronology early in the study, in the early 1960s, wolf population size was stable and moose exhibited a high twinning rate and also seemed relatively stable, so mech (1966) suggested that wolf predation was keeping moose density (about 1/km2) below the level at which food supply might be limiting. by the early 1970s, however, it was evident that moose had increased in the 1960s (krefting 1974, peterson 1977), reaching a level (almost 3/km2) in the early 1970s where nutrition was poor, at least during severe winters in 1969-1972 (peterson 1975). the wolf population expanded during 1969-1980 and moose density declined in 1972-1982 as wolf predation intensified. wolves briefly reached in 1980 the highest year-round density documented for wolves in nature up to that time (almost 0.1/ km2). as wolf numbers grew it was obvious that calf survival was negatively affected (fig. 2), prompting population decline. taking stock of the situation in the mid-1970s, peterson (1977) interpreted the moose decline as a response to habitat deterioration as post-fire successional forests (regenerating after fires in 1936 and 1948) matured and moose became more dependent on older forests over 100 years old. peterson (1977) asserted that density-dependent mechanisms prompted the moose decline, although wolf predation probably accelerated it. in short, wolf predation was considered largely compensatory, and not ultimately responsible for the moose decline. this was essentially a bottom-up interpretation, with the logical prediction being that moose density would stabilize at a new, lower level dictated by habitat, where it would remain as long as new forest disturbance did not intervene. however, the moose population quickly grew again after the wolf population crashed in 1980-1982 (coincident with the arrival of canine parvovirus; peterson et al. 1998). the 1981 cohort of moose calves, coincident with the wolf decline, was proportionately among the largest ever seen at isle alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 221 royale, constituting about one-fifth of the moose population in mid-winter (fig. 2). peterson and page (1983) acknowledged that wolf predation, not deficient habitat, had been limiting moose population growth – evidently predation loss was not simply compensatory. significantly, balsam fir trees throughout the winter range of isle royale moose exhibited lagged oscillations in growth that mirrored the inverse of moose density – when wolves were high, moose declined, prompting growth of the forest (mclaren and peterson 1994). throughout the 1980s and early 1990s, with low wolf numbers, the moose population grew almost without interruption (winter ticks were implicated in high mortality in 1989; delgiudice et al. 1997). wolves were themselves mysteriously maintained at a low level by lingering effects of disease, inbreeding, stochastic demography, or some combination of these factors (peterson et al. 1998, peterson 1999). by 1995 the moose population had grown to exceed 4/ km2 and there was ample evidence of severe undernutrition in winter. twin calves were rarely seen, and moose phenotype reflected food shortage (peterson 1995). although calves were growth-retarded, most nevertheless survived their first winter to live on as adults; density-dependence was reflected in moose morphology but not in population dynamics. the winter of 1995-1996 was the most extreme in a century (delgiudice 1998, post et al. 1999), with early winter storms, persistent deep snowcover, and cold temperatures that delayed the arrival of spring by about 6 weeks. moose began dying of starvation by february 1996 and about 80% of the population perished in the next 3 months, reducing moose density once again to about 1/km2. a dieoff of this scale also happened on isle royale in 1934, following the initial irruption of moose after colonization (mech 1966). spatial heterogeneity while temporal variation in the isle royale chronology is striking, there is also spatial heterogeneity in this ecosystem. the east end of the island, by virtue of its glacial history, has thin soils and more forest disturbance caused by wind, resulting in more light reaching the forest floor (mclaren and janke 1996). the resulting higher production of moose forage at the east end contrasts with conditions at the west end, where deep soils support old and tall deciduous forests that heavily shade the forest floor (fig. 3). balsam fir (abies balsamea), a key winter forage species for moose fig. 2. wolf population size and moose estimated population size, isle royale national park, 1959-1998 (upper panel), and the proportion of calves (~6 months of age) in the moose population (lower panel). each annual estimate of % calves is an average of a field-based estimate (mean of all available counts for each cohort, ranging from summer ground counts to aerial counts in autumn or winter) and an estimate based on population reconstruction (details at www.isleroyalewolf.org). space and time in predation dynamics – peterson et al. alces vol. 39, 2003 222 (risenhoover 1987) is regenerating at high density at the east end but not at the west end. somewhat paradoxically, this browse species receives proportionately less damage by moose herbivory in the thin soils at the east end of the island (mclaren 1996). we used dendrochronology to determine the pattern of fir growth as a key component of the isle royale trophic system. fir saplings and small trees growing with minimum competition for light were sampled from the west and east ends of the island. for each individual stem, ring widths were measured in cross-sections from the base of the stem. each ring-width series was indexed following the method described in chouinard and filion (2001). heights of fir stems on the west end ranged from 90200 cm, while on the east end heights were 200-600 cm. all trees we sampled were from sites previously studied by mclaren and peterson (1994). tree-ring growth, which may index abundance of fir forage, differs substantially across different portions of isle royale in some years (fig. 3a). spatial variation in fir growth may be the most significant aspect of spatial heterogeneity in the vegetation-moose-wolf system on isle royale. is spatial heterogeneity in balsam fir manifested in the dynamics of higher trophic levels? to gain a preliminary understanding for how spatial variation in fir could be manifest in moose population dynamics, we consider how calf production is influenced by the abundance of fir forage. using our data (figs. 2 and 3a) and multiple linear regression we obtained the following model: ct = 0.15 – 3.3×10 -5mt-1 + 4.6×10 -2ft-1, (1) where percent calf production in the current year (ct) is dependent on moose abundance (m) and fir tree-ring index (f) in the previous year. for this model ft-1, represents tree ring growth, averaged across all of isle royale. the coefficients for moose (p = 0.03) and fir (p = 0.03) are statistically fig. 3. upper panel: index of fir growth (see text) for eight trees located on the east half of isle royale and for eight trees located on the west half of isle royale. middle panel: two predictions of percent calves based on equation (1) and the hypothetical assumptions that island-wide fir growth is characterized by fir growth that actually characterizes only the east half of the island (open circles) and only the west half of the island (closed circles). lower panel: the difference between the two predicted percent calf alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 223 significant, and this model explains approximately 20% of the variation in calf production. the residuals of this model do not appear to deviate substantially from normal (p = 0.83), nor do they appear to be autocorrelated (durbin-watson statistic = 0.93). for this model, interannual variation in tree ring growth accounts for approximately 10% of the variation in calf production. consider, hypothetically, what the temporal dynamics of calf production would be like if fir growth across the entire island were like it has been on just the west end of isle royale, or just the east end of isle royale. to explore this hypothetical scenario, we predicted 2 time series of percent calf production using equation (1), except that for one time series we replaced ft-1 with a series of values representing growth at the west end of isle royale, and for the other series, we replaced ft-1 with a series of values representing growth at the east end of isle royale. these predicted percent calf time series are depicted in figure 3b, and the absolute difference between these time series is depicted in figure 3c. in absolute terms the difference between the two time series appears minor. however, the differences are highly autocorrelated. for example, for 11 consecutive years (1968-78) and for 9 consecutive years (199098) the difference is positive. such a pattern could lead to substantial differences in moose population abundance. investigation beyond the scope of this manuscript is required to accurately understand the extent to which spatial heterogeneity in fir could give rise to biologically significant spatial heterogeneity in moose population dynamics. nevertheless, our hypothetical example suggests that further investigation could reveal important insights. a convincing example of the importance of spatial differences in habitat emerged as moose population levels at opposite ends of isle royale diverged as a result of the moose die-off in 1996 when many moose survived in the thick fir stands at the east end but perished at the west end (r. o. peterson and j. a. vucetich, unpublished data). by 1999, 20 of the 25 wolves present were also supported by moose at the east end, so the bottom-up pulse of productivity was manifest at all 3 trophic levels. thus, a long-term and large-scale pattern of soil development established as glaciers retreated was manifested in a rather peculiar historic fashion as a once-per-century severe winter impacted a moose population poised at a historic high population density. this example of the influence of spatial heterogeneity in fir dynamics highlights the need for an improved understanding of meso-scale spatial heterogeneity in wolf-ungulate dynamics across north america (cf. orians et al. 1997). what’s scale got to do with it? with data on isle royale wolves and moose from any single 5-year period of the last 40 years, it would be possible to support almost any interpretation of their interaction, not unlike the fabled 10 blind men describing an elephant. thus varying interpretations of predator-prey dynamics on isle royale (vanballenberghe 1987) arose in part from the slow rate of change as a system initially interpreted as being in equilibrium (mech 1966, peterson 1977) has since exhibited long-term oscillatory tendencies (peterson et al. 1984, mclaren and peterson 1994). different ecological factors have prevailed at different times. temporal fluctuations at decadal intervals may be the norm for large-bodied ungulates and their prey, but this is much longer than our usual framework for research and management. the dilemma of scale-dependent understanding appears to be a general one for space and time in predation dynamics – peterson et al. alces vol. 39, 2003 224 animal ecology. temporal and spatial variation in population density are not wellunderstood (in relation to environment), both for ungulates and other animals in general, but these lie at the core of our understanding of population regulation (lundberg et al. 2000). mechanisms underlying population variability will be elucidated only if appropriate response variables are studied at appropriate scales (schmitz et al. 2000). long response times may induce long lags; ungulate dynamics are drawn-out over decades, and response times for woody vegetation are even longer (holt 2000). the scale of study impacts interpretation of trophic interactions as well as dynamics of single-species populations (wiens 1989). the isle royale case history provides a particularly compelling case for temporal variation in predator-prey interaction, which can now be anticipated for any regularly fluctuating system (sinclair et al. 2000). comparisons of wolf and moose status over geographical scales may not always provide mechanistic insight into predatorprey dynamics. for example, a “global” correlation between average wolf and prey density exists over a wide range of prey densities (fuller 1989). this is widely interpreted as the “numerical response” of the wolf to fluctuations in prey density (messier 1994), even as an indication of ratio-dependent predation (arditi et al. 1991), but the temporal trajectory followed by local populations conforms poorly to that which is extrapolated from the large spatial (but temporally static) global pattern. the long lives and long lags of large mammals may contribute to the poor match. what happens to the validity of models (messier 1994) and management programs (gasaway et al. 1983) when we ignore (lag-induced) temporal dynamics? large-scale geographic comparisons that ignore all temporal dynamics have also been used to assess predator-prey models of temporal dynamics. in a study from québec (crête and manseau 1996), spatial patterns were used to support prey-dependent models of predation (sensu, hairston et al. 1960) and to reject ratio-dependent models of predation (sensu, arditi and ginzburg 1989). however, it seems dubious to infer processes that occur in one dimension (time) from observations made in another dimension (space) (e.g., see also abrams 1994, lundberg and fryxell 1995, abrams and ginzburg 2000). in fact, our analysis of the temporal dynamics of isle royale wolves and moose appear to support ratio-dependent predation (vucetich et al. 2002). the ecology of moose at isle royale includes both slow and fast ecological processes, operating at large and small spatial scales. moose have virtually eliminated some plant species from the forest (e.g., taxus), and intensive foraging may eliminate regeneration of many species (e.g., abies) in the tree layer. mesoscale dynamics are dominated by cyclical fluctuations in moose and wolves with a duration of about 2 decades, a product ultimately of generation time for predator and prey. yet the forest itself has not equilibrated following the arrival of moose a century ago – at the island’s west end old fir trees that established in the canopy before moose arrived are now dying of old age without replacement, and at the east end spruce-fir stands have only recently emerged in extensive 19th century burns initiated by mineral prospectors – while both trends are superficially associated with less forage for moose, future dynamics of wolves and moose in response to these very slow changes in vegetation are not readily predicted. acknowledgements long-term studies of population dynamics of wolves, moose and balsam fir on isle alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 225 royale have been supported primarily by isle royale national park, the national science foundation (deb-9903671), earthwatch, inc., and numerous private donors. references abrams, p. a. 1994. the fallacies of “ratio-dependent” predation. ecology 75:1842-1850. , and l. r. ginzburg. 2000. the nature of predation: prey dependent, ratio dependent or neither? trends in ecology and evolution 15:337-341. arditi, r., and l. r. ginzburg. 1989. coupling in predator-prey dynamics: ratio dependence. journal of theoretical biology 139:311-326. , , and h. r. akçakaya. 1991. variation in plankton densities among lakes: a case for ratio-dependent predation models. american naturalist 138:1287-1296. bailey, t. n. 1978. moose populations on the kenai national moose range. proceedings of the north american moose conference and workshop 14:1-20. ballard, w. b., and d. g. larsen. 1987. implications of predator-prey relationships to moose management. swedish wildlife research supplement 1:581602. , and v. van ballenberghe. 1998. predator/prey relationships. pages 247273 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , j. s. whitman, and c. l. gardner. 1987. ecology of an exploited wolf population in south-central alaska. wildlife monographs 98. bangs, e., and t. n. bailey. 1980. interrelationships of weather, fire, and moose on the kenai national moose range, alaska. proceedings of the north american moose conference and workshop 16:255-274. berg, w. e., and d. w. kuehn. 1982. ecology of wolves in north-central minnesota. pages 229-247 in f. h. harrington and p. c. paquet, editors. wolves of the world: perspectives of behavior, ecology, and conservation. noyes, park ridge, new jersey, usa. bergerud, a. t. 1980. a review of the population dynamics of caribou and wild reindeer in north america. proceedings of the international reindeer/caribou symposium 2:556-581. . 1983a. the natural population control of caribou. pages 14-61 in f. l. bunnell, d. s. eastman, and j. m. peek, editors. symposium on natural regulation of wildlife populations. proceedings number 14. forest, wildlife and range experiment station, university of idaho, moscow. . 1983b. caribou, wolves and man. trends in ecology and evolution 3:6872. , and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. journal of wildlife management 32:729-746. , w. wyett, and b. snider. 1983. the role of wolf predation in limiting a moose population. journal of wildlife management 47:977-988. boyce, m. s., and e. m. anderson. 1999. evaluating the role of carnivores in the greater yellowstone ecosystem. pages 265-283 in t. w. clark, a. p. curlee, s. c. minta, and p. m. kareiva, editors. c a r n i v o r e s i n e c o s y s t e m s : t h e yellowstone experience. yale university press, new haven, connecticut, usa. cairns, a. l., and e. s. telfer. 1980. habitat use by four sympatric ungulates in boreal mixedwood forest. journal of space and time in predation dynamics – peterson et al. alces vol. 39, 2003 226 wildlife management 47:977-988. carbyn, l. n. 1983. wolf predation on elk in riding mountain national park, manitoba. journal of wildlife management 47:963-976. caughley, g., and c. j. krebs. 1983. are big mammals simply little mammals writ large? oecologia 59:7-17. cederlund, g., and g. markgren. 1987. the development of the swedish moose population, 1970-1983. swedish wildlife research supplement 1:55-61. chouinard, a., and l. filion. 2001. detrimental effects of white-tailed deer browsing on balsam fir growth and recruitment in a second-growth stand on anticosti island, quebec. ecoscience 8:199-210. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1:553-563. . 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67:373-380. , and m. manseau. 1996. natural regulation of cervidae along a 1000 km latitudinal gradient: change in trophic dominance. evolutionary ecology 10:51-62. crichton, v. f. 1977. hecla island— manitoba’s answer to isle royale. proceedings of the north american moose conference and workshop 13:191-199. delgiudice, g. d. 1998. surplus killing of w h i t e t a i l e d d e e r b y w o l v e s i n northcentral minnesota. journal of mammalogy 79:227-235. , r. o. pe t e r s o n , and w. m. samuels. 1997. trends of winter nutritional restriction, ticks and numbers of moose on isle royale. journal of wildlife management 61:895-903. detling, j. k. 1998. mammalian herbivores: ecosystem-level effects in two grassland national parks. wildlife society bulletin 26:438-448. eberhardt, l. l. 1997. is wolf predation ratio-dependent? canadian journal of zoology 75:1940-1944. ehrlich, p. r., and l. c. birch. 1967. the balance of nature and population control. american naturalist 101:97-107. e l t o n , c. 1927. animal ecology. macmillan, new york, new york, usa. etzenhouser, m. j., m. k. owens, d.e. spalinger, and d. e. murden. 1998. foraging behavior of browsing ruminants in a heterogeneous landscape. landscape ecology 13:55-64. farnell, r., and j. mcdonald. 1988. the influence of wolf predation on caribou mortality in yukon’s finlayson caribou herd. proceedings of the north american caribou workshop 3:52-70. fretwell, s. d. 1987. food chain dynamics: the central theory of ecology? oikos 50:291-301. fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52:14-21. , and a. r. e. sinclair. 1988. causes and consequences of migration by large herbivores. trends in ecology and evolution 9:237-241. fuller, t. k. 1989. population dynamics of wolves in northcentral minnesota. wildlife monographs 105. _____, and l. b. keith. 1980. wolf population dynamics and prey relationships in northeastern alberta. journal of wildlife management 44:583-602. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildalces vol. 39, 2003 peterson et al. – space and time in predation dynamics 227 life monographs 120. , r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. haber, g. c. 1977. socio-ecological dynamics of wolves and prey in a subarctic ecosystem. ph.d. thesis, university of british columbia, vancouver, british columbia, canada. hairston, n. g., sr., f. e. smith, and l. b. slobodkin. 1960. community structure, population control, and competition. american naturalist 94:421-425. hampton, b. 1997. the great american wolf. henry holt and company, new york, new york, usa. holt, r. d. 2000. trophic cascades in terrestrial ecosystems: reflections on polis et al. trends in ecology and evolution 15:444-445. kelsall, j. p. 1968. the migratory barrenground caribou of canada. canadian wildlife service monograph series number 3. ottawa, ontario, canada. krebs, c. j. 1995. two paradigms of population regulation. wildlife research 22:1-10. krefting, l. 1974. the ecology of the isle royale moose with special reference to the habitat. agricultural experiment station technical bulletin 297. university of minnesota, minneapolis, minnesota, usa. larsen, d. g. 1982. moose inventory in the southwest yukon. alces 18:142167. lindemann, r. 1942. the trophic-dynamic aspect of ecology. ecology 23:399418. lundberg, p., and j. m. fryxell. 1995. expected population density versus productivity in ratio-dependent and preydependent models. american naturalist 146:154-161. , e. ranta, j. ripa, and v. kaitala. 2000. population variability in space and time. trends in ecology and evolution 15:460-464. mclaren, b. e. 1996. plant-specific response to herbivory: simulated browsing of suppressed balsam fir on isle royale. ecology 77:228-235. , and r. a. janke. 1996. seedbed and canopy cover effects on balsam fir seedling establishment in isle royale national park. canadian journal of forest research 26:782-793. , and r. o. p e t e r s o n. 1994. wolves, moose and tree rings on isle royale. science 266:1555-1558. mech, l. d. 1966. the wolves of isle royale. u.s. national park service fauna series 7:1-210. . 1986. wolf numbers and population trend in the central superior national forest, 1967-1985. u.s. department of agriculture forest service research paper no. nc-270. north central forest experiment station, st. paul, minnesota, usa. , and l. d. frenzel jr. 1971. ecological studies of the timber wolf in northeastern minnesota. u.s. department of agriculture forest service research paper nc-52. north central forest experiment station, st. paul, minnesota, usa. , and p. d. karns. 1977. role of the wolf in a deer decline in the superior national forest. u.s. department of agriculture forest service research paper nc-148. north central forest experiment station, st. paul, minnesota, usa. , r. e. mcroberts, r.o. peterson, and r. e. page. 1987. relationship of deer and moose populations to previous winters’ snow. journal of animal ecology 56:615-627. , t. j. meier, j. w. burch, and l. g. space and time in predation dynamics – peterson et al. alces vol. 39, 2003 228 adams. 1995. patterns of prey selection by wolves in denali national park, alaska. pages 231-244 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a c h a n g i n g w o r l d . c a n a d i a n circumpolar institute occasional publication no. 35, edmonton, alberta, canada. , m . e . n e l s o n , a n d r . e . mcroberts. 1991. effects of maternal and grandmaternal nutrition on deer mass and vulnerability to wolf predation. journal of mammalogy 72:146151. mercer, w. e., and f. manuel. 1974. some aspects of moose management in newfoundland. naturaliste canadien 101:657-671. messier, f. 1991. the significance of limiting and regulating factors on the demography of moose and white-tailed deer. journal of animal ecology 60:377393. . 1994. ungulate population models with predation: a case study with the north american moose. ecology 75:478-488. , and m. crête. 1985. moose-wolf dynamics and the natural regulation of moose populations. oecologia 65:503512. murdoch, w. w. 1966. community structure, population control, and competition. american naturalist 100:219-226. nelson, m. e., and l. d. mech. 1986. white-tailed deer numbers and population trend in the central superior national forest, 1967-1985. u.s. department of agriculture forest service research paper nc-271. north central forest experiment station, st. paul, minnesota, usa. nygrén, t. 1987. the history of moose in finland. swedish wildlife research supplement 1:49-54. oksanen, l. 1983. trophic exploitation and arctic phytomass patterns. american naturalist 122:360-367. , s. fretwell, j. arruda, and p. niemela. 1981. exploitation ecosystems in gradients of primary productivity. american naturalist 118:240-261. oosenbrug, s. m., and l. n. carbyn. 1985. wolf predation on bison in wood buffalo national park. unpublished report. canadian wildlife service, edmonton, alberta, canada. orians, g.h., et al. (13 authors). 1997. wolves, bears, and their prey in alaska: biological and social challenges in wildlife management. national academy press, washington, d.c., usa. page, r. e. 1989. the inverted pyramid: ecosystem dynamics of wolves and moose on isle royale. ph.d. thesis, michigan technological university, houghton, michigan, usa. paine, r. t. 1980. food webs: linkage, interaction strength, and community structure. journal of animal ecology 49:667-685. . 2 0 0 0 . p h y c o l o g y f o r t h e mammalogist: marine rocky shores and mammal-dominated communities—how different are the structuring processes? journal of mammalogy 81:637-648. pastor, j., r. moen, and y. cohen. 1997. spatial heterogeneities, carrying capacity, and feedbacks in animal-landscape interactions. journal of mammalogy 78:1040-1052. peek, j. m., d. l. ulrich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. o. 1975. wolf response to increased moose vulnerability in isle royale. proceedings of the north american moose conference 11:344368. alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 229 . 1977. wolf ecology and prey relationships on isle royale. u.s. national park service scientific monograph series 11. washington, d.c., usa. . 1995. the wolves of isle royale—a broken balance. willow creek press, minocqua, wisconsin, usa. . 1999. wolf-moose interaction on isle royale: the end of natural regulation? ecological applications 9:10-16. , and r. e. page. 1983. wolfmoose fluctuations at isle royale national park, michigan, u.s.a. acta zoologica fennica 174:251-253. , n. j. thomas, j. m. thurber, j. a. vucetich, and t. a. waite. 1998. population limitation and the wolves of isle royale. journal of mammalogy 79:828-841. , j. d. woolington, and t. n. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88. pimlott, d. h. 1967. wolf predation and ungulate populations. american zoologist 7:267-278. , j . a . s h a n n o n , a n d g . b . kolenosky. 1969. the ecology of the timber wolf in algonquin provincial park. ontario department of lands and forests research paper (wildlife) no. 87. toronto, ontario, canada. polis, g. a., a. l. w. sears, g. r. huxel, d. r. strong, and j. maron. 2000. when is a trophic cascade a trophic cascade? trends in ecology and evolution 15:473-475. post, e., r. o. peterson, n. c. stenseth, and b. e. mclaren. 1999. ecosystem consequences of wolf behavioural response to climate. nature 401:905-907. potvin, f. 1988. wolf movements and population dynamics in papineaulabelle reserve, quebec. canadian journal of zoology 66:1266-1273. power, m. e. 2000. what enables trophic cascades? commentary on polis et al. trends in ecology and evolution 15:443444. risenhoover, k. l. 1987. winter foraging strategies of moose in subarctic and boreal forest habitats. ph.d. thesis, michigan technological university, houghton, michigan, usa. rolley, r. e., and l. b. keith. 1980. moose population dynamics and winter habitat use at rochester, alberta, 19651979. canadian field-naturalist 89:4752. rosenzweig, m. 1968. net primary productivity of terrestrial communities: prediction from climatological data. american naturalist 102:67-74. . 1969. why does the prey isocline have a hump? american naturalist 103:81-87. . 1971. the paradox of enrichment: destabilization of exploitation ecosystems in ecological time. science 171:385-387. . 1973. exploitation in three trophic levels. american naturalist 107:275294. schmitz, o. j., p. a. hambaeck, and a. p. beckerman. 2000. trophic cascades in terrestrial systems: a review of the effects of carnivore removals on plants. american naturalist 155:141-153. schwartz, c. c., and a. w. franzmann. 1991. interrelationships of black bears to moose and forest succession in the northern coniferous forest. wildlife monographs 113. seip, d. r. 1991. predation and caribou populations. proceedings the north american caribou workshop 5:46-52. . 1992. factors limiting woodland caribou populations and their interrelationships with wolves and moose in southeastern british columbia. canadian journal of zoology 70:1494-1503. space and time in predation dynamics – peterson et al. alces vol. 39, 2003 230 . 1995. introduction to wolf-prey interactions. pages 179-186 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, occasional publication no. 35, edmonton, alberta, canada. shipley, l. a., a. w. illius, k. danell, n. t. hobbs, and d. e. spalinger. 1999. predicting bite size selection of mammalian herbivores: a test of a general model of diet optimization. oikos 84:5568. sinclair, a. r. e. 1985. does interspecific competition or predation shape the african ungulate community? journal of animal ecology 54:899-918. , c. j. krebs, j. m. fryxell, r. turkington, s. boutin, r. boonstra, p. seccombe-hett, p. lundberg, and l. oksanen. 2000. testing hypotheses of trophic level interactions: a boreal forest ecosystem. oikos 89:313-328. singer, f. j., and j. dalle-molle. 1985. the denali ungulate-predator system. alces 21:339-358. , d. m. swift, m. b. coughenour, and j. d. varley. 1998a. thunder on the yellowstone revisited: an assessment of management of native ungulates by natural regulation, 1968-1993. wildlife society bulletin 26:375-390. , l. c. zeigenfuss, r. g. cates, and d. t. barnett. 1998b. elk, multiple factors, and persistence of willows in national parks. wildlife society bulletin 26:419-428. spalinger, d. e., and n. t. hobbs. 1992. mechanisms of foraging in mammalian herbivores: new models of functional response. american naturalist 140:325348. stenlund, m. h. 1955. a field study of the timber wolf (canis lupus) on the superior national forest, minnesota. minnesota department of conservation technical bulletin no. 4. st. paul, minnesota, usa. thomas, d. c. 1995. a review of wolfcaribou relationships and conservation implications in canada. pages 261-273 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, occasional publication no. 35, edmonton, alberta, canada. van ballenberghe, v. 1987. effects of predation on moose numbers: a review of recent north american studies. swedish wildlife research supplement 1:431-460. , and w. ballard. 1994. limitation and regulation of moose populations: the role of predation. canadian journal of zoology 72:2071-2077. , and . 1998. population dynamics. pages 223-245 in ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , a. w. erickson, and d. byman. 1975. ecology of the timber wolf in northeastern minnesota. wildlife monographs 43. vucetich, j. a., r. o. peterson, and c. l. schaefer. 2002. the effect of prey and predator densities on wolf predation. ecology 83:3003-3013. white, c. a., c. e. olmsted, and c. e. kay. 1998. aspen, elk, and fire in the rocky mountain national parks of north america. wildlife society bulletin 26:449-462. white, r. g. 1983. foraging patterns and their multiplier effects on productivity of northern ungulates. oikos 40:377384. wiens, j. a. 1989. spatial scaling in ecology. functional ecology 3:385-397. alces vol. 39, 2003 peterson et al. – space and time in predation dynamics 231 sl at e is la nd s c ar ib ou 6 0 0 0 0 0 12 se ip (1 99 1) n or w ay c ar ib ou 3. 5 0 1 0 0 1 7 se ip (1 99 1) n ew fo un dl an d c ar ib ou 8. 5 0 1 0 0 1 17 se ip (1 99 1) so ut h g eo rg ia c ar ib ou 2 0 1 0 0 1 4 se ip (1 99 1) q ue sn el l ak e c ar ib ou 0. 03 1 0 1 0 2 0. 06 se ip (1 99 1) o nt ar io c ar ib ou 0. 03 1 1 1 0 3 0. 06 se ip (1 99 1) sa sk at ch ew an c ar ib ou 0. 03 1 1 1 0 3 0. 06 se ip (1 99 1) s. f in la nd m oo se 0. 4 0 1 0 0 1 2. 4 n yg re n (1 98 7) a lb er ta m oo se 0. 8 0 1 1 0 2 4. 8 r ol le y an d k ei th (1 98 0) sw ed en m oo se 1. 5 0 1 0 0 1 9 c ed er lu nd an d m ar kg re n (1 98 7) e lk is la nd , a lb er ta m oo se 1. 5 0 1 0 0 1 9 c ai rn s a nd t el fe r ( 19 80 ) n ew fo un dl an d m oo se 1. 8 0 1 0 0 1 10 .8 b er ge ru d an d m an ue l ( 19 68 ), m er ce r a nd m an ue l ( 19 74 ), fr yx el l e t a l. (1 98 8) se w ar d pe n, a k m oo se 0. 4 0 1 0 1 2 2. 4 g as aw ay et al . ( 19 92 ) r id in g m t, m an ito ba m oo se 0. 8 1 1 0 0 2 4. 8 c ar by n (1 98 3) h ec la is la nd , m an ito ba m oo se 1 1 1 0 0 2 6 c ric ht on (1 97 7) g as pe si e, q ue be c m oo se 2 0 0 1 0 1 12 c rê te (1 98 9) is le r oy al e, m i m oo se 2 1 0 0 0 1 12 pe te rs on (1 99 9) a is hi hi k, y uk on m oo se 0. 1 1 1 0 1 3 0. 6 l ar se n (1 98 2) k lu an e l ., y uk on m oo se 0. 1 1 1 1 1 4 0. 6 l ar se n (1 98 2) d en al i, a la sk a m oo se 0. 2 1 0 0 1 2 1. 2 si ng er an d d al le -m ol le (1 98 5) n el ch in a b as in , a k m oo se 1 1 1 0 1 3 6 g as aw ay et al . ( 19 92 ) k en ai p en , a k m oo se 1. 1 1 1 1 1 4 6. 6 b ai le y (1 97 8) , p et er so n et al . (1 98 4) , s ch w ar tz an d fr an zm an n (1 99 1) s. q ue be c m oo se 0. 6 1 0 1 0 2 3. 6 po tv in (1 98 8) r id in g m t, m an ito ba m oo se 0. 8 1 0 1 0 2 4. 8 c ar by n (1 98 3) s. c en tr al o nt ar io m oo se 0. 3 1 0 1 0 2 1. 8 b er ge ru d et al . ( 19 83 ) sw q ue be c m oo se 0. 3 1 0 1 0 2 1. 8 m es si er a nd c rê te (1 98 5) a pp en di x 1. d at a us ed in f ig ur e 1. d en si ty / b la ck b ro w n n o. d ee req . l oc at io n sp ec ie s km 2 w ol f h um an be ar be ar pr ed . de ns . r ef er en ce space and time in predation dynamics – peterson et al. alces vol. 39, 2003 232 n . a lb er ta m oo se 0. 2 1 1 1 0 3 1. 2 o os en br ug a nd c ar by n (1 98 5) n e a lb er ta m oo se 0. 2 1 1 1 0 3 1. 2 fu lle r a nd k ei th (1 98 0) g m u 20 a -a k m oo se 0. 2 1 1 0 1 3 1. 2 g as aw ay et al . ( 19 83 ) d en al i, a la sk a m oo se 0. 2 1 0 0 0 1 1. 2 h ab er (1 97 7) n e m n m oo se 0. 6 1 1 1 0 3 3. 6 m ec h an d fr en ze l ( 19 71 ), pe ek et al . ( 19 76 ), n el so n an d m ec h (1 98 6) n e m n m oo se 0. 7 1 1 1 0 3 4. 2 v an b al le nb er gh e e t a l. (1 97 5) e . c en tr al o nt ar io m oo se 0. 2 1 0 1 0 2 1. 2 pi m lo tt et al . ( 19 69 ) e . c en tr al o nt ar io w td ee r 3. 1 1 0 1 0 2 3. 1 pi m lo tt et al . ( 19 69 ) s. q ue be c w td ee r 3 1 0 1 0 2 3 po tv in (1 98 8) e . c en tr al o nt ar io w td ee r 5. 8 1 0 1 0 2 5. 8 pi m lo tt et al . ( 19 69 ) n e m n w td ee r 5. 1 1 1 1 0 3 5. 1 v an b al le nb er gh e e t a l. (1 97 5) n c en tr al m n w td ee r 3. 5 1 1 1 0 3 3. 5 st en lu nd (1 95 5) n c en tr al m n w td ee r 6 1 1 1 0 3 6 b er g an d k ue hn (1 98 2) n c en tr al m n w td ee r 6. 2 1 1 1 0 3 6. 2 fu lle r ( 19 89 ) a nt ic os ti is la nd , p q w td ee r 15 0 1 0 0 1 15 a . c ho ui na rd (p er s. c om m un .) r id in g m t, m an ito ba w td ee r 0. 3 1 0 1 0 2 0. 3 c ar by n (1 98 3) n e m n w td ee r 0. 6 1 1 1 0 3 0. 6 m ec h (1 98 6) , n el so n an d m ec h (1 98 6) n e m n w td ee r 3. 5 1 1 1 0 3 3. 5 m ec h an d fr en ze l ( 19 71 ), pe ek et a l. (1 97 6) , n el so n an d m ec h (1 98 6) w in d c av e n p b is on 2. 6 0 1 0 0 1 20 .8 d et lin g ( 19 98 ) y el lo w st on e n p el k 14 .3 0 1 0 1 2 42 .9 si ng er et al . ( 19 98 a) r oc ky m ou nt ai n n p el k 12 .5 0 1 0 0 1 37 .5 si ng er et al . ( 19 98 b) r id in g m t, m an ito ba el k 1. 2 1 0 1 0 2 3. 6 c ar by n (1 98 3) w in d c av e n p el k 3. 3 0 1 0 0 1 9. 9 d et lin g ( 19 98 ) b an ff /j as pe r n p el k 6. 5 0 0 0 1 1 19 .5 w hi te et al . ( 19 98 ) b an ff /j as pe r n p el k 1. 5 1 0 0 1 2 4. 5 w hi te et al . ( 19 98 ) ... co nt in ue d a pp en di x 1. d at a us ed in f ig ur e 1. d en si ty / b la ck b ro w n n o. d ee req . l oc at io n sp ec ie s km 2 w ol f h um an be ar be ar pr ed . de ns . r ef er en ce 4105.p65 alces vol. 40, 2004 rea et al. mineral licks and land management 161 considerations for natural mineral licks used by moose in land use planning and development roy v. rea1, dexter p. hodder2, and kenneth n. child3 1ecosystem science and management program, university of northern british columbia, 3333 university way, prince george, bc, canada v2n 4z9, email: reav@unbc.ca; 2john prince research forest, p.o. box 2378, fort st. james, bc, canada voj 1po, email: hodderd@unbc.ca; 3systems north, 6372 cornell place, prince george, bc, canada v2n 2n7, email:kchild@shaw.ca abstract: despite an increasing body of knowledge about the predictable use and functional role of naturally occurring mineral licks in the ecology of ungulates such as moose (alces alces), no documents have been published that discuss the importance of implementing management guidelines aimed to protect these habitat features. we reviewed the literature on the biophysical attributes of mineral lick sites and their use by moose to illustrate the importance of licks and outline criteria that may serve to help in the development of guidelines to protect these land features. we canvassed the provinces and territories of canada to ascertain whether any regulatory framework for identifying, classifying, and protecting mineral licks existed. despite appeals for lick protection from several authors, few jurisdictions recognize mineral licks as a special habitat feature and none appear to base their guidelines for protecting licks on ecological principles. we also found no evidence for the existence of a set of standardized guidelines that can be used by planners and managers to ensure the protection of licks. we incorporated ecological and biophysical aspects of mineral licks into a field checklist to identify and classify mineral licks used by moose, and developed a preliminary draft of a management procedure to enable their protection. alces vol. 40: 161-167 (2004) key words: alces alces, development, forestry, habitat feature, mineral pool, mineral spring, moose, reserve, resource use, salt lick, sodium hunger, ungulate mineral licks are unique and important habitat features important in the ecology of moose (alces alces) and other ungulates (ayeni 1971, kreulen 1985, klaus and schmid 1998). unlike dry earth exposures and rock face mineral licks that are used commonly by goats (oreamnos sp.) and sheep (ovis spp.), mineral licks used by moose are generally characterized by well worn trails leading to wet muddy springs or seepage areas that contain dense track concentrations (tankersley and gasaway 1983, jones and hanson 1985). these areas, also referred to as muck licks, are also used by deer (odocoileus spp.) and elk (cervus sp.) and are thought to be extremely sensitive to impacts from land d e v e l o p m e n t a c t i v i t i e s ( w e e k s a n d kirkpatrick 1976, reger 1987, bechtold 1996, dormaar and walker 1996). however, standardized guidelines for field identification, rating the ecological importance of licks, and establishing protective measures for these sites remain uncirculated. we reviewed the literature to summarize use patterns of mineral licks by moose and to ascertain the importance of mineral licks in the ecology of moose. we also reviewed the works of authors appealing for lick protection and canvassed the provinces and territories of canada to determine the current policies and guidelines used for mineral licks and land management – rea et al. alces vol. 40, 2004 162 protecting mineral licks. our objectives were to determine if such a framework existed and identify those criteria required to construct a rating system to facilitate field identification and classification of mineral licks for purposes of protection. ecological role of licks much speculation exists as to why moose and other animals use mineral licks (kreulen 1985, dormaar and walker 1996, klaus and schmid 1998). it is believed that animals visit licks for, among other things, mineral supplementation, soils to aid digestion, and water and social gathering (fraser and hristienko 1981, jones and hanson 1985, risenhoover and peterson 1986, couturier and barrette 1988, heimer 1988). licks are used by moose predominantly from dusk until dawn (fraser and reardon 1980, tankersley and gasaway 1983, couturier and barrette 1988), most often in late spring (fraser and hristienko 1981, tankersley and gasaway 1983, couturier and barrette 1988, filus 2002) and mid-winter (rea, hodder and child, unpublished data), and to lesser degrees at other times of the year. moose use mineral licks in a predictable pattern, obtaining resources from the soil and water of these features. mineral licks and other sources of concentrated sodium may influence the spatial and temporal structure of moose populations (panichev et al. 2002). the health of some moose herds has been reported to be dependent on the presence of and regular access to mineral licks (best et al. 1977). since land management activities may disrupt the integrity of mineral licks and possibly impact ungulate populations (weeks and kirkpatrick 1976, dormaar and walker 1996), several authors have recommended protective measures for licks be integrated into land use policy (best et al. 1977, tankersley and gasaway 1983, reger 1987, bechtold 1996, dormaar and walker 1996, klaus and schmid 1998). regulatory status in canada no jurisdictions are cited in the literature as having management guidelines to safeguard mineral licks from land development activities. despite a lack of such discussion in the literature, 4 of 13 jurisdictions that we contacted across canada recognize the importance of natural licks and have drafted guidelines to ensure mineral licks are considered in land management plans. alberta recognizes mineral licks and provides management suggestions on how to treat these features. while emphasizing that a buffer zone is required, it is recommended that it be one “sight distance” (government of alberta 1994). the definition of a site distance is subjective and open to interpretation, making field application difficult. british columbia identifies a “mineral lick” or “wallow” as a wildlife habitat feature. such features are protected to different degrees on a regional basis at the discretion of the local environmental authorities (government of british columbia 2004). ontario recommends a minimum buffer of 120 m around mineral licks for moose with the recognition that some development and/or extraction activities (i.e., forest harvesting) may occur under special circumstances within the buffer area. unlike other jurisdictions, ontario recommends a sitespecific approach to establishing buffers around a lick site that considers the forest stand and other landscape characteristics (e.g., local hydrology and topography). this includes designing the shape and extent of the buffer zone to ensure the integrity of the site and safe access for moose (ontario ministry of natural resources 1988). quebec legislation defines a lick narrowly as a swamp, spring, or body of water alces vol. 40, 2004 rea et al. mineral licks and land management 163 that contains mineral salts in concentrations greater than 3 parts per million of potassium and greater than 75 parts per million of sodium. management guidelines dictate that these sites, regardless of site specific attributes, retain a 100 m wide undeveloped reserve zone around the lick (government of quebec 2004). no other jurisdictions in canada appear to have formal management guidelines for considering mineral licks, although there may be uncirculated policies and procedures that exist for identification and protection of these sites. some jurisdictions have regulations for managing “habitat features” but are only legislated into management guidelines if the species(s) using that feature is threatened or endangered as in saskatchewan (government of saskatchewan 2003), or special management recommendations are made on a case by case basis as in the yukon territory (yukon department of renewable resources 1996) and nova scotia (anthony p. duke, manager wildlife resources, nova scotia department of natural resources, personal communication). as a result, there appears to be no set of standardized, easy-to-implement guidelines available for resource managers in canada or elsewhere to use that would be helpful in delineating considerations for mineral licks in land use planning and development activities. classifying mineral licks although what constitutes a mineral lick is understood, a comprehensive understanding of use by moose and a procedure to rank the importance of these areas to moose is less clear. assessing certain attributes in the field should indicate whether a site is a functional mineral lick. the same attributes could also be used to determine and rank the relative importance of the site for moose. a site with well worn trails, denser track concentrations, and a more extensive lick area, for example, is likely more important to animals than a small seepage area containing few tracks and an inconspicuous trail network. one method that could be used to identify and classify mineral licks could employ identification of site attributes. a field checklist could be used to identify and describe site attributes commonly associated with mineral licks used by moose (table 1). this procedure would include both quantitative and qualitative measurements, but would not be too complicated, onerous, or timeconsuming for field crews. importantly, certain of these attributes could also be used to assess the impact of any activity. management guidelines there are at least three aspects to consider when managing or regulating distable 1. key site attributes for identifying, and developing a site identification/classification system for wet mineral licks used by moose. the degree to which site attributes are evident may vary seasonally (see text). site attribute reference wet muddy area or seepage dormaar and walker 1996 animal sightings or sign (e.g. pellets, hedged browse, tree rubs, muddy vegetation, bed sites) fraser and hristienko 1981, jones and hanson 1985 dense track concentrations tankersley and gasaway 1983, jones and hanson 1985 exposed mineral soils with clays or organic materials chamberlin et al. 1977, jones and hanson 1985, bechtold 1996 trail convergence fraser and hristienko 1981, tankersley and gasaway 1983, jones and hanson 1985 trail use (i.e., wear or compaction) fraser and hristienko 1981, tankersley and gasaway 1983 evidence of human activities (e.g., bullet casings, hunting blinds, animal remains, etc.) observations by authors , mineral licks and land management – rea et al. alces vol. 40, 2004 164 turbance around mineral licks: (1) protection of the mineral lick site; (2) maintaining the integrity and function of the hydrological system feeding the lick; and (3) minimizing disturbance in surrounding areas during peak visitation times. rating the importance of the mineral lick for moose is the first in a series of steps that allows for its consideration in land development planning. how to best protect the site and maintain its integrity depends on several factors including the sensitivity of all species using the lick, the biophysical factors of the site, and the type of development planned for the area. correctly identifying all species using the lick is important since misidentifying or neglecting to identify threatened species will influence protective measures necessary to mitigate disturbance impacts (reger 1987). reserve zones (buffers) or the like could be used to mitigate disturbance to licks and should be assigned in accordance with the importance of the lick to the wildlife species using the lick, the intensity of use, and the occurrence of similar features across the landscape. lick protection guidelines should also encompass lick site trail networks, hydrological features, nearby thermal and security cover, and adjacent foraging sites (wiles and weeks 1986). although the relative importance of mineral licks to moose is known to wildlife managers, their value as a land feature may be less apparent to land use planners and developers. therefore, conveying the ecological importance of mineral licks to land managers is key to developing and implementing guidelines for protecting mineral licks. specifically, integrating ecological principles within a management framework (table 2) could provide direction and flexibility when prescribing protective measures for mineral licks. for example, development could occur in late fall or early spring when moose activity at mineral licks is minimal (tankersley and gasaway 1983, couturier and barrette 1988, fraser and hristienko 1981, rea, hodder and child, unpublished data) and be carried out during mid-day hours since moose use mineral licks predominantly at night (fraser and reardon 1980, tankersley and gasaway 1 9 8 3 , c o u t u r i e r a n d b a r r e t t e 1 9 8 8 , silverberg et al. 2002). such strategies could reduce stress and unneeded energy expenditures for moose that are sensitive to disturbance (couturier and barrette 1988, silverberg et al. 2002), especially during the winter months (colescott and gillingham 1998). an integrated management approach of this type would help ensure that the integrity of the feature is protected, that the ecological value of the site is maintained, and that land development proceeds in an appropriate fashion. the final step in this integrated management process is to monitor the impact of prescriptions and subsequent development activity on the biophysical attributes of the table 2. management considerations related to the ecological characteristics and role of mineral licks used by moose. ecology m anagement seasonal use a void seasonal act ivit y p eaks (documentat ion/observat ions) daily use a void p eaks in daily use (observations) t olerance t o dist urbance g auge habit uation to human act ivit y (observat ions) t rail sy stem prot ect : machine free z ones should include habit at /t rails soils use and biop hy sical asp ects of lick function t est soils for suscep tibility t o disturbance, comp act ion, and erosion. water sources of lick prot ect : eart h moving activity should not disrup t hy drological flow of lick veget ation cover requirement s m aint ain cover and vegetat ion p roximat e t o lick alces vol. 40, 2004 rea et al. mineral licks and land management 165 site, site use, and activity patterns of moose. monitoring and assessment are imperative in the continual process of developing and modifying guidelines, and allow for feedback during the process (fig. 1). both sitespecific and regional management approaches will benefit from adequate assessment of management prescriptions developed to protect mineral licks. summary we do not fully understand the importance of mineral licks to moose or how land development may impact mineral lick function or influence moose activity patterns at these features. our findings indicate that a set of standardized guidelines for protecting licks is currently needed. systematic identification and classification of mineral licks using a field checklist would facilitate the development of an objective field procedure. broad implementation and testing at several sites would help justify its application. additionally, a set of draft procedures by which managers can start to consider and incorporate these data into management plans for prescribing appropriate lev 1. i.d. site as a mineral lick 2. describe site features (e.g., field card) 3. rank importance of site to animals (using data/expert opinion) 4. establish level of protection (e.g., buffer) 5. proceed with land development 6. monitor impacts 7. adjust field card, ranking system or buffer assignments accordingly fig. 1. process recommended for determining and assigning the appropriate level of protection for mineral licks threatened by land development activity. els of protection for mineral licks is presented. finally, the adoption of an adaptive management style that allows for a finetuning of the management framework in response to monitoring and site assessments is advocated. we recommend that research focus on monitoring moose use of licks and measuring biophysical attributes at lick sites throughout the range of moose. these data could then be used to develop a standardized set of guidelines to help planners and managers implement needed mitigation measures for licks in areas where development is proposed. until such a framework is developed and our understanding of lick function is more complete, a conservative approach, which protects the integrity and security of lick sites for animals from human activities and development, is presently advisable (bechtold 1996, dormaar and walker 1996). acknowledgements we would like to thank all of those government personnel who went out of their way to help us establish what the regulatory status for considering mineral mineral licks and land management – rea et al. alces vol. 40, 2004 166 licks in the context of land development is across canada. we also thank s. grainger of the j. prince research forest for reviewing an earlier draft of this manuscript, and the unbc field camp class of 2004 and j. black for research assistance. references ayeni, j. s. o. 1971. mineral licks: a literature review. obeche 7:46-53. bechtold, j. p. 1996. chemical characterization of natural mineral springs in northern british columbia, canada. wildlife society bulletin 24:649-654. best, d. a., g. m. lynch, and o. j. ronstad. 1977. annual spring movements of moose to mineral licks in swan hills. proceedings of the north american moose conference and workshop 13:215-228. chamberlin, l. c., h. r. timmermann, b. snider, f. dieken, b. l. loescher, and d. fraser. 1977. physical and chemical characteristics of some natural licks used by big game animals in northern ontario. proceedings of the north american moose conference and workshop 13:200-214. colescott, j. h., and m. p. gillingham. 1998. reaction of moose to snowmobile traffic in the greys river valley, wyoming. alces 34:329-338. couturier, s., and c. barrette. 1988. the behaviour of moose at natural mineral springs in quebec. canadian journal of zoology 66:522-528. dormaar, j. f., and b. d. walker. 1996. elemental content of animal licks along the eastern slopes of the rocky mountains in southern alberta, canada. canadian journal of soil science 76:509512. filus, i. a. 2002. moose behavior at salt licks. alces supplement 2:49-51. fraser, d., and h. hristienko. 1981. activity of moose and white-tailed deer at mineral springs. canadian journal of zoology 59:1991-2000. _____, and e. reardon. 1980. attraction of wild ungulates to mineral-rich springs in central canada. holarctic ecology 3:36-40. government of alberta. 1994. alberta timber harvest planning and operating ground rules. government of alberta, edmonton, alberta, canada. government of british columbia. 2004. forest and range practices act – government actions regulation, section 9(1)(c). government of british columbia, victoria, british columbia, canada. government of quebec. 2004. regulation respecting wildlife habitats – an act respecting the conservation and development of wildlife, division 1 (11). government of quebec, quebec city, quebec, canada. government of saskatchewan. 2003. saskatchewan activity restriction guidelines for sensitive species in natural habitats. government of saskatchewan, regina, saskatchewan, canada. heimer, w. e. 1988. a magnesium-driven hypothesis of dall sheep mineral lick use: preliminary results and management relevance. proceedings of the biennial symposium of northern wild sheep and goat council 6:269-279. jones, r. l., and h. c. hanson. 1985. mineral licks, geophagy, and biochemistry of north american ungulates. the iowa state university press, ames, iowa, usa. klaus, g., and b. schmid. 1998. geophagy at natural licks and mammal ecology: a review. mammalia 62:481-497. kreulen, d. a. 1985. lick use by large herbivores: a review of benefits and banes of soil consumption. mammal review 15:107-123. ontario ministry of natural resources. 1988. timber management guidelines alces vol. 40, 2004 rea et al. mineral licks and land management 167 for the provision of moose habitat. ontario ministry of natural resources, toronto, ontario, canada. panichev, a. m, o. y. u. zaumyslova, and v. v. aramilev. 2002. the importance of salt licks and other sources of sodium in the ecology of the ussuri moose. alces supplement 2:99-103. reger, r. d. 1987. survey of mineralsrelated information for selected mineral licks, matanuska valley moose range, alaska. public data file 87-9. division of geological and geophysical surveys, university of alaska, fairbanks, alaska, usa. risenhoover, k. l., and r. o. peterson. 1986. mineral licks as a sodium source for isle royale moose. oecologia 71:121-126. silverberg, j. k., p. j. pekins, and r. a. robertson. 2002. impacts of wildlife viewing on moose use of a roadside salt lick. alces 38:205-211. tankersley, n. g., and w. c. gasaway. 1983. mineral lick use by moose in alaska. canadian journal of zoology 61:2242-2249. weeks, h. p., and c. m. kirkpatrick. 1976. adaptations of white-tailed deer to naturally occurring sodium deficiencies. journal of wildlife management 40:610-625. wiles, g. j., and h. p. weeks. 1986. movements and use patterns of whitetailed deer visiting natural licks. journal of wildlife management 50:487-496. yukon department of renewable resources. 1996. habitat protection guidelines for moose. yukon departm e n t o f r e n e w a b l e r e s o u r c e s , whitehorse, yukon, canada. 144 our thanks to the following individuals who served as referees for alces volume 56. each paper was reviewed by at least 2 referees who judged its appropriateness for publication and provided editorial assistance. ed addison, ecolink science, aurora, on alan arsenault, wood, environment & infrastructure, saskatoon, sk eric bergman, colorado parks and wildlife, fort collins, co mark boyce, university of alberta, edmonton, ab michelle carstensen, minnesota department of natural resources, forest lake, mn vince crichton, manitoba conservation (retired), winnipeg, mb nick decesare, montana fish, wildlife & parks, missoula, mt glenn delguidice, minnesota department of natural resources, forest lake, mn steve destefano, usgs cooperative fish & wildlife research unit, amherst, ma goran ericsson, swedish university of agricultural sciences, uppsala, se forest hayes, university of montana, missoula, mt murray lankester, lakehead university (retired), courtenay, bc robby marcotte, trent university, peterborough, on brian mclaren, lakehead university, thunder bay, on ron moen, university of minnesota, duluth, mn peter pekins, university of new hampshire, durham, nh roy rea, university of northern british columbia, prince george, bc art rodgers, ontario ministry of natural resources, thunder bay, on bill samuel, university of alberta (retired), edmonton, ab jason smith, north dakota game and fish department, jamestown, nd william severud, university of minnesota, st. paul, mn editorial review committee alces37(1)_189.pdf alces37(2)_421.pdf 131 46th north american moose conference and workshop jackson, wyoming 23-26 may 2011 wondered why we weren’t staying at the lodge. a barbeque lunch was provided on the north shore of lake jackson at the university of wyoming-nps research center’s amk ranch in grand teton national park. the center sits at the base of the teton range and offered participants breath-taking views of the tetons and ice breaking up on jackson lake. after lunch, dr. hank harlow, director of the center and noted ecophysiologist at the university of wyoming, gave a short history about the ranch and related research. dr. terry kreeger (wgfd), renowned expert, followed with an informative workshop on chemical immobilization of wildlife. moose, elk, and bison were observed at several locations on this field trip that marked the start of beautiful local weather. the banquet was held wednesday evening when amanda mcgraw, recent m.s. graduate of the university of minnesota, received the newcomer award, and bill samuel, of winter tick fame and long-time conference attendee and recently retired from the university of alberta, received the senior travel award. the members of the organizing committee received the order of alces with kind and insightful words from dr. pete pekins, chief editor alces. the highlight of the evening was when dr. kjell danell of sweden graciously received the 2011 distinguished moose biologist award for his many contributions in research and management of moose. the evening ended with live music provided by the jared rogerson band; during the day, jared is a biologist with the wgfd. this highly diverse conference featured 3 plenary presentations, 22 contributed presentations, 5 posters, a capstone lecture, and field trip. we are grateful to our sponsors including (in alphabetical order) advance telemetry systems (ats), hayden-wing associates, jackson hole outfitters and guides association, jackson hole wildlife foundation, lotek wireless, national museum of wildlife art, nature mapping, rocky mountain elk foundation, teton science school, vectronic aerospace, virginian lodge, western ecosystems technology (west), wildlife heritage foundation of wyoming, the wildlife society-wyoming the 46th north american moose conference and workshop was held at the virginian lodge in jackson hole, wyoming, on 23-26 may 2011. the conference was hosted by the moose working group of the wyoming game and fish department, with extensive help from the wyoming chapter of the wildlife society and the teton science school. there were 100 registered delegates from throughout north america (13 states and 6 canadian provinces), as well as germany, norway, and sweden. their broad experience resulted in vigorous discussions about moose research, management, and biology. due to an extremely wet may and flooding across several states, travel to and from jackson was an adventure for several participants. welcoming remarks were provided by dr. fred lindzey, chairman of the wyoming game and fish commission and professor emeritus, university of wyoming. contributed presentations followed, covering a broad range of topics including harvest, productivity of populations, parasites, and climate change. a public plenary session entitled “managing predator-prey systems” was held tuesday evening at the national museum of wildlife art, and was preceded by an open tour of the museum for conference attendees. invited speakers were three outstanding large predator ecology specialists, mike jimenez (usfws), dr. bob garrott (montana state university-bozeman), and rod boertje (alaska department of fish and game), who discussed the usfws wolf reintroduction program in the yellowstone region, specific research on wolf-elk/bison interactions in yellowstone, and alaskan wolf/bear-prey systems, respectively. their informative presentations provided a strong scientific foundation for subsequent public discussion focused on predator management issues and balancing the new and dynamic large predator-prey systems in the greater yellowstone area. wednesday featured an educational and sight-seeing field trip north of jackson where participants viewed winter moose range in grand teton national park and the willow flats behind the beautiful, historic jackson lake lodge; some 132 chapter, wyoming biologist’s association, wyoming game warden’s association, and wyoming outfitters and guides association. we also thank all the participants who provided items for the auction. the conference was a tremendous success and we are pleased that so many enjoyed first-time viewing experiences including yellowstone national park, the jackson elk feeding grounds, wolves, bison, “sage-brush moose”, displaying sage grouse, spring elk migration, and wolf predation. co-chairs: tim thomas and steve kilpatrick host: wyoming game and fish department location: jackson hole, wyoming date: 23-26 may 2011 number of delegates/participants: 100 205 distinguished moose biologist past recipients 1992 not presented 1991 charles c. schwartz, alaska dept. of fish and game, soldotna, alaska. 1990 rolf peterson, michigan technological university, houghton, michigan. 1989 warren b. ballard, alaska dept. of fish and game, nome, alaska. 1988 vince f. j. crichton, manitoba dept. of natural resources, winnipeg manitoba. and michel crête, ministère du loisir, de la chasse et de la péche, service de la faune terrestre, québec, pq. 1987 w. c. (bill) gasaway, alaska dept. of fish and game, fairbanks, alaska. 1986 h. r. (tim) timmermann, ontario ministry of natural resources, thunder bay, ontario. 1985 ralph ritcey, fish and wildlife branch, kamloops, british columbia. 1984 edmund telfer, canadian wildlife service, edmonton, alberta. 1983 albert w. franzmann, alaska division of fish and game, soldotna, alaska. 1982 a. (tony) bubenik, ontario ministry of natural resources, maple, ontario. 1981 patrick d. karns, minnesota division of fish and wildlife, grand rapids, minnesota. and al elsey, ontario ministry of natural resources, thunder bay, ontario. in 1974, prior to the establishment of the distinguished moose biologist award, the group recognized the pioneering moose research of the late laurits (larry) krefting, u.s. fish and wildlife service, with an individual award. 2009 kenneth n. child, prince george, british columbia. 2007 kris j. hundertmark, university of alaska fairbanks, fairbanks, alaska. 2006 kristine m. rines, new hampshire fish and game department, new hampton, new hampshire. 2005 w. m. (bill) samuel, university of alberta, edmonton, alberta. 2004 w. eugene mercer, wildlife division, st. john's, newfoundland. 2003 arthur r. rodgers, ontario ministry of natural resources, thunder bay, ontario. 2002 bernt-erik sæther, norwegian university of science and technology, trondheim, norway. 2001 r. terry bowyer, university of alaska, fairbanks, alaska. 2000 gerry m. lynch, alberta environmental protection, edmonton, alberta. 1999 william j. peterson, minnesota department of natural resources, grand marais, minnesota. 1998 peter a. jordan, university of minnesota, st. paul, minnesota. 1997 margareta stéen, swedish university of agricultural sciences, uppsala, sweden. 1996 vic van ballenberghe, u.s. forest service, anchorage, alaska. 1995 not presented 1994 james m. peek, university of idaho, moscow, idaho. 1993 murray w. lankester, lakehead university, thunder bay, ontario. 181 editorial review committee our thanks to the following individuals who served as referees for alces volume 46. each paper was reviewed by at least 2 referees who judged its appropriateness for publication and provided editorial assistance. cedric alexander vermont fish and wildlife, st. johnsbury, vt john ball swedish university of agricultural sciences, umea, sweden james bridgland parks canada, ingonish beach, ns vince crichton manitoba conservation, winnipeg, mb steve destefano u.s. geological survey, ma cfrwu, university of ma, amherst, ma gordon eason ontario ministry of natural resources, wawa, on andy edwards 1854 treaty authority, duluth, mn william faber central lakes college, brainerd, mn william foreyt washington state university, pullman, wa shawn haskell maine department of inland fisheries and wildlife, bangor, me ian hatter ministry of environment, victoria, bc murray lankester lakehead university (retired), thunder bay, on mark lomolino suny-esf, syracuse university, syracuse, ny james maskey, jr. university of north dakota, grand forks, nd anthony musante usda-aphis, concord, nh kim poole aurora wildlife research, nelson, bc margo pybus alberta fish and wildlife division, edmonton, ab roy rea university of northern british columbia, prince george, bc kris rines nh fish and game department, new hampton, nh william samuel university of alberta (retired), edmonton, ab david scarpitti mass. division of fisheries and wildlife, westborough, ma tim thomas wy game and fish, jackson, wy shripad tuljapurkar stanford university, stanford, ca 178 michael w. schrage distinguished moose biologist 2010 recipient the distinguished moose biologist award was presented to mike schrage at the 45th north american moose conference and workshop, international falls, minnesota, june 23-26, 2010 in recognition of his many contributions to our understanding of moose biology and management. mike received his b.s. from the university of idaho in 1990, where he learned a great deal about moose from dr. jim peek (dmb, 1994) who also served as his undergraduate advisor. mike went on to virginia tech and received his m.s. in 1994 while studying black bear. after receiving his m.s. mike became the wildlife biologist for the fond du lac band of lake superior chippewa based out of cloquet, minnesota in 1995. his management responsibility is for all wildlife species and hunting and trapping activities on the fond du lac reservation and in their treaty areas of northeast and eastcentral minnesota. among the various projects are population surveys, habitat management projects and liaison with other government and tribal agencies and private organizations where policy and actions may impact fond du lac’s treaty rights. however, mike’s main interest is moose. although mike does not need to publish as a wildlife biologist for the fond du lac, he has been a co-author on peer-reviewed publications on moose in minnesota and isle royale, and has contributed data and effort towards many minnesota dnr publications on moose. in addition, mike has authored numerous internal reports for the fond du lac band, and reports to granting agencies for various contracts. mike has been and is an integral member of the group of managers and scientists working on moose in northeastern minnesota. each year he participates in the moose harvest planning with the dnr and other native american bands. he has manned check stations every year since the mid-1990’s and is part of the moose aerial survey in january. he has given numerous presentations to the public and has had frequent contact with many media outlets. he is a principal investigator on the vhf telemetry project that identified the higher mortality rates of adult moose in northeastern minnesota over the last decade. mike has also spent countless hours conducting over 110 field necropsies of moose in ne minnesota. mike has also assisted dr. peter jordan (dmb 1998) and dr. rolf peterson (dmb 1990) with their research projects on isle royale national park. mike helped organize the 35th north american moose conference on the grand portage reservation in 1999 and the 45th namc in international falls, minnesota. clearly mike has been a tireless advocate for moose management for 16 years. the north american moose conference and workshop is proud to recognize the professional experience of mike schrage, recipient of this year’s distinguished moose biologist award. alces vol. 34 (1), (1998) i dr. albert w. (al) franzmann, age 78, of soldotna, alaska, died unexpectedly at his winter home in green valley, arizona on february 13, 2009. everyone who ever knew al feels a deep sense of loss, but we are extremely grateful for having known him. al was a pioneer in wildlife management, particularly in moose management, and cared deeply about our wildlife resources, hunting and trapping, and future wildlife management. al was a mentor and friend to many and words alone cannot express how he affected our lives, careers, and just how much he meant to us. what follows is dr. franzmann’s obituary, which he wrote. his accomplishments and life are well represented. albert wilhelm (al) franzmann was born in hamilton, ohio, the son of william and louise (schlichter) franzmann who both preceded him in death. al graduated in 1948 from ross township high school in butler county, ohio. he then entered the ohio state university and was awarded membership in phi zeta national veterinary medicine honorary fraternity and was awarded the gamma in memoriam albert w. franzmann d.v.m., ph.d, dipl.a.c.z.m. july 1930 – february 2009 plaque as outstanding senior veterinary medical student. he was a member of alpha gamma rho (agriculture) and alpha psi (veterinary) fraternities. al married donna grueser on december 13, 1953. they were a devoted couple in marriage for 55 years. their son karl was born in 1955 and daughter louise in 1959. following graduation, he served for two years as captain in the united states air force veterinary corps stationed at mcconnell air force base in wichita kansas. from 1956 until 1959, he was in a partnership veterinary practice in tiffin, ohio. from 1959 until 1968 he operated a farm animal practice near hamilton, ohio. he was active in the butler county, cincinnati, ohio and american veterinary medical associations. the era of the family farm was coming to a close during the 1960’s in the hamilton area and this was the aspect of veterinary medicine that al loved. he had to make a choice whether to practice on companion animals, move to where family farms were still viable, or diversify. al chose to diversify and to pioneer the field of wildlife alces vol. 34 (1), (1998) ii veterinary medicine. in 1968, he entered the university of idaho and in 1971 graduated with a doctor of philosophy degree in forestry science based on his research on rocky mountain bighorn sheep physiology. there he was awarded national defense and education act and a national wildlife federation fellowship. he was elected to the xi sigma pi national forestry and phi sigma national biological sciences honorary fraternities. in 1972, the franzmann family moved to soldotna, alaska, where al became a research biologist with the alaska department of fish and game and director of the moose research center. his research produced over 250 publications. he was appointed affiliate associate professor of wildlife biology at the university of alaska, fairbanks, and the institute of arctic biology. dr. franzmann was active in professional wildlife organizations such as the wildlife society (certified wildlife biologist); wildlife disease association (council and emeritus member); american association of wildlife veterinarians (founder president, council member); world association of wildlife veterinarians (organizer): and the american association of zoo veterinarians. al was selected by the i.u.c.n. species survival commission to their deer, bear and veterinary specialty groups. al and donna formed great north enterprises, inc., in 1978 and were the sole distributors of muskol products in alaska until 1983. in 1986, al was elected to the board of directors of the hamilton tool co, in hamilton, ohio. al received recognition for his accomplishments as recipient in 1983 of the distinguished moose biologist award “in recognition for outstanding contributions to the field of moose management” and the einarson award “in recognition of long-standing unselfish dedication and professionalism to wildlife resources.” he was awarded the first honorary diplomat in the american college of zoological medicine in 1990 “ in recognition as a specialist with extensive experience who has provided important service to and achieved eminence in the field of zoological medicine." the alaska bow hunters presented him an award in 1993 “in recognition and appreciation for many years of work in wildlife research and management.” in 1996 he received an emeritus award from the wildlife disease association “in recognition for meritorious contributions to the study and understanding of disease of wildlife." the moose research center, that al directed from1972 until 1987, was awarded the group achievement award in 1992 “for outstanding achievements benefiting wildlife and objectives of the wildlife society." in 1997, he received the distinguished alumnus award from the ohio state university college of veterinary medicine “in recognition of his eminence as a veterinarian who has achieved a record of outstanding contributions in the advancement of veterinary medicine." in 2001 al became an honor roll member of the american veterinary medical association and was give the lifetime conservation award by the kenai chapter of the safari club international. upon al’s retirement in 1987, he pursued international wildlife veterinary consulting as a director of the international wildlife veterinary service, inc. he worked on projects in india, nepal, china, indonesia (irian jaya), argentina, sweden, poland, zimbabwe, south africa, namibia and several provinces in canada and states in the united states. he compiled and edited the book entitled ecology and management of the north american moose that was published in 1998. al was appointed by governor wally hickel to the alaska board of game (19921995). he was elected to the board of directors of the alaska outdoor council and the alaska fish and wildlife conservation fund. in 1999 he was elected to the board of directors of the alaska challenger center for space science technology. al’s avocations included hunting, fishing, gardening, golf, travel and photography. he had over 100 photographs published and received several photographic awards. he was a life member of the issac walton league, the nature conservancy, the national rifle association, the national wildlife federation, and the alaska outdoor council. he was a regular member of many other conservation, wildlife, veterinary and civic organizations. he was politically and socially conservative and supported those efforts. his family, friends and co-workers knew him as a dedicated professional who loved his work; he would comment that “he could not believe that he was getting paid for doing such neat things." he was recognized world-wide as a pioneer in bridging the veterinary and wildlife professions. in recognition of this, he was inducted in 2004 alces vol. 34 (1), (1998) iii into the university of idaho hall of fame “for his leadership and contributions in the field of wildlife veterinary research.” al was a dedicated and loving husband, father and grandfather who valued his family above all else. al is survived by his wife donna, his son karl and wife lisa and their children jessicca and her husband brian walsh of arizona, katherine and jacob franzmann of sterling, alaska; and his daughter louise billaud and husband jean-paul and their son keran of dublin, virginia; his sister elizabeth harding and brother fredrick preceded him in death. albert w. (al) franzmann 137 43rd north american moose conference and workshop prince george, british columbia 2 7 june 2007 order of alces 2007 in recognition of much appreciated service related to organizing the annual meeting, a special certificate, entitled the order of alces, was first presented at the 22nd north american moose conference and workshop. the recipients of this certificate in 2007 were chris johnson, roy rea, jennifer studney, (university of northern british columbia), jeremy ayotte, ken child, ian hatter, doug heard, glen watts (b.c. ministry of environment), dale seip (b.c. ministry of forests and range), mari wood (peace/williston fish & wildlife program), dan aitken (college of new caledonia), john deal (canfor), and dexter hodder (john prince research forest). nation, and unbc. thanks to local businesses and individuals who donated auction items and wild game meat for the banquet. most importantly, a special thanks to jennifer studney, unbc conference & events services, for shouldering the burden of administering the conference details. co-chairs: roy rea and ken child host: university of northern british columbia location: university of northern british columbia 3333 university way prince george, bc, v2n 4z9 date: june 2-7, 2007 number of delegates/participants: 102 as the mountain pine beetle infests pine forests throughout bc, the question is: what is the impact on wildlife, in particular moose? this set the theme for the 43rd north american moose conference: moose in a changing landscape. the 43rd annual moose conference and workshop was held at the university of northern british columbia (unbc), june 2-7, 2007. unbc hosted the conference with the bc ministry of forests and range, the bc ministry of environment, the peace/williston fish and wildlife compensation program, and the john prince research forest. the conference was attended by 102 delegates from canada, the united states, and finland who met to exchange ideas and share new developments in moose research, management, and biology. conference sessions addressed: population dynamics, management, distribution, habitat relationships, science policy, landscape change, vegetation dynamics, vehicle collisions, and mitigation. a pre-conference field trip to the john prince research forest was enjoyed by 9 delegates. delegates enjoyed a one-day field trip east of prince george to view an ancient forest stand of interior red cedar, learn about the dynamics of the mountain pine beetle infestation of pine stands and plantations, visit a moose observation facility near a major highway, and discuss on-going research on vegetation control to mitigate moose-vehicle collisions. kristine rines, recipient of the distinguished moose biologist award in 2006, gave the keynote address about the challenges of the introduction and history of moose management in the state of new hampshire. the 2007 recipient of the distinguished moose biologist award was dr. kris j. hundertmark of the university of alaska fairbanks. an ice-breaker and barbeque encouraged social interaction amongst delegates, spouses, and invited guests. a lively auction at the banquet provided a spirited highlight for all and the expertise of the auctioneer paid dividends to the coffers of alces. we are grateful to our many sponsors whose financial contributions made this conference a success: association of professional biologists of bc, bc hydro, bc ministry of environment, bc ministry of forests and range, canadian forest products, city of prince george, cn railway, dunkley lumber, guide outfitters association of british columbia, habitat conservation trust fund, insurance corporation of bc, integrated land management bureau, john prince research forest, peace/williston fish and wildlife compensation program, spruce city wildlife association, tl’azt’en alces vol. 45, 2009 baskin – moose populations in russia 1 status of regional moose populations in european and asiatic russia leonid m. baskin institute of ecology and evolution, 33 leninsky prospect, moscow, russia 117071 abstract: the moose population in russia contains 4 subspecies and peaked at >800,000 animals in 1990. a substantial population decline of >50% occurred in european russia between 1990 and 2002; populations in asiatic russian have remained more stable. this decline was influenced by the relationships among population densities of moose and humans, available forest habitat, and exploitation of moose. in general, fluctuations in moose populations were lower in areas with more forest habitat and lower human density. alces vol. 45: 1-4 (2009) key words: alces alces, distribution, decline, fluctuation, human impacts, moose, populations, russia. there are an estimated 14.4 million km2 of moose habitat in russia. four subspecies of moose have been documented in russia including 1) alces alces alces linnaeus 1758 found in european russia, the ural mountains, western siberia, and the altai mountains, 2) a. a. pfizenmayeri zukowski 1910 that occupies the area east of the yenisey river to the chersky mountain range, 3) a. a. buturlini (chernyavsky and zhelesnov 1982) that is distributed throughout northeast siberia, and 4) a. a. cameloides milne-edwards 1867 that occupies the amur region and the sikhote-alin mountain range (chernyavsky and zheleznov 1982, heptner 1989) (fig. 1). the various ecoregions that moose inhabit include the tundra, subarctic, subarctic-regime mountains, warm continental, warm continental-regime mountains, prairie, prairie-regime mountains, and temperate steppe (bailey 1998). the population density of moose is influenced by human population density and associated hunting pressure. moose populations in european russia declined in the late 20th century (baskin 1998) and are predicted to continue declining in the early 21st century (lomanov 1995). the state centre for game animal control (borisov et al. 1992, danilkin 1999, lomanov and lomanova 1996, 2000, 2004, lomanova 2007) has provided useful data to interpret population dynamics of moose in russia, and specific regional factors have fig. 1. the range and population estimates (in thousands) of the 4 subspecies of moose found in russia (from lomanova 2007). moose populations in russia baskin alces vol. 45, 2009 2 been identified that influence moose populations. in addition, national data about human population density and distribution, and regional forest inventories have been used to analyze anthropogenic factors that influence moose populations in russia. the highest estimated moose population (833,000) in russia occurred in 1990 (fig. 2). the 2007 population was estimated to be about 600,000 with the overall populations in european and asiatic russia about equal (fig. 1 and 2). there was a substantial decline in the overall moose population from 1991 to 2002, with greatest decline (>50%) in european russia; the asiatic population remained stable at about 300,000 (fig. 2). since 2002, the population has increased in european russia and remained reasonably stable in asiatic russia (fig. 2). current (2007) regional population densities vary; those in european russia are generally higher than those in asiatic russia (fig. 3). concurrent with increasing populations in european russia, moose populations in southern regions have also recovered, particularly in foreststeppe habitats and neighboring ukraine and kazakhstan (erzhanov 2008, minoranskiy et al. 2009, volokh 2009). the fluctuation in moose density appears to correlate with the amount of forest cover available in different regions. for example, dramatic fluctuations in moose population density have been negligible in yakutia that has ample forest habitat and low human population density (boeskorov et al. 2008). areas with high forest cover tend to have the highest moose density and more population stability (fig. 4 and 5). of 45 regions in european russia with higher human population density, 42 experienced population declines from 1990 to 2002; conversely, 36 regions had population increases and only 6 declined from 2002 to 2007 (fig. 4). of 28 regions in asiatic russia with relatively low human population density, only 5 had population declines and 22 had increases from 1990 to 2002; 14 regions had increases and 11 declines from 2002 to 2007 (fig. 5). both population declines and increases in asiatic russia were much smaller on a relative scale than those in european russia (fig. 4 and 5). the relationships among moose population density, forest habitat availability, human population density, and exploitation of moose populations require further study to evaluate and predict the future fig. 3. the estimated regional population densities of moose in russia (moose/km2; from lomanova 2007). fig. 2. the estimated moose populations in european and asiatic russia, 1970-2007 (from borisov et al. 1992, danilkin 1999, lomanov and lomanova 1996, 2000, 2004, lomanova 2007). alces vol. 45, 2009 baskin – moose populations in russia 3 of regional russian moose populations. references bailey, r. g. 1998. ecoregions: the ecosystem geography of the oceans and continents. springer, berlin, germany. baskin, l. m. 1998. moose conservation in ecosystems of eastern europe. alces 34: 395-407. boeskorov, g. g., a. a. krivoshapkin, i. m. fig. 4. the relative change in moose density in 45 regions of european russia; regions are listed left to right by increasing amount of forest area. open boxes represent moose density in 2002 minus moose density in 1990; shaded boxes represent moose density in 2007 minus moose density in 2002. fig. 5. the relative change in moose density in 28 regions of asiatic russia; regions are listed left to right by increasing amount of forest area. open boxes represent moose density in 2002 minus moose density in 1990; shaded boxes represent moose density in 2007 minus moose density in 2002. moose populations in russia baskin alces vol. 45, 2009 4 okhlopkov, a. l. popov, i. i. mordosov, v. t. sedalishchev, n. g. solomonov, and v. g. tikhonov. 2008. the current state of moose in the sakha republic (yakutia). 6th international moose symposium, yakutsk, russia, 14-20 august, 2008. abstract only. borisov, b. p., l. a. gibet, and j. p. gubar’. 1992. status of hunting grounds and numbers of main wild animals in rsfsr. glavokhota, moscow, russia. (in russian). chernyavsky, f. b., and n. k. zheleznov. 1982. moose distribution and systematics in the siberian north-east. byulleten moskovskogo obshchestva ispytatelei prirody otdelenie biologicheskoe 87: 25-32. (in russian with english summary). danilkin, a. a. 1999. olen’i (cervidae). geos, moscow, russia. (in russian). erzhanov, n. t. 2008. peculiarities of moose ecology (alces alces linnaeus, 1758) in central kazakhstan. 6th international moose symposium, yakutsk, russia, 1420 august, 2008. abstract only. heptner, v. g. 1989. taxonomy, geographic distribution, and geographic variation. pages 19-1058 in v. g. heptner and n. p. naumov, editors. mammals of the soviet union, volume 1. model press, leiden, russia. (in russian). lomanov, i. k. 1995. regularities of moose population dynamics and distribution in european russia. central’naja nauchnoissledovatel’skaja laboratoriaja glavokhota rsfsr, moscow, russia. (in russian). _____, and n. v. lomanova. 1996. moose. pages 31-50 in tsentralnaya nauchnoissledovatelskaya laboratoriya okhotnichego khozyaistva i zapovednikov, moscow, russia. (in russian). _____, and _____. 2000. moose. game animals of russia 2: 13-23. (in russian). _____, and _____. 2004. moose. game animals of russia 6: 12-23. (in russian). lomanova, n. v. 2007. moose. game ani-game animals of russia 8: 13-21. (in russian). minoranskiy, v. a., v. v. sidelnikov, and e. i. simanovich. 2009. history and status of moose in the rostov region, russia. alces 45: 21-24. volokh, a. m. 2009. history and status of the population dynamics of moose in the steppe zone of ukraine. alces 45: 5-12. 105 the seasonality of a migratory moose population in northern yukon dorothy cooley1,5, heather clarke1, shel graupe2,6, manuelle landry-cuerrier3, trevor lantz4, heather milligan1, troy pretzlaw3,7, guillaume larocque3,8, and murray m. humphries3 1department of environment, yukon government, whitehorse, yt, canada, y1a 2c6; 2vuntut gwitchin government, old crow, yt, canada, y0b 1n0; 3natural resource sciences, macdonald campus, mcgill university, ste-anne-de-bellevue, qc, canada h9x 3v9; 4environmental studies, university of victoria, victoria, bc, canada, v8w 2y2; 5teslin tlingit council, teslin, yt, canada, yoa 1b0; 6sgs canada inc, edmonton, ab, canada, t6e 4n1; 7parks canada, annapolis, ns, canada, b0t 1b0; 8québec centre for biodiversity science, montréal, qc, canada, h3a 1b1 abstract: at the northern edge of their north american range, moose (alces alces) occupy treeline and shrub tundra environments characterized by extreme seasonality. here we describe aspects of the seasonal ecology of a northern yukon moose population that summers in old crow flats, a thermokarst wetland complex, and winters in surrounding alpine habitat. we collared 19 moose (10 adult males and 9 adult females) fitted with gps radio-collars in old crow flats during summer, and monitored their year-round habitat use, associated environmental conditions, and movements for 2 years. seventeen of 19 moose were classified as migratory, leaving old crow flats between august and november and returning in april to july, and spent winter in alpine habitats either northwest (n = 8), west (n = 4), or southeast (n = 5) of old crow flats. the straight-line migration distance between summer and winter ranges ranged from 59 to 144 km, averaging 27 km further for bulls than cows. in summer, 18 of 19 moose situated their home ranges in and around drained lake basins and shallow lake habitats within old crow flats. in winter, moose at elevations < 400 m selected for river, shrub, or drained lake habitats, whereas those at elevations >600 m selected for shrubby valley bottoms near lakes and rivers within home ranges dominated by alpine tundra. moose at high elevations marginally reduced their exposure to cold extremes due to the prevalence of thermal inversions, but cold avoidance was not a strong driver of habitat selection, including for moose at low elevations. stable isotope signatures of moose hair, aquatic plants, and terrestrial plants were consistent with a year-round, shrub-dominated diet characterized by slight habitatand season-associated dietary differences. local knowledge of the vuntut gwitchin first nation predicted several of our major results, including 1) summer home range fidelity, 2) selection of lakeshore habitats, 3) use of drained lake basins, 4) dietary reliance on shrubs and emergent vegetation, and 5) responses to contemporary environmental changes. although the core habitat of this moose population, including the winter ranges of its 3 subpopulations, is well protected by a variety of special management units, parks, and protected areas in yukon and alaska, pronounced climate warming is dramatically impacting this thermokarst wetland. coordinated monitoring, management, and conservation of this unique landscape, moose population, and socio-ecological system is warranted. alces vol. 55: 105–130 (2019) key words: alaska-yukon, alpine, cervidae, habitat selection, migration, seasonality, subsistence, thermal ecology, traditional ecological knowledge, winter moose (alces alces) are broadly distributed across the boreal forests of north america and scandinavia (telfer 1984), but much recent research has focused on populations at the periphery of their range beyond or at the margins of the boreal forest. at the southern edge of the north american range, a recent period of recolonization and range migratory moose in yukon – cooley et al. alces vol. 55, 2019 106 expansion (wattles and destefano 2011) appears to be transitioning into population decline in certain areas due to impacts of climate change, disease, parasites, and human harvest (lenarz et al. 2009, van beest et al. 2012, decesare et al. 2015, monteith et al. 2015, jones et al. 2019). conversely, certain populations at the northern edge are expanding range and increasing in abundance at the forest-tundra transition zone (hayes and barichello 1986, jung et al. 2009, wald and nielson 2014). less is understood about habitat and climate conditions that constrain seasonal distribution of forest-tundra moose populations (tape et al. 2016), specifically, the ecological determinants of their northern range limit in north america. moose inhabit highly seasonal environments throughout their range and are generally well-adapted to endure highly seasonal environments. for example, in summer they use aquatic habitats extensively for foraging, cooling, and insect relief (timmermann and mcnicol 1988), whereas in winter their large size counteracts deep snow (telfer and kelsall 1984) and cold temperatures (renecker and hudson 1986). seasonal extremes can potentially impact moose populations, particularly as climate change increases summer and autumnal ambient temperature (ta); e.g., increasing summer ta influences habitat selection in norway (van beest et al. 2012) and longer, warmer autumns increase parasitic infestations in the northeastern united states (jones et al. 2019). at the northern edge of the range, the duration and severity of winters, the brevity and intensity of summers, and the overall magnitude of seasonal environmental variation presumably influence seasonal movement patterns and habitat use and preferences. although most moose populations are not migratory, localized movements between summer and winter ranges are common (timmermann and mcnicol 1988). if seasonal habitat use includes changes in elevation, moose typically occupy lowest elevations in late winter and higher elevations in summer, autumn, and early winter (hauge and keith 1981, jenkins and wright 1987). home ranges might be expected to be smaller in summer and larger in winter given that the quantity and quality of forage is higher in summer than winter (timmermann and mcnicol 1988), and home range typically increases as habitat quality declines (van beest et al. 2011, bjørneraas et al. 2012). however, snow depth >60 cm impedes movement (renecker and schwartz 1998) and restricts winter home range size (houston 1968, loisa and pulliainen 1968, phillips et al. 1973), with size declining from early to late winter as snow accumulates (goddard 1970, van ballenberghe and peek 1971, phillips et al. 1973, thompson and vukelich 1981). seasonal shifts in habitat use and movements of certain populations are consistent and long enough to be classified as migration (leresche 1974, pulliainen 1974, van ballenberghe 1977, mauer 1998, demarchi 2003, white et al. 2014, singh et al. 2016, rolandsen et al. 2017), ranging from < 25 km (ball et al. 2001) to >150 km (mauer 1998). the timing of migration varies by population, but typically occurs in late summer and early fall prior to the breeding period, with return around spring thaw. for example, fall migration preceded breeding and was unrelated to snow depth in alaska (gasaway et al. 1983), while spring migration occurred after snow depth dissipated to < 16 cm in sweden (ball et al. 1999). old crow flats (van tat) is an expansive wetland complex located north of the arctic circle in northern yukon and is important, traditional territory of the vuntut gwitchin (people of the lakes) living in old crow. likewise, moose are an important traditional food for the vuntut gwitchin alces vol. 55, 2019 migratory moose in yukon – cooley et al. 107 (schuster et al. 2011), but are harvested infrequently in old crow flats because they are common there only in summer (mossop 1975) and seldom encountered during the autumn, winter, and spring harvesting seasons. this seasonal habitat use is described by local traditional knowledge: “in the spring it’s known from way back [long ago] that moose, they start from the higher ground and go toward the river [and old crow flats]. but during the winter, they stay up in the hills and creeks in the mountains.” (vuntut gwitchin first nation and smith 2010: 215). consistent with this observation, moose radio-collared in late winter in the eastern portion of the brooks range, alaska spent summer in the western portion of old crow flats (mauer 1998). further, during spring and summer, moose in old crow flats feed on vegetation in the drained margins of lakes: “[the best places to hunt moose were] around where you call a dry lake, around the lakes. in summertime, even in wintertime, they stay in one place. they have a trail to where they feed around the lake. you see the grass” (vuntun gwitchin first nation and smith 2010: 71). yeendoo nanh nakhweenjit k’atr’ahanahtyaa (ynnk; taking care of the land for the future) was a community-initiated and community-led international polar year project motivated by local observations of rapid landscape change in old crow flats, including warmer temperatures, low water levels, lake drainages, and increased shrub growth (technical working group and the management committee [twgmc] 2006, wolfe et al. 2011). major findings of this collaborative research project were that 1) old crow flats is warmer now than at any time in the past 300 years (porter and pisaric 2011), 2) catastrophic lake drainages were 4–5 × more frequent from 1972–1990 and 1991–2009 than from 1951–1972 (lantz and turner 2015), 3) shoreline stability is compromised by high water level, wave action, ice push, and the presence of ice wedges (roy-léveillée and burn 2010), 4) vegetative cover surrounding lakes determines whether hydrological processes are dominated by snowmelt or rainfall (turner et al. 2014) which in turn determines lake productivity and other limnological characteristics (balasubramaniam et al. 2015), and 5) shrub succession in drained basins is proceeding along 2 major trajectories dictated by moisture level (lantz 2017). here we report the findings of moose research conducted as part of the ynnk projexct. we combined local knowledge with that data obtained from gps radio-collars to describe moose movements and habitat use in old crow flats relative to season, ecosystem change, subsistence use, and habitat protection. from local knowledge, we hypothesized that this moose population would be migratory and express seasonally divergent habitat selection. we predicted that moose in summer would select highly productive, low elevation wetlands within old crow flats, where they would use and prefer early succession, shrubby habitats within and around drained lake basins. in winter, we predicted moose would select shrubby alpine habitats and, due to the prevalence of thermal inversions, these higher elevation habitats would be characterized by warmer air temperatures than lower elevations in old crow flats. given local knowledge that moose are common in old crow flats in summer but not winter, we radio-collared moose in mid-summer, and tracked the consistency, timing, and spatial extent of movements from old crow flats to determine the location and habitat characteristics of winter ranges. migratory moose in yukon – cooley et al. alces vol. 55, 2019 108 seasonality in movement patterns and habitat selection are likely to be key features of moose ecology at the northern range edge, given the magnitude of seasonal variation and its broad effects on resources, predators, thermal stress, and landscape movement. this research was designed to expand the limited knowledge concerning migration and the seasonal ecology of populations at the northern edge of moose range, while considering habitat protection, harvest vulnerability, and climate change sensitivity of a culturally-important northern yukon moose population. study area old crow flats is a 6,170 km2 wetland complex of international significance (ramsar convention 2004) comprised of 40% water (russell et al. 1978) and containing ~9,000 shallow lakes (turner et al. 2010, lantz and turner 2015) situated within a low elevation basin (<300 m asl) surrounded by mountainous uplands. although above the arctic circle in a zone of continuous permafrost (roy-léveillée et al. 2014) and spanning the forest-tundra transition zone, old crow flats is a highly productive wetland system (smith et al. 2004, mossop 2015). the shallow, flatbottomed lakes are highly productive (allenby 1989, smith et al. 2004) and surrounded by a mixed community of tall and dwarf shrubs and herbaceous vegetation, with conifer woodlands concentrated around rivers and creeks (turner et al. 2014). predators of moose region include wolves (canis lupus), grizzly bears (ursus arctos horribilis), and black bears (ursus americanus) (north yukon planning commission 2009). drained lake basins resulting from catastrophic drainages and gradual declines in water level are common across old crow flats and range in age from <10 to >11,000 years (ovenden 1986, lauriol et al. 2009, lantz and turner 2015). catastrophic drainages result from lakes elevated above and close to incised streams, in combination with unstable shorelines often related to permafrost degradation, ice push, heavy precipitation and/or wave action that can cause lakes to drain rapidly (lantz and turner 2015). vuntut gwichin observers have reported vegetation changes in old crow flats and note an increase in the number of lakes draining and drying (wolfe et al. 2011). drainage events have been detected from aerial photos and satellite imagery (lantz and turner 2015) and observed directly during the study period (wolfe and turner 2008). the climate is characterized by long, cold winters (mean january temperature = −31°c) and short, warm summers (mean july temperature = 15°c) with annual precipitation ~257 mm, with 100 mm as snow (turner et al. 2010). in general, low elevation localities like old crow flats have warmer summers and longer growing seasons than the surrounding mountains which are cooler in summer with longer periods of snow cover. in non-summer months, temperature inversions characterized by warmer air at higher and cooler air at lower elevations are a common feature in most arctic regions including northern yukon and the northeast interior of alaska (bradley et al. 1992, bourne et al. 2010). methods study animals ten male and 9 female (2 with calf) adult moose were captured between 31 july and 4 august 2007 and fitted with gps radio collars (gps 4400mtm, lotek, newmarket, ontario) programmed to record location, collar temperature, and elevation every 4 (n = 3) or 5 h (n = 16). fourteen collars were recovered from recaptured animals in august 2009; satellite transmissions from the other alces vol. 55, 2019 migratory moose in yukon – cooley et al. 109 5 animals excluded temperature and elevation data (fig. 1). for collar deployment, moose were first located from a fixed-wing aircraft, then approached and darted from a helicopter. immobilization was achieved with a mixture of carfentanil and xylazine or medetomidine (telazoltm) and ketamine; naltrexone and antisedantm were used as reversal agents. to avoid stress, chase times were limited and averaged just under 3 min. once safely immobilized, each was blindfolded and provided oxygen; body temperature, heart rate, breathing rate, blood oxygen level, and blood pressure were monitored continuously. each received a radio-collar and ear tags to provide identification by researchers and hunters. blood, fecal, and hair samples were collected and body condition was assessed via ultrasound measurement of rump fat; total handling time was typically 14–20 min. after the reversal procedure, each was monitored until regaining consciousness and on its feet, typically <20 min. no capture mortality was documented. migration, home range, and habitat selection the seasonal timing, spatial extent, and directionality of migration was quantified using the net-squared displacement approach of bunnefeld et al. (2011) as calculated with the adehabitat package (calenge 2006) for fig. 1. individual tracks from gps radio-collar locations generated from 10 bull and 9 cow moose in old crow flats. the inset panel represents the direction and maximum displacement (km) between summer and winter locations for each moose, including 8 animals that migrated to the northwest (white circle outline), 5 to the west (gray circle outline), and 6 to the southeast (black circle outline), as well as 2 non-migratory cows (outlined icons associated with yellow and dark purple tracks). the maximum elevation (m) used by each moose is indicated next to each icon. migratory moose in yukon – cooley et al. alces vol. 55, 2019 110 r (r core team 2016). this approach examines the linear distance between the first summer location and all subsequent relocations to assess individual trajectory of displacement in space and time. migratory phases are characterized by progressive changes (sustained increase/decrease over time) in displacement, whereas ranging phases of seasonal residents are characterized by displacement plateaus. migratory individuals were defined as those exhibiting a plateau-increase and plateau decrease pattern of displacement over time scales (4 phases summing to 1 year) and distance (> home range diameter) consistent with seasonal migration. to assess seasonal timing, we defined the start of a migration phase (autumn departure and spring return) for a given individual as the date when its displacement plateau first transitioned into a sustained increase/decrease, and the end of the migration phase as the first date of the plateau. relocations were classified as ranging after the end of a migration phase (prior to the start of the next phase). to quantify migration distance, we measured the maximum displacement distance between a ranging location in june, july, or august 2007 and a ranging location recorded in december 2007 or january-february 2008. migration direction was defined as the cardinal direction between these maximally-displaced summer and winter locations. as an alternative measure of migration distance, we also estimated the straight-line distance between the last summer ranging location and the first winter ranging location. we also estimated migratory path lengths as the sum of location-to-location distances from the last summer ranging location to the first winter ranging location; movement rates were based on path lengths travelled per unit time. finally, the relative straightness of migratory routes was estimated as linear displacement/path length, again based on the last summer and first winter ranging locations. we estimated seasonal home ranges using single minimum convex hulls in qgis (qgis development team 2016) with ranging locations recorded in winter 2007–2008 and summer 2008. we focus on these 2 consecutive seasons because they provided the most individuals (the same 7 bulls and 7 cows) with complete data. the average number of locations used to calculate individual home ranges was 376 (range = 128–544) in winter 2007–2008 and 362 (range = 92–551) in summer 2008. the location numbers reflect differences in the pre-set fix rate (3 collars recorded location every 4 h, 11 collars every 5 h), missed fixes (4% average, range = 0–34% in winter 2007–2008; 5% average, range = 0–40% in summer 2008), and certain animals commencing migration during the summer (n = 7) and winter sampling periods (n = 9). we controlled for this source of variation by including the number of locations/ season/individual as a covariate in our analysis of gender and seasonal differences in home range size, and calculating home ranges with a standardized number of locations (n = 200) that was intermediate of the minimum and maximum number of locations per individual/season. to assess habitat use, we aggregated 2 terrain classifications developed for old crow flats (turner et al. 2014, lantz and turner 2015) into a simplified classification extended across a larger spatial area that encompassed the winter range outside old crow flats (fig. s1). in addition to the land cover classes described in turner et al. (2014), we included 3 other classes (glacier ice, tussock tundra, and lush vegetation/ shrub thicket) frequently found at higher elevations around old crow flats, and a separate class for drained lake basin which is hypothesized as important moose habitat (lantz and turner 2015). these land cover classes were combined into 7 aggregated alces vol. 55, 2019 migratory moose in yukon – cooley et al. 111 habitat categories (table s1): forest, lake, river, shrub, tundra, barren, and drained. we used these habitat categories to assess second-order (home range habitat within study area) and third-order (habitat use within the home range) habitat selection (johnson 1980) of moose in winter and summer. for second-order selection, we considered the entire study area as the available habitat for all individuals (design ii in manly et al. 2007). we emphasize this second-order analysis because: 1) given the distances and variable timing of the migration movements we observed, it was reasonable to assume that any individual moose could have moved anywhere in the study area within days or weeks, and 2) this scale of selection relates most directly to our seasonal hypothesis because it compares the habitat characteristics of occupied home ranges to those available if an animal had not migrated, or migrated to a different location. because the second-order selection is likely complemented by third-order selection of specific habitat features, we considered a 100-m buffer radius around each ranging location as the used habitat and the associated individual home range as the available habitat (design iii in manly et al. 2007). using log-likelihood chi-squared statistics in the adehabitat package (calenge 2006) for r (r core team 2016), we computed selection ratios to test whether individuals used habitats in proportion to availability and whether selection was similar for all individuals. influence of air temperature this analysis evaluated the hypothesis that moose capitalize on thermal inversions by moving to higher elevations in winter to avoid cold temperature extremes in old crow flats. because different moose occupied a range of elevations at any point in time (except in summer), relating collar temperature to elevation provided a means to assess the prevalence and strength of inversions within the study region, and the potential use of elevation in defining the thermal environment occupied by moose. specifically, we predicted that in winter, i) temperatures recorded by collars (tc) in winter would be positively correlated with elevation, ii) the relationship between tc at high elevations and ambient temperature (ta) at low elevations would be characterized by a breakpoint, below which high elevation tc would decline less than ta at low elevation, iii) if moose demonstrated temperature selection through comparison of used versus available ta, the strongest selection would occur at high elevation when ta was coldest, and iv) moose were more likely to move from low to high elevation when a decline in ta coincided with development of a thermal inversion. data available during the study were primarily limited to hourly measurements of ta at the old crow airport (elevation 251 m, hereafter ta-oc) and tc measured by 14 of the radio-collars. in addition, a temporary weather station positioned within the central portion of old crow flats (67.903995 °n, −139.746503 °w, elevation 308 m) recorded hourly measurements of ta from june 2008 to september 2011. hourly ta-oc was highly correlated with that measured at the temporary station in old crow flats (ta-ocf) on 5 august 2008 to 2 august 2009 (r = 0.986, n = 8712, p < 0.0001); 88 and 98% of observations differed < 5 and < 10°c, respectively (fig. s2a). the pattern of residual variation around this correlation indicated that summer temperatures were slightly warmer at old crow airport than within old crow flats (e.g., when ta-oc = 25°c, predicted ta-ocf = 22.7°c), with winter temperatures more similar (e.g., when ta-oc = −40°c, predicted ta-ocf = −41.0°c). the largest deviations that occasionally exceeded 10°c (1.8% of migratory moose in yukon – cooley et al. alces vol. 55, 2019 112 observations) tended to occur in winter when ta-oc was between −5 and −30°c (fig. s2b). we tested prediction i) by assessing the correlation between tc and elevation from 1 october to 30 april when moose occupied a wide range of elevations (coefficient of variation > 30%). the basis of this comparison was the range of elevations occupied by different animals and the corresponding tc at a given point in time. we assessed this relationship 6 times per day (tc was recorded every 4–5 h) using simple linear regression where slope represented the change in tc by elevation, hereafter defined as l for lapse rate in oc/1000 m; sample size was the number of moose with measured tc and elevation during the 4–5 h interval. because l was likely to change over time as thermal inversions form and dissipate (bradley et al. 1992), we used a generalized additive model (gam) to assess how l varied over time in relation to time of day (fitted as a cyclic cubic spline function), day of year (also fitted as a cyclic cubic spline function) interacting with elevation (fitted non isotropically using the “tensor product” smoothing function which performs better when covariates are not on the same scale), and habitat category (fitted as a factor) in the package mgcv (wood 2017) for r (r core team 2016). we tested prediction ii) by assessing the relationship between tc recorded on 14 moose and ta-oc. for each we examined the ta-oc threshold at which the relationship between tc and ta-oc decouples by fitting a 2-phase linear regression using the package segmented (muggeo 2017) for r (r core team 2016). we then assessed whether the position and slope of the low temperature divergence was related to the maximum elevation occupied by moose in winter. we tested prediction iii) by interpreting tc as the used temperature and comparing it to available tas which were estimated using a combination of ta-oc and the tc recorded at various elevations. specifically, ta at a given elevation ex, at a given time ty was estimated with the following equation: t e e la oc t x oc t, y y( )= + −− where ta-oc is the temperature at old crow airport, eoc is the elevation of old crow airport (251 m), and lty is the estimated lapse rate at time ty in units oc/1000 m. we estimated ta for 42 ex (25 m increments from 225 to 1250 m) at 6 ty per day from 3 august 2007 to 5 august 2009. for ty between 1 october and 30 april, we used the gam described above to estimate l as the slope of the relationship between tc and elevation recorded by 14 radio-collars (excluding time blocks with < 3 recordings) using a fixed time and habitat category (i.e., 1200 h in tundra habitat). for ty between 1 may to 30 september when inversions rarely occur (bradley et al. 1992), and most moose occupied only low elevations, we adjusted ta-oc to reflect the average relationship (intercept and slope) with tc recorded on 3 moose at similar elevations (table s2), and assumed l was equal to the dry adiabatic lapse rate of −9.8°c/1000 m. finally, to test prediction iv), we combined information on the temporal pattern of thermal inversions with moose movements. this included an analysis of individual mid-day movements as a function of tc, using the minimum distance (m) moved during the same 4or 5-hour period ending between 1000 and 1400 h each day, excluding the migration phase of the annual movement cycle. a second analysis assessed if, when weather changed from inversion conditions to standard lapse rates or vice versa, moose moved to higher or lower elevation to minimize their exposure to temperature variation. alces vol. 55, 2019 migratory moose in yukon – cooley et al. 113 stable isotopes and diet we assessed moose diets by comparing δ13c and δ15n signatures measured in moose guard hairs collected at summer captures with those measured in potential forages within old crow flats, capitalizing on the distinct δ13c signatures of terrestrial and aquatic plants, and δ15n differentiation across terrestrial plants (milligan et al. 2010). submerged aquatic vegetation included siberian water-milfoil (myriophyllum sibiricum, milf), alpine pondweed (potamogeton alpinus, pwda), white-stem pondweed (p. praelongus, pwdw), and richardson’s pondweed (p. richardsonii, pwdr). emergent aquatic plants sampled included yellow pond-lily (nuphar polysepala, lily), northern bur-reed (sparganium hyperboreum, brrd), white (water) sedge (carex aquatilis, sdge) and water horsetail (equisetum fluviatile, hrtl). terrestrial plants sampled included tealeaf willow (salix pulchra, wlt), feltleaf willow (s. alaxensis, wlf), unidentified willow species (s. spp., wls), alder (alnus viridis, ald), and dwarf birch (betula glandulosa, dbr). guard hairs were collected from 12 moose (7 bulls, 5 cows) during the 2007 summer capture. metabolically inert tissues such as hair reflect the diet during the period of growth and retain this dietary signature in chronological sequence (darimont and reimchen 2002, ayliffe et al. 2004). the base portion of the hair represents the most recent dietary assimilation and the tip at an earlier period (darimont and reimchen 2002). given that moose have a single annual molt in april-may (franzmann and schwartz 1997) which precedes spring melt in our study area, and that hair grows incrementally through summer into autumn, the base portion should represent the summer diet and the tip the diet prior to spring melt (darimont and reimchen 2002). we cut the guard hairs into 3 sections (base, middle, and tip) and assumed the base was reflective of the summer diet and the tip the late winter diet (milligan 2010). although we did not know specifically where moose were prior to sampling, given the annual fidelity to winter and summer ranges, we assigned individuals to the summer and winter ranges they occupied post-collar deployment. guard hair samples were prepared for continuous-flow isotope ratio mass spectrometry analysis at the water and aquatic sciences research program at university of victoria using a costech 4010 elemental analyser coupled to a therm delta v mass spectrometer. we incorporated diet-to-hair fractionation values for mammalian herbivores based on meta-analyses of captive animal experiments from the literature and estimates for moose (+3.0‰ for δ13c and +2.7‰ for δ15n; mccutchan et al. 2003, sponheimer et al. 2003a, 2003b, vanderklift and ponsard 2003, tischler 2004, schwertl et al. 2005). this correction involved subtracting 3.0‰ for δ13c and 2.7‰ for δ15n to account for dietary fractionation. we used isotopic mixing models (isosource; phillips and gregg 2003) to estimate the contribution of shrubs and macrophytes in the diets. plans with similar isotopic signatures were grouped together to meet the criteria for distinct dietary sources isotopic modeling (gannes et al. 1997, phillips and gregg 2003). the 4 categories of plants were 1) willow (bark and leaves; salix spp.), 2) feltleaf willow/alder/birch (bark and leaves; s. alaxensis, betula glandulosa, and alnus crispa), 3) a composite of emergent aquatic plants that represented emergent and submergent plants with similar isotopic values (including carex sp., c. utriculata, c. aquatilus, comarum palustre, equisetum fluviatile, sparganium sp., and s. hyperboreum), and 4) submerged aquatic plants (myriophyllum sibiricum, potamogetonzosterifolium, p. richardsonii, migratory moose in yukon – cooley et al. alces vol. 55, 2019 114 p. pusillus, p. alpinus, p. praelongus). in our isosource models, we examined all possible combinations of the models using source increments of 1‰ and mass balance tolerance values of 0.1%, which incorporate uncertainty to the models with a magnitude similar to measurement error and source variability in isotopic values (phillips and gregg 2003). results migration, home range, and habitat selection seventeen of 19 moose were classified as migratory (fig. 2a) moving to the northwest (n = 8), west (n = 4), or southeast (n = 5) from old crow flats, and wintering at high (x = 733, range = 588–902 m), intermediate (x = 436, range = 387–502 m), and low elevations (x = 347, range = 292–488 m), respectively (fig. 1); the 2 non-migratory animals remained in or near the southeastern portion of old crow flats throughout the year. in each winter range, the maximum elevation occupied by bulls averaged 70–125 m higher than that of cows. the timing and duration of the autumn migration was highly variable, with the first moose leaving old crow flats on 4 august and the last on 30 november. a peak in movement occurred in mid-september during a 15-day window (2–17 september) when 9 of 17 moose were migrating (fig. 2b). the timing of spring migration back to old crow flats also varied, with the first moose returning on 14 may and the last on 14 july. in general, spring migration peaked during a 15-day window (9 and 24 may) when 6 animals were migrating (fig. 2b). the duration of the autumn migration varied between 2 and 201 days (x = 45 d, n = 17), and in spring between 2 and 102 days (x = 39 d, n = 14) (fig. 2b). there were no gender differences in migration start dates or durations in autumn (f1,15 = 0.676, p = 0.424; f1,15 = 2.245, p = 0.155), or spring end dates or durations in spring (f1,10= 0.006, p = 0.937; f1,10 = 0.621, p = 0.449). migration distance (measured as maximum straight-line displacement distance from a summer and winter ranging location) varied from 59 to 144 km (n = 15 moose with complete winter data), and was 27 km longer (f1,13 = 5.861, p = 0.030) for bulls (x = 111 ± 25 km (sd), n = 9) than cows (x = 84 ± 13 km (sd), n = 6) (fig. 2c). migration distance based on straight-line displacement distance measured from the last summer ranging location to the first winter ranging location varied from 26 to 135 km among the same 15 animals, and averaged 33 km further for bulls (84 ± 26 km) than cows (51 ± 22 km). migratory path lengths travelled by these moose (the sum of location to location distances from the last summer ranging location to the first winter ranging location) varied from 39 to 326 km, and averaged 81 km further for bulls (165 ± 96 km) than cows (84 ± 44 km). daily migratory path lengths were typically < 10 km/d, but the fastest migrants (2 cows and 1 bull) traveled ~20 km/d. the relationship between migratory path length and linear displacement, indicative of the relative straightness of a migratory route, averaged 0.65 (range = 0.34–0.91) and was somewhat more variable in bulls (range = 0.34–0.91) than cows (range = 0.52–0.89). annual path lengths (total distance traveled in 365 consecutive days; n = 12 moose [2 non-migratory] with location data > 1 year in duration) ranged between 501 and 1099 km (708 ± 183 km). the 2 non-migratory cows had ~ 25% shorter path lengths (522 and 568 km) than the migratory cows (710 ± 134 km). home range size differed by season (f1,24 = 6.913, p = 0.015) but not by sex (f1,24 = 0.001, p = 0.975), with the significant correlation between estimated home range and alces vol. 55, 2019 migratory moose in yukon – cooley et al. 115 the number of relocations (f1,24 = 5.928, p = 0.022) accounted for as a covariate. home range was smaller in summer (48 ± 33 km2, range = 2–117 km2) than winter (170 ± 164 km2, range = 13–674 km2) across the population and by individual (tpaired = −2.7885, df = 13, p = 0.0154). based on significant coefficient estimates [log10(hrs (km2)) = 0.93177 + 0.0017666 * (number of locations) + 0.4799 * (0 summer, 1 winter)], the predicted home range size (200 relocations) in summer and winter was 19 and 58 km2, respectively. home range elevation was lower in summer (307 ± 10 m) than fig. 2. migratory pattern, distance, phenology, and sex differences for moose radio-collared in the study area. panel (a) presents the displacement (km) between the first summer location and all subsequent relocations over time for 19 moose grouped by migration direction and elevation (northwest-high elevation, nw/high); west-intermediate elevation, w/int); southeast-low elevation, se/low), by sex, and ordered by displacement within these categories. two nonmigratory cows are presented separately (outlined icons) and included in the se/low group. light and dark gray shaded areas represent 3-month summer and winter seasons, respectively. panel (b) is a circular plot showing the timing and duration of autumn (white lines, n = 17) and spring migration (black lines, n = 12) in the first year of the study; radial lines indicate the first day of each month. panel (c) indicates sex differences in migration distance expressed as maximum displacement (km) for 9 bulls and 6 cows with complete winter data. the closed circle indicates 1 bull that migrated a shorter distance (59 km) than most, but was classified as migratory; the open circles are the 2 non-migratory cows. migratory moose in yukon – cooley et al. alces vol. 55, 2019 116 winter (648 ± 270 m) across the population (f1,26 = 22.295, p < 0.001) and by individual (tpaired = −4.6651, df = 13, p < 0.001). moose used resources differently in winter and summer at the second-order of habitat selection, and the “average” level of selection was strong (summer xl2² – xl1² = 901,326, 6 df, p < 0.001, winter xl2² – xl1² = 110,303, 6 df, p < 0.001). in summer, moose consistently used home ranges in drained lake basins and lake habitats, except for 1 animal in the southeast that used shrub and river habitat (fig. 3, top left panel). in contrast, second-order selection in winter was individually and regionally variable (fig. 3, top right panel). moose at high elevation in the northwest established home ranges in barrens, river, shrub, and tundra habitats, whereas moose at intermediate (west) and low elevations (southeast) had home ranges in forested, shrub, or lake habitats, except 1 animal using drained lake basins. at the third-order of habitat selection, habitat use in winter and summer was also selective and strong overall (summer xl2² = 37,059, 81 df, p < 0.001, winter xl2² = 23,898, 57 df, p < 0.001). in summer, fig. 3. selection ratios for 7 habitat categories in summer (left) and winter (right) at a) the second order scale (home range) and b) the third order scale (locations within home range). colors indicate selection ratios for moose wintering in the northwest at high elevation (white, n = 7), in the southeast at low elevation (black, n = 5 including the 2 non-migratory cows), and in the west at intermediate elevation (grey, n = 2). alces vol. 55, 2019 migratory moose in yukon – cooley et al. 117 third-order habitat selection was inconsistent among moose, with some most likely to use locations in drained lake basins, and others in river habitats or forested, barren, or tundra habitat (fig. 3, bottom left panel). in contrast, third-order habitat use was individually and regionally consistent in winter. most selected strongly for river habitats, particularly those at high elevation in the northwest. the few outliers included single animals in forested habitat at high (northwest) and intermediate elevation (southeast), and 2 animals in drained lake basins at low elevation in the southeast (fig. 3, bottom right panel). drained lake basins were used preferentially during summer at all selection scales. at the broadest level, moose concentrated summer activity in old crow flats where drained lake basins are a major landscape feature. at the second order of selection, drained lake basins averaged 6% (0–25%, n = 14) of home ranges but only 0.2% of available habitat in the entire study area and 3% of old crow flats (table s1). conversely, drained lake basins were not strongly or consistently selected for at the third order of selection across the population (fig. 3). however, 26–46% of summer locations for 5 animals were in drained lake basins, 2–40 × higher than expected based on availability in their home ranges. influence of air temperature in 2008, ta-oc ranged between a maximum of 27°c on 23 june 2008 at 1800 h and a minimum of −49°c on 31 january 2008 at 1100 hr. similarly, between 1100 and 1300 h on 31 january, tc averaged −42°c (range = −40 to −43°c) for 7 moose located at < 500 m elevation; tc was much warmer (range = −26 to −27°c, fig. 4a) for 2 moose at > 800 m elevation. this inversion of increasing temperature with increasing elevation (r2 = 0.8, p < 0.001) was not constant over time. for example, by 24 february 2008, tc at high elevation (600–800 m) was cooler (ave. = −23, range = −22 to −25°c) than at low elevation (<350 m) (ave. = −16, range = −18 to −13°c; r2 = 0.6, p = 0.001, fig. 4a). fitting a gam to tc identified a significant interaction of date and elevation, consistent with strong winter inversions when ta in december-januaryfebruary averaged 9°c warmer at 1250 m than at 250 m, and standard lapse rates in autumn when ta in september-october was approximately 6°c cooler at 1250 m than 250 m. however, the patterns of variation in tc at different elevations suggested that even in mid-winter, conditions frequently alternated between strong and weak inversions, and standard lapse rates (fig. 4b). comparing ta-oc to tc for 3 moose that remained at low elevation throughout the study period indicated that tc closely tracked variation in ta-oc with additional influence of season (consistent with changes in pelage insulation), daynight (consistent with solar radiation), land cover (consistent with thermal cover), and ta (controlling for all other variables; fig. s3). winter occupation of high elevations permitted moose to marginally reduce their exposure to cold extremes, while their consistent summer use of low elevations resulted in exposure to warm extremes. when available temperatures t( )a e t, ,x y were between −30 and −40°c, it was 3.5°c warmer, on average, for the highest than lowest elevation moose (fig. 5). extremely cold temperatures (tc < −40°c; n = 1646 records) were rarely measured for moose at elevations > 800 m (8% of 1646 records), but were more frequent at < 600 m (62% of 1646 records). two-phase regression analysis of tc (across the range of occupied elevations) and ta-oc indicated that for 7 high elevation moose, tc began to decouple from ta-oc at an average ta-oc threshold of −13.8°c (range = −22.5 to migratory moose in yukon – cooley et al. alces vol. 55, 2019 118 −3.2°c); below this threshold, for every 10°c decline in ta-oc, tc declined on average 5.8°c (range = 4.9 to 6.6°c). in contrast, for 7 low and intermediate elevation moose, the decoupling of tc and ta-oc occurred at a lower threshold (average = −23.3°c, range = −46.8 to −5.4°c) and was stronger; for every 10°c decline in low elevation ta-oc below the 2-phase threshold, tc declined on average 10.7°c (range = 6.2 to 32.4°c) (fig. s4). nevertheless, the magnitude of cold avoidance was small relative to the range of temperatures moose were potentially exposed to across the 225–1250 m elevation profile within the annual range of the study population (fig. s4). the tc was consistently intermediate within this range, including during cold extremes and among moose at all elevations (fig. s4). the only exception was when tc was consistently close to the maximum available during the warmest summer temperatures, because nearly all moose were located at low elevations where temperatures were warmest. finally, the mid-day movement distance of moose (restricted to ranging phases of the annual movement cycle) did not vary consistently with tc (fig. s5) at any elevation. in winter, when conditions reversed from strong inversions to strong standard lapse fig. 4. elevation-related variation in winter air temperatures in and around old crow flats, yukon as measured by radio-collars (tc; n = 13). panel (a) presents the relationship between tc and elevation (m) between 1200 and 1600 h on 31 january 2008, the coldest day of winter 2007–2008 which was characterized by a strong thermal inversion (open circles); the same relationship is presented between 800 and 1200 h on 24 february, a warmer day characterized by a strong standard lapse rate (closed circles). panel (b) presents seasonal prevalence and finer scale temporal variation in this relationship from 1 october 2007 to 30 april 2008. the vertical axis indicates the lapse rate ( lty ; change in tc /1000 m in elevation) with positive values indicating thermal inversions and negative values indicating standard lapses. the solid line indicates the daily linear relationship between 1000 and 1400 hr. the dashed line indicates values predicted for 1200 h in tundra habitat by a gam model including time of day, day interacting with elevation, and habitat category. the first and second vertical dotted lines indicate the 2 days presented in panel (a). alces vol. 55, 2019 migratory moose in yukon – cooley et al. 119 rates or from standard lapse to inversion conditions, moose did not move in elevation to minimize tc variation or avoid temperature extremes (fig. s6). stable isotope and diet summer and late winter diets of moose, as inferred by stable isotope signatures of hair samples relative to vegetation, were consistent with an annual shrub-dominated diet, especially willow, dwarf birch, and/or alder (alnus crispa) (fig. 6). submerged aquatic vegetation had enriched δ13c relative to terrestrial vegetation, with emergent aquatic vegetation intermediate of the two. terrestrial vegetation differed primarily in δ15n enrichment, with all aquatic vegetation characterized by similar δ15n signatures intermediate of the least and most enriched terrestrial vegetation. in general, moose had low enrichment of δ13c and δ15n, with mixing models indicating a diet dominated by terrestrial shrubs (ca. 60% including various willow species, dwarf birch, and/or alder), followed in importance by emergent vegetation (ca. 25% mostly water sedge and horsetail) and submerged vegetation (<15% including potamogeton and myriophyllum). in summer, δ13c signatures in moose were consistent with the importance of terrestrial versus aquatic vegetation, whereas in late winter, most variation was in δ15n fig. 5. the relationship between tc (used) and ta (available) throughout the study area and study period; these measurements are presented for every 4-h block from 3 august 2007 to 5 august 2009 (situated on the x-axis) according to ambient temperature (ta-oc) recorded at old crow airport (251 m elevation) during the same 4-h block. available temperatures (grey circles) were estimated for each 4-h block across 25 m increments from 225 to 1250 m (see methods and table s2). black lines indicate used temperatures based on the relationship between tc and ta-oc using a 2-phase linear regression. the relationships for 2 moose are presented; one occupied the highest annual average elevation (dotted black line) and one the lowest annual average elevation (dashed black line) relative to all moose (see fig. s4). the black solid line is the 1-to-1 line. the corner boxes indicate temperatures below the estimated lower and upper critical temperatures of moose in winter (<−35°c) and summer (>15°c), respectively (renecker and hudson 1986). migratory moose in yukon – cooley et al. alces vol. 55, 2019 120 signatures consistent with the importance of terrestrial species. in summer specifically, moose in lake-dominated home ranges had slightly higher δ13c signatures indicating higher use of aquatic vegetation compared to moose in forest-dominated home ranges with δ13c signatures indicating higher use of terrestrial shrubs (f1,10 = 5.381, p < 0.05) (fig. 6). in late winter when aquatic plants are unavailable, moose occupying home ranges predominated by alpine habitat had δ15n signatures indicating higher use of feltleaf willow, dwarf birch, and/or alder than moose with home ranges in forest-dominated habitat which had δ15n signatures indicating higher use of tealeaf willow (f1,10 = 2.083, p = 0.18) (fig. 6). bulls and cows spending late winter in forest-dominated home ranges or summering in lake-dominated home ranges had similar δ13c and δ15n signatures. discussion the old crow flats moose population is characterized by seasonal movement patterns and habitat selection, reflective of the extreme seasonality of the subarctic-to -arctic landscape where this population lives. the landscape transitions from a highly productive, food-rich wetland in summer to a winter landscape dominated by lake ice (shallow lakes comprise 40% of the surface area), snow, and cold air drainage. accordingly, moose concentrate activity in old crow fig. 6. the δ13c and δ15n signatures measured in moose hair and vegetation collected in the study. the hair signatures are adjusted by a fractionation constant (see methods). vegetation signatures are for plants collected in and around old crow flats as measured by milligan (2010) and include 4–6 species of submerged aquatic plants (dark grey), 4 species of emergent aquatic plants (white), and 5 species of terrestrial plants (light grey) (species codes in methods). ellipses show standard deviations (75% ci) for the mean isotopic values of the three vegetation categories. alces vol. 55, 2019 migratory moose in yukon – cooley et al. 121 flats during summer, arriving there around spring thaw and leaving in late summer for autumn and winter ranges in alpine valleys widely dispersed around old crow flats. old crow flats is already a large landscape covering > 5,600 km2, and moose migration into the surrounding mountains extends the ecological reach of old crow flats to an area 14 × larger based on spatial locations (100% minimum convex polygon) of the 19 study moose captured and radio-collared in old crow flats. this behavior renders old crow flats a central foraging place (sensu (boyd et al. 2014)) that moose return to year after year from multiple distant wintering sites. the annual movement patterns indicate that moose summering in old crow flats are mostly migratory, with ~ 90% migrating 50–150 km between summer and winter ranges. the average and maximum migration distances (bulls = 111 and 144 km; cows = 84 and 98 km) were somewhat shorter than those for the same study population captured and radio-collared on their winter range and estimated with the same method (average = 123 km, maximum = 196 km; mauer 1998). most moose displayed inter-annual consistency in migration timing, fidelity to winter and summer ranges, and migratory routes between these ranges. migration tracks were relatively straight (average linear displacement/path length = 0.65) given the topographical complexity of the landscape, possibly reflecting limited effects of terrain and land cover on migratory paths and travel rate and warranting further quantitative assessment. both our and mauer’s (1998) studies confirm that the summer population of moose in old crow flats lies at the extreme of moose migratory behaviour in north america (leresche 1974, van ballenberghe 1977, demarchi 2003, white et al. 2014). migration from the north atlantic coast in norway to sweden’s inland boreal forests is another example of long distance movement (31–171 km), yet opposite in that moose move from high elevations in summer to low coastal elevations in winter (bunnefeld et al. 2011). moose returned to old crow flats coincident with spring breakup, but many leave old crow flats well before the onset of autumn freeze up and prior to the september and early october breeding season (mauer 1998). although bulls migrated farther than cows, the timing of movement (onset, halfway, and completion) was similar for both sexes. it is possible that a nutritional decline in forage quality occurs late in the growing season in old crow flats, and moose seek higher quality forage at higher elevations where spring green-up and autumn senescence occur later, and moose density and herbivory are low in summer. in summer, moose situate their home ranges where more lakes and drained lakes are available in old crow flats than at higher elevations. the isotope composition of moose hair collected in old crow flats was consistent with a year-round, shrub dominated diet supplemented with aquatic vegetation in summer, particularly for moose using substantial lake habitat. analysis of fecal samples from the study moose yielded similar results (milligan 2010). these data indicate that moose in old crow flats use and select drained lake basins, but also that use and selection varies individually and seasonally. a previous resource selection function analysis of these data indicated that, within the home range scale, moose in old crow flats are more likely to use areas closer to water with a higher proportion of upright shrubs, higher diversity of vegetation types in the vicinity, and drained lake basins if present (clarke et al. 2017). home range size was much smaller in summer than winter, suggesting that availability of high quality forage in summer reduces space migratory moose in yukon – cooley et al. alces vol. 55, 2019 122 use more so than deep snow in winter (phillips et al. 1973, timmermann and mcnicol 1988, van beest et al. 2011). overall, the shrub-dominated diet supplemented by aquatic vegetation in summer conforms with the typical year-round diet of most moose populations (timmermann and mcnicol 1988). the moose population that summers in old crow flats breaks into 3 largely distinct subpopulations during winter. one subpopulation summers on the west and north sides of old crow flats and migrates mainly northwest to higher elevations (600–900 m), where they join moose from the upper coleen drainage in alaska, the firth river drainage in canada and alaska, and as far as the kongakut river in alaska (mauer 1998). a second subpopulation summers on the west and south sides of old crow flats and migrates mainly west to moderately higher elevations (400–500 m) within the lower drainages of the coleen river in alaska. the migration of these 2 subpopulations was described by mauer (1998) who noted that the westward migrating subpopulation occasionally reaches the upper reaches of the sheenjek river drainage 50 km west of the coleen river drainage. a third subpopulation using the eastern portion of old crow flats in summer either remains resident year-round locally in little flat where forested habitat comprises 26% of land cover compared to 12% elsewhere in old crow flats, or migrates southeasterly towards low elevation winter ranges along the porcupine, driftwood, and bell rivers. if nonmigratory moose are considered a distinct subpopulation, then the summer population in old crow flats consists of 4 subpopulations – 3 migratory and 1 resident. moose at high elevations in winter, especially the northwest subpopulation, selected for shrub cover close to rivers and streams within long and narrow home ranges situated at the bottom of alpine valleys. moose at low elevations in winter, especially the southeast subpopulation, occupied variably shaped home ranges that tended to be close to streams or rivers and composed of more forested and shrub habitat than surrounding areas. stable isotope analysis indicated that moose in alpine riparian habitats had δ15n signatures indicative of greater use of low enrichment shrubs like feltleaf willow and dwarf birch than moose occupying home ranges predominated by forest habitat with δ15n signatures indicating higher use of tealeaf willow. these results are consistent with previous research with alaska-yukon moose indicating greater use and preference of higher-growing, more nutritious feltleaf willow than lower-growing, less nutritious tealeaf willow, except in forested habitats where tealeaf willow can be the only abundant shrub (risenhoover 1989). our findings are consistent with mauer’s (1998) speculation that seasonal migration from old crow flats to surrounding higher elevations allows moose to avoid the coldest winter temperatures at low elevations where average snowpack may also be deeper and winds more substantial. we lack the data on elevationand land cover-specific snow accumulations across the study area to directly evaluate the possibility of snow avoidance. the frequency and magnitude of inversions detected from tc aligned well with patterns detected by meteorological studies using weather balloons (bourne et al. 2010). on most days, the tc for moose at 1000 m was 5–15°c warmer than that for moose at 350 m. the lower critical temperature of moose in alberta is estimated as about −35°c (renecker and hudson 1986), and large body size and cold acclimatization likely provides increased thermoregulatory ability for this alaska-yukon subspecies. we found no evidence that moose moved more on the coldest days or responded to the alces vol. 55, 2019 migratory moose in yukon – cooley et al. 123 dynamic nature of inversions by moving up when temperatures were warmer at high elevations or down when warmer at lower elevation. furthermore, several moose remained at low elevation throughout winter, exposing themselves to the consistently colder conditions there. even in the extreme winter climate that prevails in this subarctic environment, thermoregulation is likely a minor cost of the daily energy budget of moose in all but the coldest conditions. although we focus here on snow and temperature as an explanation for seasonal moose migration, spatial variation in predation risk may be an additional factor influencing migration and habitat selection in this region. for example, the risk of predation by wolves and bears may be minimal in old crow flats during the open water season, whereas the risk of wolf predation in winter may be lessened by occupying isolated alpine valleys. speculation regarding seasonand habitat-specific predation risk warrants further examination. vuntut gwitchin knowledge of moose in old crow flats predicted our habitat selection and dietary results, including inter annual home range fidelity in summer, habitat preference for locations in and around drained lake basins and close to other water bodies, and dietary reliance on shrubs. other research in old crow flats also confirms local observations regarding recent climate warming (porter and pisaric 2011), increased frequency of lake drainages in this landscape (lantz and turner 2015), shrub growth trajectories associated with these drainages (lantz 2017), and permafrost degradation particularly along exposed shorelines (turner et al. 2014, roy-léveillée and burn 2016). while an assessment of climate change impacts on moose habitat in old crow flats was beyond the scope of this study, our results suggest that moose benefit from proliferation of shrubs in drained lake basins that are a major contemporary feature of the old crow flats. further, our results reinforce the concept that the continuity of a habitat mosaic within old crow flats is important in providing moose spatial proximity to water and access to a diversity of vegetation, including seasonally high use of aquatic vegetation. from a habitat protection and conservation perspective, the core habitat and winter ranges of the 3 subpopulations of the old crow flats moose population is well protected. the old crow flats wetland complex is protected under vuntut gwitchin first nation final agreement category a lands, vuntut national park, and the old crow flats special management plan, with lands “to be protected and managed in a manner that permanently protects the ecological integrity of the flats, including its diversity, its fish and wildlife populations and its habitats from activities that could reduce the land’s capability, while maintaining access to this area by vuntut gwitchin citizens for traditional and current harvesting of fish and wildlife resources.” (twgmc 2006: 38). the entirety of the migratory routes and winter ranges of the northwest and west subpopulations is protected in canada by vuntut national park and ivvavik national park and in alaska by the arctic national wildlife refuge. although the migratory routes and winter range of the southeast population are largely outside of category a lands or the old crow flats special management unit (except for the driftwood river), they are designated as low development zones within the north yukon regional land use plan, indicative of locations with “high ecological and heritage/cultural values” where “maintaining ecological integrity, protecting heritage and cultural resources, and minimizing land use impacts is the priority.” (north yukon planning commission 2009). migratory moose in yukon – cooley et al. alces vol. 55, 2019 124 in terms of moose harvest, vuntut gwitchin have subsistence harvest rights across nearly all of this population’s canadian distribution including vuntut national park and (contingent on the inuvialuit harvester agreement) ivvavik national park. the southeastern subpopulation is subject to licensed harvest in the yukon, and subsistence and recreational harvest is permitted within the arctic national wildlife refuge in alaska. however, given the remoteness and seasonal movement patterns of this population, hunting pressure and harvest are limited. summer and autumn boat travel to old crow flats is often prevented by low water levels in the old crow river, and by the time hunters can reach old crow flats by snow and ice, most moose have migrated to winter ranges where they remain until travel by snow and ice is impossible; hunters do have winter access to moose that remain resident in the southeast. furthermore, the winter ranges of the northwestern and western subpopulations are largely outside of canada, and too distant and mountainous for easy access from the community of old crow. however, the southeast subpopulation is likely accessible to boat-based hunters in autumn especially along the porcupine river, and in winter throughout much of its winter range. any recreational and subsistence moose harvet occurring in the eastern side of the arctic national wildlife refuge, whether in autumn or winter, likely includes moose summering in old crow flats. accordingly, monitoring population and harvest trends in yukon, particularly around the porcupine river and the southeastern portion of old crow flats, and in alaska within the arctic national wildlife refuge, is important for the conservation status and management of this population. the ynnk old crow ipy project offers a model of community leadership in research, including how indigenous knowledge and scientific research can be combined to identify knowledge needs, then broaden ecological knowledge of locally important landscapes and wildlife populations (wolfe et al. 2011, brunet et al. 2014). the gwitchin project title, yeendoo nanh nakhweenjit k’atr’ahanahtyaa translates, roughly, as “taking care of the land for the future.” the local knowledge and research described here reveals the enormity of the land base supporting a single moose population, how extreme seasonality in climate manifests as extreme seasonality in behaviour and habitat use, and the resulting complexity of climate change impacts on moose and moose habitat at the periphery of the species range. the uniqueness of the landscape of old crow flats and surrounding uplands, the migratory moose population that resides there, the vuntut gwitchin’s reliance on this landscape and its wildlife resources, and the magnitude of locally observed and anticipated climate change impacts combine to motivate continued international monitoring and conservation of these vital landscapes, natural resources, and human-nature relationships in a collaborative “taking care of the land for the future.” acknowledgements we thank the people of old crow, especially the elders and hunters, for allowing us to learn together about their lands and wildlife, for hosting us in their community, for their permission to deploy collars on moose, and for keeping us safe on their lands. w. josie, d. matthiesson, s. foss, b. bell, and m. williams provided critical assistance and insight in framing the initial collaboration and guiding it from the idea stage through fieldwork. r. ward provided useful insights into moose ecology, helped with the project design, and generously loaned us collars on short notice. m. oakley and m. kienzler ably assisted with moose alces vol. 55, 2019 migratory moose in yukon – cooley et al. 125 captures, collar deployments, and recoveries. the many researchers and graduate students involved in other components of the old crow ipy project were critical to the success of the overall project and helped us to situate moose ecology within the physical and biological landscape of old crow flats. finally, referees and editors provided helpful comments on the manuscript. references allenby, r. j. 1989. clustered, rectangular lakes of the canadian old crow basin. tectonophysics 170: 43–56. doi:10.1016/ 0040-1951(89)90102-9 ayliffe, l., t. e. cerling, t. robinson, a. wes, m. sponheimer, b. passey, j. hammer, b. roeder, m. -d. dearing, and j. r. ehleringer. 2004. turnover of carbon isotopes in tail hair and breath co2 of horses fed an isotopically varied diet. oecologia 139: 11–22. doi:10.1007/ s00442-003-1479-x balasubramaniam, a. m., r. i. hall, b. b. wolfe, j. n. sweetman, and x. wang. 2015. source water inputs and catchment characteristics regulate limnological conditions of shallow subarctic lakes (old crow flats, yukon, canada). canadian journal of fisheries and aquatic sciences 72: 1058–1072. doi:10.1139/ cjfas-2014-0340 ball, j. p., g. ericsson, and k. wallin. 1999. climate changes, moose and their human predators. ecological bulletins 47: 178–187. _____, c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose alces alces ranges in northern sweden. wildlife biology 7: 39–48. doi:10.2981/ wlb.2001.007 bjørneraas, k., i. herfindal, e. j. solberg, b. e. saether, b. van moorter, and c. m. rolandsen. 2012. habitat quality influences population distribution, individual space use and functional responses in habitat selection by a large herbivore. oecologia 168: 231–243. doi:10.1007/ s00442-011-2072-3 bourne, s. m., u. s. bhatt, j. zhang, and r. thoman. 2010. surface-based temperature inversions in alaska from a climate perspective. atmospheric research 95: 353–366. doi:10.1016/j.atmosres.2009. 09.013 boyd, c., a. punt, h. weimerskirch, and s. bertrand. 2014. movement models provide insights into variation in the foraging effort of central place foragers. ecological modelling 286: 13–25. doi:10.1016/j.ecolmodel.2014.03.015 bradley, r. s., f. t. keimig, and h. f. diaz. 1992. climatology of surface-based inversions in the north american arctic. journal of geophysical research: a t m o s p h e r e s 9 7 : 1 5 6 9 9 – 1 5 7 1 2 . doi:10.1029/92jd01451 brunet, n., g. hickey, and m. m. humphries. 2014. understanding community researcher partnerships in the natural sciences: a case study from the arctic. journal of rural studies 36: 247–261. doi:10.1016/j.jrurstud.2014.09.001 bunnefeld, n., l. borger, b. van moorter, c. m. rolandsen, h. dettki, e. j. solberg, and g. ericsson. 2011. a model-driven approach to quantify migration patterns: individual, regional and yearly differences. journal of animal ecology 80: 466–476. doi:10.1111/j.1365 -2656.2010.01776.x calenge, c. 2006. the package “adehabitat” for the r software: a tool for the analysis of space and habitat use by animals. ecological modelling 197: 516–519. doi:10.1016/j.ecolmodel.2006.03.017 clarke, h., d. cooley, m. m. humphries, m. landry-cuerrier, and t. lantz. 2017. summer habitat selection by moose on the old crow flats. yukon fish and wildlife branch report mr 17-01. whitehorse, yukon, canada. darimont, c. t., and t. e. reimchen. 2002. intra-hair stable isotope analysis implies migratory moose in yukon – cooley et al. alces vol. 55, 2019 126 seasonal shift to salmon in gray wolf diet. canadian journal of zoology 80: 1638–1642. doi:10.1139/z02-149 decesare, n. j., j. r. newby, and j. m. ramsey. 2015. a review of parasites and disease impacting moose in north america. intermountain journal of sciences 21: 62. demarchi, m. w. 2003. migratory movements and home range size of moose in the central nass valley, british columbia. northwestern naturalist 84: 135–141. doi:10.2307/3536539 franzmann, a. w., and c. c. schwartz. 1997. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. gannes, l. z., d. m. o’brien, and c. martinez del rio. 1997. stable isotopes in animal ecology: assumptions, caveats, and a call for more laboratory experiments. ecology 78: 1271–1276. d o i : 1 0 . 1 8 9 0 / 0 0 1 2 9 6 5 8 ( 1 9 9 7 ) 0 7 8 [1271:siiaea]2.0.co;2 gasaway, w. c., r. o. stephenson, j. l. davis, p. e. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84: 1–50. goddard, j. 1970. movements of moose in a heavily hunted area of ontario. the journal of wildlife management 34: 439–445. doi:10.2307/3799030 hauge, t. m., and l. b. keith. 1981. dynamics of moose populations in northeastern alberta. the journal of wildlife management 45: 573–597. doi:10.2307/3808692 hayes, r., and n. barichello. 1986. wolf, moose, muskoxen and grizzly bear observations on the yukon north slope. yukon department of environment, whitehorse, yukon, canada. houston, d. b. 1968. the shiras moose in jackson hole, wyoming. technical bulletin number 1. grand teton natural history association, moose, wyoming, usa. jenkins, k. j., and r. g. wright. 1987. dietary niche relationships among cervids relative to winter snowpack in northwestern montana. canadian journal of zoology 65: 1397–1401. doi:10.1139/z87-220 johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61: 65–71. doi:10.2307/1937156 jones, h., p. j. pekins, l. e. kantar, d. ellingwood, i. sidor, a. lichtenwalner, and m. o’neil. 2019. mortality assessment of calf moose during successive years of winter tick epizootics in new hampshire and maine. canadian journal of zoology 97: 22–30. doi:10.1139/ cjz-2018-0140 jung, t. s., t. e. chubbs, c. g. jones, f. r. phillips, and r. d. otto. 2009. winter habitat associations of a low-density moose (alces americanus) population in central labrador. northeastern naturalist 16: 471–480. doi:10.1656/045.016.n313 lantz, t. c. 2017. vegetation succession and environmental conditions following catastrophic lake drainage in old crow flats, yukon. arctic 70: 177–189. doi:10.14430/arctic4646 _____, and k. turner. 2015. changes in lake area in response to thermokarst processes and climate in old crow flats, yukon. journal of geophysical research: biogeosciences 120: 513–524. doi:10.1002/2014jg002744 lauriol, b., d. lacele, s. labreque, c. r. duguay, and a. telka. 2009. holocene evolution of lakes in the bluefish basin, northern yukon, canada. arctic 6: 212–224. doi:10.14430/arctic133 lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. the journal of wildlife management 73: 503–510. doi:10.2193/ 2008-265 leresche, r. 1974. moose migrations in north america. naturaliste canadien 101: 393–415. alces vol. 55, 2019 migratory moose in yukon – cooley et al. 127 loisa, k., and e. pulliainen. 1968. winter food and movements of two moose (alces alces l.) in northeastern finland. annales zoologici fennici 5: 220–223. manly, b., l. mcdonald, d. l. thomas, t. l. mcdonald, and w. p. erickson. 2007. resource selection by animals: statistical design and analysis for field studies. springer science & business media, new york, new york, usa. mauer, f. 1998. moose migration: northeastern alaska to northwestern yukon territory, canada. alces 34: 75–81. mccutchan, j. h., w. m. lewis, c. kendall, and c. c. mcgrath. 2003. variation in trophic shift for stable isotope ratios of carbon, nitrogen, and sulfur. oikos 102: 378–390. doi:10.1034/j.1600-0706. 2003. 12098.x milligan, h. e. 2010. seasonal and spatial variability of aquatic and terrestrial feeding by moose in old crow flats, yukon. special report. environment yukon, whitehorse, yukon, canada. milligan, h. e., t. d. pretzlaw, and m. m. humphries. 2010. stable isotope differentiation of freshwater and terrestrial vascular plants in two subarctic regions. ecoscience 17: 265–275. doi:10.2980/ 17-3-3282 monteith, k. l., r. w. klaver, k. r. hersey, a. a. holland, t. p. thomas, and m. j. kauffman. 2015. effects of climate and plant phenology on recruitment of moose at the southern extent of their range. oecologia 178: 1137–1148. doi:10.1007/s00442-015-3296-4 mossop, d. h. 1975. moose counts – old crow 1974 and 1975. environment yukon, whitehorse, yukon, canada. _____. 2015. the changing ecology of the old crow flats. yukon research centre, yukon college, whitehorse, yukon, canada. muggeo, v. m. r. 2017. regression models with break-points / change-points estimation. r package version 37. nation, v. g. f., and s. smith. 2010. people of the lakes: stories of our van tat gwich’in elders/googwandak nakhwach’anjoo van tat gwich’in. university of alberta, edmonton, alberta, canada. north yukon planning commission. 2009. north yukon regional land use plan. whitehorse, yukon, canada. http:// www.emr.gov.yk.ca/rlup/north-yukonregional-land-use-plan.html (accessed september 2019). ovenden, l. 1986. vegetation colonizing the bed of a recently drained thermokarst lake (illisarvik), northwest territories. canadian journal of botany 64: 2688–2692. doi:10.1139/b86-354 phillips, d. l., and j. w. gregg. 2003. source partitioning using stable isotopes: coping with too many sources. oecologia 136: 261–269. doi:10.1007/ s00442-003-1218-3 phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. the journal of wildlife management 37: 266–278. doi:10.2307/ 3800117 porter, t. j., and m. f. pisaric. 2011. temperature-growth divergence in white spruce forests of old crow flats, yukon territory, and adjacent regions of northwestern north america. global change biology 17: 3418– 3430. doi:10.1111/j.1365-2486.2011. 02507.x pulliainen, e. 1974. seasonal movements of moose in europe. le naturaliste canadien 101: 379–392. qgis development team. 2016. qgis geographic information system. open source geospatial foundation project. http://qgis.osgeo.org (accessed september 2019). r core team. 2016. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. https://www.r-project. org/ (accessed september 2019). http://www.emr.gov.yk.ca/rlup/north-yukon-regional-land-use-plan.html http://www.emr.gov.yk.ca/rlup/north-yukon-regional-land-use-plan.html http://www.emr.gov.yk.ca/rlup/north-yukon-regional-land-use-plan.html http://qgis.osgeo.org https://www.r-project.org/ https://www.r-project.org/ migratory moose in yukon – cooley et al. alces vol. 55, 2019 128 ramsar convention. 2004. the list of wetlands of international importance. ramsar secretariat, gland, switzerland. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. doi:10.1139/z86-052 _____, and c. c. schwartz. 1998. nutrition and energetics. pages 403–440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution, washington, dc, usa. risenhoover, k. l. 1989. composition and quality of moose winter diets in interior alaska. the journal of wildlife management 53: 568–577. doi:10.2307/ 3809178 rolandsen, c. m., e. j. solberg, b. e. saether, b. v. moorter, i. herfindal, and k. bjorneraas. 2017. on fitness and partial migration in a large herbivore–migratory moose have higher reproductive performance than residents. oikos 126: 547–555. doi:10.1111/ oik.02996 roy-léveillée, p., and c. burn. 2010. permafrost conditions near shorelines of oriented lakes in old crow flats, yukon territory. pages 1509–1516 in conference proceedings of geo, 12–16 september, 2010 calgary, alberta, canada. _____, and _____. 2016. a modified landform development model for the topography of drained thermokarst lake basins in fine-grained sediments. earth surfaces processes and landforms 41: 1504–1520. doi:10.1002/esp.3918 _____, _____, and i. d. mcdonald. 2014. vegetation-permafrost relations within the treeline ecotone near old crow, northern yukon, canada. permafrost and periglacial processes 25: 127–135. doi:10.1002/ppp.1805 russell, d., d. mossop, and c. goodfellow. 1978. remote sensing for waterfowl nesting and nesting habitat in old crow flats, yukon territory, canada. pecora iv–conference, sioux falls, south dakota, usa. schuster, r. c., e. e. wein, c. dickson, and h. m. chan. 2011. importance of traditional foods for the food security of two first nations communities in the yukon, canada. international journal of circumpolar health 70: 286–300. doi:10.3402/ijch.v70i3.17833 schwertl, m., k. auerswald, r. schaufele, and h. schnyder. 2005. carbon and nitrogen stable isotope composition of cattle hair: ecological fingerprints of production systems? agriculture, ecosystems & environment 109: 153–165. doi:10.1016/j. agee.2005.01.015 singh, n. j., a. m. allen, and g. ericsson. 2016. quantifying migration behaviour using net squared displacement approach: clarifications and caveats. plos one 11: e0149594. doi:10.1371/journal.pone. 0149594 smith, c., j. meikle, and c. roots. 2004. ecoregions of the yukon territory: biophysical properties of yukon landscapes. parc technical bulletin no. 04-01. agriculture and agri-food canada, summerland, british columbia, canada. sponheimer, m., t. robinson, l. ayliffe, b. passey, b. roeder, l. shipley, e. lopez, t. cerling, d. dearing, and j. ehleringer. 2003a. an experimental study of carbon-isotope fractionation between diet, hair, and feces of mammalian herbivores. canadian journal of zoology 81: 871– 876. doi:10.1139/z03-066 _____, _____, _____, b. roeder, j. hammer, b. passey, a. west, t. cerling, d. dearing, and j. ehleringer. 2003b. nitrogen isotopes in mammalian herbivores: hair δ15n values from a controlled feeding study. international journal of osteoarchaeology 13: 80–87. doi:10.1002/oa.655 tape, k. d., d. d. gustine, r. w. ruess, l. g. adams, and j. a. clark. 2016. range expansion of moose in arctic alces vol. 55, 2019 migratory moose in yukon – cooley et al. 129 alaska linked to warming and increased shrub habitat. plos one 11: e0152636. doi:10.1371/journal.pone.0152636 telfer, e. s. 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145–182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. _____, and j. p. kelsall. 1984. adaptation of some large north american mammals for survival in snow. ecology 65: 1828– 1834. doi:10.2307/1937779 thompson, i. d., and m. f. vukelich. 1981. use of logged habitats in winter by moose cows with calves in northeastern ontario. canadian journal of zoology 59: 2103–2114. doi:10.1139/z81-287 timmermann, h. r., and j. g. mcnicol. 1988. moose habitat needs. the forestry chronicle 64: 238–245. doi:10.5558/ tfc64238-3 tischler, k. b. 2004. aquatic plant nutritional quality and contribution to moose diet at isle royale national park. m. s. thesis. michigan technologial institute, houghton, mi, usa. turner, k. w., b. b. wolfe, and t. w. d. edwards. 2010. characterizing the role of hydrological processes on lake water balances in the old crow flats, yukon territory, canada, using water isotope tracers. journal of hydrology 386: 103– 117. doi:10.1016/j.jhydrol.2010.03.012 _____, _____, _____, t. c. lantz, r. i. hall, and g. larocque. 2014. controls on water balance of shallow thermokarst lakes and their relations with catchment characteristics: a multi-year, landscape scale assessment based on water isotope tracers and remote sensing in old crow flats, yukon (canada). global change biology 20: 1585–1603. doi:10.1111/ gcb.12465 twgmc (technical working group and management committee). 2006. old crow flats van tat k’atr’anahtii special management area management plan. vuntut gwitchin government and yukon environment, old crow, yukon, canada. http://www.env.gov.yk.ca/publications-maps/documents/old_crow_ flats.pdf (accessed september 2019). van ballenberghe, v. 1977. migratory behavior of moose in southcentral alaska. transactions of the 13th international congress of game biologists 13: 103–109. _____, and j. m. peek. 1971. radiotelemetry studies of moose in northeastern minnesota. journal of wildlife management 35: 63–71. doi:10.2307/ 3799872 van beest, f. m., i. m. rivrud, l. e. loe, j. m. milner, and a. mysterud. 2011. what determines variation in home range size across spatiotemporal scales in a large browsing herbivore? journal of animal ecology 80: 771–785. doi:10.1111/j.1365-2656.2011.01829.x _____, b. van moorter, and j. m. milner. 2012. temperature-mediated habitat use and selection by a heat-sensitive northern ungulate. animal behaviour 84: 723–735. doi:10.1016/j.anbehav.2012. 06.032 vanderklift, m. a., and s. ponsard. 2003. sources of variation in consumer-diet δ 15 n enrichment: a meta-analysis. oecologia136: 169–182. doi:10.1007/ s00442-003-1270-z wald, e. j., and r. m. nielson. 2014. estimating moose abundance in linear subarctic habitats in low snow conditions with distance sampling and a kernel estimator. alces 50: 133–158. wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. white, k. s., n. l. barten, s. crouse, and j. crouse. 2014. benefits of migration in relation to nutritional condition and predation risk in a partially migratory moose population. ecology 95: 225–237. doi:10.1890/13-0054.1 http://www.env.gov.yk.ca/publications-maps/documents/old_crow_flats.pdf http://www.env.gov.yk.ca/publications-maps/documents/old_crow_flats.pdf http://www.env.gov.yk.ca/publications-maps/documents/old_crow_flats.pdf migratory moose in yukon – cooley et al. alces vol. 55, 2019 130 wolfe, b. b., m. m. humphries, m. f. pisaric, a. m. balasubramaniam, c. r. burn, l. chan, d. cooley, d. g. froese, s. graupe, and r. i. hall. 2011. environmental change and traditional use of the old crow flats in northern canada: an ipy opportunity to meet the challenges of the new northern research paradigm. arctic 64: 127. doi:10.14430/arctic4092 _____, and k. w. turner. 2008. near-record precipitation causes rapid drainage of zelma lake, old crow flats, northern yukon territory. meridian spring edition: 7–12. wood, s. n. 2017. mixed gam computation vehicle with automatic smoothness estimation. r package version 291. 4304.pdf alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose 49 diagnosing parelaphostrongylosis in moose (alces alces) murray lankester1,2, wm. peterson3, and oladele ogunremi4,5 1retired, professor emeritus, lakehead university, 955 oliver road, thunder bay, on canada p7b 5e1; 3retired, minnesota department of natural resources, 4541 lake creek road, troy, mt 59935, usa; 4centre for animal parasitology, canadian food inspection agency, saskatoon laboratory, 116 veterinary road, saskatoon, sk, canada s7n 2r3 abstract: thirty-six moose (alces alces) reported as acting abnormally were examined in northwestern ontario and adjacent northeastern minnesota in 1986 – 2000. thirty-four typically had little fear of humans, remained in an area for some time, and showed clinical signs of neuromotor incoparelaphostrongylus tenuis p. tenuis typical neurological signs was found in 14 more, but examination was impossible or incomplete for 9 p. tenuis for p. tenuis the sample (21/34) and 10 were judged underweight. the remaining 2 moose in the sample, although p. tenuis. moose with adult p. tenuis (u = 20, p the p. tenuis p. tenuis exposure experienced by moose populations sharing range with infected white-tailed deer. key words: alces, meningeal worm, moose disease, moose sickness, parelaphostrongylosis, parelaphostrongylus tenuis parelaphostrongylosis is a disease in moose (alces alces) and other ungulates caused by a neurotropic nematode, parelaphostrongylus tenuis, spread by white-tailed deer (odocoileus virginianus) which is the ester 2001). in extreme cases, infected moose may show a pronounced posterior weakness fected may lack fear of humans or show only slight, transitory signs such as unsteady gait experimental work (lankester 2002), the 2present address: 101-2001 blue jay place, courtney, bc, canada v9n 3z6 road, ottawa, on, canada k2h 8p9 diagnosing parelaphostrongylosis in moose – lankester et al. alces vol. 43, 2007 parasite burden, the age of the infection, and possibly the host’s immunological familiarity with the parasite. in the wild, only the be reported. yet, because of their large size, careful post-mortem examination of animals showing signs is a daunting task and may be typical suite of recognizable neuromotor signs accurately predicts p. tenuis infection in moose has not been examined thoroughly. losis has been known for many years but important aspects of its pathogenesis and impact on moose populations remain unclear. an early experiment demonstrated that when to calf moose, p. tenuis can cause a rapidly derson 1964). as well, naturally infected in the cranium, making it tempting to think that moose were particularly susceptible. the only where they were almost totally isolated study of contemporary moose populations the impact of parelaphostrongylosis on moose is likely more subtle and complex (whitlaw and lankester 1994a,b). moose currently persist in many areas of eastern north america where deer densities (whitlaw and lankester 1994b). when deer 2, p. tenuis may cause only low, and marginally limiting mortality (karns 1967, lenarz and kerr 1987, whitlaw and lankester 1994a, dumont and p. tenuis marked moose population declines (whitlaw and lankester 1994a) and could again with numbers. the current and future impact of parelaphostrongylosis on moose populations may be underestimated because of our limited exposed animals. this paper describes the clinical manifesperiod in northeastern minnesota and northprocedures and tools used to determine which had parelaphostrongylosis. methods by the authors (murray lankester and wm. of 14 years (1986 – 2000). a description of taken on occasion. when possible, the age of animals was estimated by tooth eruption and wear. on the basis of their size alone, 7 animals whose teeth could not be examined were simply categorized as adults (i.e., conducted and the head, feces, and serum (in 4 instances) collected and frozen until examsurface inspected using a stereomicroscope at 16x. within the cranium, the inner surface of the dura was examined for worms and all searched for adult p. tenuis using the method and decanted, and the sediment examined for initially using the classical baermann funnel technique and in later years (after 1994), the rester and lankester 1997). numbers of alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose no. date examined location1 sex age (years) abnormal signs p. tenuis meningitis2 feces 13 9/22/1986 mn yes ? ? 0 2 12/7/1986 mn yes 0 ? ? 3 mn m 1 yes 0 ? 0 4 9/24/1987 mn yes ? 0 11/17/1988 mn 3+ yes yes 0 6 12/26/1988 m 0.6 yes ? 2.8/g 7 1/10/1989 tb a yes 0 yes ? 8 2/7/1989 mn yes 0 yes 0 9 6/9/1989 tb m 1+ yes yes 2.2/g 10 2/3/1991 a yes yes 0.3/g 114 3/26/1991 m 2+ yes ? ? 0 12 mn m 1 yes ? 0 13 6/18/1991 mn m yes 0 yes 0 14 12/2/1991 tb a yes 0 yes 0 2/13/1992 8+ yes 0 ? 0 16 2/17/1992 mn m a yes 1m yes 0 174 3/16/1992 mn ? a yes ? ? 0 18 3/19/1992 mn m 0.8 dead yes 0 19 9/7/1992 mn yes yes 0 20 10/24/1992 mn yes 0 no 0 21 12/3/1992 3+ yes 0 no 0 22 12/9/1992 a yes 0 ? 0 236 3/19/1993 mn 2 atypical 0 no 0 243 1/1/1994 mn m 0.6 yes ? ? 0 4/18/1994 ke 1+ yes yes 0.2/g 26 ??/06/94 sgpk a yes 0 no ? 276 6/18/1994 mn 4 atypical 0 no ? 28 7/4/1994 mn 14 yes 0 (cord only) yes 0 29 7/7/1994 1+ yes yes 0 (cranium+) minnesota, 1986-2000. 1 2 3 4 lens or cornea of 1 or both eyes opaque. 6 7 diagnosing parelaphostrongylosis in moose – lankester et al. alces vol. 43, 2007 fresh fecal material. an enzyme-linked immunosorbent assay (elisa) was conducted on serum from 4 clinically abnormal moose and on 4 control moose from an area without p. tenuis. the method was similar to that and natural parelaphostrongylosis in moose (ogunremi et al. 2002a) and white-tailed an anti-igg conjugate produced according to ogunremi et al. (2002a). mean ages and mean number of adult worms found include statistical analysis of mean ages was carried out with the mann-whitney (u) test. results marais, minnesota (n ontario (n ontario (n = 10) in 1986 – 2000 (table l). animals (#23 and 27) were described as being otherwise showed normal gait and coordinaand 10 of the 36 were judged to be underweight or undersize for their age. the mean age of 29 animals for which tooth-age estimates reported in all months of the year including the remaining seasons. moose included 1 or more of the following: lack of fear upon human approach (13 moose), remaining in an area for an extended period (13, including the 1 found dead), walking or swimming in circles (8), inability to stand (incoordination) (7), partial paralysis manitilted to one side (4), head and neck turned posteriorly (torticollis) (6), standing with legs no. date examined location1 sex age (years) abnormal signs p. tenuis meningitis2 feces 30 8/3/1994 mn 4 yes yes 0 313 8/20/1994 ke m 1+ yes ? ? 0.12/g 32 10/19/1994 sgpk m yes yes 0 337 mn m 0.9 yes yes 0 (cranium+) 34 ke yes 0 no 0 7 6/14/1999 tb 1 yes yes 0.7/g 36 1/6/2000 mn yes 0 no 0 northeastern minnesota, 1986-2000. 1 2 3 4 lens or cornea of 1 or both eyes opaque. 6 7 alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose positioned wide apart (wide base stance) or twitching (nystagmus) (3), unsteady gait (2), knuckling of lower limb joints or stumbling of these signs were not collected for further examination. of 13 animals described as remaining in the same area for extended periods, the initial sighting while one was reportedly present in the area for up to 10 months. adult p. tenuis were found in the cranium parelaphostrongylosis. an additional animal its feces identical to those of p. tenuis. the mean age of 14 animals with adult p. tenuis (u = 20, p = 0.006). worms in the cranium were most frequently located on the surface of the brain or 3 animals, portions of worms penetrated into brain tissue. worms in 2 moose were located the mean number of adult p. tenuis found moose with adult worms in the cranium were but none in feces. all moose with adult p. tenuis in the cranium (except 3 unsuitable for detailed and across the surface of the pia-arachnoid the piarachnoid had a whitish cloudy appearance and loose patches of yellowish exudate were often seen on the brain surface. an cloudy meningitis but neither adult worms in one or both eyes (table 1). all heads had been frozen and thawed before examination. histologically, the eyes of #36 were considered wildlife health centre, saskatoon), despite the lens of one eye appearing enlarged and opaque. the eyes of the remaining 4 moose were not examined histologically. had neurogical signs typical of p. tenuis infection except #23 that was killed because it was p. tenuis exposure. the highest titre (1,140 units occurred in a 10-month-old calf (#33) with 1 adult worm in the cranium, but a lower titre (140). its serum sample was contaminated with ingesta (titre < 10 units). discussion thirty-four of the animals examined here showed a reasonably consistent set of clinical signs not unlike those described by some of the earliest workers studying moose sickness (thomas and cahn 1932, lamson 1941, stage of infection (anderson 1964, lankester 2002). animals recently infected with only can appear lethargic, walk with occasional diagnosing parelaphostrongylosis in moose – lankester et al. alces vol. 43, 2007 knuckling, stumbling, or an unsteady gait. unaffected. other animals, presumably with manifestations of p. tenuis infection. signs include marked paresis or weakness of the hind-quarters, wide-base stance needed to maintain balance, circling, head tilted or turned ing, or being unable to stand. total paralysis nystagmus, or twitching of the eyes, often seen suggest a balance disorder. animals showing nial nematode infections, for example, moose in sweden with elaphostrongylus alces and caribou in newfoundland with e. rangiferi cal of parelaphostrongylosis. interestingly, an p. tenuis. on the basis of typical neuromotor signs, this diagnosis could be corroborated in only 22 p. tenuis p. tenuis no head examined), and no worms or cloudy remaining 12 moose with typical neuromotor infection on examination (3) or examination was incomplete (9). these results are comparable to those of smith and archibald (1967) and gilbert (1974) who both found worms in it is impossible to determine whether their able p. tenuis. losis in moose of north america is possible when long slender nematodes can be located within the cranium. male worms can be p. tenuis is expected in that location (e. rangiferi in moose of newfoundland is an exception). but conducting a reliable cranial examination requires appropriate necropsy tools and facilities and some experience on the part of the muscle worm, p. andersoni, also becomes moose of the study area since it is not known in deer of northeastern minnesota (peterson and p. tenuis can now be distinguished from those of p. andersoni (and other related species) using molecular diagnostic methods (pcr-polymerase chain reaction and sscpsingle-strand conformation polymorphism) 2006). accumulation of lymphocytes, plasma cells and eosinophils is a known feature of p. tenuis infection (anderson 1964, smith et al. 1964, smith and archibald 1967, lankester 1974) examination. thus, brain tissue must be in in formalin or quickly frozen. animals found dead in the warmer months, or heads left for periods in the sun before freezing, will be unto yellowish) of the meninges, especially in the pia-arachnoid membrane against the brain alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose the presence of loose, yellowish-red accumulain the subdural space, on top of, or beneath the brain. meningitis is most easily appreciated and thawing and any degree of post-mortem change decreases the likelihood of recognizing it. meningitis resulting in cloudiness of the examined moose with p. tenuis in the cranium mal moose in which no worms were located. either worms were missed in the latter group of killing worms in the cranium and abnormal worms is gone (lankester 2002, ogunremi an eosinophilic meningoencephalitis along the diagnosis of parelaphostrongylosis in the p. tenuis (ogunremi et al. 2002a). used here, it correctly present in the cranium. of equal interest was of p. tenuis could be found. without the elisa p. tenuis diagnosis could only of typical clinical signs. although atypical, of this study, infections in more clinically the p. tenuis sk (contact: gail.krohn@pds.usask.ca). proximately 1 ml) is required for the elisa. although serum ideally is separated from be directed at the chest to cause bleeding into examination). vehicle-killed animals may between the ribs on the “down side” of the animal will release the pooled blood that can be drained into a container (plastic bag or jam sawn open or accessed through the diaphragm be scooped into a container, although samples badly contaminated with rumen material are room temperature for a few hours allowing a clot to form and then chilled in a refrigerator pipetting the straw-coloured serum into small by red blood cells, the serum will be red and “bloody” and the contaminating red blood serum or allowed to settle and the watery, occur during sample collection and preparation in which case the serum sample will be red or reddish. as long as such samples were promptly frozen after preparation, testing can still be successfully carried out. whole blood put directly into the freezer may not be used in the elisa. results reported here contribute to our understanding of p. tenuis infection rates in feces was used by lenarz and kerr (1987) to moose were infected with p. tenuis. their in minnesota by karns (1977) who found diagnosing parelaphostrongylosis in moose – lankester et al. alces vol. 43, 2007 kerr (1987) arbitrarily chose a more conserinfected moose. in our study, determining the all moose exhibiting typical neuromuscular is not appreciably different from that chosen moose with less conspicuous, or sub-clinical signs, were also included, the larger sample denominator would further lower the expected and thereby increase the predicted rates of p. tenuis infection in the population. in addition, it is important to recognize that these are “point in time” estimates that ignore the may be short but ongoing. this could further increase estimates of annual morbidity and mortality. samples because of the habit of sick animals staying for long periods in one area. thirteen of 36 abnormal moose examined here were 1 – 2 weeks. this can bias fecal collections. studies in the same areas (upshall et al. 1987, mccollough and pollard 1993). animals with parelaphostrongylosis often spend considerable time in a limited area depositing pellets whether shedding occurs more in one season other parasitic diseases are known to be important to moose health but the pattern of their epizootics differs from that expected of meningeal worm. ticks (dermacentor albipictus) kill moose, usually in late winter and hair loss and numerous carcasses being found following a “ticky“ winter (samuel 2004, dependent on moose densities and on weather only 1or 2 winters and are not dependent on densities of co-habiting deer. in contrast, historical moose declines in which p. tenuis was thought to play a role, were characterized peaked (whitlaw and lankester 1994a). the fascioloides magna) is a parasite whose impact on moose is not fully understood but has recently been implicated in a moose decline in northwestern minnesota (murray et al. 2006). in contrast tion in moose will be related to densities of in moose, but instead requires deer or wapiti (and the presence of appropriate aquatic snail, lymnaea spp.) for dissemination. they are go unnoticed if present and do not occur in although some progress has been made in recognizing meningeal worm infections in and other diseases such as winter ticks in limiting moose numbers, continues to be a challenge. results reported here demonstrate that the display of a typical set of neuromotor clinical signs is a reasonably good indication alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose nosis, and the possible detection of sub-clinical animals, can be expected using the p. tenuis elisa. only by accurately identifying all moose compromised by infection will the true impact of this disease on moose populations be understood. work of this type has begun in minnesota (doncarlos et al. 2002). acknowledgements saskatoon, and m. lankester. the excellent laboratory skills of ms. lily macdonald, mr. shaun dergousoff, and mrs. jocelyn vidal, the elisa and testing of samples, are also gratefully acknowledged. references anderson, r. c. 1964. neurologic disease in moose infected experimentally with pneumostrongylus tenuis from white-tailed deer. veterinary pathology 1:289-322. benson, d. a. chilton, n. b., f. huby-chilton, m. w. lankester, and a. a. gajadhar a method for extracting genomic dna differentiation of elaphostrongylus spp. from parelaphostrongylus spp. by pcr assay. journal of veterinary diagnostic clark, r. a., and r. t. bowyer. 1986. occurrence of protostrongylid nematodes in sympatric populations of moose and white-tailed deer in maine. alces 22:313322. doncarlos, m., r. o. kimmel, j. s. lawrence, and m. s. lenarz. 2002. summary of wildlife research findings 2002. unpublished report, section of wildlife, minnesota department of natural resources, st. paul, minnesota, usa. dumont, a., and m. créte. 1996. the meningeal worm, parelaphostrongylus tenuis, a marginal limiting factor for moose, alces alces, in southern quebec. canadian forrester, s. g., and m. w. lankester. 1997. extracting protostrongylid nematode gilbert, f. f. 1974. pneumostrongylus tenuis journal of wildlife management 38:4246. gogan, p. j. p., k. d. kozie, e. m. olexia, and n. s. duncan. 1997. ecological status of moose and white-tailed deer at voyageurs national park, minnesota. alces 33:187201. huby-chilton, f., n. b. chilton, m. w. lankester, and a. a. gajadhar. 2006. single-strand conformation polymorphism (sscp) analysis as a new diagnostic the elaphostrongyline (nematoda: protokarns, p. d. 1967. pneumostrongylus tenuis in deer in minnesota and implications for moose. journal of wildlife management 31:299-303. _____. 1977. deer-moose relationships with emphasis on parelaphostrongylus tenuis. minnesota wildlife research quarterly 37:40-61. lamson, a. l. 1941. maine moose disease studies. m.sc. thesis, department of maine, usa. lankester, m. w. 1974. parelaphostrongylus tenuis (nematoda) and fascioloides magna (trematoda) in moose of southeastern manitoba. canadian journal of diagnosing parelaphostrongylosis in moose – lankester et al. alces vol. 43, 2007 _____. 2001. extrapulmonary lungworms of in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. press, ames, iowa, usa. _____. 2002. low-dose meningeal worm (parelaphostrongylus tenuis) infections in moose (alces alces). journal of wildlife lenarz, m. s., and k. d. kerr. 1987. an parelaphostrongylus tenuis as a source of mortality of moose. annual report of the minnesota department of natural resources, populations and research unit, grand rapids, minnesota, usa. mccollough, m. a., and k. a. pollard. 1993. parelaphostrongylus tenuis in maine moose and the possible influence of faulty baermann procedures. journal murray, d. l., e. c. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. influence of pathogens, nutritional deficiency, and climate change on a declining southern moose population. wildlife monographs 166. ogunremi, o., m., w. lankester, s. dergou-, s. dergousoff, and a. a. gajadhar. 2002a. detection of anti-parelaphostrongylus tenuis antibodies in experimentally infected and free-ranging moose (alces alces). journal of wildlife diseases 38:796-803. _____, _____, and a. a. gajadhar. 2002b. immunodiagnosis of experimental parelaphostrongylus tenuis infection in elk. canadian journal of veterinary research 66:1-7. _____, _____, j. kendall, and a. gajadhar. 1999a. serological diagnosis of parelaphostrongylus tenuis infection in white-tailed deer and identification of a potentially unique parasite antigen. jour_____, _____, s. loran, and a. gajadhar. products and somatic worm antigens for the serodiagnosis of experimental parelaphostrongylus tenuis infection in white-tailed deer. journal of veterinary peterson, w. j., and m. w. lankester. 1991. aspects of the epizootiology of parelaphostrongylus tenuis in a white-tailed deer population. alces 27:183-192. pybus, m. j. 149 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. second edition. usa. samuel, w. m. 2004. white as a ghost: alberta naturalists, edmonton, alberta, canada. ics of winter ticks and die-offs of moose. alces 43:39-48. slomke, a. m., m. w. lankester, and w. j. peterson dynamics of parelaphostrongylus tenuis in white-tailed deer. journal of wildlife smith, h. j., and r. m. archibald. 1967. moose sickness, a neurological disease of parasite, elaphostrongylus tenuis. canadian veterinary journal 8:173-177. _____, _____, and a. h. corner. 1964. elaphostrongylosis in maritime moose and deer. canadian veterinary journal thomas, j. e., and d. g. dodds. 1988. brainworm, parelaphostrongylus tenuis in moose alces alces, and white-tailed deer odocoileus virginianus thomas, l. j., and a. r. cahn. 1932. a new disease in moose. i: preliminary report. alces vol. 43, 2007 lankester et al. diagnosing parelaphostrongylosis in moose journal of parasitology 18:219-231. upshall, s. m., m. d. b. burt, and t. g. dilworth. 1987. parelaphostrongylus tenuis in new brunswick: the parasite in white-tailed deer (odocoileus virginianus) and moose (alces alces). journal whitlaw, h. a., and m. w. lankester. the effects of parelaphostrongylosis on moose populations. canadian journal of zoology 72:1-7. _____, and _____. 1994b. the co-occurrence of moose, white-tailed deer and parelaphostrongylus tenuis in ontario. canadian f:\alces\supp2\pagema~1\rus 30s alces suppl. 2, 2002 zaitsev – structure of moose populations 137 structure of moose (alces alces) populations in russia with special reference to communication distances vitaliy a. zaitsev institute of evolutionary animal morphology and ecology, russian academy of science, 117071, moscow, russia abstract: the structure of moose populations was studied in the southern subzone of taiga and mixed forests. moose were distributed irregularly in small groups. some migrations were recorded. the distance between moose in compact groups (5–9.3 m) and the distance between groups were compared to orientation vectors a–k, representing one of the parameters of the inner activity rhythms. the distances and the vectors correspond to the definition of the critical levels of the development of natural systems. the spatial pattern of moose distribution was revealed and a canonic distribution model developed. alces supplement 2: 137-141 (2002) key words: activity rhythms, behavior, distances between moose, ecology, hierarchical structures, natural systems, population structure, social groupings t h e s p a t i a l s t r u c t u r e s o f m o o s e populations are diverse and their most important properties should be determined. this study examines spatial structures of a moose population and shows regular utilization of habitats by dispersing moose. a model is suggested that may be used to search moose habitats, estimate moose population size, and predict the distribution of moose without a complete census. study area the behavior and ecology of moose (alces alces alces) was studied in the yaroslavl, kostroma, and moscow regions of russia. vegetation is at the southern limits of taiga and boreal mixed forests, including birch forests, aspen forests, willow thickets, and also spruce and mixed forests. moose density ranged from 0.3 to 2.2 moose per km2 in the yaroslavl and kostroma regions (300 km2). moose density in the kostroma region was 0.2–0.9 moose per km2. in sikhote–alin forests, moose (alces alces cameloides milne– edwards 1867) density was only 0.01–0.2 moose per km2. localized high density concentrations were near salt licks. methods moose were observed throughout the year and monitored day and night. a night vision device was used when it was dark. we recorded the distribution of tracks, bedding sites, locations of mating areas, camps, and fecal pellet groups. results at yaroslavl, moose tracks are found throughout the entire area, but moose concentrate at certain sites in the forest. a total of 5 main groups occupy an area of 2– 10 km2 in the 300 km2 forest (naumov 1967). each group consisted of 1 8–year– old male, 2 1–4–year–old males, 3 females, and 2 calves. all members of groups were observed in close proximity day and night in winter. yearlings and 2–3–year–old moose migrate from one forest to another in spring and summer. cows concentrated at the sites of future calving in spring. approximately 67–85% of shrub layer stems showed alces suppl. 2, 2002 zaitsev – structure of moose populations 141 in which the activity rhythm is manifested. asymmetry features of the distribution of moose over the brief period of time in the course of movement result in symmetrical structures in the distribution of tracks, and, on the whole, to a fairly regular utilization of the habitats. it is possible to use the model to search moose habitats, estimate moose population size, and predict the distribution of moose without a complete census. references filonov, k. p. 1983. moose. lesnaya promyshlennost, moscow, russia. (in russian). heptner, w. g., and a. a. nasimovitch. 1967. der elch. ziesmen verlag, wittenberg–lutherstadt, germany. (in german). knorre, e. p. 1959. ecology of moose. proceedings of the pechoro-ilych state reserve 7:5–22. (in russian). naumov, n. p. 1967. population structure and population dynamics of terrestrial vertebrates. zoologichesky zhurnal 46:1470–1486. (in russian). yazan, y. p. 1961. biological features and u t i l i z a t i o n o f m i g r a t i n g m o o s e populations in the pechora taiga. proceedings of pechora–ilychskii state reserve 9:114–201. (in russian). zaitsev, v. a. 1991. kabarga sikhote– alin. ecology and behavior. nauka, moscow, russia. (in russian). zhirmunsky, a. v., and v. n. kuzmin. 1990. critical levels in the development of systems. nauka, leningrad, russia. (in russian). untitled-1 i in memoriam roy clayton anderson april 1926 – august 2001 professor roy c. anderson, re nowned professor of parasitology at the university of guelph and world recognized researcher in the fi eld of wildlife disease died august 27, 2001 in guelph, ontario. a memorial service held in the university of guelph arboretum, september 3, was widely at tend ed by close friends, academic col leagues and former graduate students who gathered with roy’s family to celebrate his ac com plish ments and to relate how his enthusiasm for research and scientifi c rigor, his love for writing, his joie de vivre and continued friend ship affected their lives. above all, his sense of humour and mastery of story-telling that brought fun and laughter will be sorely missed. roy always considered himself to be a member of the north american moose group and attended several annual meet ings. “moosers”, more than any other scientifi c group, recognized the importance of one of his biggest discoveries – iden ti fy ing the cause of the mysterious neurologic disease known for many years simply as “moose sickness”. like many great dis cov er ies, his was part serendipity and part good de tec tive work. while working one summer at the lake sasejewun wildlife research sta tion, algonquin park, ontario, a col league showed roy a worm he’d found in the cranium of a white-tailed deer killed by a car. this introduction led professor anderson to a long and productive series of scientifi c studies progressively revealing the importance of this parasite, known as the “meningeal worm” or parelaphostrongylus tenuis. first, he showed that p. tenuis has to develop in land snails or slugs before the larvae are infective. deer picked up these land gastropods accidentally while feeding on herbs and grasses close to the ground. the larval worms are released from the tissue of the snail and penetrate the true stomach (abomasum). from there they migrate to the spinal cord and penetrate the nerve tissue where they mature. after about 90 days, they fi nally emerge from the spinal cord and spend the rest of their lives between the membranes (meninges) cov er ing the brain. in this location they can reach a length of almost 9 cm! virtually all deer in the east (but not western north america) have this ii worm and amazingly, none show any sign of disease. this was all new to science and very exciting, but more im por tant ly, this knowledge led roy to formulate an even more productive hypothesis. interest in moose was growing with the publication of randolph peterson’s book in 1955 and studies of moose and moose man age ment by douglas pimlott at u. of t. and other researchers in eastern canada. as well, there were renewed questions about an unexplained neurologic disease of moose in the maritimes and minnesota and its possible link to moose population declines. roy knew of other nematodes like p. tenuis (in rats of the south pacifi c and cattle of china and korea) that matured in the cen tral nervous system of their “normal host” without causing disease but became path o gen ic when transmitted to unfamiliar hosts (like humans and horses). he suspected that p. tenuis of white-tails might behave this way if it had the opportunity to infect moose, an abnormal host. opportunities to do so increased as deer numbers grew in the early 1900s in response to low-snow win ters and extensive cutting of mature conifer forests. increased densities of deer brought with them a parasite with which moose had no prior experience. his suspicions about p. tenuis were reinforced when he plotted known cases of moose sickness on a map and discovered that they occurred only where the ranges of moose and deer overlapped. and conversely, cases did not occur, for example, in moose in newfoundland, on anticosti island, or isle royale, where deer were absent. his hypothesis was put to the test by experimentally infecting 2 moose calves with larvae of the meningeal worm from deer. interestingly, these calves were obtained from the chapleau district of north ern ontario with the assistance of mr. vince crichton sr., father of the moose call editor. both calves developed the classical signs of staggers, hind-quarter weakness and paresis typical of moose sick ness. these experimental fi ndings were quickly confi rmed by roy and others who found p. tenuis in the spinal cords and brains of wild moose showing similar dis ease signs. finally, more than 50 years after the fi rst reports of moose sickness, the causative agent had been identifi ed. it was perhaps fi tting that roy shared his discovery with moose biologists at the north american moose conference, con vened that year as the 1st international symposium on moose ecology, quebec city, in 1974. his presentation followed im me di ate ly after the keynote address given by dr. randolph peterson. at subsequent annual moose conferences, roy extended his ideas about the possible role of moose sickness in the decline of moose populations over parts of their range in eastern north america and speculated that p. tenuis might similarly limit the present day distribution of caribou and elk which he also demonstrated ex per i men tal ly to be susceptible to infection. despite his enthusiasm for these ideas, roy anderson was a careful scientist and me tic u lous writer who spent his career teach ing students that even the most imaginative hypothesis, no matter how logical and how well it seemed to explain things, ultimately has to be tested and “proved” by ex per i men ta tion. roy was not a fi eld biologist and not really in a position to be able to do this. in fact, he didn’t see his fi rst infected sick moose alive in the wild until traveling in northern minnesota with bill peterson and myself in 1995. although anderson pop u lar ized the idea that p. tenuis might have caused historical moose population declines and limited the eastward distribution of oth er cervids, he continually urged fi eld par a si tol o gists to prove it. this diffi cult task has been undertaken by a few of us, including several hard working graduate students. we’ve made progress and now have a much improved understanding of the fi eld biology of the parasite and it’s path o genic i ty in moose, but a defi nitive measure of the impact of meningeal worm awaits the use of new tools such as iii the newly developed blood test for p. tenuis infection in moose. although roy was undoubtedly best known among north american wildlife bi ol o gists for discovering the cause of moose sickness, his other research and academic contributions were also remarkable. be gin ning with his ph.d. work in algonquin park, ontario, on ornithofi laria fallisensis of waterfowl, he developed a life-long in ter est in the spirurid nematodes, particularly the fi larioids and the acuarioids. his orig i nal contributions to our understanding of the taxonomy, systematics, and transmission of these and related groups earned him an international reputation among nematologists and parasitologists. roy was born (april 26, 1926) in camrose, alberta, canada, a small com mu ni ty on the prairies. here he grew up with an appreciation of the environment and the wildlife that inhabited it. while in high school, roy became an avid “birder” under the tutelage of frank l. farley and learned to recognize birds by sight and by sound. he indulged this hobby throughout his life. on graduating from high school, roy entered the navy and served during world war ii in communications on a corvette. after the war, he married phyllis, had two sons, doug las and michel, and enrolled in the biology program at the university of alberta. upon graduation (1950), he went to the uni ver si ty of toronto as a graduate student stud y ing under dr. a. murray fallis and received his m.sc. (1952) and the ph.d. (1956). this was followed by post-doctoral studies with basil goodey at the rothamsted ex per i men tal station, uk, with j.j.c. buckley at the london school of hygiene and trop i cal medicine, and with professor alain chabaud at the college de france, paris. roy returned to canada becoming a mem ber of the staff of the ontario research foundation (1958). in 1965 he was ap point ed professor of invertebrate zoology at the then fl edgling university of guelph and served as chair of zoology (19791989) and as acting dean of the college of biological science (1971 and 1977-78). he remained in the department until retirement in 1991 and continued as university pro fes sor emeritus, working daily in the offi ce on books and manuscripts. roy’s ease with writing, which he tire less ly tried to impart to his students, pro duced an outstanding legacy of published research. he was the sole or co-author, or main advisor, of 269 peer-reviewed sci en tifi c papers. included in these were de scrip tions of 81 new species and 8 new genera. the 10 volume c.i.h. keys to the nematodes (edited with drs. a. chabaud and s. wilmot) set a new standard for nematode classifi cation. he also authored or co-authored 11 chapters in books and 5 books, including his beloved nematode par a sites of vertebrates their development and transmission (cabi publishing). the 2nd edition of this classic work appeared in spring of 2000 and includes all species of parasitic nematodes of which something is known of their development and trans mis sion (almost 600 spp. and 3,200 ref er enc es). dr. anderson was a strong supporter of the canadian society of zoologists. he served as its 2nd and 1st vice-presidents and president (1975-76). he chaired the parasitology section (1982-83) and nom i nat ing committee (1978-79), was convener of the annual meeting at the university of guelph (1975), and on the organizing com mit tee for icopa v, toronto (1982). roy was the parasitology section’s wardle medalist (1988), received honourary mem ber ship in the section (1998), and recently was made an honourary member of the society (2001). almost as a right of pas sage for the 14 m.sc. and 15 ph.d. students he trained was the expectation that each would become a society member and present a paper at an annual meeting of the csz. service to other professional societies included executive positions in the amer i can society of parasitologists (vice-pres i dent 1977-78) and the wildlife disease association (president 1981-1983) as well as membership on iv numerous committees. he served as associate editor, canadian journal of zoology (196878); assistant editor, journal of parasitology (usa) (1968-72); editor-in chief for the classifi cation of the nematode parasites of vertebrates, commonwealth agricultural bureaux (1972-1984); co-editor, systematic parasitology (1978-2001); and on the editorial boards of annales de parasitologie humaine et comparée, paris (1989-94), proceedings of the helminthological society of wash ing ton (1984-2001), folia parasitologia, prague (1986-96), helminthological abstracts, se ries a (1988-2001), and cabi publishing. awards for his contributions to par a si tol o gy and the training of students included the henry baldwin ward medal, american society of parasitologists (1968); sigma xi award for excellence in research, guelph chapter (1973); distinguished service award (1978) and emeritus member (1993), wildlife disease association; robert arnold wardle award/medal, canadian society of zoologists (1988); mentor award, amer i can society of parasitologists (1997); and director’s award and lifetime member, friends of algonquin park, ontario (1992). no greater tribute can be bestowed by peers on a parasitologist than to have their name assigned to a valid new species; dr. anderson has had a total of 14 species and 2 genera named in his honour. regrettably, roy’s untimely death (a young 75) prevented the north american moose group from formally recognizing his contributions to our knowledge of moose diseases and for his role in training several generations of fish and wildlife students at the university of guelph. nonetheless, many will have fond memories of hearing one of roy’s presentations at an annual moose conference, enjoying his company on postconference fi eld trips or simply calling him up on the spur of the moment for an opinion about some peculiar condition seen in a moose. despite international fame in the academic world, roy loved most of all to spend time with biologists whose knowledge stemmed from years of working in the fi eld. moose managers had his genuine respect. fittingly, roy’s con tri bu tions to canadian science are being com mem o rat ed by the the college of biological science, university of guelph in es tab lish ing an annual lecture in parasitology named the roy c. anderson memorial lecture. on behalf of all of its members, the north american moose group was one of the fi rst to make a fi nancial contribution to this me mo ri al. the endowed fund still has some growing to do and personal donations (tax-deductible) to the university of guelph, guelph, ontario, canada n1g 2w1, ac count # 801801 “the roy c. anderson memorial lecture series” are welcome. contact pwoo@uoguelph.ca if you need clarifi cation or more information on the lecture series. murray w. lankester 31 a review of methods to estimate and monitor moose density and abundance remington j. moll1, mairi k. p. poisson1, david r. heit1, henry jones2, peter j. pekins1, and lee kantar3 1department of natural resources and the environment, university of new hampshire, 56 college road, durham, nh 03824, usa; 2new hampshire fish & game department, 200 main street, new hampton, nh 03256, usa; 3maine department of inland fisheries and wildlife, b state street, bangor, me 04401, usa abstract: acquiring accurate and precise population parameters is fundamental to the ecological understanding and management and conservation of moose (alces alces). moose density is challenging to measure and often estimated using winter aerial surveys; however, numerous alternative approaches exist including harvest analysis, public observation, unpiloted aerial system (uas) surveys, and camera trapping. given recent developments in a number of field and analytical techniques, there is value in reviewing and synthesizing the strengths and limitations of monitoring methods to best evaluate their respective tradeoffs in management scenarios. we reviewed 89 studies that included 131 estimates or indices of moose density. as expected, aerial surveys were the most common method of obtaining a moose density estimate (58%) followed by use of public data (e.g., harvest records = 27%); more recent studies employed novel methods including uas. most estimates (64%) failed to account for imperfect detection of moose (i.e., “sightability”) and this tendency has not improved over time. density estimates ranged from <0.1 to 10.6 moose/km2 (average = 0.7) and population precision, as measured by the 90% confidence interval, ranged from 6.5 to 120.0% of the density estimate (average = 37.4%). correlations among estimates obtained for the same populations varied widely, with r2 values ranging from 0.02 to 0.99 (average = 0.58). our review indicates that: 1) methods to estimate moose density have been dominated by aerial surveys but are diversifying, 2) precision of density estimates has been highly variable and on average lower than broadly accepted target benchmarks, and 3) many methods did not account for sightability and presumably underestimated moose density. we reflect on these trends and discuss how emerging methods, including camera trapping, uas surveys, and integrated population modeling (ipm) can complement and improve traditional approaches. we suggest that no single “best” method exists, but rather the best method is one that accounts for sightability bias and yields target precision at reasonable cost, which vary by jurisdiction and goal. alces vol. 58: 31–49 (2022) key words: abundance, alces alces, density, distribution, management, occurrence, survey introduction obtaining accurate and precise estimates of population parameters is fundamental to the ecological understanding, effective management, and conservation of wildlife (skalski et al. 2005, sinclair et al. 2006, silvy 2012). such parameters are particularly important for moose (alces alces) because this species is often managed for harvest (jensen et al. 2020), has both negative and positive economic impact (e.g., vehicle collisions and ecotourism, respectively; storaas et al. 2001, silverberg et al. 2003, sample et al. 2020), and is hypothesized to be susceptible to global climate change (murray et al. 2006, jensen et al. 2020). the population parameters required to manage and conserve moose vary by location and context, but typically include monitoring moose populations – moll et al. alces vol. 58, 2022 32 density, survival, recruitment, and composition (gasaway et al. 1986, krausman 2002, van ballenberghe and ballard 2007). estimates of density (hereafter, the term density includes the related term abundance) and proxies thereof are arguably the most fundamental because management is frequently focused on maintaining moose populations at specific densities over time (leopold 1933, franzmann and schwartz 2007). the estimated global moose population is ~2.2 million, with roughly half in eurasia and half in north america (timmermann and rodgers 2017, jensen et al. 2020). recent reviews have highlighted variation in population dynamics across management jurisdictions, with some populations declining and others increasing or stable (timmermann and rodgers 2017, jensen et al. 2020). population dynamics are complex and geographically varied, but broadly reflect habitat composition, forest management, abiotic environmental conditions, hunter harvest, predation pressure, and parasites (boutin 1992, messier 1994, rempel et al. 1997, solberg et al. 1999, musante et al. 2010, jones et al. 2017, pekins 2020). climate change has an underlying influence on these factors as well as moose behavior and susceptibility to parasites and disease (joly et al. 2012, tape et al. 2016, montgomery et al. 2019, pekins 2020), underscoring the need for techniques that accurately monitor moose populations over time (van ballenberghe and ballard 2007, jensen et al. 2020). moose density is challenging to estimate and monitor for logistical, financial, and ecological reasons. moose are highly mobile and inhabit large ranges that make monitoring difficult by enlarging the scale required for adequate sampling (krebs 2006, singh and milner-gulland 2011, harris et al. 2015) and rendering survey efforts expensive (bontaites et al. 2000, peters et al. 2014, boyce and corrigan 2017). behavior frequently reduces detection or “sightability” because moose use dense forest cover, avoid human disturbance (including activities associated with surveying populations), and are mostly crepuscular or nocturnal when active (frid and dill 2002, harris et al. 2015). further, there are practical difficulties associated with surveying populations in remote regions with high topographic relief (van ballenberghe and ballard 2007, kellie et al. 2019). thus, monitoring moose populations is an “evolving art” (krebs 2006, p. 367) that is shaped by the advent and implementation of emerging technologies and methods (boyce and corrigan 2017, oyster et al. 2018, mcmahon et al. 2021). winter aerial surveys are a common method used to estimate and monitor moose density, and often conducted by helicopter with observers counting moose on snow (gasaway et al. 1985, 1986, van ballenberghe and ballard 2007, timmermann and rodgers 2017). many jurisdictions have employed surveys for decades as the backbone of monitoring efforts (e.g., alaska, usa, and alberta, canada, alberta environment and parks 2016, kellie et al. 2019). traditionally conducted with a simple single or double-observer method, surveys have evolved to include a distance sampling approach (gasaway et al. 1986, alberta environment and parks 2016, oyster et al. 2018). challenges associated with moose monitoring include high cost, accounting for sightability bias, and danger to aviators and observers (sasse 2003, peters et al. 2014, oyster et al. 2018). alternatively, where sufficient harvest occurs, moose populations have been monitored using harvest data analysis (solberg et al. 1999, skalski et al. 2005, decesare et al. 2016). hunter observations have also been used as a cost-efficient index for moose density, although such observations require calibration with other data to achieve reliability (bontaites et al. 2000, boyce and alces vol. 58, 2022 monitoring moose populations – moll et al. 33 corrigan 2017). less common approaches include snow tracking, pellet surveys, camera trapping, and aerial surveys using unpiloted aerial systems (uas) (bobek et al. 2005, krester et al. 2016, pfeffer et al. 2018, mcmahon et al. 2021). the challenges and importance of obtaining accurate moose density estimates are clear and best management requires continuous assessment and adaptation to evolving methodology. thus, it is valuable to review and synthesize current methods and evaluate their associated strengths and limitations to identify and navigate their trade-offs. here, we review methods used to estimate and monitor moose density with an emphasis on studies that directly analyzed a method of monitoring moose to derive a density estimate or index. we summarize the results of our literature survey, the limitations of each method, and discuss future direction in monitoring and estimating moose density. methods we performed a literature survey to review methods for estimating and monitoring moose density with an emphasis on studies with management application. in may 2022 we used the web of science to search all collections using the following boolean string of terms: title: (“alces alces” or moose) and topic: (abundance or density) and topic: (management); the search yielded 453 studies. we eliminated irrelevant studies including those focused on other species, purely mathematical or simulation-based studies, and reviews. we only retained studies that were primary sources and omitted those using density estimates from other sources. we retained studies that used densities to test ecological hypotheses, but only if they met the other criteria and included sufficient details regarding study design and density estimation methods. we included the term “management” in the search to emphasize studies relevant to practitioners and to help eliminate irrelevant studies (e.g., those that used moose density estimates from other sources). for each relevant study, we recoded information according to the framework described below (fig. 1). we recorded the study location, the spatial scale at which inference was desired (km2; typically, the study area), and the method(s) employed to estimate or monitor the target population density. we then classified each method as follows. we first identified the goal type (n = 3) of the method according to timmermann and buss (2007): 1) a census, where an attempt was made to count all animals within an area, 2) a sample, where inference regarding density was achieved through sampling and statistical analysis, or 3) an index, where a relative measure representing density was desired. next, we classified the primary method as either aerial, ground-based, or based on public observations. aerial methods included fixed-wing aircraft, helicopters, and unpiloted aerial systems (uas, or “drones”); ground methods included pellet counts, snow track surveys, and camera trapping; public methods included the use of harvest data and public observations (e.g., hunter or citizen science observations). for all methods, we recorded whether the survey design was systematic, random, stratified random, or non-random (e.g., a survey targeting a specific management area). for aerial methods we also recorded whether the sampling scheme was conducted using block searches (an area-based design where “blocks” might also be called “quadrats”, “plots”, or “sample units”) or strip-transects (a line-based design). for public observations, we recorded whether data reporting was mandatory (typical for harvest data) or voluntary. we recorded the timing (i.e., seasons) of each method, the duration of data collection monitoring moose populations – moll et al. alces vol. 58, 2022 34 (in years), and whether each method accounted for the imperfect detection of individuals (i.e., an individual is present but not detected; mackenzie et al. 2002), often referred to as sightability (gasaway et al. 1986). we note that the duration of data collection might not always reflect the duration of monitoring programs, but rather is indicative of the dataset’s use for a particular analysis. thus, this value represents the duration of data collection used to inform the scientific literature rather than the true duration of moose monitoring efforts. seasons were set as fall (september through november), winter (december through april), and spring-summer (may through august). we also recorded whether the method distinguished the age or sex of individual moose. for aerial and ground surveys, we considered a study to have accounted for sightability if it formally accounted for undetected individuals, including the use of double-observation methods, mark-resight methods, distance sampling methods, or previously calculated sightability correction factors. for public observations, we considered a study to have accounted for sightability if it formally corrected for imperfect reporting (e.g., a moose was observed but not reported) or undocumented harvest. we then used a logistic regression to evaluate whether there was a temporal trend in accounting for sightability, where a binary response of accounting for sightability was modeled as a function of the publication year; p < 0.05 was the threshold for inference. when reported, we recorded the mean population estimate and converted it to moose density (individuals per km2). we acknowledge the substantial variation in moose densities across their circumpolar range. we report the densities of the studies in our review to quantify this variation broadly and provide to context for the various monitoring methods, which might be informed by local moose densities (e.g., tailoring survey methods for a low-density population; hinton et al. 2022). fig. 1. the framework used to organize and summarize methods used to estimate and monitor moose (alces alces) density and abundance. see text for detailed descriptions of categories. uas = unpiloted aerial system. alces vol. 58, 2022 monitoring moose populations – moll et al. 35 we recorded the precision of that estimate following gasaway et al. (1986), according to the equation: u p p ˆ ˆ − (1) where u is the upper value of the 90% confidence interval and p̂ is the estimated population value (density or abundance). this precision metric is widely recognized and cited as a target benchmark for management decisions (e.g., timmermann 1993, bontaites et al. 2000, peters et al. 2014). for studies that reported multiple density estimates and precisions (e.g., across management units or years), we recorded the average density and precision across all estimates. in cases where only the standard error was reported for precision, we converted this error into a confidence interval by multiplying it by 1.645 (the z-value for a 90% confidence interval). lastly, we recorded whether a study formally compared one population estimate to an estimate produced via another method, and if so, the pearson’s correlation coefficient (r) or r2 value of each such comparison. results a total of 89 (20%) of 453 studies returned by the literature search met the criteria for review (see appendix 1), with most (74%) conducted in north america and the remainder in fennoscandia and eurasia (fig. 2). of these, 65 provided details regarding the spatial scale of desired inference that ranged from as small as 6.0 km2 to >13.6 million km2 (see appendix 1); the average and median values were 278,000 km2 and 3,456 km2, respectively, indicating that the distribution of these scales was non-normal and heavily right-skewed. the 89 studies contained 131 estimates or indices used to monitor moose density. however, the proportional statistics below fig. 2. geographic distribution of studies included in a literature survey of methods used to estimate and monitor moose (alces alces) density and abundance conducted in may 2022. note that the sum of studies here exceeds the total number of studies reviewed because several studies were conducted in more than one country. monitoring moose populations – moll et al. alces vol. 58, 2022 36 omit certain studies lacking sufficient detail to adequately characterize the methodology (i.e., nas in a given category were dropped; see appendix 1). across all methods, the most common goal type (n = 79, 61%) was a sample or statistical representative of the broader population density; less common were the index (n = 29, 22%) and census (n = 22, 17%) goal types. a single method employed a cohort analysis that combined multiple data types (see appendix 1). of the three primary methods, aerial surveys were most common (n = 76, 58%), followed by public reporting (n = 35, 27%) and ground surveys (n = 19, 15%) (fig. 3). helicopters (n = 35, 46%) and fixed-wing aircraft (n = 26, 34%) were the most common flight modes, in which block (n = 56, 76%) and strip-transect surveys (n = 17, 23%) were used exclusively, other than a single exception, being their combination. harvest (n = 20, 57%) and public observation (n = 14, 40%) were the most common public reporting methods with mandatory (n = 19, 56%) and voluntary (n = 15, 44%) reporting used in both. pellet counts (n = 10, 53%) and direct observation (n = 6, 32%) were the most common ground survey methods; snow tracking methods (n = 2, 11%) were used less frequently (fig. 3). the most common survey design was non-random (n = 60, 49%), followed by stratified random (n = 37, 30%), systematic (n = 19, 15%), and random (n = 7, 6%). most (67%) non-random survey designs were associated with public observations, while all stratified random survey designs, except one, were associated with aerial methods (table 1; appendix 1). approximately onethird of all methods formally accounted for sightability (n = 45, 36%). the application of a sightability correction for density estimates in studies did not change linearly over time (β = 0.004, df = 122, p = 0.77; fig. 4). fig. 3. primary and secondary methods used to estimate and monitor moose (alces alces) density and abundance according to a literature survey conducted in may 2022. primary methods are represented in the central circle; associated secondary methods are represented in the outer circle. one method employed a cohort analysis that combined multiple methods, and thus was not included in this figure. fig. 4. model predictions of the probability that a study estimating or monitoring moose (alces alces) densities accounted for sightability, or the imperfect detection of individual moose, as a function of the year the study was published. the slope of the model is not significantly different from zero (p = 0.83). model was fit to data collected from studies returned from a literature survey of methods used to estimate and monitor moose (alces alces) density and abundance conducted in may 2022. alces vol. 58, 2022 monitoring moose populations – moll et al. 37 helicopter surveys were the only method that accounted for sightability most (59%) of the time (table 1). fifty-four studies reported 68 mean population densities ranging from <0.1 to 10.6 moose/km2 (average = 0.7; fig. 5a). the precision of 30 estimates reported or determinable in 24 different studies ranged from 0.07 to 1.20 moose/km2 (average = 0.37; fig. 5b); half of these estimates were larger than the 0.25 target benchmark suggested by gasaway et al. (1986). of the 131 estimates, 87% (n = 114) were conducted during a specific season, most in winter (n = 63, 55%) and fall (n = 31, 27%); spring-summer estimates were uncommon (n = 8), as were year-round estimates (n = 5). aerial estimates more frequently occurred in winter (80% of the time), while public estimates typically occurred in fall (77% of the time; table 1). eighty-one estimates (62%) distinguished age (calf vs. adult) and/or sex of moose; however, those studies distinguishing sex did not necessarily distinguish age or vice versa (see appendix 1). study duration ranged from 1 to 73 years. twenty-four methods (18%) employed a single year of study, typically a single season; 38 methods (29%) were employed for two to five years, 17 (13%) for 6–10 years, 15 (11%) for 11–20 years, and 15 for >20 years. twenty-two studies did not report duration or varied so greatly in duration and/or spatial coverage (due to funding limitations or other barriers) that they were omitted from the summary (see appendix 1). seventeen studies provided pearson’s correlation coefficient (r) or r2 values to compare methods. the most frequent comparisons included moose observations by hunters, hunter success rate, density estimates from aerial surveys, and harvest data analysis (see appendix 2). the r2 values from these comparisons ranged from 0.02 to 0.99 (average = 0.58). sample size for these values ranged from 6 to 111 (average = 22). table 1. a summary of the number of methods used to estimate and monitor moose (alces alces) density and abundance according to a literature survey conducted in may 2022. aerial ground public fixed-wing helicopter uas pellets tracks harvest obs. goal census 8 9 0 0 1 0 2 index 2 0 0 3 0 8 14 sample 16 26 1 7 0 12 4 survey design non-random 6 7 0 4 1 18 18 random 3 1 0 1 0 1 0 strat. rand. 12 16 0 0 0 1 0 systematic 3 9 1 4 0 0 1 timing fall 2 0 0 0 0 15 12 spr-su 1 0 1 6 0 0 0 winter 17 32 0 2 1 1 1 other 5 1 0 1 0 0 6 sight-ability yes 13 20 1 2 0 2 1 no 12 14 0 8 1 16 19 uas = unpiloted aerial system, obs. = observation, stat. rand. = stratified random. monitoring moose populations – moll et al. alces vol. 58, 2022 38 the lowest r2 values were associated with comparisons between hunter harvest per unit effort and minimum population counts from aerial surveys, and between change-in-ratio abundance estimates and rates of moosetrain collisions (appendix 2). a comparison between hunter harvest per unit effort and an aerial density estimate yielded the highest correlation using a non-linear model fit with small sample size (n = 6; appendix 2). the complete references for reviewed studies are provided in appendix 1 available on the alces website https://alcesjournal.org/ index.php/alces. discussion as expected, this literature survey revealed a historical reliance on aerial methods to fig. 5. the mean density of moose/km2 (panel a) and the density estimate precision (panel b) according to a literature survey of methods used to estimate and monitor moose (alces alces) density and abundance conducted in may 2022. for clarity, one outlier density estimate of 10.6 moose/km2 from a local-scale study was omitted in panel a (see appendix 1). https://alcesjournal.org/index.php/alces https://alcesjournal.org/index.php/alces alces vol. 58, 2022 monitoring moose populations – moll et al. 39 estimate moose density. many authors noted limitations to conducting an ideal aerial survey, especially financial and logistical costs and weather (e.g., peek 1974, nygrén and pesonen 1993, bowyer et al. 1999, harris et al. 2015, kellie et al. 2019). although sightability bias was acknowledged frequently as an issue, almost two-thirds of density estimates across studies did not account for it (fig. 4). we also found high variation in the precision of density estimates, with only half achieving the benchmark management target of gasaway et al. (1986). landscape, habitat, and social factors all appeared to play a role in method choice. for example, aerial surveys were particularly common in alaska, usa, where habitat is more open and individuals tend to cluster across the landscape, whereas harvest and hunter-based approaches were common in scandinavia where harvest is high and reporting is strong (appendix 1). finally, several recent studies highlighted new technologies and modern analytical developments that hold promise for moose population monitoring, although key questions and challenges remain regarding their reliability and overall efficacy (table 2). surprisingly, the critical issue of sightability bias was usually unaccounted for, with no evidence of improvement over time (fig. 4), despite its influence on the accuracy and precision of moose population estimates (evans et al. 1966, caughley 1974, gasaway et al. 1986, peters et al. 2014, harris et al. 2015). multiple studies using aerial methods have documented declines in moose sightability with increasing canopy cover, especially in conifer patches. for example, using thermal drone surveys, mcmahon et al. (2021) found that sightability declined from near 100% in open habitat to < 25% in 75% canopy cover, and peters et al. (2014) estimated sightability along transects to be as low as 46% during helicopter flights over non-ideal snow conditions. regarding public data, sightability can influence physical observation rates or manifest as imperfect reporting or undocumented harvest (e.g., illegal harvest). in these cases, the number of moose present within or removed from the population would be underestimated, and could be a common scenario. such underestimation is likely pervasive for moose and other species where imperfect detection of individuals is the norm (caughley 1974, mackenzie et al. 2002, stephens et al. 2006). therefore, a key area of improvement for moose monitoring methods is to explicitly and rigorously account for sightability bias. studies accounting for sightability bias varied in approach, ranging from extensive field efforts and sophisticated statistical modeling (e.g., oyster et al. 2018) to correction factors based upon previous work in the same study area using similar methods (e.g., bontaites et al. 2000). to increase sightability, aerial surveys are typically conducted in snow conditions that accentuate the contrast between moose and their background environment (gasaway et al. 1986); however, this reliance on snow cover is problematic from a scheduling perspective. the expectation is that this problem will worsen presuming that climate change reduces the length and timing of snow cover, especially along the southern range of moose (bormann et al. 2018, kellie et al. 2019, jensen et al. 2020). another approach to quantify sightability in aerial methods is to search for a subset of vhfor gps-collared individuals and quantify their detection probabilities as a function of habitat or environmental conditions (e.g., peters et al. 2014). this approach is effective but usually part of a separate ecological study given the substantial costs of capture, collars, and tracking individual moose. an approach generally absent in our literature survey, but common for other taxa, is using monitoring moose populations – moll et al. alces vol. 58, 2022 40 repeated surveys in space or time to quantify detection probabilities (e.g., as in occupancy modeling; mackenzie et al. 2002, tyre et al. 2003). for moose, this method could be applied by using repeated sampling over time with camera traps, with multiple observers in ground-based surveys or among the public, or by flying transects more than once in rapid succession such that population closure assumptions are reasonably met (adams et al. 1997, rota et al. 2009). depending on context, such strategies are a cost-effective alternative to using radiomarked animals. however, we note that sightability is method-specific and the correction from one method might not apply to another. whatever the method, it is crucial to consider sightability bias because a failure to do so likely results in density underestimation and leads to bias in other types of inference (e.g., wildlife-habitat relationships; see kéry et al. 2010). for example, the early successional habitat that moose prefer is associated with decreased sightability. therefore, a survey that does not separate the observational effects of habitat on moose detection (i.e., a negative effect) from the ecological effects (i.e., a positive effect) table 2. a comparative summary of the predominant methods used to estimate and monitor moose (alces alces) density and abundance. method advantages limitations outlook fixed-wing and helicopter surveys easily understood and trusted by many stakeholders; sometimes considered the “gold standard” logistical and financial costs; danger to aviators; often dependent on weather and snow, and sensitive to terrain variation useful method in jurisdictions where resources enable consistent surveys; overall prevalence likely to decline as other methods develop camera trapping relatively low field effort and expense; non-invasive must process voluminous image data efficiently; analytical methods are rapidly changing increase in prevalence as analytical methods continue to develop harvest data analysis inexpensive; produces reliable estimates of population when appropriate statistical methods are applied limited to areas where harvest is substantial; bias in non-random hunter behavior must be accounted for a foundational method when harvest is substantial pellet and track surveys relatively inexpensive; non-invasive requires calibration with other density estimates to achieve reliability; low precision; field intensive in terms of person-power; tracking requires snow pellet surveys could become more powerful when combined with genetic analyses; tracking a useful low-tech index at local scales public or hunter observations inexpensive; engages multiple stakeholders study design must be carefully considered to avoid reporting bias; indices do not always track with population dynamics viable management tool when calibration using other methods occurs regularly uass less expensive and safer than traditional aerial surveys spatial extent and locations of surveys often limited by regulations; unproven as broadscale monitoring method; non-trivial initial purchase costs current applications are most effective for local scale studies; regulatory changes could precipitate rapid increase in capacity uas = unpiloted aerial system. italicized methods indicate those appearing in recent literature. alces vol. 58, 2022 monitoring moose populations – moll et al. 41 risks conflating observation error with ecological inference. indices, especially those based on hunter observation and success rate, were a relatively common approach in tracking moose population trends (appendix 2). indices are often employed because they are cheaper – sometimes much cheaper – than more field intensive and potentially hazardous methods such as helicopter surveys (sasse 2003, krebs 2006). however, indices must be carefully calibrated and regularly reviewed to ensure reliability (caughley 1974, bontaites et al. 2000, hatter 2001, ueno et al. 2014). although several papers reported high correlations between an index and population density, sample sizes were often small (e.g., n < 10) and a number of comparisons had only moderate correlation (see appendix 2). for example, moose observations made by hunters in maine, usa exhibited only a modest correlation with densities obtained from aerial helicopter surveys (r2 = 0.32; n = 13; kantar and cumberland 2013). recent work in ontario, canada showed that the relationship between harvest and abundance varies by age and sex class, thereby highlighting the need to calibrate indices by demographic categories (priadka et al. 2020). this study also found non-linear relationships between harvest effort, harvest, and abundance, suggesting that harvest might underestimate abundance when harvest effort is high. these observations emphasize that such indices must account for non-randomness and change in hunter activity relative to spatial coverage, effort, participation rate, weather, and technique, as well as potential non-linear relationships with hunter effort, harvest success, and moose abundance (fryxell et al. 1988, kantar and cumberland 2013, larson et al. 2014, priadka et al. 2020). more broadly, our literature survey suggests there could be untapped utility to use indices (e.g., hunter observations) in more comprehensive integrated population models (ipms). such models combine multiple data sources to improve the precision of population parameter estimates, thereby improving management efficiency (schaub and abadi 2011, arnold et al. 2018). for example, månsson et al. (2011) used simulations to demonstrate how the combination of hunter observations and pellet counts could more accurately inform management than more expensive aerial surveys. similarly, marrotte et al. (2021) integrated aerial survey data with hunter-reported data into a single model to estimate moose population trends in relation to harvest and predation. each of these data sources had unique limitations; the aerial surveys were infrequent but had higher accuracy while the hunter reports were more frequent but less standardized. the ipm developed by marrotte et al. (2021) partially overcame these challenges, resulting in increased confidence in overall moose population trends. the target precision for moose density estimates was only achieved half of the time in the studies we analyzed (fig. 5b; appendix 1). in their seminal paper, gasaway et al. (1986) suggested a target precision with a confidence interval width of ±25% of the population estimate; however, the rationale for this target was not provided and they noted that higher precision was desirable, but often prohibitively expensive (p. 5). these authors also endorsed a stratified random sampling design as a means to achieve target precision, which was an approach used in 50% of the aerial surveys we reviewed. nonetheless, certain studies found that stratification did not increase precision, especially in low-density populations (e.g., crete et al. 1986). the expanding suite of remote-sensing products (e.g., land cover and digital elevation maps) is making strict stratification less important than a random monitoring moose populations – moll et al. alces vol. 58, 2022 42 design that captures the full range of spatial conditions in a study area (fletcher and fortin 2018). spatial modeling using remotely-sensed covariates can improve accuracy and precision, while enabling prediction of moose density beyond sample units (ver hoef 2008, michaud et al. 2014). the degree to which target precision (gasaway et al. 1986) is acceptable for management objectives is an open and context-dependent question. ideally, initial target precision would be informed by power analyses and simulation, then updated with field data in an adaptive management framework (steidl et al. 1997, lyons et al. 2008). the utility of this approach was demonstrated by boyce et al. (2012), where tradeoffs between infrequent but accurate aerial surveys and frequent but less accurate kill-per-unit-effort harvest data were mathematically explored using population projection analysis. simulation can also inform study design by enabling researchers and managers to estimate required sample sizes for particular precisions (hinton et al. 2022). for example, gasaway et al. (1986) recommended an initial aerial survey to stratify landscapes by coarse-scale moose population densities. however, these flights can be expensive, thus simulation using known, assumed, or preliminary data (e.g., habitat suitability) represent an attractive alternative to traditional gasaway-type stratification. precision can also be improved through the use of ipms (marrotte et al. 2021; discussed further below). it is often difficult to monitor wildlife population trends over sufficient periods because of the many sources of variation in ecological systems and the short-term funding cycles that support monitoring efforts (field et al. 2007). indeed, population monitoring efforts with other taxa are often biased in site selection or not conducted long enough to reliably detect population trends (fournier et al. 2019, white 2019). likewise, many studies uncovered by our literature survey lasted just one year, or when they occurred across multiple years were often limited by cost that influenced the survey area in any given year (appendix 1). such constraints emphasize the need to creatively and effectively design monitoring programs that inform stated management objectives in a given landscape context. in locations with substantial harvest, cohort analysis and statistical population reconstruction represent powerful methods that can offer high accuracy and precision across broad spatiotemporal scales (solberg et al. 1999, skalski et al. 2005). in addition, the impacts of sightability bias related to unknown sources of harvest should be considered (skalski et al. 2005, timmermann and rodgers 2017). for low-density populations or those with little to no harvest, multiple methods that complement each other and account for sightability bias are paramount and will likely co-evolve with technological and analytical developments. in particular, the advantages of advanced statistical modeling coupled with robust geospatial data should be used to “get the most” out of sparse data that are expensive to collect for low-density populations. hinton et al. (2022) provide an example of this approach by combining non-linear generalized additive models, adaptive sampling, and informative geospatial covariates to monitor a low density population over a large (~25,000 km2) landscape in new york, usa. we conclude that several methods deserve additional consideration, and pending evaluation, could be implemented increasingly in moose density estimation and monitoring, foremost camera traps, uas, and ipm. technological advancements have increased wildlife detection capability of camera traps and improved their field reliability (burton et al. 2015, alces vol. 58, 2022 monitoring moose populations – moll et al. 43 steenweg et al. 2017). statistical models that estimate population density with camera trap data have also advanced (kéry and royle 2015, gilbert et al. 2021b), with several capable of estimating habitat relationships used to predict population density in non-sampled areas (moeller et al. 2018, nakashima et al. 2020), thereby enabling broader monitoring across spatiotemporal scales. camera traps can also potentially collect information on sex ratio and age structure of moose. although no study in the literature survey used camera traps, studies are being implemented in the northeastern united states (r. j. moll and h. jones, pers. comm.) and appear in recent literature (see pfeffer et al. 2018). likewise, uas surveys of wildlife populations are becoming more common due to technological advances related to field reliability and sensor capability (linchant et al. 2015, witczuk et al. 2018). recent uas surveys for moose in minnesota yielded high accuracy at a cost lower than traditional aerial surveys; albeit, scale (time and area) and flight regulations remain as issues (mcmahon et al. 2021). nonetheless, uas can provide aerial surveys at lower cost and without the potential danger of traditional flights (sasse 2003), and will be especially useful for local scale surveys (mcmahon et al. 2021). ipms represent an underused approach to improve moose monitoring by enhancing analytical power, reducing bias, and increasing the precision of population estimates (schaub and abadi 2011). for example, ipms can be used to combine known-fate data from collared individuals, counts from aerial surveys, and occupancy data collected by hunter observations to estimate population density and demographic parameters such as survival (zipkin and saunders 2018). the ultimate goals of ipms are usually to estimate abundance over time (i.e., population trends) and obtain demographic parameters that are traditionally difficult to estimate from a single data source (e.g., immigration). many moose monitoring programs seek these parameters, particularly population change over time, to inform management. to take advantage of the power of ipms, researchers and managers might first build a process-based model of the target population using standard stageor age-structured population matrices (boyce et al. 2012). then, multiple, independent datasets could be jointly analyzed to estimate parameters of that model, which would in turn be used to predict and project population trends under management scenarios. ipms require that some parameters be shared among projects with different datasets. examples of such shared parameters in moose monitoring efforts might include survival (e.g., from capture-recapture or telemetry data) and recruitment rates (e.g., from aerial counts or harvest data; zipkin and saunders 2018). while use of ipms has increased dramatically in other applications and taxa (schaub and abadi 2011, arnold et al. 2018, zipkin and saunders 2018, gilbert et al. 2021a), they are rarely used for moose (although see recent implementation by marrotte et al. 2021). we direct readers interested in using ipms for moose monitoring to the comprehensive reviews and texts published in recent years for additional details regarding data requirements and analytical implementation methods (schaub and abadi 2011, zipkin and saunders 2018, schaub and kéry 2021). moose and moose managers face a myriad of environmental and conservation challenges in the 21st century, and using accurate and reliable population information will be paramount in management decisions. the broad variation in range, habitat, environmental conditions, and population density precludes a single survey method that can monitoring moose populations – moll et al. alces vol. 58, 2022 44 address each jurisdictional goal or need. while aerial surveys are often described as the best method for population estimation and monitoring (e.g., peters et al. 2014, boyce and corrigan 2017), we suggest that the “best” method is case-specific and meets an acceptable target precision while accounting for sightability at a reasonable cost. fortunately, the technological and analytical toolboxes available to researchers and managers have never been fuller, and developments continue. we emphasize the judicious adaptation and evaluation of new methods and approaches to address context-specific needs and objectives, and likewise encourage coordinated efforts across jurisdictions and spatial scales. acknowledgements we thank the university of new hampshire and the national science foundation graduate fellowship program for supporting our research. we thank s. richard for assistance with the literature survey. partial funding was provided by the new hampshire agricultural experiment station. this is scientific contribution number 2941. this work was supported by the usda national institute of food and agriculture hatch project 1024128. we thank b. patterson, an anonymous reviewer, and b. mclaren for comments that improved the manuscript. references adams, k. p., p. j. pekins, k. a. gustafson, and k. bontaites. 1997. evaluation of infrared technology for aerial moose surveys in new hampshire. alces 33: 129–140. alberta environment and parks. 2016. aerial ungulate surveys using distance sampling techniques – protocol manual. alberta environment and parks, edmonton, alberta, canada. arnold, t. w., r. g. clark, d. n. koons, and m. schaub. 2018. integrated population models facilitate ecological understanding and improved management decisions. journal of wildlife management 82: 266–274. doi: 10.1002/ jwmg.21404 bobek, b., d. merta, p. sulkowski, and a. siuta. 2005. moose recovery plan for poland: main objectives and tasks. alces 41: 129–138. bontaites, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69–75. bormann, k. j., r. d. brown, c. derksen, and t. h. painter. 2018. estimating snow-cover trends from space. nature climate change 8: 924–928. doi: 10.1038/s41558-018-0318-3 boutin, s. 1992. predation and moose population dynamics: a critique. journal of wildlife management 56: 116–127. doi: 10.2307/3808799 boyce, m. s., p. w. j. baxter, and h. p. possingham. 2012. managing moose harvests by the seat of your pants. theoretical population biology 82: 340–347. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–89. boyce, m. s., and r. corrigan. 2017. moose survey app for population monitoring. wildlife society bulletin 41: 125–128. doi: 10.1002/wsb.732 burton, a. c., e. neilson, d. moreira, a. ladle, r. steenweg, j. t. fisher, e. bayne, and s. boutin. 2015. wildlife camera trapping: a review and recommendations for linking surveys to ecological processes. journal of applied ecology 52: 675–685. doi: 10.1111/ 1365-2664.12432 alces vol. 58, 2022 monitoring moose populations – moll et al. 45 caughley, g. 1974. bias in aerial survey. journal of wildlife management 38: 921–933. doi: 10.2307/3800067 crete, m., l. rivest, h. jolicoeur, j. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose with observations made from fixed-wing aircraft in southern quebec. journal of applied ecology 23: 751–761. doi: 10.2307/2403931 decesare, n. j., j. r. newby, v. j. boccadori, t. chilton-radandt, t. their, d. waltee, k. podruzny, and j. a. gude. 2016. calibrating minimum counts and catch-per-unit-effort as indices of moose population trend. wildlife society bulletin 40: 537–547. doi: 10.1002/wsb.678 evans, c. d., w. a. troyer, and c. j. lensink. 1966. aerial census of moose by quadrat sampling units. journal of wildlife management 30: 767–776. doi: 10.2307/3798283 field, s. a., p. j. o’connor, a. j. tyre, and h. p. possingham. 2007. making monitoring meaningful. austral ecology 32: 485–491. doi: 10.1111/j.1442-9993. 2007.01715.x fletcher, r. j., and m.-j. fortin. 2018. spatial ecology and conservation modeling: applications with r. springer nature switzerland, cham, switzerland. fournier, a. m. v., e. r. white, and s. b. heard. 2019. site-selection bias and apparent population declines in longterm studies. conservation biology 33: 1370–1379. doi: 10.1111/cobi.13371 franzmann, a. w., and c. c. schwartz, editors. 2007. ecology and management of the north american moose. 2nd edition. university press of colorado, boulder, colorado, usa. frid, a., and l. dill. 2002. human-caused disturbance stimuli as a form of predation risk. ecology and society 6: 11. doi: 10.5751/es-00404-060111 fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52: 14–21. doi: 10.2307/3801050 gasaway, w. c., s. d. dubois, and s. j. harbo. 1985. biases in aerial transect surveys for moose during may and june. journal of wildlife management 49: 777–784. doi: 10.2307/3801711 _____., _____, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska 22: 1–99. gilbert, n. a., b. s. pease., c. m. anhaltdepies, j. d. j. clare, j. l. stenglein, p. a. townsend, t. r. van deelen, and b. zuckerberg. 2021a. integrating harvest and camera trap data in species distribution models. biological conservation 258: 109147. doi: 10.1016/j. biocon.2021.109147 _____, j. d. j. clare, j. l. stenglein, and b. zuckerberg. 2021b. abundance estimation of unmarked animals based on camera-trap data. conservation biology 35: 88–100. doi: 10.1111/cobi.13517 harris, r. b., m. atamian, h. ferguson, and i. keren. 2015. estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty. alces 51: 57–69. hatter, i. w. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces: 37: 71–77. hinton, j. w., r. e. wheat, p. schuette, j. e. hurst, d. w. kramer, j. h. stickles, and j. l. frair. 2022. challenges and opportunities for estimating abundance of a low-density moose population. the journal of wildlife management 86: e22213. doi: 10.1002/jwmg.22213 jensen, w., r. v. rea, c. e. penner, j. r. smith, e. v. bragina, e. razenkova, l. balciauskas, h. bao, s. bystiansky, s. monitoring moose populations – moll et al. alces vol. 58, 2022 46 csányi, z. chovanova, g. done, k. hackländer, m. heurich, g. jiang, a. kazarez, j. pusenius, e. j. solberg, r, veeroja, and f. widemo. 2020. a review of circumpolar moose populations with emphasis on eurasian moose distributions and densities. alces 56: 63–78. joly, k., p. a. duffy, and t. s. rupp. 2012. simulating the effects of climate change on fire regimes in arctic biomes: implications for caribou and moose habitat. ecosphere 3: art36. doi: 10.1890/ es12-00012.1 jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. kantar, l. e., and r. e. cumberland. 2013. using a double-count aerial survey to estimate moose. alces 49: 29–37. kellie, k. a., k. e. colson, and j. h. reynold. 2019. challenges to monitoring moose in alaska. alaska department of fish and game, juneau, alaska, usa. kéry, m., b. gardner, and c. monnerat. 2010. predicting species distributions from checklist data using site-occupancy models. journal of biogeography 37: 1851–1862. doi: 10.1111/j.1365-2699. 2010.02345.x kéry, m., and j. a. royle. 2015. applied hierarchical modeling in ecology: analysis of distribution, abundance and species richness in r and bugs. volume 1. prelude and static models. elsevier, san diego, california, usa. krausman, p. r. 2002. introduction to wildlife management. pearson, upper saddle river, new jersey, usa. krebs, c. j. 2006. mammals. pages 351– 369 in w. sutherland, editor. ecological census techniques: a handbook. 2nd edition. cambridge university press, cambridge, england. krester, h., m. glennon, a. whitelaw, a. hurt, k. pilgrim, and m. schwartz. 2016. scat-detection dogs survey low density moose in new york. alces 52: 55–66. larson, l. r., r. c. stedman, d. j. decker, w. f. siemer, and m. s. baumer. 2014. exploring the social habitat for hunting: toward a comprehensive framework for understanding hunter recruitment and retention. human dimensions of wildlife 19: 105–122. doi: 10.1080/10871209.2014.850126 leopold, a. 1933. game management. charles scribner’s sons, new york, new york, usa. linchant, j., j. lisein, j. semeki, p. lejeune, and c. vermeulen. 2015. are unmanned aircraft systems (uass) the future of wildlife monitoring? a review of accomplishments and challenges. mammal review 45: 239–252. doi: 10.1111/ mam.12046 lyons, j. e., m. c. runge, h. p. laskowski, and w. l. kendall. 2008. monitoring in the context of structured decision-making and adaptive management. journal of wildlife management 72: 1683–1692. doi: 10.2193/2008-141 mackenzie, d. i., j. d. nichols, g. b. lachman, s. droege, j. a. royle, and c. a. langtimm. 2002. estimating site occupancy rates when detection probabilities are less than one. ecology 83: 2248–2255. doi: 10.1890/0012-9658 (2002)083[2248:esorwd]2.0.co;2 månsson, j., c. e. hauser, h. andrén, and h. p. possingham. 2011. survey method choice for wildlife management: the case of moose alces alces in sweden. wildlife biology 17: 176–190. doi: 10.2981/10-052 marrotte, r. r., b. r. patterson, and j. m. northrup. 2021. harvest and density-dependent predation drive long-term population decline in a northern ungulate. ecological applications e2629. alces vol. 58, 2022 monitoring moose populations – moll et al. 47 doi: 10.22541/au.162210989.948595 93/v1 mcmahon, m. c., m. a. ditmer, e. j. isaac, s. a. moore, and j. d. forester. 2021. evaluating unmanned aerial systems for the detection and monitoring of moose in northeastern minnesota. wildlife society bulletin: 45: 312–324. doi: 10.1002/wsb.1167 messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478–488. doi: 10.2307/1939551 michaud, j. s., n. c. coops, m. e. andrew, m. a. wulder, g. s. brown, and g. j. m. rickbeil. 2014. estimating moose (alces alces) occurrence and abundance from remotely derived environmental indicators. remote sensing of environment 152: 190–201. doi: 10.1016/j.rse.2014.06.005 moeller, a. k., p. m. lukacs, and j. s. horne. 2018. three novel methods to estimate abundance of unmarked animals using remote cameras. ecosphere 9: e02331. doi: 10.1002/ecs2.2331 montgomery, r. a., k. m. redilla, r. j. moll, b. van moorter, c. m. rolandsen, j. j. millspaugh, and e. j. solberg. 2019. movement modeling reveals the complex nature of the response of moose to ambient temperatures during summer. journal of mammalogy 100: 169–177. doi: 10.1093/jmammal/gyy185 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. bartnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. doi: 10.2193/0084-0173 (2006)166[1:pndaci]2.0.co;2 musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185–204. doi: 10.2981/09-014 nakashima, y., s. hongo, and e. f. akomookoue. 2020. landscape-scale estimation of forest ungulate density and biomass using camera traps: applying the rest model. biological conservation 241: 108381. doi: 10.1016/j.biocon.2019.108381 nygrén, t., and m. pesonen. 1993. the moose population (alces alces l.) and methods of moose management in finland, 1975–89. finnish game research: 45: 45–53. oyster, j. h., i. n. keren, s. j. k. hansen, and r. b. harris. 2018. hierarchical mark-recapture distance sampling to estimate moose abundance. journal of wildlife management 82: 1668–1679. doi: 10.1002/jwmg.21541 peek, j. a. 1974. initial response of moose to a forest fire in northeastern minnesota. the american midland naturalist 91: 435–438. doi: 10.2307/ 2424334 pekins, p. j. 2020. metabolic and population effects of winter tick infestations on moose: unique evolutionary circumstances? frontiers in ecology and evolution 8: 1–13. doi: 10.3389/ fevo.2020.00176 peters, w. m. hebblewhite, k. g. smith, s. m. webb, n. webb, m. russell, c. stambaugh, and r. b. anderson. 2014. contrasting aerial moose population estimation methods and evaluating sightability in west-central alberta, canada. wildlife society bulletin 38: 639–649. doi: 10.1002/wsb.433 pfeffer, s. e., r. spitzer, a. m. allen, t. r. hofmeester, g. ericsson, f. widemo, n. j. singh, and j. p. g. m. cromsigt. 2018. pictures or pellets? comparing camera trapping and dung counts as methods for estimating population densities of ungulates. remote sensing in ecology and conservation 4: 173–183. doi: 10.1002/rse2.67 monitoring moose populations – moll et al. alces vol. 58, 2022 48 priadka, p., g. s. brown, b. r. patterson, and f. f. mallory. 2020. sex and age-specific differences in the performance of harvest indices as proxies of population abundance under selective harvesting. wildlife biology 2020: 1–11. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timbermanagement and natural-disturbance effects on moose habitat: landscape evaluation. the journal of wildlife management 61: 517–524. doi: 10.2307/3802610 rota, c. t., r. j. fletcher, r. m. dorazio, and m. g. betts. 2009. occupancy estimation and the closure assumption. journal of applied ecology 46: 1173–1181. doi: 10.1111/j.1365-2664.\2009.01734.x sample, c., r. v. rea, and g. hesse. 2020. tracking mooseand deer-vehicle collisions using gps and landmark inventory systems in british columbia. alces 56: 49–61. sasse, d. b. 2003. job-related mortality of wildlife workers in the united states, 1937–2000. wildlife society bulletin 31: 1015–1020. schaub, m., and f. abadi. 2011. integrated population models: a novel analysis framework for deeper insights into population dynamics. journal of ornithology 152: s227–s237. doi: 10.1007/s10336 010-0632-7 _____, and m. kéry. 2021. integrated population models: theory and ecological applications with r and jags. academic press, cambridge, massachusetts, usa. silverberg, j. k., p. j. pekins, and r. a. robertson. 2003. moose responses to wildlife viewing and traffic stimuli. alces 39: 153–160. silvy, n. j., editor. 2012. the wildlife techniques manual: management. the johns hopkins university press, baltimore, maryland, usa. sinclair, a. r. e., j. m. fryxell, and g. caughley. 2006. wildlife ecology, conservation, and management. 2nd edition. blackwell publishing, malden, massachusetts, usa. singh, n. j., and e. j. milner-gulland. 2011. monitoring ungulates in central asia: current constraints and future potential. oryx 45: 38–49. doi: 10.1017/ s0030605310000839 skalski, j. r., k. e. ryding, and j. j. millspaugh. 2005. wildlife demography: analysis of sex, age, and count data. elsevier academic press, new york, new york, usa. solberg, e. j., b.-e. saethert, o. strand, and a. loison. 1999. dynamics of a harvested moose population in a variable environment. journal of animal ecology 68: 186–204. doi: 10.1046/j.1365-2656.1999.00275.x steenweg, r., m. hebblewhite, r. kays, j. ahumada, j. t. fisher, c. burton, s. e. townsend, c. carbone, j. m. rowcliffe, j. whittington, j. brodie, j. a. royle, a. switalski, a. p. clevenger, n. heim, and l. n. rich. 2017. scaling-up camera traps: monitoring the planet’s biodiversity with networks of remote sensors. frontiers in ecology and the environment 15: 26–34. doi: 10.1002/ fee.1448 steidl, r. j., j. p. hayes, and e. schauber. 1997. statistical power analysis in wildlife research. the journal of wildlife management 61: 270–279. doi: 10.2307/3802582 stephens, p. a., o. y. zaumyslova, d. g. miquelle, a. i. myslenkov, and g. d. hayward. 2006. estimating population density from indirect sign: track counts and the formozov-malyshev-pereleshin formula. animal conservation 9: 339– 348. doi: 10.1111/ j.1469-1795.2006. 00044.x storaas, t., h. gundersen, h. henriksen, and h. andreassen. 2001. the economic alces vol. 58, 2022 monitoring moose populations – moll et al. 49 value of moose – a review. alces 37: 9–107. tape, k. d., d. d. gustine, r. w. ruess, l. g. adams, and j. a. clark. 2016. range expansion of moose in arctic alaska linked to warming and increased shrub habitat. plos one 11: 1–12. doi: 10.1371/journal.pone.0160049 timmermann, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29: 35–46. _____, and m. e. buss. 2007. population and harvest management. pages 559– 616 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. 2nd edition. university press of colorado, boulder, colorado, usa. _____, and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. tyre, a. j., b. tenhumberg, s. a. field, d. niejalke, k. parris, and h. p. possingham. 2003. improving precision and reducing bias in biological surveys: estimating false-negative error rates. ecological applications 13: 1790–1801. doi: 10.1890/02-5078 ueno, m., e. j. solberg, h. iijima, c. m. rolandsen, and l. e. gangsei. 2014. performance of hunting statistics as spatiotemporal density indices of moose (alces alces) in norway. ecosphere 5: art13. doi: 10.1890/ es13-00083.1 van ballenberghe, v., and w. b. ballard. 2007. population dynamics. pages 223– 246 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. 2nd edition. university press of colorado, boulder, colorado, usa. ver hoef, j. m. 2008. spatial methods for plot-based sampling of wildlife populations. environmental and ecological statistics 15: 3–13. doi: 10.1007/ s10651-007-0035-y white, e. r. 2019. minimum time required to detect population trends: the need for long-term monitoring programs. bioscience 69: 26–39. doi: 10.7287/ peerj.preprints.3168v4 witczuk, j., s. pagacz, a. zmarz, and m. cypel. 2018. exploring the feasibility of unmanned aerial vehicles and thermal imaging for ungulate surveys in forests – preliminary results. international journal of remote sensing 39: 5504–5521. doi: 10.1080/01431161.2017.1390621 zipkin, e. f., and s. p. saunders. 2018. synthesizing multiple data types for biological conservation using integrated population models. biological conservation 217: 240–250. doi: 10.1016/j.biocon.2017.10.017 p129-138_3931.pdf alces vol. 41, 2005 bobek et al. moose recovery plan in poland 129 a moose recovery plan for poland: main objectives and tasks 1 1 2 3 1department of ecology, wildlife research, and ecotourism, pedagogical university of cracow, podbnezie 3, 30-054 krakow, poland; 2polish hunting association, nowowiejska 10, 00-653 warsaw, poland; 3department of animal nutrition, agricultural university, mickiewicza 24/28, 30-059 kraków, poland abstract: hunting statistics showed that moose (alces alces) numbers in poland declined from 5,400 animals in 1991 to 1,718 in 2000. a nation-wide ban on moose hunting was imposed in 2001 in response to this decline in moose abundance. the main purpose of this paper is to outline a moose collecting data on population demographic variables, and understanding moose habitat preferences. during 1998-2002 in the forest habitat of north-eastern poland (total area: 311,400 ha) a line intercept snow track index and plot sampling were used to estimate moose population numbers at 276 animals. it was shown that the population census in this area carried out by hunters in this period through a guess-estimate method overestimated the moose population by 46.0%. research in augustowska forest (110,200 ha) shows that the autumn recruitment rate was 64.4 calves per 100 cows, and the ratio of cows to bulls was 1.34. analysis of moose population dynamics during 4 hunting seasons (1998-2001) shows that the maximum sustainable harvest is about 30% of population numbers estimated in february. preferred habitats in bog and wet sites dominated by deciduous and mixed forests. the decline in moose populations in poland over 20 years was caused by overestimation of population numbers and over-harvest. it is suggested that a moose recovery program in poland should be started by locating 2 large moose management/conservation units where moose population numbers should be estimated by reliable methods, and sustained harvest would then maintain a viable moose population. at the same time, forestry in moose wintering areas should stimulate deciduous browse production as well as providing estimates of forest damage caused by moose using different standards than those applied in lowland commercial forests. alces vol. 41: 129-138 (2005) key words: habitat selection, moose, over-harvest, poland, population census, recovery plan, recruitment rate moose (alces alces) in poland experienced ing the 20th century. world war i drastically reduced the moose population in the country, and only the joint efforts of wildlife biologists, foresters, and hunters allowed its restoration. by 1937 there were estimated to be 1,130 moose in poland. world war ii again threatened the existence of the moose population, leaving only 10-15 animals by 1945. the species was then declared protected, and an intensive reintroduction program was undertaken using a large enclosure in kampinos hunting statistics, the population was estimated at 425 animals in 1965 and 3,250 in 1975. harvesting of moose started in 1967, with progressively increasing numbers of animals harvested in each year. moose continued to increase however, and the population was estimated at 6,181 individuals in 1981. winter concentrations of moose in some local forest areas caused over-browsing of tree species considered desirable by the timber industry. the harvest quota was therefore increased to 1,115 animals for the 1981/82 hunting season moose recovery plan in poland bobek et al. alces vol. 41, 2005 130 and 1,613 in the 1982/83 season. this harvest reduced the population to 4,900 moose in march 1983 (bobek and morow 1987). then, during the next 6 hunting seasons, the harvest quotas were reduced to 1,250 animals per year. during 1987-1991 hunting statistics showed an increase in numbers of moose from 4,100 animals to 5,400 individuals. therefore during 1989-91 hunting seasons, the harvest quota was increased to approximately 1,660 animals per year. this hunting pressure badly affected population dynamics of moose over the next decade. hunting statistics showed a serious decrease of moose numbers to 1,718 animals in 2000. acmoose in poland in march 2000, 800 of which parków narodowych 2000). a nation-wide ban on moose hunting was imposed in 2001 in response to this decline in moose abundance. it is expected that the ban will remain in effect until moose numbers recover. the main purpose of this project is to outline a recovery plan for moose in poland. the proposal has three main objectives: (1) to moose population numbers in particular forest areas – these areas constitute the core habitat of the moose population in north-eastern poland; (2) estimation of moose population demographic variables necessary to manage a sustainable harvest; and (3) understanding moose habitat preferences. study area research on the moose population was carried out in 4 large forests (fig. 1) situated in north-eastern poland: augustowska forest (110,200 ha), borecka forest (63,800 ha), piska forest (124,500 ha), and romincka forest (13,100 ha). the augustowska and piska forests are typical lowland forests where scots pine (pinus silvestris) is a dominant tree species. the terrain of borecka and romincka forests is hilly and the main tree species are scots pine, spruce (picea excelsa), and oak (quercus robur). the early glacial landscape moraines and numerous lakes, including œniardwy lake, the largest in poland. bison (bison bonasus), moose (alces alces), red deer (cervus elaphus), wild boar (sus scrofa), and roe deer (capreolus capreolus) are the largest wild ungulates in the region. well-established populations of wolves (canis lupus) and lynx (lynx lynx) occur in the eastern part of the region. the climate of the study area is continental, with snow cover during 75-100 days per year (lencewicz and kondracki 1964, kondracki 1998). methods moose abundance moose density and numbers were estimated using the “carpathian method” (bobek et al. 2001) for comparison with estimates carpathian method is based on the relationship between absolute population density (n/1,000 ha), as an independent variable, and a snow track density index (t/km/day) as a dependent variable. this relationship was measured using 44 sampling plots of 400-500 ha each in the augustowska forest. during winter 1998 a team of 6 trackers, working in pairs, and 3540 observers were used to sample the plots. plot sampling began with 2 trackers marking and clearing all the moose tracks along line transects inside the plot. the following day, 2 pairs of trackers searched the perimeter of the plot and cleared all moose tracks encountered. next, 2 trackers counted all new moose tracks left along the same transect during the previous 24 hours. after this had been done, observers entered the plot and each person searched 10-15 ha per hour, recording all moose seen. observers recorded the sex, age, group size, and sex-age composition of the group for all moose seen. observers also recorded the exact time of each observation and the direction of alces vol. 41, 2005 bobek et al. moose recovery plan in poland 131 the animals’ escape. after the observers had completed their search, 2 pairs of trackers retraced their path around the perimeter of the plot, recording all new tracks left by moose entering or leaving the plot, along with the location and the number of tracks. the observations of the trackers and observers were compiled on one map. timespace analysis was used to avoid double counting of moose, estimate the number of animals entering and leaving the plot, and the plot during the count period. the number of animals in the sampling plots was converted into population density (n1/1,000 ha) and the results were regressed against track density (t1/km/day) recorded along the track transects located in the plots. inverse prediction, using the derived regression equation, then allowed us to calculate moose population density (y) from the track density index (x). the carpathian method was subsequently used to estimate moose density in the augustowska, borecka, and romincka forests in february 1998 and in the piska forest in february 2002. fresh tracks (left during the previous 24 hours) were recorded along the line transects laid out along forest roads passable by trackers in off-road vehicles (at least 50 km per 10,000 ha of forest) during 5 consecutive days. the average daily track density (t2/km/day) was used to estimate moose density (n2/1,000 ha) and the number of moose per forest inventory unit (1,000-2,000 ha). the number of moose calculated from the inventory units were summed separately fig. 1. moose study areas in northeast poland during 1998-2002. 1 – augustowska forest, 2 – borecka forest, 3 – piska forest, 4 – romincka forest. baltic sea germany czech republik slovakia ukraine byelorus lithuania russia 1 2 3 4 warsaw moose recovery plan in poland bobek et al. alces vol. 41, 2005 132 for each of the 4 forest research areas. moose abundance estimated for each forest area using the carpathian method was compared with the traditional guess-estimates developed for the same areas by hunters and the forest service. during the following 4 years, estimation of moose population numbers through the new method was continued in augustowska forest. results of the population census were accepted by the local forest and wildlife service and used to calculate harvest quotas. population composition and harvest population sex ratio and the number of calves per female in autumn were obtained by direct observation of animals in augustowska forest during the years 1999 and 2000. the observations were carried out by the authors and by well-experienced game managers from the local forest service. the forest and wildlife service provided data on number, sex, and age of harvested moose in augustowska forest. habitat selection to determine moose habitat selection, track transect lines along roads in augustowska forest were precisely drawn on forest maps. using a car odometer and forest maps, the location of all moose tracks were recorded together with forest types and age classes (10or 20-year intervals) in which they occurred. then, using forest maps, the length of forest types and forest age classes along the line transects were measured and ability of forest habitats. habitat selection by moose was tested according to cherry (1996) intervals (bailey 1980). population simulations computer simulations were used to gain a better understanding of moose population dynamics in poland (bobek 2002). simulaabundance estimates. the annual recruitment rate was assumed to be 30% of the population size in february; i.e., after the hunting season. this assumption is based on reliable moose abundance and harvest data for augustowska forest. this forest area is part of the most important moose habitat in north-eastern poland. the recruitment rate of the moose population in augustowska forest can also be considered as representative of the second core area of moose population in polesie lubelskie (eastcentral poland), because of a similar habitat, climate, and predation. the simulation represents 2 different separated periods: 1981-92 and 1995-2001. these periods cannot be compared because some of the hunting districts existing from 1992 to 1995, where a large portion of the moose population lived, were discontinued when biebrza national park, polesie national park, and narew national park were created. results relationship between track indices and moose abundance (carpathian method) moose and track counts were conducted in a total of 44 sampling plots (the total area of which was 21,410 ha) in the augustowska forest (table 1). moose were observed in 12 plots and tracks were recorded in 3 plots where no moose were seen. no moose or tracks were recorded in the remaining 29 plots and these were not included in our regression analysis. the moose density ranged from 0 to 8.25 animals per 1,000 ha (fig. 2). the track index varied between 0.00 and 0.90 tracks/km/day. the relationship between moose population density (n/1,000 ha) and track density index mula: t/km/day = 0.20 * tan (0.14*n/1000 ha) r2 = 0.56, n = 15. alces vol. 41, 2005 bobek et al. moose recovery plan in poland 133 inverse prediction, using the derived regression equation, allowed calculation of population density from the track density index according to the following formula: y = (7.17) * arctan (5.07 * x) r2 = 0.56, n = 15, where, y is population density (n/1000 ha) and x is the track index (tracks/km/day). e s t i m a t i o n o f m o o s e p o p u l a t i o n numbers the relationship between tracks and moose density developed for the augustowska forest was used to estimate moose numbers in the 4 forest areas studied (table 2). in addition to the estimated 180 in the augustowska forest, there were 28 moose in the borecka forest, 56 in the piska forest, and 12 in the romincka forest, for a total estimate of 276 animals in the 4 forests studied. average population density was 0.88 moose/1,000 ha and ranged from 0.44/1,000 ha in the borecka and romincka forests to 1.63/1,000 ha in augustowska forest. moose population estimates based on the carpathian method were much lower than table 1. estimation of moose population size by the line intercept track index in augustowska forest (110, 200 ha), north-eastern poland. data were collected by using 225 line transects of a total length of 735 km in february 1998. days of tracking 1 2 3 4 5 mean number of tracks 37 32 36 35 34 34.8 tracks/km/day 0.050 0.043 0.049 0.048 0.046 0.047 population density (n/1000 ha) 1.74 1.51 1.68 1.67 1.56 1.63 population size (n) 192 167 185 184 172 1801 1 fig. 2. relationship between snow track index (x) and population density (y) of moose in augustowska forest. y = (7.17)*arctan(5.07*x), r2 = 0.56, n = 15. -1 0 1 2 3 4 5 6 7 8 9 10 -0,05 0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35 moose recovery plan in poland bobek et al. alces vol. 41, 2005 134 guess-estimates prepared by the forest wildlife service and the polish hunting associahunting statistics consistently overestimated moose abundance in the studied forests by an average of 46.0%. overestimates in individual forests ranged from 8.2% in the piska forest to 128.6% for the borecka forest. moose population dynamics in the augustowska forest of 227 moose observed in augustowska forest in the autumns of 1999 and 2000, 41.7% were cows, 31.7% bulls, and 26.8% calves. the autumn recruitment rate was 64.4 calves per 100 cows and the ratio of cows per one bull was 1.34. according to harvest data, 39.2% of the 166 moose harvested in augustowska forest between 1998 and 2001 were cows, 32.5% were bulls, and 28.3% calves. hunting pressure on bulls and calves was therefore slightly higher than their occurrence in the population. a harvest rate of 31-33% of the winter population caused a small decline in population numbers (table 3). a 25.0% harvest rate led to an increase in population numbers. it is evident that the ban on moose hunting in the 2001/2002 season caused a 31.0% increase in moose numbers in this area. habitat selection the spatial distribution of 348 sets of erences for forest habitats occupying wet and bog sites, which were dominated by deciduous, mixed deciduous, and mixed coniferous forest types (table 4). moose avoided fresh (i.e., mesic soil conditions) coniferous forest types. the other 5 forest types; fresh deciduous, fresh mixed deciduous, fresh mixed coniferous, wet, and bog coniferous were used in proportion to their availability. young forest age classes (11-20 years old) and stands older than 100 years were preferred by moose, while age class 81-100 years was avoided (table 4). discussion mates of moose abundance, developed by the forest wildlife service and the polish hunting association, for our 311,600 ha study area were too high. one likely reason is the non-objective (guess-estimate) nature of the round observations and the personal intuition of game managers. another related reason may be the seasonal migratory patterns of moose in the region. in poland, moose winter mainly pielowski 1979), but during spring and summer a substantial part of the moose population name of forest year forest area sampled (ha x 103) population density (moose/1000 ha) population size (n) a b augustowska 1998 110.2 1.63 180 (67-191) 263 borecka 1998 63.8 0.44 28 (23-33) 64 piska 2002 124.5 0.45 56 (46-66) 61 romincka 1998 13.1 0.92 12 (11-13) 15 total/mean 311.6 0.881 276 403 1 weighted mean. table 2. moose population estimates and densities estimated by (a) the line intercept track index and intervals are given in parentheses. alces vol. 41, 2005 bobek et al. moose recovery plan in poland 135 districts, migrating moose may be mistaken for resident individuals. it is highly probable that similar mistakes were made throughout poland. these errors led to overestimates of moose abundance, overharvesting, and ultimately, declines in moose population simulation results (bobek 2002) indicate that if there had been 6,200 moose in poland in 1981, the harvest of 1,115 and 1,613 animals in the 1981/82 and 1982/83 hunting seasons would not have caused a destatistics claimed (fig. 3). there were probably no more than 4,400 moose in poland in 1982. according to population simulation results, the harvest during the 1980s caused only a small decline of moose numbers in poland. during the late 1980s and 1990s, harvesting 1,650-1,670 animals per year resulted in a large decline in the moose population, which there was an increase in the moose population in poland (fig. 3). it is likely that at the end of moose hunting season in march 2000 there were only about 800 moose in all hunting estimate of 1,917 (fig. 3). taking into consideration moose occurring in national parks, the total moose population in poland probably numbers about 1,250 animals. the goal of the polish moose recovery program is to create both a viable moose population and a harvest strategy designed to maintain a stable population. however, the lack of research on moose in poland is a serious problem. it is therefore highly recommended that a national research program be created. it should include the following measures: 1. establish 2 core areas that could be used for both management and conservation of the moose population. they should include wetland, which is preferred by moose during the growing season, as well as forest habitats, which are moose wintering areas. in north-eastern poland, such conditions occur in 240,000 ha of the wetlands of biebrza national park and neighbouring large forest areas. the other core area, polesie lubelskie (about 110,000 ha), should be located at the polish-ukrainian border and include the wetlands of polesie national park and adjacent forest areas. 2. in 2 areas, viable moose population size must be calculated. a sustainable harvest may take place only outside of national parks and it should be lower than recruitment rate. these 2 populations will act as source populations for rebuilding the moose population elsewhere in poland. moose population objectives for these areas should be developed with input from the polish hunting association, the state year population size annual harvest (n3) harvest rate population growth rate n1 n2 n3 / n1 n3 / n2 n1 /n 1998 180 246 59 0.33 0.24 0.98 1999 176 240 44 0.25 0.18 1.11 2000 195 266 63 0.32 0.24 0.93 2001 182 249 0 0.00 0.00 1.31 2002 238 325 table 3. annual variation in the moose population and harvest in the augustowska forest in northeastern poland. n1 is the population size estimated using the line intercept track index in february. n2 is the calculated population size before hunting season in september. n /n is the population growth rate. moose recovery plan in poland bobek et al. alces vol. 41, 2005 136 table 4. winter habitat selection by moose in augustowska forest estimated by spatial distribution of soil conditions. category of habitat proportion of tracks available bailey’s 95% in habitat proportion of simultaneous forest types fresh coniferous 0.433 0.546 0.358-0.509 (-) fresh mixed coniferous 0.279 0.251 0.212-0.349 wet mixed coniferous 0.069 0.047 0.035-0.114 fresh deciduous 0.014 0.016 0.002-0.041 fresh mixed deciduous 0.054 0.045 0.025-0.096 wet and bog coniferous 0.032 0.021 0.010-0.066 and wet mixed deciduous alder 0.037 0.028 0.014-0.074 forest meadows 0.029 0.031 0.009-0.062 forest age class (years) 0-10 0.035 0.026 0.011-0.068 21-40 0.232 0.228 0.161-0.290 41-6 0.232 0.248 0.161-0.290 61-80 0.227 0.254 0.156-0.284 81-100 0.109 0.167 0.066-0.164 (-) forest service, and national parks. it will are an attractive target for hunters and ecotourists. at the same time, they cause serious damage to young forest plantations. therefore, it would be necessary to establish moose wintering areas that would be mainly deciduous and mixed deciduous forests growing on wet and bog sites. in the moose wintering areas, acceptable levels of forest damage caused by moose should be much higher than in lowland commercial forest. forest management in winter moose concentration areas should result in an increase in the biomass of winter browse and should be based on natural forest succession. 3. after the moose population targets for these areas have been reached, harvest quotas should be based on reliable population estimates and annual population recruitment rates. estimation of populaalces vol. 41, 2005 bobek et al. moose recovery plan in poland 137 tion numbers should be based on snow tracking along line transects. these line transects have been successfully used for a few years for estimating the numbers of big game animals in north-eastern poland. it appears that the maximum sustainable harvest under polish conditions is around 30% of population numbers estimated after each hunting season. this is higher than that applicable to the moose population inhabiting scandinavia (solberg et al. 1999) and similar to that for moose living in the baltic countries (baleishis et al. 1998). polish guidelines currently recommend a high harvest rate for calves and young bulls. this policy has caused the disappearance of males older than 10 years, which are an attractive target for hunters and ecotourists. appropriate proportions of calves and young males in the harvest should be decided on through computer simulation models. these models should include the age structure of males. acknowledgements this research was supported by the international institute of ecology, limited, and grant g-1328/kzz/00-03 from agricultural university, kraków, poland. references bailey, b. j. r. 1980. large simultaneous probabilities based on transformation of cell frequencies. technometrics 22:583589. baleishis, r., p. bluzma, a. ornicans, and j. tonisson. 1998. the history of moose in baltic countries. alces 34:339-345. bobek, b , merta, r. paszkiewicz, a. pawlak, m. and t.zajac. polski 1:17-21. numbers inhabiting hunting districts, population harvest and result of simulation moose population dynamics (bobek et al. 2002). for more explanation see text. 6,200 4,800 1,917 4,414 2,438 800 1,6501,6701,610 0 1000 2000 3000 4000 5000 6000 7000 1980 1985 1990 1995 2000 years nu m be r o f a ni m al s hunting statistics results of simulation harv est moose recovery plan in poland bobek et al. alces vol. 41, 2005 138 , and k. morow. 1987. present status of the moose (alces alces) in poland. swedish wildlife research supplement 1:69-70. , and p. su kowski. 2002. moose recovery program in poland. page 22 in abstracts 5th international moose conference oyer, august 4-9, 2002. cherry, s. dence interval methods for habitat useavailability studies. journal of wildlife management 60:653-658. owski, r., and z. pielowski. 1979. , and 1992. dynamics and management of moose population in the biebrza river valley. pages 631-635 in b. bobek, k. perzanowski, and w.l. regelin, editors. global trends in wildlife management. proceedings kraków-warszawa, poland. , and . 1998. dlaczego 11:12-14. kondracki, j. 1998 polski. pwn warszawa, poland. krajowy zarzad parków narodowych. dowych. warszawa, poland. lencewicz, s., and j. kondracki. 1964. gepoland. morow, k. 1975. moose population characteristic and range use in augustowska forest. ekologia polska 3:493-506. 1969. reproduction and dynamics of moose (alces alces l.) population in the kampinos national park. ekologia polska a. 17(37):709-718. and a. loison. 1999. dynamics of a harvested moose population in a variable environment. journal of animal ecology 68:186-204. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice p63-74_4015.pdf alces vol. 41, 2005 selby et al. threat to sustainability 63 threats to the sustainability of moose management in finland ashley selby, leena petäjistö, and terhi koskela finnish forest research institute, unioninkatu 40a, 00170 helsinki, finland abstract: the large population of moose (alces alces l.) in finland has resulted in increased demographic and socio-economic conditions. pre-conditions for hunting club membership occur in and moose hunting. alces vol. 41: 63-74 (2005) key words: the moose (alces alces l.) is the largand its population in finland has increased estimated pre-hunting season number in 2002 was about 180,000 animals, and the posthunting population in winter 2002-2003 was 113,000 – 125,000 animals (finnish game and fisheries research institute 2004) (fig. 1). the desired size of the moose population, owners for moose damage (e.g., helle et al. 1987), and measures to reduce moose-related road accidents (haikonen and summala 2000) are topics of public concern in finland. the knowledge of the interested parties – the debate to address this lack of knowledge, the finnish forest research institute has conthe main interested parties. first, the structure 100,000 hunters. a mean input of 16.5 huntthreat to sustainability – selby et al. alces vol. 41, 2005 64 until now there has been a strong correlation between the number of moose to be decline in farm numbers and the weakening demographic situation in rural areas, this paper were not interested in taking part in moose the moose hunters wished that the hunt was not so time-consuming. 2003), 60% of forest owners were found to because of moose damage. forest owners who were also moose hunters were of the opinion that the moose population was too large, and damage in their forests. forest owners who concerning control of the moose population the number of annual moose hunting permits is determined. opinion that the moose population was too large. one-third of respondents considered that the moose population could be 20% less than in winter 2004 (95,000), while 10% of respondents considered that the moose popu0 20000 40000 60000 80000 100000 120000 140000 19 80 -1 19 82 -3 19 84 -5 19 86 -7 19 88 -9 19 90 -1 19 92 -3 19 94 -5 19 96 -7 19 98 -9 20 00 -1 20 02 -3 moose winter population moose harvest alces vol. 41, 2005 selby et al. threat to sustainability 65 ment associations. one of the few remaining social institutions in rural finland. the continuing decline in the and the aging of the remaining rural population trends in finnish municipalities for the period 1975 – 2030. apart from the aging of the rural population concomitant with a low birth rate and out-migration, a striking feature of the demographics is the concentration of the future population in the region surrounding the capito other urban centres. natural population growth and the effects of migration will cause the rural areas to lose between 20 and 40% of their population during the period 2000 – 2030. the relationship between these demographic effects and the current distribution of moose hunting clubs is shown in figure 2. the areas with the greatest concentration of moose the greatest effects of demographic change populated areas and 15% of the population and haapanen 2002). the international trade in agricultural goods has forced radical changes in finnish farms has declined from about 130,000 to remaining farms could be as few as 40,000 (niemi and pietola 2005). changes in farming and other economic and demographic changes in rural areas will also affect the forest ownership structure that in turn can affect moose management. in addition, the widespread est regeneration on abandoned farmland that is associated with the decline of farming and ideal browsing conditions for moose (heikkilä now in the 60+ age-group, while a further 45% are in the 40 – 59 age-group. these age other than that in which their forest is located. tance of 125 km to their forests (karppinen et al. 2002). institutional arrangements for moose hunting. bership of moose hunting clubs in 2002 and the assessment is made from the standpoint of club leaders. club leaders are responsible threat to sustainability – selby et al. alces vol. 41, 2005 66 to the issue of moose hunting permits. it is methods 5,200 moose hunting clubs that submitted reports for the 1999 moose hunting season. tionnaire was not mailed again to clubs that geneous groups of cases based on selected club membership class percentages formed the cluster membership was assigned to each case was applied. this test is a measure of how is that there is no association between row and test has been applied to cell counts. results membership structure the largest single group are the local 10% of members. the two smallest groups were non-local landowners (about 10%) and -150000 -100000 -50000 0 50000 100000 150000 200000 250000 300000 1 2 3 4 0 10 20 30 40 50 60 population growth and migration effects moose hunting club distribution 1: sparsely populated areas 2: core rural areas 3: urban-rural interaction 4 4: capital conurbation p op ul at io n ch an ge h un tin g cl ub d is tr ib ut io n, % the current distribution of moose hunting clubs. alces vol. 41, 2005 selby et al. threat to sustainability 67 account for 71% of moose hunting club members. differences in club membership with respect to hunting methods (hunting with 12 = 14.6, p = 0.26, n = 319). hunting clubs and associations of clubs are socio-economic, and socio-cultural, etc. the composition of club and association memberthe moose hunting clubs was constructed. membership typology 5 clusters (table 1) that are interpreted as follows: mainly local residents. few local landowners. the cluster contained 54 clubs (16.9%). mainly others. contain a small proportion of local residents. the cluster contained 21 clubs (6.6%). mainly local landowners. although a small proportion of non-local landowners and local residents are present. this was the second largest cluster and contained 101 clubs (31.7%). mainly local landowners and their relatives and friends. residents and other landowners can also be found in this group. the cluster consisted of mainly local landowners and local residents. local residents, with the occasional non-local landowners. this was the largest cluster, the 105 clubs accounting for 32.9% of clubs in membership trends 1 residents others landowners landowners and their landowners and residents f2 p local landowners 12.9 3.1 72.6 29.5 38.1 227.5 0.000 local inhabitant (nonlandowning) 81.7 7.4 10.7 10.2 28.6 254.6 0.000 other landowner (not 1.2 1.5 10.1 9.1 15.1 12.1 0.000 2.4 0 5.3 48.2 7.8 170.4 0.000 2.1 88.5 2 2.8 3.5 549.7 0.000 n 54 21 101 38 105 tion. 1 2 the f 2 threat to sustainability – selby et al. alces vol. 41, 2005 68 58% founded in the 1960s, 21% in the 1970s, 2). of the remaining clubs, 13% reported an increase in membership and 16% reported a ( 8 = 3.36, p = 0.91, n = 327) regardless of ferences in membership trends were found in the different membership groups (table most stable membership – 82% reported no greatest increase in membership at some three greatest decrease in membership was found consistent with current rural trends that are depopulation. membership constraints ers reported that applications for membership this suggests that more new members could membership than were admitted. had more applicants than openings (17%). ing the openings for membership applications 8 = 20.72, p = 0.008, n = 316). it is not uncommon in social institutions for there to be pre-conditions for group memreact to a set of pre-conditions for membership common pre-conditions were domicile in the hunting area (19%), land ownership in the and an understanding of the hunting culture as 1 change in membership residents others landowners landowners and their landowners and local residents total2 increase 11.1 38.1 11.9 7.9 10.5 12.5 no change 70.4 52.4 73.3 81.6 69.5 71.2 decrease 18.5 9.5 14.9 10.5 20.5 16.3 total 100 100 100 100 100 100 n 54 21 101 38 105 319 1 2 8 = 16.34, p = 0.04, n = 319. 2 2 2 alces vol. 41, 2005 selby et al. threat to sustainability 69 of the leaders reported that their clubs did not impose pre-conditions. the imposition of pre-conditions could be a factor restricting the growth of clubs, but the relationship between pre-conditions and changes in club size as such ( 12 = 10.98, p = 0.53, n = 317). the imposed pre-conditions for membership local landowners were the least open (less than 20% were without pre-conditions). domicile logical result. land ownership in the hunting of hunting culture were also common prepre-condition, which suggests that the tacit of the recommendation of a club member in club member as an important pre-condition is based on comradeship. were failed pre-conditions (63%), lack of trust 1 preconditions residents others landowners landowners and their landowners and local residents total2 no preconditions 30.2 27.8 19.2 21.6 20.6 22.3 land ownership and domicile in hunting district 5.7 0 14.1 13.5 15.7 12.3 domicile in hunting district 22.6 0 16.2 13.5 24.5 18.8 land ownership in hunting district 0 5.6 30.3 16.2 14.7 16.8 hunting district 1.9 5.6 8.1 8.1 2.9 5.2 recommendation of club member 5.7 22.2 4 13.5 4.9 6.8 other (e.g., hunting 34 38.9 8.1 13.5 16.7 17.8 total 100 100 100 100 100 100 n 53 18 99 37 102 309 no preconditions 30.2 27.8 19.2 21.6 20.6 22.3 1 2 24 = 68.22, p = 0.000, n = 309. 2 2 threat to sustainability – selby et al. alces vol. 41, 2005 70 (26%), and a desire to limit club size (18%). tions concerning the reasons for restricting membership. of these, the most common ported that maintaining good fellowship was club is founded on comradeship in personal done so in order to maintain good fellowship. ( 8 = 11.09, p = 0.21, n = 314). continuity and sustainability two-thirds (65%) of club leaders in this age of members was in the 50 – 60 age group. in the distribution of age classes between ( 12 = 18.80, p = 0.094, n = 318). considered that the increase in the propormembers were now fewer (22%) or because 1 residents others landowners landowners and their landowners and local residents total2 not important 20 9.1 35 22.2 50 33.3 46.7 27.3 27.5 22.2 22.2 27.5 important 33.3 63.6 37.5 55.6 27.8 39.2 total 100 100 100 100 100 100 n 15 11 40 18 36 120 1 2 8 = 13.26, p = 0.10, n = 120. 2 2 2 alces vol. 41, 2005 selby et al. threat to sustainability 71 1 aging problem residents others landowners landowners and their landowners and local residents total2 membership 14.8 35 18.6 19.4 26.9 21.9 2-age has caused hunt demanding 7.4 10 17.5 11.1 13.5 13.2 sub-total 1+2 22.2 45 36.1 30.5 40.4 35.1 of club 29.6 20 27.8 19.4 21.2 24.4 age of members has not brought changes in 48.1 35 36.1 50 38.5 40.5 sub-total 3+4 77.7 55 63.9 69.4 59.7 64.9 grand total 100 100 100 100 100 100 n 54 20 97 36 104 311 1 2 12 = 11.98, p = 0.45, n = 311. 2 the remaining 65% considered that aging new members had maintained the mean age of members. the effects of aging were most often found landowner component also show signs of aging tent with aging of the farming population. new membership was reported to offset aged). from the standpoint of the future social renewal of moose hunting clubs, the answers threat to sustainability – selby et al. alces vol. 41, 2005 72 showed the greatest interest were also those ( 9 = 29.0, p = 0.000, n = 323). discussion the rationalization of finnish agriculture will result in a drastic reduction in the number farmland will be either reforested or left to tions indicate that the rural areas will continue in the annual moose hunt, while the remainthe period from october to december. this late autumn and winter conditions, often with low temperatures and deep snow. hunting the current membership situation is stable, but warning signals seem to be present. the largest single group of moose hunters are in hunting, as reported elsewhere (vikberg et the aging process because the clubs in which age of members. another cause for concern is the closed nature of moose hunting clubs – a fact that is strict membership pre-conditions it would seem that landowners are seeking to preargument, koskela (2004) has found that the 2 1 people in hunting residents others landowners landowners and their landowners and local residents total2 15.1 4.8 9 7.9 8.6 9.5 one or two 49.1 38.1 63 52.6 64.8 58.4 none 26.4 28.6 26 31.6 24.8 26.5 9.4 28.6 2 7.9 1.9 5.7 total 100 100 100 100 100 100 n 53 21 100 38 105 317 1 2 12 = 32.64, p = 0.001, n = 317. 2 alces vol. 41, 2005 selby et al. threat to sustainability 73 commercialisation of hunting rights. new moose hunting clubs are also rare, fact that the current moose hunting legislation is weighted towards landowners. this, and for new club formation. of moose hunting in its present form. on the ownership becomes concentrated in absentee increased demand for moose hunting opporor who are otherwise interested in hunting. to accommodate this new demand, the pre-condito enable new urban-based moose hunting clubs to operate. references finnish game and fisheries research institute accessed december 2005. finnish statistical yearbook of forestry. 2004. finnish forest research institute. haikonen, h., and h. summala heikkilä land. _____, and s. härkönen (alces alces stands in relation to the characteristics 27:127-143. heikkinen taloustieteen laitos. helle, t., h. pajuoja, and k. nygrén. 1987. forest economics 29:7-26. karppinen, h., h. hänninen, and p. ripatti. 852. koskela mahdollisuus? unpublished report to foundation. helsinki, finland. _____, and t. nygrén 1999. suomen riista 48:65-79. m a a s e u t u p o l i t i i k a n y h t e i s t y ö ry h m ä . 2000. ihmisten maaseutu tahdon marsden, t., j. murdoch, p. lowe, r. munton, and a. flynn. 1993. constructing the land. oore, d. s., and g. p. mccabe. 1999. introduction to the practice of statistics. third new york, new york, usa. ormont rural. pages 21-44 in london, u.k. nevalainen aapanen. 2002. ikäänthreat to sustainability – selby et al. alces vol. 41, 2005 74 niemi, j., and k. pietola. 2005. structural in j. niemi and j. ahlstedt, editors. finnish agriculture and rural industries 105a. petäjistö _____, j. aarnio, p. horne, t. koskela, and a. selby _____, _____, _____, a. selby, and t. koskela selby, a., and l. petäjistö. 1994. field afforestation in finland in the 1990s: _____, _____, and t. koskela. 2003. field _____, _____, _____, and j. aarnio. 2005. 51: 69-82. user guide. spss incorporated, chicago, illinois, usa. vikberg, p., r. orava svendsberg. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 4310.pdf alces vol. 43, 2007 bridgland et al. moose on cape breton island 111 moose on cape breton island, nova scotia: 20th century demographics and emerging issues in the 21st century james bridgland1, tony nette2, charlie dennis3, and derek quann1 1parks canada, ingonish beach, ns, canada, b0c 1l0; 2nova scotia department of natural resources, 136 exhibition street, kentville, ns, canada, b4n 4e5; 3unama’ki institute of natural resources, po box 8096, eskasoni, ns, canada, b1w 1c2 abstract: presumed extirpated in the early 1900s, moose were re-introduced to cape breton island by the federal park service in the late 1940s. after 25 years of gradual growth the population expanded rapidly following a spruce budworm outbreak in the midto late-1970s, yielding a large huntable population by the mid-1980s. continued growth of the herd has presented a number of management challenges and opportunities to the province of nova scotia, the local first nations, and parks canada, each seeking to maintain sustainable moose numbers from different perspectives. presented here is a history of population growth and exploitation of moose on cape breton in the latter 20th century, the evolution of cooperative management of the herd, and emerging management issues. alces vol. 43: 111-121 (2007) key words: cape breton island, cooperative management, moose, nova scotia, population growth, unama’ki early history and re-introduction of moose moose (alces alces americana) were native to cape breton island, nova scotia, and were a major source of food, clothing, and tools as well as an important spiritual and cultural totem for the unama’ki (cape breton island) mi’kmaq for 2,000 years prior to the arrival of europeans in the 1500s (davis and browne 1996). as elsewhere in eastern north america, the arrival of europeans on cape breton brought commerce, especially for hides, which resulted in harvests much larger than previously required by first nations; reports of several commentators from 1600 to 1800 indicated a repeated pattern of apparent over-harvest and dramatic population decline followed by recovery. in the mid1600s nicolas denys observed that moose, formerly abundant, had been reduced to the point that “the indians . . . have abandoned the the wherewithal for living.” (ganong 1908). nevertheless, a letter (undated but likely from is quoted by cautley (1934): “in the year 1729 upwards of 10,000 moose were killed by indians and foreign hunters, merely for their skins and the carcasses left to rot in the woods. they are now scarce.” peterson (1955) cited a record of moose presence on cape breton from 1784. fletcher (1884) reported “moose, once numerous, are now seldom seen.” reporting on the suitability of northern cape breton island for the establishment of a national park, r.w. cautley (1934) wrote “at the present time there are a number of whitetail deer within the site, but very few moose and no caribou.” r.m. anderson, who visited northern cape breton in 1924, reported that moose had been considered extinct for several years but that there was memory in ingonish of large hunts in the 1880s (clarke 1942). the province attempted an introduction of 2 adults and 5 calves from mainland nova scotia to inverness county in 1928–29 (peterson 1955), but to little effect (benson and dodds 1977). on the recommendation of moose on cape breton island bridgland et al. alces vol. 43, 2007 112 clarke (1942), the newly created cape breton highlands national park (the park) attempted to re-introduce moose to the park following the second world war. this introduction the national parks branch to restore a former indigenous species to a park. nevertheless, the rationale for the introduction appears to have been mostly to provide a large mammal to attract park visitors (maceachern 2001). the introduced animals came from elk island national park in central alberta which was suffering from an overabundance of moose and elk in the 1930s and 1940s (lothian moose, these were the western subspecies a. a. andersoni. whether moose actually were extirpated from cape breton at the start of the last century is moot. while the extirpation has been attributed to over-harvest (clark 1942, peterson 1955, cameron 1958), one has to question on practical grounds the notion that hunting alone could have resulted in the elimination of the species, given the ruggedness of the terrain, the severe winter weather conditions, and the available hunting tools and techniques of the era. it is possible that sweeping environmental factors, such as forest succession through the 19th century, also contributed to the decline of moose. while denys describes the plateau of the mid-1600s as dominated by “firs intermingled with a few little birches” (ganong 1908), and bentley and smith (1956) surmised years, there is evidence that great expanses burned in the late 1700s (bridgland et al. 1995). the resulting second growth would have provided ample moose habitat through the 19th century, but the quality of this habitat may have been reduced as the forest returned abies balsamea), which was the case in the early 1900s when fernow (1912) characterized 42% of victoria county and 15% of inverness county as “virgin” conifer forest; forest which upon examination by nichols (1918) turned out to be uniformly 70 years old. benson and dodds (1977) doubted that moose would have been completely absent from cape breton island for the half century prior to the re-introduction project. they suggested that even if all cape breton moose had been extirpated at some point, the narrow strait of canso, which separates cape breton from the mainland, could hardly be considered a serious impediment to moose movement. indeed, the nova scotia department of natural resources (nsdnr) has reports in the past 25 years of moose swimming from isle madame on southern cape breton to guysborough county on the mainland and back. in mid-august of 1947, 5 cows, 1 female calf, and 2 male calves were released at roper brook on the east side of the park (cameron 1958, lothian 1976). these were sighted through the following winter ranging from the aspy valley at the north end of the park to ingonish and a point 25 miles south of the park (kelsall 1948). another 10 alberta moose, 5 bulls and 5 cows, were released at the same location in june of 1948 (cameron 1958, lothian 1976). between september 1948 and march 1950 there were 21 records of sightings or tracks in the park, including 3 mortalities and 3 new calves (boyer 1950). these and sightings in 1952 and 1954 (cameron 1958) showed a wide dispersal throughout the park and beyond. sightings of moose were sporadic through the late 1950s amounting to < 10 animals per year, but increased through the 1960s regularly surpassing 20 sightings after 1964, reaching 57 in 1969 (warden service, cape breton highlands national park, unpublished wildlife observation data). park began in 1970 with aerial monitoring of caribou (rangifer tarandus) re-introduced in 1968 and 1969. while caribou numbers plummeted, moose reported in these surveys rose steadily from 2 in 1970 to 66 in 1975 (macdonald and buchanan 1975). a switch alces vol. 43, 2007 bridgland et al. moose on cape breton island 113 shortened from 35 to 9 hours, 45 moose were recorded. population monitoring and management the eastern spruce budworm (choristoneura fumiferana) is a defoliator of balsam (martineau 1984). it returns cyclically approximately every 30 years, with the severity of the outbreak largely determined by the condition of the forest at the time. outbreaks on cape breton island occurred in the mid-1840s, the 1890s and 1911–15 (nsdlf 1977). an outbreak, starting in 1974 and lasting until the early 1980s, occurred when most mesic sites on the plateau were dominated by almost pure years, resulting in an average mortality of 87% (maclean and ostaff 1989). second growth by vigorous growth of white birch (betula papyrifera). birch, arguably more palatable to moose (peterson 1955), was accompanied by a dramatic increase in moose numbers. a helicopter survey of the whole park in 1977 (couchie and baldwin 1977, prescott 1979), while focussed on distribution rather than abundance, was seen but indicated by track concentrations. substituting the observed average group size of 2 animals for each “unassociated” track concentration, the population was estimated at a minimum of 215 moose. due to the cost of surveying by helicopter and the apparent growth of the herd, the park stepped back from annual surveys. the 1980 survey followed the same methodology as used in 1977 with the exception that the multiplier for “unassociated” tracks was reduced to 1 moose per track concentration (warden service cape breton highlands national park 1985) that was based on park land regions (eer 1978). approximately 10% each of the acadian and boreal land regions, where moose were known to concentrate in winter, were surveyed with 37 randomly selected survey blocks measuring 2 km2. this survey yielded a population estimate for the entire interval ranging from 678 to 1573 moose. one consequence of the spruce budworm outbreak was the development in the late 1970s of an extensive road network to allow salvage of damaged wood from the plateau south of the park. with this new access, the size of the cape breton herd was better realized; a 1978 survey on provincial crown lands south of the park estimated a minimum of 163 moose. in 1980 the province of nova scotia established a moose management zone in northern cape breton (zones 1, 2, and 3, fig. 1) and opened the area for an experimental, limited hunt with 60 licenses awarded by lottery (pulsifer and nette 1995). fifty licenses were issued for a second experimental hunt in 1981. harvest results indicated that the population could support a limited hunt and from 1986 to 2002, 200 licenses were awarded annually for all of victoria and inverness counties outside the park (zones 1–4; fig. 1). the season was limited to a single week from 1986 to 1992; in 1993 it was expanded to 2 weeks with no change in the number of licenses issued (200; pulsifer and nette 1995). through the 1980s, information for managing the moose harvest was primarily based on hunter success rates. hunter success through this period averaged 78%, ranging from 57 to 93%. despite the known expansion of the population to southwest cape breton island, hunting effort was largely restricted to the southern highlands between the cabot trail and the park (fig. 1). in 2003 the number of licenses issued was increased to 310. the 2-week hunt was moose on cape breton island bridgland et al. alces vol. 43, 2007 114 split into 2, single week hunts, and licenses to better distribute the hunt geographically as well as temporally. in 2004 a third single week hunt with an additional 25 licenses was established in december for the area north of the park; in 2005 this hunt issued 35 licenses for a total of 345 licenses. aerial surveys were conducted both north and south of the park in 1987–1993 (survey areas; fig. 1) using parallel, half-mile-wide transects spaced 2 miles apart (scott 1976); observed animals were sexed and aged. estimates for the combined population of these 2 survey areas ranged from 1,848 in 1989 to 2,940 in 1993. by the early 1990s the park was concerned about the size of the moose herd and its impacts on forest composition. recognizing that the herd was not restricted by agency boundaries, nsdnr was approached and an informal arrangement was made for park staff to assist as spotters on the provincial surveys in return for coverage inside the park. in 1992 and 1993 these cooperative transect-based surveys were carried out mostly on provincial lands north and south of the park and in 1994 were restricted to the park. the 1994 results indicated a park population of 2,016 animals at a density of 2.23 moose/km2 (thompson 1995). the switch from loran-c to gps navigation provided the ability to more accurately map the distribution of animals. no survey was done in 1995 or 1996, but in 1997 the park and province succeeded in surveying the entire northern cape breton study area from baddeck to the northern tip of the island with the transect method. the population estimate was 2,018 moose for the entire area. despite this success, both agencies were frustrated by the effort involved in surveying at an awkward time of year. winter lize before march, and there is only a narrow window in which to survey with snow cover for optimal tracking and moose sightability. a frustration arose from the transect methof population change from one survey to the next open to considerable speculation since the trends in the single number estimate were far from consistent. in 1998 the park and nsdnr abandoned (1986) which had become standard elsewhere. the blocks and routes to navigate within them, northern cape breton was divided into 205 survey units measuring 0.06 degrees longitude by 0.04 degrees latitude (average 20 km2). in an blocks into density classes was done based on knowledge from the previous transect surveys, of animals were removed from the census proper to a separate survey done in spring, fig. 1. moose survey areas (north, park, and south) and nsdnr moose management zones 1-4 (gray) located on cape breton island, nova scotia. provincial protected wilderness areas are indicated by hatching. alces vol. 43, 2007 bridgland et al. moose on cape breton island 115 supplemented by a second survey in fall. as it was a trial year, the 1998 survey was restricted to the northern half of the study area that included the park, provincial land north south of the park. the estimated population interval of ± 89%. the 1999 survey expanded coverage to the complete study area and increased the number of blocks surveyed; it produced a population estimate of 1,438 moose ± 46%. these results caused re-examination of the methodology, and after considering a number of possible contributing factors, it was survey more units, and invest in a pre-survey first nations harvest for millennia moose were a most important resource for the mi’kmaq yielding meat, hides for clothing and footwear, bones for tools, and sinew for twine and rope. the ability to kill a large animal such as moose was one of the rights of passage of mi’kmaq boys to manhood (reeves and mccabe 1998). while its use by mi’kmaq declined through the 20th century due to low numbers of moose in the policies of the federal government promoting assimilation, abandonment of traditional game laws introduced by the province in the 1920s. the latter were applied to mi’kmaq as well as non-natives and led to the gabriel sylliboy case of 1929 where the treaties were held to be of no force and effect by the nova scotia provincial court when the grand chief was charged and convicted of hunting muskrat out of season. with passage of the constitution act of treaty rights over subsequent federal and provincial legislation, came interest among first nations in exploring and establishing what those rights implied. through the 1980s a number of test cases related to natural resource use began to appear across canada. the sparrow case, which stemmed from a 1984 in a supreme court ruling that natives have that native access came second only to needs for conservation and perpetuation of the resource (supreme court of canada 1990). as a result of the denny, paul, and sylliboy case, the nova scotia court of appeal established in 1990 that the mi’kmaq had the right to communication). with the return of huntable populations of moose to cape breton island, native harvest activity resumed in tandem with the provincially regulated hunt. to assert the mi’kmaq right to hunt moose, a protest hunt was undertaken in 1988 which resulted in charges being laid on 14 mi’kmaq. in 1989 the province entered into conservation and safety agreements with the confederacy of mainland micmacs, the native council of nova scotia, and the union of nova scotia indians allowing mi’kmaq to hunt moose based on sport licenses issued agreement was renewed the following year, but with the sparrow ruling in the supreme court and the denny, paul, and sylliboy ruling in the provincial court of appeal, this agreement became irrelevant and the charges laid over the 1988 protest hunt were dropped. the marshall ruling (supreme court of scotia, determined that aboriginal resource use could not be restricted to non-commercial subsistence harvest, but could encompass resale to ensure a moderate livelihood. the extent to which these decisions might apply to terrestrial wildlife became a subject of moose on cape breton island bridgland et al. alces vol. 43, 2007 116 disagreement between the government of nova scotia and the mi’kmaq leadership. their impact on the prohibition of hunting in the park was equally problematic for parks canada. one consequence of these rulings in cape breton was that, in addition to a traditional native subsistence hunt, a small number of natives embarked on harvesting moose for commercial sale of the meat to natives and non-natives. though still illegal for a non-mi’kmaq hunter to hunt without a license issued by the province, a small number of natives started guiding unlicensed nonnative hunters, on the assumption they could “share” hunting related treaty rights. these developments in the late 1990s were viewed with some concern by both native and nonnative communities. prior to european settlement, the mi’kmaq formed a loose confederacy of semi-nomadic family-groups, stretching from the gaspé to maine to southwest newfoundland, organized hunting territory. unama’kik (cape breton island) was 1 of 7 districts of the mi’kmaq district other than their own at the invitation of the saqamaw or chief of that district, the invitation being extended when consistent with netukulimk, the mi’kmaq philosophy of care and respect for the land (barsh 2002). through the 1990s there were reports of individuals both from unama’ki bands and from mainland nova scotia bands engaged in increasingly large harvests of cape breton moose for the sale of meat throughout the province. this raised concern of the unama’ki leadership over the potential impact of unchecked commercial harvest on the sustainability of the herd because they viewed the native right to hunt as a community right rather than an individual right, and the unama’ki bands as the rightful stewards of the moose on their traditional hunting grounds. a set of draft guidelines to manage the mi’kmaq hunt was drawn up by charles webb and tuma young in the mid-1990s for the eskasoni fish and wildlife commission. these guidelines focussed on issues of hunter safety, conservation and management, eligibility, and culturally appropriate use. while they envisaged management and enforcement by unama’ki justice system for resolving disputes and infractions, the guidelines had little authority beyond moral suasion. consequently, the principal tool available for enforcement was the willingness of the native community to support hunters subscribing to these guidelines against prosecution under provincial or federal regulations. emergence of co-management in 1999, at the invitation of the newly formed unama’ki institute of natural resources (uinr), talks began among uinr, nsdnr, and parks canada to address management of the cape breton moose herd. the province and parks canada were at this to accurately census the herd. all parties had an interest in the sustainability of the harvest and all were concerned about the size of the native harvest. the unama’ki bands viewed the herd as primarily a first nations’ resource with the marshall case still under appeal at the supreme court, and with the cape breton herd remaining abundant, the province focussed its enforcement on safety related matters of the mi’kmaq hunt. parks canada was concerned with environmental degradation from allterrain vehicle use, public safety issues, and hunting, especially unregulated hunting within the park. these talks spawned 2 initiatives among the 3 agencies: a cooperative 5-year population and dispersal study and a cooperative moose management committee. the research program was established through formal agreement between senior alces vol. 43, 2007 bridgland et al. moose on cape breton island 117 management at the park and the renewable resources branch of nsdnr, and a less formal agreement with uinr. it was guided by a technical steering committee of biologists from parks canada and nsdnr who drew on staff from the 3 agencies. the research focussed on obtaining: (1) accurate data on population size, productivity, and survival to enable population modelling; (2) data on habitat use to determine distribution and impacts of moose on the forests of the cape breton plateau; and (3) information on patterns of habitat selection, dispersal, and seasonal migration. the study area was the same as that used in the late 1990s surveys, all of cape breton island north of the cabot trail (fig. 1). the approximately 3,900 km2 area was comprised of 3 distinct landscapes (survey areas; fig. 1) differing in forest and wildlife management regime. the north survey area is predominantly provincial crown land, much of it protected as a provincial wilderness area in which there is no active forest management, hunting is permitted, but very limited motorized access. the middle or park survey area has a policy of passive forest management and general prohibition of hunting. the south survey area is largely nova scotia crown land, most of it under forest lease to a nearby paper mill. this area is actively managed for wood dense network of woods roads, and receives a large amount of hunter effort, both native and non-native. full aerial population surveys were con(90% c.i. of ± 20%) was achieved, and every second year thereafter. to improve accuracy, and the size of the survey blocks was reduced to 2 minutes longitude by 1 minute latitude (average 4.73 km2) to survey more blocks with the same effort. initially sex and age surveys were conducted annually in spring and fall, then spring only. forty vhf and 14 gps radio-collars were used variously on calves, sub-adults, and adults to track calf survival and juvenile dispersal, seasonal migration, home range, and landscape level habitat use. preliminary results, based on the whole herd over the entire study area, indicated wide estimated carrying capacity of 5,000–6,000 moose. further analysis is required to determine if this pattern is consistent in each survey area, or if it varies with different regimes of forest management and moose harvest. it was evident that moose were not uniformly distributed throughout the study area in late winter, and aggregated in roughly the same areas each winter. the preliminary results also indicated that there may be relatively little dispersal between management areas and that seasonal movement and home range size are highly variable among individuals. finally, that hunting and carrying capacity affect both adult sex ratios and calf production within the survey areas. of the three survey areas, the south survey area (management zones 2b and 3; fig.1) was the most heavily hunted, had the lowest bull:cow ratios, and the highest calf:cow ratios and twinning rates. the objective of the cooperative management committee has been essentially to establish real and effective mi’kmaq management of the mi’kmaq moose harvest on unama’ki lands, to ensure both the continued sustainability and culturally responsible use of the harvest. two priorities were education in the native communities on issues affecting the sustainability of the mi’kmaq moose harvest, and development of a mi’kmaq moose harvest plan supported by all native hunters to obtain a better estimate of the harvest. the 1990s guidelines were revisited as the basis of a management plan, but it became apparent that there were questions of eligibility and allocation that required political resolution that was beyond the mandate of this technical moose on cape breton island bridgland et al. alces vol. 43, 2007 118 committee. who was to be eligible to hunt moose under these guidelines? questions arose concerning status and place of residence. foreseeing that it might eventually be necessary to implement a quota, how was the quota to be allocated equitably? equally daunting was the absence of an existing mechanism for ensuring compliance. to raise awareness of the need to manage the native hunt, uinr undertook a survey of traditional ecological unama’ki bands. concurrently a program for voluntary reporting of harvest data, including the collection of jaws, was established to educate native hunters on the importance of hunter reporting and to improve understanding of the size and demographics of the native harvest. participation in the voluntary reporting program has been slow to build and accurate harvest data remains a large gap in management information. in 2005 uinr was mandated by the mi’kmaq grand council and the unama’ki council of elders to develop and draft a management plan, and moose management has gained prominence in current tri-partite negotiations between the mi’kmaq, the nova scotia, and canadian governments aimed at bringing relevant modern interpretation to treaties that date from the 18th century. as part of this process, uinr has been commissioned by the assembly of nova scotia mi’kmaq chiefs, which represents all 13 nova scotia bands, to off reserve, status or non-status – to canvass their views on how the native hunt should be be put forward in the tri-partite process. emerging management issues continued viability of the cape breton moose population is of paramount importance to all 3 agencies. the herd is important to the economy of cape breton through providing opportunities for eco-tourism and hunting, as well as providing first nations with an opportunity to revive their cultural heritage with respect to moose and forge modern and sustainable interpretations of treaty rights. as the moose population has risen and densities have remained high, it has been easy to meet the needs of all 3 agencies. concern remains, however; can these population levels persist the population decline and apparent extirpation of moose in the late 1800s. despite its apparent robustness, the current cape breton moose population exhibits tooth breakage and anomalous behaviours such as osteophagy and excessive bark stripping that have been interpreted as indicators of nutritional stress (clough et al. 2006). more important may be density dependent impacts of the herd on its own habitat including large areas of the plateau not recovered from the budworm outbreak of the 1970s. national parks policy calls for minimal human interference with natural processes (parks canada 1983, 1994), consequently, no effort was made to prevent or manage the infestation and the subsequent expansion of the moose population. by the mid-1990s it was apparent that moose were browsing preferentially on white birch, mountain ash (sorbus americana the expanding population was substantially impacting these preferred species (basquill and thompson 1997, broaders 1998). by the early 2000s white birch regeneration and other deciduous shrubs were browsed to the point that grasses (particularly calamagrostis canadensis) dominated large tracts, preventing germination and suppressing growth of tree seedlings. as well as reducing the habitat quality for moose, this halting or reversal of forest succession poses severe problems for marten (martes americana) and lynx (lynx lynx). both are provincially endangered species limited to northern cape breton, and both require structurally complex, closed-canopy forest. the impact of moose browsing on forest succession also affects other species alces vol. 43, 2007 bridgland et al. moose on cape breton island 119 including birds and small mammals that occupy post-disturbance boreal ecosystems from mid-successional through climax forest. of the 3 survey areas, the forest landscape of the park, where hunting is not permitted, is probably the most compromised, followed by the remoter, more inaccessible areas of the provincial wilderness area north of the park. while parks canada policy allows for strictly controlled active management where needed, it remains to be determined what manner and level of intervention would be possible and effective in reducing the moose population to allow recovery of the over-browsed forest in the park. assuming that natural succession the capacity of that habitat to support moose is unknown. however, the population would certainly be smaller than currently exists, and all 3 agencies would have to decide how fewer moose could be sustained and shared. alternatively, if it is not possible to reduce the density of moose in the park and the provincial wilderness area to the north, change to forest composition and biological diversity. reduced habitat quality could well lead to density-dependent responses including decline in nutritional condition, productivity, and population. it remains to be seen if collaborative management will succeed in maintaining a healthy moose population over the long-term, or if we are destined to again in northern cape breton. acknowledgements we thank eric zscheile of the kwilmuk viewing an earlier draft of the manuscript and providing useful insight. figure 1 was produced by geordon harvey of cape breton highlands national park. references barsh. r. l. 2002. netukulimk past and present: mikmaw ethics and the atlantic fishery. journal of canadian studies, trent university 37(2):15-42. basquill, s., and r. thompson. 1997. moose (alces alces) browse availability and utilization in cape breton highlands national park. parks canada, technical reports in ecosystem science 010. benson, d. w., and g. d. dodds. 1977. the deer of nova scotia. department of lands and forests, province of nova scotia, halifax, nova scotia, canada. bentley, p. a., and e. c. smith. 1956. the forests of cape breton in the seventeenth and eighteenth centuries. proceedings of the nova scotia institute of science, halifax, nova scotia, canada 24(1):1-15. boyer, g. f. 1950. moose at cape breton highlands national park. unpublished typewritten report. dominion wildlife service. cape breton highlands national park library, ingonish beach, nova scotia, canada. bridgland, j., r. cook, r. power, and b. pardy. 1995. fire history of northern cape breton: gis analysis of biophysical data. pages 398-406 in t. b. herman, s. bondrup-nielsen, j. h. m. willison, and n. w. p. munro, editors. ecosystem monitoring and protected areas. proceedings of the 2nd international conference on science and the management of protected areas, dalhousie university, halifax, nova scotia, canada. broaders, t. 1998. effects of moose browsing on plant growth and succession at cape breton highlands national park. unpublished report. cape breton highlands national park library, ingonish, nova scotia, canada. cameron, a. w. 1958. the mammals of the islands in the gulf of st. lawrence. national museum of canada, bulletin 154. cautley, r. w. 1934. report on examination of sites for a national park in the province of nova scotia. report to j. b. harkin, moose on cape breton island bridgland et al. alces vol. 43, 2007 120 commissioner, national parks of canada, department of the interior, ottawa. cape breton highlands national park library, ingonish, nova scotia, canada. clarke, c. h. d. 1942. investigation of cape breton highlands national park, (with appended mammals of cape breton highlands national park by r. m. anderson). national parks bureau. ottawa. mimeo cape breton highlands national park library, ingonish, nova scotia, canada. clough, m., m. zentilli, h. g. broders, and t. nette. 2006. elemental composition of incisors in nova scotia moose: evaluation of a population with abnormal incisor breakage. alces 42:55-64. couchie, d., and w. baldwin. 1977. 1977 aerial ungulate survey, cape breton highlands national park. unpublished m.s. report. cape breton highlands national park library, ingonish, nova scotia, canada. davis, d. s., and s. browne, editors. 1996. the natural history of nova scotia. nova scotia museum, nimbus, halifax, nova scotia, canada. (eer) eastern ecological research limited. 1978. ecological land classification, cape breton highlands national park. cape breton highlands national park library, ingonish, nova scotia, canada. fernow, b. e. 1912. forest conditions of nova scotia. commission of conservation canada, ottawa, ontario, canada. fletcher, h. 1884. report on the geology of northern cape breton. geological and natural history survey of canada. dawson brothers, montreal, quebec, canada. ganong, w. f., translator and editor. 1908. the description and natural history of the coasts of north america (acadia) by nicolas denys, 1672. champlain society, toronto, ontario, canada. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, 22. kelsall, j. p. 1948. moose investigation, cape breton highlands national park, may 26 to june 3, 1948. memorandum and report to the chief, dominion wildlife service, department of mines and resources, ottawa, ontario, canada. lothian, w. f. 1976. a history of canada’s national parks. volume 1. indian and northern affairs, parks canada, ottawa, ontario, canada. macdonald, j. d., and b. buchanan. 1975. ungulate survey 1975, cape breton highlands national park. unpublished report. cape breton highlands national park library, ingonish, nova scotia, canada. maceachern, a. 2001. natural selections: national parks in atlantic canada, 19351970. mcgill-queen’s university press, kingston, ontario, canada. maclean, d. a., and d. p. ostaff. 1989. patterns of balsam fir mortality caused by an uncontrolled spruce budworm outbreak. canadian journal of forest research 19:1087-1095. martineau, r. 1984. insects harmful to forest trees. canadian forestry service and multiscience publications limited supply and services canada, ottawa, ontario, canada. nichols, g. e. 1918. the vegetation of northern cape breton island, nova scotia. transactions of the connecticut academy of arts and sciences, yale university press, new haven, connecticut, usa. volume 22:249-467. (nsdlf) nova scotia department of lands and forests. 1977. nova scotia’s spruce budworm situation ... history, status and strategies. nova scotia department of lands and forests, halifax, nova scotia, canada. alces vol. 43, 2007 bridgland et al. moose on cape breton island 121 parks canada. 1983. parks canada policy. environment canada and supply and services canada. ottawa, ontario, canada. _____ 1994. parks canada guiding principles and operational policies. department of canadian heritage. ottawa, ontario, canada. paul, d. n. 1993. we were not the savages. nimbus, halifax, nova scotia, canada. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. prescott, w. h. 1979. cape breton highlands national park, aerial ungulate survey, 1977. canadian wildlife service, atlantic region, manuscript report. sackville, new brunswick, canada. pulsifer, m. d., and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31:209219. reeves, h. m., and r. e. mccabe. 1998. of moose and man. pages 1-75 in a. w. franzman and c. c. schwarz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. scott, c. j. 1976. nova scotia moose: a new inventory technique. m.sc. thesis. acadia university, wolfville, nova scotia, canada. supreme court of canada. 1990. r. v. sparrow, [1990] 1 s.c.r. 1075. _____. 1999. r. v. marshall, [1999] 3 s.c.r. 533. thompson, r. 1995. 1994 aerial moose survey, cape breton highlands national park. unpublished report. cape breton highlands park library, ingonish, nova scotia, canada. warden service cape breton highlands national park. 1980. cape breton highlands national park aerial ungulate survey 1980. unpublished report. cape breton highlands national park library. ingonish, nova scotia, canada. wentzell, n. 1985. aerial moose census, cape breton highlands national park, march 1985. unpublished report. cape breton highlands national park library, ingonish, nova scotia, canada. 4212(89-109).pdf alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 89 complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada karen beazley1, mark ball2, lisa isaacman1, scott mcburney3, paul wilson2, and tony nette4 1school for resource and environmental studies, dalhousie university, 6100 university avenue, suite 5010, halifax, ns, canada b3h 3j5; 2 3canadian cooperative c1a 4p3; 4nova scotia department of natural resources, wildlife division, 136 exhibition street, abstract: in 2003, the eastern moose (alces alces americana) on mainland nova scotia was declared an endangered species under the nova scotia endangered species act. subsequently, as required by the act, a recovery team was established and the recovery planning process was initiated. very early in this process, it was recognized that developing a recovery strategy for this moose population was tion size, structure, reproduction, and mortality are not current for the population, and the assessment methodologies are inconsistent. the ability to evaluate potential factors limiting the population is hindered by a lack of information, primarily in the subject areas of genetic structure, health, illegal effect relationships, as well as verifying the potential cumulative and synergistic effects of the factors impacting the moose population. answering these questions is challenging and will require substantial requisite data. until the information gaps can be addressed, it is prudent to adopt a precautionary and adaptive approach to the recovery of this species. alces vol. 42: 89-109 (2006) key words: access, climate change, demographics, disease, disturbance, fragmentation, genetics, habitat suitability, population viability, road effects beginning in the late 1700s, and accelerating through the 1800s, the southern range limits of eastern moose (alces alces americana) retreated north from most new england states bontaites and gustafson 1993, morris and elowe 1993, vecellio et al. 1993). in new brunswick and nova scotia, moose met a similar, although delayed, fate (nova scotia department of natural resources, unpublished data). moose numbers in pre-european mainland nova scotia may have been in the range of 0.38 moose/km2 animals, but declined to a low of several thousand by 1825 (peterson 1955). a 3-year closed hunting season enacted in 1874 was followed shortly by a prohibition on snaring and hunting with dogs; restrictions that may have saved scotia 1934). greater legislative protection and increased enforcement allowed moose to slowly increase throughout the mainland; by 1908, their numbers may have rivaled those of the pre-european era. moose in nova scotia declined through the 1920s and 1930s and the hunting season was closed in 1938 (benson and dodds 1977). in eastern and northeastrecovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 90 ern mainland nova scotia, where densities season was opened from 1964 to 1974, closed in 1975 and 1976, and re-opened from 1977 to 1981 (nova scotia department of natural resources, unpublished data). the season has been closed since 1981 due to concerns about declining moose numbers. recent surveys on mainland nova scotia, although somewhat abated, has continued in spite of the hunting season closure and moose probably now only number between 1,000 and 1,200 animals (parker 2003; nova scotia department of natural resources, unpublished data). moose 1800s, reintroduced in the late 1940s (from moose transplanted from elk island national park in alberta), and are common throughout the northern two-thirds of the island, where hunting is currently permitted (nova scotia department of natural resources, unpublished data). in 2003, following the preparation of a status report (parker 2003), mainland nova scotia moose were declared ‘endangered’ under the nova scotia endangered species act. in accordance with an endangered status under the act, the moose receives full protection, which includes a prohibition on all hunting, a recovery team has been established, and a recovery plan is currently in preparation. this paper describes some of the challenges facing the recovery planning process for the endangered mainland nova scotia and gaps in knowledge about, population dynamics and cause-effect relationships with potential threats and limiting factors related to habitat, viability, genetics, health, and climate change. demographics and viability demographics for over 200 years, the highly fragmented and discontinuous moose habitat remaining in the area of the chignecto isthmus, at the border between nova scotia and new brunswick, is believed to have minimized movement of moose between the two provinces. this near absence of connectivity has resulted in these different genetic characteristics (see genetics section). similarly, there does not appear to 3 remaining localized groups1 of moose on mainland nova scotia and the reintroduced cape breton island population, which currently numbers about 8,000 animals (nova scotia department of natural resources, unpublished data) (fig. 1). efforts by nova scotia department of natural resources biologists have been less than successful in achieving an overall population estimate with on mainland nova scotia. however, based on of moose remaining on the mainland has been estimated at between 1,000 and 1,200 (parker 2003; nova scotia department of natural suggest that cow:bull ratios of moose on mainland nova scotia are within normal values (dodds 1963, benson and dodds 1977, brannen 2004; nova scotia department of natural resources, unpublished data). however, far as moose management efforts in the province have focussed on determining abundance and distribution, hunter success rates, kill per area, 1we use the term localized groups rather than subpopulations because there are currently insufdistinct populations or subpopulations; from a as individuals within a given larger population level of relatedness; since this is not clearly more general term, localized groups. alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 91 ments of hunter-harvested animals (nova scotia department of natural resources, unpublished data). according to the limited and varied sources available, there appear to be widely differing values on calf numbers reported both spatially and temporally. prior to 1956, abundance of moose in nova scotia was determined through hunter success rates and observation reports, wildlife and forest ranger reports, and casual observation (peterson 1955, dodds 1963, benson and dodds 1977). big game hunter moose observation reports for eastern mainland nova scotia between 1950 – 1952 indicated that the number of calves to 100 cows was 14 in 1950, 47 in 1951, and 32 in 1952 (dodds 1963). observational data submitted by “unsuccessful” eastern mainland moose hunters in 1964, 1977, and 1978 indicated that the number of calves to 100 cows in those years were 14.5, 38, and 50, respectively, in that region (nova scotia department of natural resources, unpublished data). according to wildlife sanctuary warden moose observation reports for the period 1934-1953, the number of calves to 100 cows for the 4 largest wildlife sanctuaries on mainland nova scotia (fig. 1) ranged from a low of 15 in the waverley sanctuary during the period 1949-1953, to a high of 62 in the tobeatic wildlife management area during the period 1934-1938 (dodds 1963). vukelich (1977) analysed the reproductive tracts of moose of known age, fig. 1. general location and size of largest remaining localized groups of moose on mainland nova island populations, and mainland nova scotia localized groups has not been found. recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 92 which were killed by hunters in the eastern mainland in 1973 and 1974. fifty percent of yearling cows were found to be pregnant (all carrying single fetuses), while 85.3% of adult cows were pregnant (25% with single fetuses and 75% carrying 2 fetuses). according to this study, “comparison of the nova scotia data with that from other regions indicate that the moose in nova scotia are as productive as any other north american moose herd aerial survey of moose in nova scotia was conducted by d. w. benson in the late winter of 1956 (nova scotia department of natural resources, unpublished data). however, little data were gathered on calf:cow ratios in early surveys since they focused on density and distribution. recent moose surveys on mainland nova scotia have observed so few animals that little insight could be gained on tion of: (1) a january 1993 aerial survey of the tobeatic area, which documented 50 calves per 100 cows (nova scotia department of natural resources, unpublished data); and (2) a 1999-2001 study conducted in the same area, which documented 22 calves per 100 cows (brannen 2004). the limited and infrequent data available long-term rates of reproduction, calf survival, recruitment, or overall population structure of moose on mainland nova scotia. there is no apparent consistent pattern and, with insufcalf representation between years within the same localized group of moose and between different localized groups remain unclear. acquiring adequate data on this subject will be challenging and costly. conducting moose surveys in nova scotia is arduous, not only because of low moose densities on the mainland, but also due to unpredictable weather conditions (including frequent winter rain storms), poor snow conditions near coastal areas, and sporadic availability of government-owned helicopters (pulsifer and nette 1995). population viability a viable population is one that will conreproductive rates are higher than or equal to mortality rates (salwasser et al. 1984, newmark 1985). a minimum viable population (mvp) is the population size below which high, but above which the probability of over a given period of time (shaffer 1981, samson 1983, lehmkhul 1984, gilpin and soulé 1986). population viability analysis is a method used to determine mvp, which can be used to identify threatened populations and quantitative targets of population size for recovery efforts (salwasser et al. 1984, shaffer 1990, boyce 1992). population viability analyses are data intensive since viability is a function of genetic (i.e., to prevent inbreeding depression), demographic (i.e., birthrate, mortality, reproductive age, and fecundity), environmental (i.e., stochasticity), and spatial (i.e., distribution) factors (for an overview see snaith and beazley 2002). nova scotia to conduct viability analyses even using simple computer-based models such as vortex (for an assessment of such models see lindenmayer et al. 1995). nonetheless, it is important to understand viability issues since the remnant localized groups of moose in mainland nova scotia are small (< 500 individuals per group) and thus may not be viable. although small founder groups are available (timmermann and mcnicol 1988) and some have grown to widely distributed populations (pulsifer 1995, basquille and thompson 1997, wangersky 2000), since heterozygosity is likely reduced (as is the case in newfoundland and cape breton, nova scotia), long-term viability may be compromised due alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 93 to reduced genetic variability (broders at al. 1999). while moose in mainland nova scotia formerly constituted a continuous population, remnant localized groups are currently isolated from each other by distances of 200-300 km, highways, and areas of low habitat suitability (snaith and beazley 2002). thus, it is unlikely (at least one reproductively successful migrant per generation [soulé 1980, brussard 1985, reed et al. 1986, beier 1993]) for these groups to function and persist as a metapopulation (i.e., localized groups loosely associated by 1970, fahrig and merriam 1994, beissinger and westphal 1998]). conducted a rough estimate of moose mvp using a 50/500 rule-of-thumb proposed by franklin (1980) and soulé (1980) as an alternative to data-intensive analysis. the 50/500 rule suggests that effective or ideal breeding populations (ne) of 50 and 500 individuals are required to maintain short-term (decades) and long-term (centuries) viability, respectively, based on inbreeding considerations alone. to estimate the size of census populations (n) required to support ne for moose, an average 10:1 ratio of n:ne was assumed from ratios reported by ryman et al. (1981) and arsenault (2000). accordingly, census populations for shortand long-term viability were calculated as 500 and 5,000 individuals, respectively. although such coarse estimates are subject to criticism, several empirical, modeling, and genetic studies have found consistent results, and thus they may serve as a general guideline in the absence of more precise data (brussard 1985, samson et al. 1985, lande 1987, thomas 1990, henriksen 1997, belovsky et al. 1999). these estimates indicate that current isolated, localized groups of less than 500 individuals each are unlikely to persist over the shortterm. if connectivity among localized groups is such that they do collectively function as a metapopulation, at less than a total of 5,000 individuals, they are unlikely to persist over the long-term. nonetheless, considerable population viability and in devising strategies tirpation caused by the small size and isolated locations of localized groups. genetics between landscape features and population dynamics is the impact on the movement of cant role in the persistence of a species. it which is necessary for the maintenance of genetic diversity, thus enabling individuals to adapt to their environments, while rendering the species as a whole more resilient to changes such as disease and climate (hedrick and miller 1992, wayne et al. 1992, o’brien 1994, haig 1998). the movement of individuals may also promote the persistence of populations where their net reproductive rate is less than replacement (watkinson and sutherland 1995, dias 1996, novaro et al. 2000). in such cases, population viability is dependant upon not only its own population parameters, but also the population dynamics of neighboring patches and ease of movement among these patches (caprio 2001). alternatively, the movement of individuals can also be a detriment to a speamong areas can promote the spread of disease or could potentially introduce animals into culty adapting. either of these scenarios could have devastating consequences for population recovery. in nova scotia, understanding the dispersal dynamics of resident and neighboring moose populations, subpopulations, and localized groups is essential for any potential recovery efforts. the relationship between the landscape recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 94 address using traditional methods of population and landscape ecology. however, with the advent of molecular genetic tools, new questions regarding population structure and evolutionary processes, such as genetic drift namics (manel et al. 2003). in 2002, analysis of microsatellite data derived from a limited number of moose from mainland nova scotia (n = 35) and neighboring cape breton island (n = 23) and new brunswick (n = 16) showed areas. in addition, the analysis showed little from the cumberland region with those collected from both the guysborough and tobeatic regions, presenting the likelihood of two genetically structured populations occurring on mainland nova scotia (fig. 1). measures of genetic variability of moose on mainland nova scotia showed that the levels were breton and northern ontario where there are genetic variability in moose from mainland nova scotia could be a response to 2 factors: (1) non-assortive mating; and (2) pooling of samples from locally structured populations within the geographic area, otherwise known as the wahlund effect. discriminating among raise important concerns regarding the highly fragmented nature of localized moose groups on the mainland of nova scotia. acknowledging the analytical limitations of our low sample sizes and high dispersion of sampling areas, a more comprehensive genetic analysis could be obtained with additional samples from a greater number of areas. the current low numbers of mainland moose and their obtaining additional tissue and blood samples for dna analysis. as a result, a non-invasive approach has been developed that involves the collection of moose fecal material as a source of dna. this method, which has proven highly successful for woodland caribou (ball augment the data set and could be used for future population monitoring. the development of conservation and recovery plans to increase moose movement scotia will prove to be an arduous task in light of many of the problems confronting nova scotia moose. the danger of population fragmentation and isolation increases concerns about progressive genetic deterioration, which would reduce individual health and population productivity and potentially eventually considering limitations on the available moose habitat remaining on mainland nova scotia, as well as the presence of a parasite (parelaphostrongylus tenuis) potentially fatal to moose throughout the province, facilitating the effective movement of migrant moose through either the creation of interconnecting habitat corridors or translocation of effective breeders from neighboring new brunswick originated from moose transplanted from alberta, the reintroduced cape breton moose is a distinct subspecies from those inhabiting the mainland (pimlott and carberry 1958, dodds 1975). therefore, the effectiveness of cape breton moose as breeders with the mainland population may be limited, and even if successful, would change the genetic identity of the mainland moose. with this said, a clear assessment of moose population structure, habitat and disease data, will prepare managers to identify threats to moose population dynamics and implement appropriate solutions. health the study of mortality factors, including those that are directly related to disease and alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 95 those that are not (e.g., predation, trauma, and environmental conditions), provides essential information that can lead to a greater understanding of processes contributing to the decline of any wildlife population. therefore, since 1998, opportunistic moose mortalities and euthanized sick or injured moose have been utilized to assess the health of the mainland nova scotia moose population. whenever possible, entire carcasses were necropsied at the post mortem facilities of the atlantic veterinary college, university of prince edward island or the veterinary diagnostic laboratory, nova scotia department of agriculture and fisheries. when this could not occur, necropdepartment of natural resources staff, and the entire head, spinal column, heart, lungs, liver, and kidneys, together with a section of quadriceps muscle and an intact femur from one of the hind legs were collected and frozen were live captured and transported to nova scotia department of natural resources’ wildlife park in shubenacadie, nova scotia. these individuals were held in quarantine for a 1 – 2 month period of clinical assessment, and subsequently, due to the progression of their clinical disease and a continued deterioration of their body condition, both animals were nation of the carcasses and tissues submitted for necropsy was combined with ancillary diagnostic testing, including histology and, when required, parasitology, bacteriology, and also, the incisor bar was collected for aging, and the liver and kidneys were sampled for inclusion in a cervid trace element study initiated by nova scotia department of natural resources, canadian wildlife service, and environment canada. pregnant, and 6 male moose were necropsied. class, 8 were in the 2 – 8-year-old age class, and 2 were in the > 8-year-old age class. two males were in the 0 – 2-year-old age class and 4 were in the 2 – 8-year-old age class. trauma was the most common cause of mortality. vehicular collision was the cause of death in 6 animals, another 2 were killed by gunshot, and 1 died as the result of an accidental fall from a cliff. all of the affected condition with normal skeletal muscle mass and abundant adipose tissue stores. undering that the accident causing the trauma was the result of a simple mishap. cant mortality factor affecting the population. a diagnosis of parelaphostrongylosis was p. tenuis adults, larvae, or tral nervous system (cns) neuropil. in cases where there were random multifocal, linear areas of necrosis, and/or degeneration of the cns neuropil associated with nonsuppurative were considered compatible with parasitic tracts and strongly suggestive of an infection with p. tenuis despite the absence of nematode adults, larvae, or eggs. based on these as the cause of death in 2 animals and was the most likely cause of death in an additional 5 cases. in one of the affected moose, p. tenuis eggs were present in the interstitium of the pulmonary alveolar septa, which is consistent with a patent infection. in 2000, 3 moose were euthanized because animal was uncoordinated and circling, and the other 2 were unable to rise due to hind limb paralysis. all affected individuals were adult females in very good body condition and had a diffuse, nonsuppurative encephalomyelitis. during that same year, 2 moose from the other nova scotia moose population in cape breton recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 96 were euthanized due to similar neurological abnormalities. these moose, an adult male and a yearling female, were in very good body changes in the cns. in all cases, the lesions were suggestive of a viral etiology, but the cause was not determined. additional cases have not been observed in either of the nova scotia moose populations since that time. and pulmonary hemorrhage post-collaring due to capture-related complications. the cause of death could not be determined in 2 mortalities. a small number of the moose necropsied included the presence of dictyocaulus viviparous (lungworm), dermacentor albipictus (winter tick), and antler deformities. the parasitic infections were of low intensity and only found in animals affected by neurological disease. there were 6 cases of mild dictyocaulosis and 5 cases of winter tick infestation with alopecia affecting 0 – 10% of the affected animal’s body surface, primarily the head and neck in 2 of the 6 males necropsied, 1 healthy and 1 with parelaphostrongylosis. also, similar antler deformities were observed in 3 of the 9 males necropsied from the cape breton moose population. calf survival was considered as a potential measure of population health. a concurrent study on a localized group of moose in mainland nova scotia’s tobeatic region indicated that calf survival for the radio-collared females was very low during the years 1998 – 2001 (brannen 2004). the causes of calf mortality were not investigated in this study, and it remains to be determined if calf survivorship is similar in the other localized groups of mainland nova scotia moose. the data from the cervid trace element study are analyzed in another report (pollock 2005). the kidney and liver samples were analyzed for arsenic, cadmium, cobalt, copper, lead, manganese, nickel, selenium, and zinc. kidney cadmium concentrations were high in some nova scotia moose and, relative to reference values for cattle, cobalt, copper, manganese, selenium, and zinc levels in some (pollock 2005). however, gross or microscopor cobalt, copper, manganese, selenium, and of the moose necropsied, suggesting clinical disease associated with these trace elements is not occurring in nova scotia moose. accidental random vehicular trauma is a frequent and often unavoidable cause of mortality in wildlife populations. however, in the case of endangered species, these losses may be considered unacceptable and may need to be mitigated, if possible. two of the a section of a 4-lane nova scotia provincial highway known as the cobequid pass. this piece of highway has been a common site of moose-vehicle collision since its completion through necropsy, there are 8 known moose mortalities from this area (t. nette, nova scotia department of natural resources, unpublished data). therefore, as in other jurisdictions, investigation of management options, such as strategic fencing, implementation of signage, creation of wildlife underpasses or overpasses, and reduction of speed limits, is required to address this problem. mortality factor affecting the moose population of mainland nova scotia. there is ample support in the literature demonstrating the affect of this parasite on the population dynamics of moose and other cervid species in nova scotia and elsewhere in north america (benson 1958a, b; smith et al. 1964; smith and archibald 1967; anderson 1971, 1972; thomas and dodds 1988). the refugia hypothesis (kearney and gilbert 1976) was used alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 97 white-tailed deer (odocoileus virginianus) when p. tenuis is present in the environment. this hypothesis and the role p. tenuis plays in regulating cervid populations has been challenged in the literature (nudds 1990, whitlaw and lankester 1994, bender et al. 2005), but parelaphostrongylosis is regulating moose on mainland nova scotia, and the localized groups are surviving in refugia. this evidence includes the following observations: 1. the current decline of the nova scotia mainland moose population began subsepopulation growth that occurred in the late 1920s through the early 1950s and has ing but persistently high white-tailed deer population density (benson and dodds 1977; t. nette, nova scotia department of natural resources, unpublished data); 2. the remnant localized groups of moose appear to be geographically limited to the elevated regions of the province, in areas where white-tailed deer are absent or in low density (t. nette, nova scotia department of natural resources, unpublished data); 3. the remnant localized groups of moose are also geographically limited to a granitic soil type that is not compatible with the presence of the intermediate molluscan hosts of p. tenuis (r. cameron, nova scotia department of natural resources, unpublished data). it is clear that further research is required mainland nova scotia moose population dynamics. selectively decreasing the populations of white-tailed deer or molluscan intermediate hosts of p. tenuis are management tools that could be utilized to address this problem in certain areas of critical moose habitat. antler deformities were relatively commainland and cape breton moose populations. in cervids, antler deformities have been associated with genetic abnormalities; traumatic skeletal injuries; copper, calcium, physiological factors such as failure of the antler’s testosterone receptors during growth and testosterone imbalances due to testicular atrophy (marburger et al. 1972, gogan et al. 1988, carrasco et al. 1997, tiller et al. 1997, o’hara et al. 2001). injuries to the antler or the cases. in a mortality study like this one, because it requires observation of an individual and/or his offspring with the same deformity over several consecutive years. the potential involvement of nutritional or physiological etiologies remains to be determined. based on the available data, calf mortality appears to be a problem in at least one localized group on mainland nova scotia. in other north american moose populations, calf mortality is most often associated with regulator, it is strongly suggested to require sympatry with both wolves (canis lupus) and a bear species (ursus americanus or u. arctos) (franzmann et al. 1980, ballard et al. 1981, boer 1988, crête and courtois 1997). in the province (nova scotia department of natural resources, unpublished data). historically, the eastern wolf (c. l. lycaon) was rare 1969, lohr and ballard 1996). however, the eastern coyote (canis latrans) is present in the scotia during the mid-1970s (moore and parker 1992). while coyotes are reported to be a predator of large cervids in other studies recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 98 (smith and anderson 1996), the combined impact of bear and coyote predation on the nova scotia mainland moose population has yet to be assessed. as in other cervid populations, health related stressors such as infectious diseases and nutritional problems could also affect the survivability of moose calves (smith and anderson 1996). however, the potential involvement of disease in the death of calves can only be documented through post mortem ously in this paper, from what limited data are available, reproduction and calf survival in the mainland nova scotia moose appear to vary temporally and spatially. however, it is mortality can be a major factor in the decline of a moose population. therefore, a moose calf mortality study is required to address this despite the lack of supporting evidence of clinical disease associated with trace elecopper, manganese, selenium, and zinc and/ or high levels of cadmium may impact the health of individual animals, either directly or through interactions with other factors (e.g., infectious and noninfectious diseases, harsh environmental conditions, and habitat limitations), cannot be dismissed (pollock 2005). therefore, ongoing trace element monitoring, especially in cases of natural mortality, is essential to discover the potential role of these compounds in the ongoing decline of the nova scotia mainland moose population. in conclusion, 22 animals represent a small sample size. however, the necropsy results be contributing to the decline of the mainland moose population or impeding the recovery process. continued disease surveillance, with complete necropsies when possible, integrated with additional targeted research designed to lighted above, is essential to determine the impact of health-related issues on the recovery of the endangered mainland nova scotia moose population. habitat suitability and availability moose have a relatively small home range (youmans 1999). use of selected areas is learned and traditional with the home range of individual animals encompassing adequate resources to meet their basic year-round needs of water, food, cover, and security. required habitat features for moose related to thermal relief, security, and aquatic resources, as well setting. a review of research in nova scotia and maine indicates that key habitat components vary seasonally and biogeographically distances are required to meet annual and lifehistory needs, especially for cows (prescott 1987, miller 1989, brannen 2004). winter moose range is reported as being unimpeded by average snow depth in mainland nova scotia greater than 13 m in height with 40 – 50% canopy closure (of mature softwood in years and areas with deep snow), young or successional forage stands nearby (less than 30 years since disturbance), and/or partial cuts with remaining overstory and residual stands miller 1989, brannen 2004). moose avoid poorly drained lowlands, and hardwood and hardwood-softwood stands in winter (thomp(abies balsamea), paper (betula papyrifera) and yellow birch (b. alleghaniensis), sugar maple (acer saccharum), mountain maple (a. spicatum), aspen (populus spp.), willow (salix spp.), and other shrubs (prescott 1968, alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 99 by cows of wet lowlands and softwoods, hardand softwood with canopy closure ranging from 0 to 60%, aquatic areas and licks, residual stands, areas with dense to moderate overstory for hiding and cover, and an interspersion of 1986, thompson 1987, miller 1989, brannen 2004). for calving, cows select secluded, undisturbed, poorly-drained lowlands, which have young stands nearby for food and are less than 100 m from water (leptich 1986, thompson 1987, miller 1989). minimum critical habitat area of habitat required to support a viable population or metapopulation of moose in nova scotia. minimum critical area (mca) is a rough measure of the quantity (spatial-area) of suitable habitat required to support a viable population over time, and is calculated as a function of viable population size, home range size, and moose density. it is a useful measure for setting habitat targets for species recovery and conservation. snaith and beazley (2002) conducted coarse mca calculations for moose in mainland nova scotia using two methods: (1) multiplying shortand long-term n by home range size; and (2) dividing n by density. n (500/5,000) multiplied by the mean (40 km2) annual home range size for moose in mainland nova scotia (~55 km2) (d. brannen, nova scotia department of natural resources, personal communication) and in the acadian forest region (~25 km2) (dunn 1976, crossley and gilbert 1983, crête 1987, leptich and gilbert 1989, mcnicol 1990) results in shortand long-term mcas of 20,000 and 200,000 km2, respectively. n (500/5,000) divided by mean density (0.05 / km2) of moose in mainland nova scotia (0.01 – 0.09 / km2) (pulsifer and nette 1995) results in shortand long-term mcas of 10,000 and 100,000 km2, respectively. obviously there is considerable uncertainty in such simple and coarse estimates and the range of mca calculations they produce. nevertheless, they indicate that to maintain viable populations of moose in nova scotia over both short and long terms, and that long-term viability is likely dependent upon continued habitat connectivity to new brunswick since the total area of mainland nova scotia is ~45,000 km2, which is well below long-term mca estimates. habitat suitability analysis life-history variations in habitat requirements ability and availability and to manage forests for moose. however, preliminary assessments of the availability of forage, security, thermal cover, and aquatic habitat components in mainland nova scotia have been conducted, though they are relatively coarse, localized, and/or incomplete (snaith 2001, snaith et al. 2002, brannen 2004, kwan 2005). snaith (2001) (also see snaith et al. 2002) conducted a coarse-scale habitat suitability analysis to quantify the proportional availability of 4 habitat components (forage, softwood cover, model ii. the results indicated little optimal habitat according to all 6 theoretical models 2 of very good and good habitat (snaith et al. 2002). logistic regression analysis indicated no hsi values and moose pellet presence/absence on pellet group inventory (pgi) transect lines (nova scotia department of natural resources, unpublished data); however, forage ence/absence. this obviously raises questions about the validity of the model, its assumptions, and application in mainland nova scotia importance of habitat components for moose recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 100 regions. although hsi values alone were unable to predict moose pellet presence/absence, in subsequent multivariate regression analyses, snaith et al. (2002) found statistically p < 0.05) between moose pellet presence/absence and: (1) road density alone; (2) hsi values and road density combined; (3) road density after hsi values are accounted for; and (4) 2 of the 6 hsi equations after road density is accounted for. these results must be interpreted with caution since pellets indicate only winter habitat use, and the pgi data are incomplete and contain sampling biases. nonetheless, it appears that higher road densities may reduce the effectiveness of otherwise suitable habitat for moose (see beazley et al. 2004 for a review), though between roads, habitat suitability, and moose distribution in nova scotia are not known. habitat disturbance and human access in moose or demographic changes in moose populations in response to human activities (youmans 1999). according to a review of to north american ungulates provided by youmans (1999), responses of these animals to human-induced disturbances may include areas of desirable habitat. moose have been by cross-country skiers, snowmobilers, and hunters (forman et al. 1997, jalkotzy et al. adapt to and tolerate disturbances caused by human activities that are predictable and nonthreatening (geist 1971, shank 1979, westworth et al. 1989), backcountry recreational pursuits in the form of hiking, mountain biking, angling, snowshoeing, cross-country skiing, off-highway-vehicle use, and snowmobiling are not highly predictable either spatially or temporally (shank 1979, ferguson and keith 1982, rudd and irwin 1985). although moose hunting is illegal on mainland nova scotia, poaching and hunting of other species may be major sources of human-induced disturbance. in the province are unknown, these activities not only cause a direct loss of animals, but may, even at low frequencies, effectively elevate the level of disturbance associated with other back-country human activities. the inundation of moose winter habitat by recreational snowmobilers over the past 40 years is not unique to nova scotia, nor is the more recent invasion of moose spring, summer, and fall new generation of recreational off-highway wheeled vehicles. the impact of these vehicles on nova scotia's natural environment may be tions because of the province's relatively small small privately owned parcels, and very large number of access roads and trails, combining to make regulation and control of this form of there is considerable debate on the direct and indirect effects of roads, trails, and other linear developments relative to direct mortality from vehicle collisions, habitat fragmentation and avoidance, displacement and disturbance, browse on the vegetation along road verges, they also suffer direct mortality from collisions and indirect effects resulting from the access afforded to humans and other species, which opportunities (timmermann and gollat 1982, lyon 1984, boer 1990, jalkotzy et al. 1997, gucinski et al. 2001). the fact that vehicular the mainland nova scotia moose population is supported by data discussed in the health section of this paper. roads and their use alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 101 also provide predators, such as the eastern coyote, and other species, such as white-tailed deer, access into moose habitat. facilitating deer movement into moose habitat refugia serves to decrease the spatial separation between moose and deer, and consequently increases the likelihood of transmission of p. tenuis moose (see health section). while indirect, these effects may nonetheless be substantial, especially in combination; yet, a direct causeto establish or prove. nonetheless, it is well documented that moose and other ungulates are negatively affected by roads, trails, and other linear developments (for summaries see beazley et al. 2004, jalkotzy et al. 1997, and gucinski et al. 2001). habitat loss, conversion, degradation, and fragmentation result from roads and related human activities and developments such as agriculture, forestry, mining, and urbanizasuitable habitat are eliminated or reduced to small, isolated patches that can no longer support viable localized groups of moose over time. for species that are sensitive to human activities, such as herbivores, road density is often the most accurate predictor of habitat effectiveness (lyon 1983, thiel 1985, noss et al. 1999). consequently, this factor has been counter-indicator for suitable habitat for large vertebrates and ecological integrity (forman 1995, noss 1995, rudis 1995, forman et al. 1997, noss et al. 1999). although statistical analyses have indicated a correlation between road density and moose pellet presence/absence in mainland nova scotia (snaith et al. 2002, beazley et al. 2004), it is unclear whether this is as a consequence of direct or indirect effects of the roads themselves, the access that roads provide to humans, or some other factor for which road density may serve as a surrogate, such as human settlement, population density, forest harvesting intensity, or other activities or developments. nonetheless, brannen (2004) also found that the length of primary and secondary roads in study areas in mainland nova scotia had a negative effect on moose presence. the level and type of use of the road or trail affects the nature and magnitude of the associated impact, thus limiting our ability to come to general conclusions about cause-effect relationships and potential management options. nonetheless, given the necessity other linear corridors for recreation, forestry, urban and rural developments, and many other uses, the challenges in deriving politically acceptable and defensible solutions are not insubstantial. approaches to minimize direct and indirect human-related disturbances to moose and the conversion and fragmentation public awareness and budgetary enhancements will be necessary to achieve effective tions and to undertake habitat restoration and conservation efforts. distribution and climate change the southern limit of their natural distribution on the north american continent. issues of heat stress, thermal cover, snow depth, moose/ deer separation, and temperature and precipitation changes associated with climate change raise questions about the persistence of moose at this latitude. climate change scenarios for mainland nova scotia project that by 2080 mean annual temperatures will rise from curin the southwest (i.e., tobeatic localized group) ized group) (canadian institute for climate change, canadian climate impact scenarios, cgi?scenarios). temperature rises are likely to recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 102 result in increased thermal stress, particularly in summer. consequently, moose may spend more time under thermal cover during the heat of the day and travel at night to feed, thereby and therefore potential risks to calves and calf mortality. aquatic resources, thermal cover areas and at higher-elevations are likely to become increasingly important. while further warming would be to the detriment of moose in terms of thermoregugreater abundance and distribution of deer p. tenuis. total annual days with snow are projected to be less than half of current rates in both the southwest and northeast (from 51 to 23 and 42 to 19, respectively), while days with rain are predicted to increase (canadian institute for climate change, canadian climate impact scenarios, http://www.cics.uvic. in annual days-with-snow, increases in dayswith-rain, and increases in mean winter temlower snow depths. lower snow depths may also facilitate deer movement into areas from ing habitat refugia for moose from white-tailed to p. tenuis. a direct positive relationship has been established between warmer climates, higher winter precipitation, and prevalence of a lungworm parasite similar to p. tenuis (ball et al. 2001), thus suggesting a potential increase in the prevalence of p. tenuis as a consequence of warming. potential climate-change-induced increases in direct and indirect factors, such as thermal stress and p. tenuis-related mortality, while climatic changes may ultimately limit the ability for moose recovery in nova scotia, the predictions indicate that measures for increasing thermal cover and for addressing p. tenuis-related mortality will be even more critical in the future than at present. furthermore, if genetic diversity is valued as a component of biodiversity conservation, measures to retain genetic material of localized groups of a. a. americana (i.e., as a species and as individuals hypothetically adapted for persistence at the southern limits of their contemporary range) may be warranted even if the material does not reside in mainland nova scotia. in the ability of moose to persist in mainland nova scotia, portions of the genetic material translocations of individuals to more northerly regions. regardless, the uncertainties around the projections and the consequent impacts on moose require a precautionary and adaptive approach to recovery planning. conclusion it is apparent that the number of moose in mainland nova scotia is in decline and that it may be as a consequence of one or all of several direct and indirect factors related to habitat suitability and effectiveness, human access and disturbance, population demographics and viability, genetics, health, and climate and knowledge gaps with respect to these factors and their interrelationships that present a challenge to recovery planning. while areas of apparently suitable habitat remain, there are questions as to their effectiveness for moose where road densities are high and white-tailed deer are present, since moose appear to be only occupying a small portion of otherwise suitable habitat (possibly in refugia), which is isolated from deer and limit human access and to monitor and prevent illegal activities due to the proliferation of roads, trails, and off-highway vehicles, which facilitates access into otherwise remote areas, some of which may currently be functioning alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 103 moose consist of small numbers of individuals that are isolated from each other and unlikely to be functioning as a metapopulation. these localized groups are at population levels that are well below estimates of minimum viable population size, especially for the long term, suggesting negative implications for genetic viability and an unacceptably high probability ing isolated groups through physical habitat restoration at the landscape and regional scales are limited due to physical barriers such as long distances, major highways, and the small numbers of remnant moose. opportunities for increasing the number of individuals within localized groups are hindered by various mortality factors related to trauma from vehicular collisions and gunshots, p. tenuis, icity effects, and our lack of understanding of the interrelationships among these and genetic and environmental factors. climate southern limit of their range and predictions of future climate change indicate increases in temperature and decreases in snow cover, both of which may increase heat stress and transmission of p. tenuis. moose recovery in mainland nova scotia will increasingly need to address thermal cover requirements and p. tenuis-related mortality. effect relationships between these factors and moose decline, since it is likely that the decline is the result of synergistic and cumulative effects of several, if not all, of the factors. addressing these challenges will require support to acquire the requisite data and to adopt a precautionary and adaptive approach to recovery planning. acknowledgements we thank gerry parker for contributions to the introductory section, other members of nova scotia’s american moose recovery planning team whose discussions and ideas ment, the editors of alces, and anonymous reviewers. references alexander, c. e. 1993. the status and management of moose in vermont. alces 29: 187-195. allen, a. w., p. a. jordan, and j. w. terrell models: moose, lake superior region. u.s. department of the interior biological report 82 (10.155). u.s. department of the interior, fish and wildlife service, washington, d.c., usa. anderson, r. c. 1971. neurologic disease in reindeer (rangifer tarandus tarandus) introduced in ontario. canadian journal of zoology 49: 159-166. _____. 1972. the ecological relationships of meningeal worm and native cervids in north america. journal of wildlife diseases 8: 304-310. arsenault, a. a. 2000. status and management of moose (alces alces) in saskatchewan. fish and wildlife branch technical report 00-1:1-84. saskatchewan environment and resource management. saskatchewan environment, fish and wildlife branch, regina, saskatchewan, canada. ball, m. c., m. w. lankester, and s. p. mahoney. 2001. factors affecting the distribution and transmission of elaphostrongylus rangiferi (protostrongylidae) in caribou (rangifer tarandus caribou) of newfoundland, canada. canadian journal of zoology 79: 1265-1277. _____, r. pither, m. manseau, j. clark, s.d. petersen, s. kingston, n. morrill, and p. wilson. 2007. characterization of target recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 104 nuclear dna from faeces reduces technical issues associated with the assumptions of low-quality and quantity template. conservation genetics 8: 577-586. ballard, w. b., t. h. spraker, and k. p. taylor. 1981. causes of neonatal moose calf mortality in south central alaska. journal of wildlife management 45: 335-342. basquille, s., and r. thompson. 1997. moose (alces alces) browse availability and utilization in cape breton highlands national park. parks canada technical report in ecosystem science 10:1-37. parks canada, historical properties, beazley, k. f., t. v. snaith, f. mackinnon, and d. colville. 2004. road density and potential impacts on wildlife species such as american moose in mainland nova scotia. proceedings of the nova scotian institute of science 42: 339-357. beier, p. 1993. determining minimum habitat areas and habitat corridors for cougars. conservation biology 7: 94-108. beissinger, s. r., and m. i. westphal. 1998. on the use of demographic models of population viability in endangered species management. journal of wildlife management 62: 821-841. belovsky, g. e., c. mellison, c. larson, and p. a. vanzandt 286: 1175-1177. bender, l. c., s. m. schmitt, e. carlson, j. b. haufler, and d.e. beyer, jr. 2005. mortality of rocky mountain elk in michigan due to meningeal worm. journal of wildlife diseases 41: 134-140. benson, d. a. 1958a. moose “sickness” in nova scotia – i. canadian journal of comparative medicine 22: 244-248. _____. 1958b. moose “sickness” in nova scotia – ii. canadian journal of comparative medicine 22: 282-286. _____, and g. d. dodds. 1977. the deer of nova scotia. department of lands and forests, cat/77/126/5m. department of scotia, canada. boer, a. h. 1988. moose, alces alces, calf mortality in new brunswick. canadian field-naturalist 102: 74-75. _____. 1990. spatial distribution of moose kills in new brunswick. wildlife society bulletin 18: 431-434. bontaites, k. m., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29: 163-167. boyce, m. s. 1992. population viability analysis. annual review of ecology and systematics 23: 481-506. brannen, d. c. 2004. population parameters and multivariate modeling of winter habitat for moose (alces alces) on mainland nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8: 1309-1315. brussard, p. f. 1985. minimum viable populations: how many are too few? restoration management notes 3: 21-25. caprio, m. a. 2001. source-sink dynamics between transgenic and nontransgenic habitats and their role in the evolution of resistance. journal of economic entomology 94: 698-705. carrasco l., y. fierro, j. m. sánchez-castillejo, j. hervás, j. pérez, and j. c. gómez-villamados. 1997. abnormal antler growth associated with testicular hypogonadism in red deer. journal of wildlife diseases 33: 670-672. cioffi, p. j. 1981. winter cover and browse selection by moose in maine. m.sc. thesis, university of maine, orono, maine, usa. crête, m. 1987. the impact of sport hunting alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 105 on north american moose. swedish wildlife research supplement 1: 553-563. _____, and r. courtois. 1997. limiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal of zoology (london) 242: 765-781. crossley, a., and j. r. gilbert. 1983. home range and habitat use of female moose in northern maine a preliminary look. transactions of the northeast section of the wildlife society 40: 67-75. dias, p. c. 1996. sources and sinks in population biology. trends in ecology and evolution 11: 326–330. dodds, d. g. 1963. the present status of moose (alces alces americana) in nova scotia. proceedings of the northeast wildlife conference 2: 1-40. _____. 1975. distribution, habitat and status of moose in the atlantic provinces of canada and northeastern united states. naturaliste canadien 101: 51–65. _____, e. mullen, and a. martell. 1969. nova scotia. nova scotia department of lands and forests, internal report. department of natural resources library, dunn, f. 1976. behavioural study of moose. maine department of inland fish and wildlife project w-66-r-6 job 2-1. maine maine department of inland fisheries and wildlife, augusta, maine, usa. fahrig, l., and g. merriam. 1994. conservation of fragmented populations. conservation biology 8: 50-59. ferguson, m. a. d., and l. b. keith. 1982. of moose and elk in elk island national park, alberta. canadian field-naturalist 96: 69-78. forman, r. t. t. 1995. land mosaics: the ecology of landscapes and regions. cambridge university press, cambridge, massachusetts, usa. _____, d. s. friedman, d. fitzhenry, j. d. martin, a. s. chen, and l. e. alexander. 1997. ecological effects of roads: toward three summary indices and an overview of north america. pages 40-54 in k. canter, editor. habitat fragmentation and infrastructure. minister of transport and public works and water management, delft, netherlands. franklin, i. r. 1980. evolutionary change in small populations. pages 135-150 in conservation biology: an evolutionaryecological perspective. sinauer associates, sunderland, massachusetts, usa. franzmann, a. w., c. c. schwartz, and r. o. peterson. 1980. moose calf mortality in summer on the kenai peninsula, alaska. journal of wildlife management 44: 764-768. geist, v. 1971. a behavioral approach to the management of wild ungulates. pages 413-424 in e. duffey and a. s. watt, animal and plant communities for conservation. eleventh symposium of the british ecological society, university of east anglia, norwich, july 7-9, 1970. gilpin, m., and m. e. soulé. 1986. minimum viable populations: processes of species in m. e. soulé, editor. conservation biology: the science of scarcity and diversity. sinauer associates, sunderland, massachusetts, usa. gogan, p. j. p., d. a. jessup, and r. h. barrett. 1988. antler anomalies in tule elk. journal of wildlife diseases 24: 656-652. gucinski, h., m. furniss, r. ziermer, and m. brookes. 2001. forest service roads: a states department of agriculture, forestry service, pacific northwest research station, portland, oregon, general technical report pnw-gtr-509.1. recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 106 haig, s. m. 1998. molecular contributions to conservation. ecology 79: 413-425. hedrick, p. w., and p. s. miller. 1992. rare alleles, mhc and captive breeding. pages 187-204 in v. loeschcke, j. tomiuk, and s. k. jain, editors. conservation genetics. birkhauser, basel, switzerland. henriksen and critique of minimum viable population size. fauna norvegica 18: 33-41. jalkotzy, m. g., p. i. ross, and m. d. nasserden. 1997. the effects of linear developments on wildlife: a review of services limited, prepared for canadian association of petroleum producers, calgary, alberta, canada. kearney, s. r., and f. f. gilbert. 1976. habitat use by white-tailed deer and moose on sympatric range. journal of wildlife management 40: 645-657. kwan sence of established moose (alces alces) populations in southeastern cape breton island, nova scotia. mes thesis, dalcanada. lande in demographic models of territorial populations. american naturalist 130: 624-635. lehmkhul, j. f. 1984. determining size and dispersion of minimum viable populations for land management planning and species conservation. environmental management 8: 167-176. leptich, d. j. 1986. summer habitat selection by moose in northern maine. m.sc. thesis, university of maine, orono, maine, usa. _____, and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53: 880-885. levins in m. gerstenhaber, editor. some mathematical questions in biology. american mathematical society, providence, rhode island, usa. lindenmayer, d. b., m. a. burgman, h. r. akcakaya, r. c. lacy, and h. p. possingham. 1995. a review of three models for metapopulation viability analysis - alex, ramas/space and vortex. ecological modeling 82: 161-174. lohr, c., and w. b. ballard. 1996. historical occurrence of wolves, canis lupus, in the maritime provinces. canadian fieldnaturalist 110: 607-610. lyon, l. j. 1983. road density models describing habitat effectiveness for elk. journal of forestry 81: 592-595. _____. 1984. road effects and impacts on in 7th proceedings of the forest transportation symposium, casper, wyoming, usa. december 1113, 1984. manel, s., m. k. schwartz, g. luikart, and p. taberlet. 2003. landscape genetics: combining landscape ecology and population genetics. trends in ecology and evolution 18: 189-197. marburger, r. g., r. m. robinson, j. w. thomas, m. j. andregg, and k. a. clark. 1972. antler malformation produced by leg injury in white-tailed deer. journal of wildlife diseases 8: 311-314. mcnicol, j. 1990. moose and their environment. pages 11-18 in m. buss and r. truman, editors. the moose in ontario, book 1. ontario ministry of natural resources, queen’s printer for ontario, toronto, ontario, canada. miller, b. k. 1989. seasonal movement patterns and habitat use of moose in northern new hamphire. m.sc. thesis, university of new hampshire, durham, new hampshire, usa. moore, g. c., and g. r. parker. 1992. colonization by the eastern coyote (canis latrans). pages 23-37 in a. h. boer, editor. ecology and management of the alces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 107 eastern coyote. wildlife research unit, fredericton, new brunswick, canada. morris, k., and k. elowe. 1993. the status of moose and their management in maine. alces 29: 91-97. newmark, w. d. 1985. legal and biotic boundaries of western north american national parks: a problem of congruence. biological conservation 33: 197-208. noss, r. f. 1995. maintaining ecological integrity in representative reserve networks. world wildlife fund canada, toronto, ontario, canada. _____, e. dinerstein, b. gilbert, m. gilpin, b. j. miller, j. terborgh, and s. trombulak. 1999. core areas: where nature reigns. pages 99-128 in m. e. soulé and j. terborgh, editors. continental regional reserve networks. island press, washington, d.c., usa. novaro, a. j., k. h. redford, and r. e. bodmer. 2000. effect of hunting in source-sink systems in the neotropics. conservation biology 14: 713-721 nudds, t. d. 1990. retroductive logic in retrospect: the ecological effects of meningeal worms. journal of wildlife management 54: 396-402. o’brien, s. j. 1994. a role for molecular genetics in biological conservation. proceedings of the national academy of sciences of the united states of america 91: 5748-5755. o’hara, m. t., g. carroll, p. barboza, k. mueller, j. blake, v. woshner, and c. willetto. 2001. mineral and heavy metal status as related to a mortality event and poor recruitment in a moose population in alaska. journal of wildlife diseases 37: 509-522. parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. report prepared for the nova scotia department of canada. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. pimlott, d. h., and w. j. carberry. 1958. north american moose transplantations and handling techniques. journal of wildlife management 22: 51–62. pollock, b. 2005. trace element status of white-tailed deer (odocoileus virginianus) and moose (alces alces) in nova scotia. nova scotia department of natural resources and canadian cooperative wildlife health centre, final report. department of natural resources library, prescott, w. h. 1968. a study of winter concentration areas and food habits of moose in nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. province of nova scotia. 1934. report of the department of lands and forests 1933. minister of public works and mines, pulsifer, m. d. 1995. moose herd perseveres. nova scotia conservation 19: 6-7. _____, and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. reed, j. m., p. d. doerr, and j. r. walters. 1986. determining minimum population sizes for birds and mammals. wildlife society bulletin 14: 255-261. rudd, l. t., and l. l. irwin. 1985. wintering moose vs. oil/gas activity in western wyoming. alces 21: 279-298. rudis, v. a. 1995. regional forest fragmentation effects on bottomland hardwood community types and resource values. landscape ecology 10: 291307. ryman, n., r. baccus, c. reuterwall, and m. h. smith. 1981. effective population size, generation interval, and potential loss of genetic variability in game species recovery planning for moose in nova scotia – beazley et al. alces vol. 42, 2006 108 under different hunting regimes. oikos 36: 257-266. salwasser, h., s. p. mealey, and k. johnson. 1984. wildlife population viability: a question of risk. north american wildlife and natural resources conference 49: 421-439. samson, f. b. 1983. minimum viable populations a review. natural areas journal 3: 15-23. _____, f. perez-trejo, h. salwasser, l. f. ruggiero, and m. l. shaffer. 1985. on determining and managing minimum population size. wildlife society bulletin 13: 425-433. shaffer, m. l. 1981. minimum population sizes for species conservation. bioscience 31: 131-134. _____. 1990. population viability analysis. conservation biology 4: 39-40. shank, c. c. 1979. human related behavioral disturbance to northern large mammals: a bibliography and review. report prepared for foothills pipe lines (south yukon) limited, calgary, alberta, canada. smith, b. l., and s. h. anderson. 1996. patterns of neonatal mortality of elk in northwest wyoming. canadian journal of zoology 74: 1229-1237. smith, h. j., and r. m. archibald. 1967. moose sickness, a neurological disease of moose infected with the common cervine parasite elaphostrongylus tenuis. canadian veterinary journal 8: 173-177. _____, _____, and a. h. corner. 1964. elaphostrongylosis in maritime moose and deer. canadian veterinary journal 5: 287-296. snaith, t. v. 2001. the status of moose in mainland nova scotia: population viability and habitat suitability. mes nova scotia, canada. _____, and k. beazley. 2002. application of population viability theory to moose in mainland nova scotia. alces 38: 193-204. _____, _____, f. mackinnon, and p. duinker. 2002. preliminary habitat suitability analysis for moose in mainland nova scotia, canada. alces 38: 73-88. soulé, m. e. 1980. thresholds for survival: potential. pages 151-169 in m. e. soulé biology: an evolutionary-ecological perspective. sinauer associates, sunderland, massachusetts, usa. thiel, r. p. 1985. relationship between road densities and wolf habitat suitability in wisconsin. american midland naturalist 113: 404-407. thomas, c. d. 1990. what do real population dynamics tell us about minimum viable population sizes. conservation biology 4: 324-327. thomas, j. e., and d. g. dodds. 1988. brainworm, parelaphostrongylus tenuis, in moose, alces alces, and white-tailed deer, odocoileus virginianus, of nova scotia. canadian field-naturalist 102: 639-642. thompson, m. e. 1987. seasonal home range and habitat use by moose in northern maine. m.sc. thesis, university of maine, orono, maine, usa. tiller, b. l., g. e. dagle, and l. l. cadwell. 1997. testicular atrophy in a mule deer population. journal of wildlife diseases 33: 420-429. timmermann, h. r., and gollat, r. 1982. related to season manipulation and access. alces 18: 301-328. _____, and j. g. mcnicol. 1988. moose habitat needs. forestry chronicle 64: 238-245. vecellio, g. m., r. d. deblinger, and j. e. cardoza. 1993. status and management of moose in massachusetts. alces 29: 1-7. vukelich, m. f. 1977. reproduction and proalces vol. 42, 2006 beazley et al. recovery planning for moose in nova scotia 109 ductivity of moose in nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. wangersky, r. 2000. too many moose? canadian geographic nov/dec: 44-56. watkinson, a. r., and w. j. sutherland. 1995. sources, sinks, and pseudo-sinks. journal of animal ecology 64: 126–130. wayne, r. k., n. lehman, m. a. allard, and r. l. honeycutt. 1992. mitochondrial dna variability of the gray wolf: genetic consequences of population decline and habitat fragmentation. conservation biology 6: 559-569. westworth, d., l. brusnky, j. roberts, and h. veldhuzien. 1989. winter habitat use by moose in the vicinity of an open pit copper mine in north-central british columbia. alces 25: 156-166. whitlaw, h. a., and m. w. lankester. 1994. a retrospective evaluation of the effects of parelaphostrongylosis on moose populations. canadian journal of zoology 72: 1-7. youmans, h. 1999. project overview. pages 1.1-1.18 in g. joslin and h. youmans, coordinators. effects of recreation on rocky mountain wildlife: a review for montana. committee on effects of recreation on wildlife, montana chapter of the wildlife society. (accessed march 2006). alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 81 an examination of the absence of established moose (alces alces) populations in southeastern cape breton island, nova scotia, canada karen beazley1, helen kwan1,2, and tony nette3 1dalhousie university, 6100 university avenue, suite 5010 , halifax, nova scotia, canada, b3h 3j5; 2jacques whitford ltd., environmental planning and permitting, 7271 warden avenue, markham, on, l3r 5x5; 3nova scotia department of natural resources, wildlife division, 136 exhibition street, kentville, ns, b4n 4e5, canada; e-mail: karen.beazley@dal.ca abstract: an analysis was performed on habitat-related factors for the southeastern side of cape breton island, nova scotia to investigate the continued absence of moose (alces alces) from the region. temperature and snow depth, at times, reach levels that could cause thermal stress or impede movement of moose; however, it is unlikely that these factors dictate the absence of moose. no clear relationships were established between environmental concentration levels of the heavy metals molybdenum, cadmium, copper, and lead and moose distribution; however, high concentration levels of molybdenum in the cape breton study area warrant further investigation. road density assessments showed that the study area has a higher level of road density compared to 2 mainland control sites; however, higher road density occurs in other areas in which moose persist. anthropogenic factors such as poaching were not considered influential enough to exclude moose. a forest habitat comparison analysis was performed to identify habitat features that were statistically correlated with moose presence, and then were applied in a probability model to predict moose presence in the study area. the logistic regression model used to predict the probability of moose presence was composed of positively associated forest inventory variables (softwood average maturity, hardwood average maturity, % mixed hardwood, % non-forested area, total wetland area) that best fit the data. the model identified 43% of the cape breton study area as having a high-probability weighting for moose presence. overall, this study did not reveal a clearly identifiable cause for the continued absence of moose in southeastern cape breton island. alces vol. 44: 81-100 (2008) key words: alces alces, exclusion factors, geochemistry assessment, geographical features, habitat analysis, moose, nova scotia, probability model, road density. two subspecies of moose are found in nova scotia; indigenous moose (alces alces americana) exist in localized groups in mainland nova scotia, and moose (a. a. andersoni) introduced from alberta, after the extirpation or near-extirpation of the endemic species, exist in northern cape breton island. on the mainland, we use the term localized groups as there is currently insufficient data to determine whether they qualify as distinct subpopulations. genetically, a subpopulation is defined as individuals within a given larger population that are further defined by some recognizable level of relatedness. since this is not determined throughout mainland nova scotia, we use the general term, localized groups (beazley et al. 2006). in 2003, the indigenous mainland moose was listed as ‘endangered’ under the nova scotia endangered species act, with 1000-1200 animals existing in localized groups (parker 2003). the reasons behind the decline of the mainland moose are not entirely understood (beazley et al. 2006). in contrast with their decline, the moose population in northern cape breton reached high densities while subject to the only open moose hunting season in the province, and is currently estimated as 5000 ± 1000 animals (parks canada, unpublished data) (fig. 1). absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 82 the high abundance of moose in the highlands area of cape breton is such that animals are dispersing throughout most of northern cape breton. further, hunting is not permitted in cape breton highlands national park that encompasses approximately 40% of the highlands. there is, however, a notable and unexplained absence of established moose populations in southeastern cape breton, with only occasional sightings. interestingly, black bears (ursus americanus), the only significant natural predator of moose in the province, are also absent from this same area, yet whitetailed deer (odocoileus virginianus) inhabit the area (unpublished data, nsdnr; see macmichael 2007). the absence of two wide-ranging mammals from the same geographical region suggests that there may be issues with habitat suitability, and/or with human activities such as past and current hunting and poaching. this paper examines factors that may help explain the exclusion of moose from southeastern cape breton, as well as the limited distribution and apparent continued decline of surviving remnant localized groups on mainland nova scotia. the 3 objectives of this research were to 1) determine the historical distribution of moose in nova scotia, 2) compare habitat characteristics in the cape breton study area with control sites in coastal mainland habitat to identify potential exclusion factors, and 3) identify any non-habitat associated, anthropogenic exclusion factors. methods the cape breton study area included part of richmond county and all of cape breton county, comprising 318,193 ha (fig. 2). we examined existing archival records and documented the historical distribution of moose across the province, particularly cape breton county, to determine whether moose had inhabited the study area at some time in the past. we then examined moose presence/absence data, geographical characteristics including geology and climate, geochemistry, road density, biogeographic factors, interspecific interactions, and overall forest habitat suitability in the study area, and made comparisons with the control sites in ecologically similar moose habitat on mainland nova scotia. non-habitat factors that could potentially exclude moose from the area, such as poaching and human development,were also identified. with the exception of interviews with key informants, the research was limited to analysis of existing information. for a full description of the fig. 1. current moose distribution in mainland nova scotia and cape breton island, showing core populations and lower-density/fringe-areas. remnants of the indigenous moose population (a. a. americana) exist in small, localized groups in mainland nova scotia. in contrast to the high abundance of moose (a. a. andersoni) in northern cape breton, there is a notable absence of established populations in southeastern cape breton. fig. 2. cape breton study area and mainland control sites on chebucto and chedabucto peninsulas. alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 83 study design refer to kwan (2005). historical distribution archival information was obtained from the public archives of nova scotia and the fortress of louisbourg national historic site in cape breton county. the historian at the fortress (k. donovan) provided additional information on moose presence, or the lack thereof, in the proximity of the historic site in the 1700s. relevant sources were reviewed and any indication that moose were historically present in the study area and surrounding region was noted. key sources included denys (1672), pichon (1760), smith (1801), dodd (1805), holland (1935), benson and dodds (1980), fortier (1983), pulsifer and nette (1995), and landry (1997). habitat assessment habitat assessment was performed in an attempt to determine whether suitable habitat was the limiting factor influencing moose distribution in the cape breton study area. the habitat assessment included examination and comparison of moose presence/absence, biogeographic factors, interspecific interactions, geographical features, road density, geochemical analyses, and forest composition. this coarse-scale habitat comparison used 2 mainland sites (fig. 2) as controls, which required that these sites be as ecologically similar as possible. the chebucto peninsula and the chedabucto peninsula occur within the atlantic coastal ecoregion (neilly et al. 2003) as does a large portion of the cape breton study site, such that these areas share the same coastal influence and have similar climate effects and vegetation types. both of the mainland control sites support localized groups of resident moose, indicating their suitability for comparison with the cape breton study area. moose presence in the study sites was determined by assessing the provincial pellet group inventory (pgi; unpublished data, nsdnr 1983-2003), which is a system of randomly placed 1-km-long sampling transects used to inventory deer and moose pellet groups. the pgi system, originally developed in 1983 to monitor deer population density, includes information on moose pellet group presence. pellet groups deposited after leaf-fall and throughout the winter are counted during the spring survey, and reliably indicate moose presence or absence during late fall-early spring only. the presence of moose pellets on a transect was considered indicative of moose presence, regardless of the number of pellet groups recorded. biogeographic factors were considered by visually assessing maps of the region in terms of physical distances, barriers (including human settlement), or bottlenecks in the landscape that may potentially exclude or restrict moose dispersal into the study area. interspecific interactions between moose and predators, and competitors and pests, particularly white-tailed deer, were examined by qualitatively comparing the study area with sites elsewhere in nova scotia and information in the literature. potentially relevant geographical features of the cape breton and mainland sites including climate, bedrock geology, predominant tree species, soil ph, and elevation were compared qualitatively using data derived from existing sources (table 1). sources on local features include environment canada (2004), ecological land classification for nova scotia (neilly et al. 2003), and the natural history of nova scotia, volume 1: topics and habitats, and volume 2: theme regions (davis and browne 2003a, b). comparisons of these features were conducted to identify potential variations across sites, and to assess any variations against acceptable ranges or limiting values for moose identified in the literature. road densities for the cape breton and mainland sites were extracted from a preexisting gis-based road-density classification absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 84 (1-km square grid; 6 density classes) for nova scotia (snaith 2001, snaith et al. 2002, beazley et al. 2004) and were compared visually. since road density of 0.6 km/km2 has been suggested as a threshold value above which populations of many large vertebrate species decline (forman et al. 1997), road densities for primary and secondary roads were reclassified into 2 density classes (i.e., <0.6 km/ km2 and >0.6 km/km2). the proportion of area with road densities in these 2 classes was compared in each study site. moose sightings (n = 11) from wildlife incidence reports (wir; unpublished data, nsdnr 1985-2004) in the cape breton study area were plotted against the road density values and assessed visually for spatial correlation. the geochemistry of the study area was examined for particular heavy metals known to have negative health effects in moose when they accumulate in the body or are at deficient levels, such as molybdenum, cadmium, and copper (crichton and paquet 2000, frank et al. 2000 a, b, selinus and frank 2000, pollock 2006). a biological imbalance of these metals may cause physiological abnormalities, feature chebucto peninsula (shearwater a weather station) chedabucto peninsula (stillwater sherbrooke weather station) cape breton richmond (sydney a weather station) mean elevation 71 m 115.5 m 119 m mean aspect south-facing south-facing south-facing bedrock 95% granitic with 5% metasandstones of the goldenville formation 50% granitic intrusions with 25% metasandstone units and 25% halifax formation slate south: volcanic and associated plutonic rocks (granitic) north: sandstone, presence of coalbeds predominant tree species red and black spruce red and black spruce with some balsam fir balsam fir with some black spruce and larch ph 5.0-6.5 ≤6.0 6.5-7.5 total annual precipitation 1400-1500 mm 1400-1500 mm 1400-1500 mm average total annual snowfall 200-250 cm 150-250 cm 150-300 cm snow depth maximum (1971-2000) average <10 cm; highest recorded value 84 cm average <20 cm; highest recorded value 92 cm generally moderate; <25 cm; highest recorded value 123 cm average winter (jan-feb) mean maximum temperature (1971-2000) -0.2°c; extreme high recorded 16.2°c, 1994 -0.9°c; extreme high recorded 17.5°c, 1995 -1.9°c; extreme high recorded 18°c, 2000 average winter (jan-feb) mean minimum temperature (1971-2000) -9.2°c; extreme low recorded -26.5°c, 1994 -11°c; extreme low recorded -39°c, 1985 -11.1°c; extreme low recorded -27.3°c, 1994 average summer (jul-aug) mean maximum temperature (1944-2000) 22.4°c; extreme high recorded 33.3°c, 1945 24°c; extreme high recorded 34°c, 1999 23°c; extreme high recorded 34.4°c, 1944 average summer (jul-aug) mean minimum temperature (1961-2000) 13.5°c; extreme low recorded 5.6°c, 1965 12.7°c; extreme low recorded 1.7°c, 1968 12.3°c; extreme low recorded 2.2°c, 1961 fog days 101 115 (canso) 80 table 1. comparison of selected geographical features across the mainland control sites at chebucto peninsula and chedabucto peninsula and the cape breton study area (compiled from davis and browne 2003a, b, neilly et al. 2003, environment canada 2004). alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 85 aberrant behaviour, or have toxic effect on the reproductive and central nervous systems. lead, while not reported to have negative effects on moose, has been known to cause anemia and neuropathy or encephalopathy in mammals (underwood 1971). four geochemical datasets from the nsdnr were screened for concentration levels of molybdenum, cadmium, copper, and lead. these sets included 1) a vegetation survey completed in 1991 (dunn et al. 1992 a, b), 2) stream sediment surveys completed in 1982-83 (rogers and lombard 1990) and 1986-87 (mills 1989), 3) lake sediment surveys completed in 1977-78 (bingley and richardson 1978, richardson and bingley 1980, rogers and lombard 1990) and 1983 (rogers and macdonald 1983 a-g, rogers and lombard 1990), and 4) glacial till surveys initiated in 1977 (stea and fowler 1979, 1981; stea 1982, 1983; stea and grant 1982; stea and finck 1986; turner and stea 1987a, b, 1988a, b) and 1984-89 (bonner et al. 1990). concentrations of metals were “normalized” across sample media (lake sediment, stream sediment, glacial till, and vegetation (i.e., red and black spruce bark)) in an attempt to make meaningful comparisons across the different sample media. we identified the mean value of each element in each dataset, then divided the remainder of the data into standard deviations around the mean. habitat assessment was based on vegetative features associated with mainland sites currently used by moose. habitat associations were created by linking provincial pgi data (unpublished data, nsdnr 1983-2003) that indicated moose presence/absence with forest resource inventory data that identified vegetative features (nsdnr 1999), following methods developed by mackinnon (2001) and brannen (2004). to better ensure that habitat use by moose was identified in this analysis, the plot size for data extraction was set with a 1 km buffer around each transect because a pellet group could occur at any point along a transect; this resulted in 2x3 km plots. the aim was to discover habitat preferences by comparing and contrasting quantifiable forest variables from plots where moose were present against those where moose were absent. the pgi data were statistically analyzed using logistic regression (spss, v.11.5; p<0.05) to allow for direct comparison between pgi transects with moose presence and absence. the presence/absence information was used to extract 32 habitat variables (table 2) from the forest resource inventory following mackinnon (2001) and brannen (2004). correlations between these 32 habitat variables and moose presence were then tested with a stepwise binary logistic regression analysis. the nature of the analysis forced the assumption that the pgi are spatially and temporally independent, which may not be true. the benefit of doing such an analysis is that no prior assumptions are made as to what comprises preferred habitat. from the results of the stepwise binary logistic regression analysis, 3 models were created that should be able to predict the probability of moose presence (table 3). one of these predictive models, model 3, was shown to best fit the data using 2 data fitness tests (-2 log likelihood, and nagelkerke r square) and was also found to better predict moose presence. model 3 appeared to make the most sense biologically to meet moose requirements based on the variables used to construct the model. the log (p/1-p) values produced by these equations were then converted into probability values, which fell within the range 0.0-1.0, by applying the equation, probability = exp(modelx/1 + exp(modelx), where exp(modelx) is equivalent to e (modelx) = 2.7183(logmodelx). these values were then broken into 5 categories for habitat classification (table 4). the models were then applied to the mainland control sites and the cape breton study area using a moving-window technique absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 86 developed by duinker et al. (1991, 1993) and previously applied to mainland nova scotia by snaith et al. (2002). this gis technique calculates a value for each search window, in this case a 2x3 km plot to remain consistent with the data extraction. during the model run, habitat variable description p_sw percent softwood sw_avght softwood average height (m) sw_avgmt softwood average maturity value (class) sw_svgcr softwood average crown closure p_hw percent hardwood stands hw_avght hardwood average height (m) hw_avgmt hardwood average maturity value (class) hw_avgcr hardwood average crown closure p_mswd percent mixed softwood dominant ms_avght mixed softwood average height (m) ms_avgmt mixed softwood average maturity value (class) ms_avgcr mixed softwood average crown closure p_mhwd percent mixed hardwood dominant mh_avght mixed hardwood average height (m) mh_avgmt mixed hardwood average maturity value (class) mh_avgcr mixed hardwood average crown closure p_uc percent unclassified forest sp1_dom dominant tree species in the plot sp1_prop proportion of habitat sp1 occupies sp_rich total number of tree species sp_div species diversity (shannon’s index of diversity) p_forest percent of forested area p_nonforest percent of area that is brush, rock barren, urban, barrens, agriculture, alders, miscellaneous p_clearcut percent of area occupied by clearcuts wtlnd-area total wetland area (ha) stream-len length of streams and rivers (m) lake_area total area of lakes (ha) prim_rds length (m) of primary (paved) roads other than 100 series highways secon_rds length (m) of unpaved roads trails_rds length (m) of trails, wood roads, abandoned roads, abandoned railways elev-mean average elevation (m) hab_div habitat diversity using % softwood, % hardwood, % mixed softwood, % mixed hardwood, wetlands, barrens, clearcuts, and agriculture (shannon’s index of diversity) table 2. at total of 32 habitat variables were extracted from the forest resource inventory (nsdnr 1999) through statistical analysis (logistic regression (spss, v.11.5; p<0.05)) of forest variables from pgi-transect plots with moose presence and absence. correlations between these 32 habitat variables and moose presence were then tested with a stepwise binary logistic regression analysis to create predictive models (see table 3) of the probability of moose presence in the mainland control sites and the cape breton study area. alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 87 the assessment units, or “windows”, overlap with neighboring units by 50% to permit each forest stand to contribute to the calculation several times. this is meant to reflect the possibility that moose ranges overlap the boundaries of these arbitrary windows, and an attribute of a forest stand occurring just outside 1 window may affect the value of the adjacent window. this mechanism, following snaith et al. (2002), is used to account for moose home ranges that encompass many forest stands, and the fact that moose range freely between these stands in spatial patterns difficult to predict. the roving window technique assesses a greater number of possibilities for home range composition based on the forest data. once the units are amalgamated, they show a spatial pattern of how the different classifications of habitat are distributed across the landscape. the models were applied to the cape breton and mainland sites to identify and compare the extent to which suitable habitat exists at each site. non-habitat exclusion factors informal, semi-structured interviews with key informants (n = 11) were conducted to identify potential non-habitat exclusion factors. participants included nsdnr staff, university professors, forest industry employees, and staff from museums and educational centres local to richmond and cape breton counties. common questions were asked of participants about the historical and current presence of moose in the area, whether they believe that sufficient quality and quantity of habitat exists in the area to support moose, why they think moose may be excluded from the area, and current local societal attitudes towards moose. their responses were recorded and qualitatively assessed to identify the range of factors indicated and areas of agreement model 1 mixed softwood stand average maturity, % mixed hardwood, % non-forested area, total wetland area model 1 = log (p/1-p) = ms_avgmt + p_mhwd + p_nonfor + wtlnd_area -16.993 + 1.669(mixedsw_maturity) + 0.621(%mixedhw) + 0.078(%nonforested) + 0.059(wetland area) model 2 softwood stand average maturity, mixed softwood stand average maturity, % mixed hardwood, % nonforested area, total wetland area model 2 = log (p/1-p) = sw_avgmt + ms_ avgmt + p_mhwd + p_nonfor + wtlnd_ area -28.843 + 2.315(sw_maturity) + 1.234(mixedsw_maturity) + 0.718(%mixedhw) + 0.105(%nonforested) + 0.101(wetland area) model 3 softwood stand average maturity, hardwood stand average maturity, % mixed hardwood, % non-forested area, total wetland area model 3 = log (p/1-p) = sw_avgmt + hw_ avgmt + p_mhwd + p_nonfor + wtlnd_ area -53.289 + 6.555(sw_maturity) + 1.340(hw_maturity) + 0.844(%mixedhw) + 0.158(%nonforested) + 0.161(wetland area) table 3. three predictive models (models 1-3) were created from correlations revealed through a stepwise binary logistic regression analysis of 32 habitat variables (see table 2 for habitat variable descriptions) and moose presence on pgi-transect plots. using 2 data fitness tests (-2 log likelihood, and nagelkerke r square), model 3 was shown to best fit the data and predict moose presence. probability value class rating 0.0 – 0.199 1 very low probability 0.2 0.399 2 low probability 0.4 – 0.599 3 medium probability 0.6 0.799 4 good probability 0.8 – 1.0 5 high probability table 4. probability values for moose presence were derived from log (p/1-p) values produced by models 1-3, then reclassified for rating habitat. the models were then applied to the cape breton study area and the mainland control sites to determine the percentage of land area in each probability class for moose presence (see table 5). absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 88 and disagreement. results historical distribution european settlement of nova scotia took place in the early 1600s, and by all accounts moose and caribou (rangifer tarandus) were abundant and hunted heavily for subsistence (lescarbot 1609, ducreux 1664, denys 1672). wide-scale alteration of the landscape and unrestricted hunting of these species caused the eventual extirpation of caribou and the decline of moose (denys 1672, benson and dodds 1980). historical memoirs from the fortress of louisbourg, a large french settlement established on the cape breton coast in 1713, mention a lack of fresh meat in the area by the mid-1700s (pichon 1760), suggesting that moose were absent from that area of cape breton island. white-tailed deer, while gradually expanding their range into the province, were purposefully introduced on mainland nova scotia in 1894, and spread across the province within 17 years (benson and dodds 1980). moose were eventually extirpated or nearly extirpated from cape breton, while localized groups on the mainland were fragmented and declining. this led to harvest restrictions in the mid-1800s and eventual hunting closures that continued on cape breton until 1980 when it became apparent that a reintroduction of moose had been highly successful. on the mainland, moose numbers fluctuated with intermittent periods of hunting until the last legal hunt in 1981 (benson and dodds 1980, pulsifer and nette 1995). while these and other accounts indicate that moose existed on cape breton island and mainland nova scotia, specific references to the study area do not exist. habitat assessment moose presence - moose have occupied the chebucto peninsula since establishment of the pgi in 1983 (8 of 12 pgi transects with pellet groups; 25-30 resident moose; density 1:21-25km2) (fig. 3). the chedabucto peninsula has a sparser and perhaps more recently established moose population, as pellet groups were recorded only since 1995 (2 of 17 pgi transects with pellet groups; 8-12 resident moose; density 1:218-272 km2). in contrast, no pellet groups were found on 28 pgi transects at the cape breton study area, although moose are seen occasionally. biogeographic factors - in 1947-48, 18 moose were introduced to cape breton island from alberta. these moose, and any native moose surviving in the region at that time, form the basis of the current population (broders et al. 1999). moose have thus had >50 years to re-establish in the southeastern side of cape breton, yet continue to be absent. moose expand into new territory at rates ranging from 8-24 km/year (pimlot 1953, mercer and kitchen 1968), and moose in some populations seasonally migrate 40-80 km/year (edwards and ritcey 1956, gillingham and klein 1992, van ballenberghe 1992). some moose have dispersed 100-200 km beyond their normal range (kelsall 1987). given these behaviors and distances, it seems unusual that moose have not expanded into southeastern cape breton since the introduction. bras d’or lake essentially divides cape breton into 2 parts (fig. 4). land bridges that provide the major connectivity for highways and human activity are narrow. to the south, bras d’or lake is separated from the atlantic ocean by a very slender land bridge at st. peter’s inlet, which is narrow and dotted with small islands. movement across st. peter’s inlet is not physically inhibited, but may be influenced by lack of security, disturbance, or poaching associated with the local human population. the north contains a more densely human populated and developed area, including the communities of sydney, sydney mines, and north sydney, and several major highways that might cause an avoidance response. moose would be much more visible and vulnerable alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 89 to the greater traffic in this region. in contrast, there are a number of narrow channels and bays that moose could easily cross to avoid many of these situations. moose are strong swimmers (benson 1957) and have been documented swimming to offshore islands and across large bodies of salt water such as the strait of canso (benson and dodds 1980). access to the southeastern side of cape breton should not influence moose dispersing from the north. as evidence, moose have been observed in southeastern cape breton in the past, but sighting reports typically cease abruptly (a. mclain, manager of heritage protection, halifax parks canada, ret.). these disappearances were attributed to poaching in the absence of any other reasonable explanation or indication that moose dispersed from the area (l. mcdonald, supervisor of forest services, nsdnr, pers. comm.; m. pulsifer, regional biologist, nsdnr pers. comm.; l. reeves, senior park warden, fortress of louisberg, pers. comm.). inter-species interactions -there are no natural predators of moose in the cape breton study area given the absence of black bears, and gray wolves (canis lupus) are extirpated from the province (whitaker 2006). as elsewhere, however, there is overlap between winter browse use by moose and white-tailed deer fig. 3. locations of pellet group inventory (pgi) transects at the two mainland control sites and the cape breton study area. the chebucto control site (3a) had 8 of 12 pgi transects with moose pellet groups, the chedabucto control site (3b) had 2 of 17 pgi transects with pellet groups, and 0 of 28 pgi transects in the cape breton study area (3c) had pellet groups. fig. 4. configuration of the cape breton study area, showing narrow land bridge at st. peter’s and separation of the study area from the remainder of cape breton by bras d’or lake. absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 90 (telfer 1972). studies of moose and whitetailed deer in eastern north america have often demonstrated that when one species increases in numbers, the other declines (dodds 1963, telfer 1970, benson and dodds 1980, pulsifer and nette 1995). this is usually related to “moose sickness” or parelaphostrongylosis as a result of the transmission of the nematode parasite, parelaphostrongylus tenuis, normally carried by white-tailed deer (telfer 1967a,b; peterson et al. 1996, lankester and samuel 1998), not as direct competition for resources. while it is unclear the extent to which the presence of white-tailed deer can depress moose populations through the transmission of p. tenuis, the potential mortality of moose is well documented (thomas and dodds 1988, schmitz and nudds 1994, lankester and samuel 1998, beazely et al. 2006). nonetheless, moose and deer persist together in other regions in nova scotia including the two mainland control sites, as well as other eastern canadian provinces and the northeastern united states. further, there is no evidence of moose in the area prior to establishment of deer in the region. thus, it would seem that this interaction, taken alone, cannot explain the absence of moose from the cape breton study area. in combination with climatic and geological factors, however, p. tenuis may play a role in limiting or excluding moose from the cape breton study area. beazley et al. (2006) suggest that parelaphostrongylosis may regulate moose populations on mainland nova scotia, with localized groups surviving in refugia in elevated regions of the province where whitetailed deer are absent or in low density (t. nette, nsdnr, unpublished data), or in areas with granitic soil that is not compatible with the intermediate molluscan hosts of p. tenuis (r. cameron, nsdnr, unpublished data). similar geographic and interspecific factors may be influencing the distribution of moose in cape breton; moose at high densities exist in the elevated, granitic regions of the cape breton highlands where white-tailed deer densities are low, and moose are limited in southeastern cape breton where elevations are lower and deer densities are higher. this possibility is explored in the following section with reference to relevant geographical differences. geographical comparison -the cape breton study area and the two mainland control sites fall primarily within the atlantic coastal ecoregion. however, the mainland control sites lie within the granite barrens ecodistrict, dominated by granite bedrock, much of it exposed and with acidic soils, whereas the cape breton study area lies primarily within the till plain, a low-lying area with poorly drained surface and neutral ph levels (davis and browne 2003b) (table 1). both moose and deer exist in the granitic mainland control sites, whereas in the cape breton study site deer exist in higher densities and moose are infrequent. moose exist in higher densities in other areas of the atlantic provinces with similar geology to southeastern cape breton. localized groups of moose in the cobequid mountains, antigonish-pictou highland, and cape breton highlands all fall within the avalon terrane in which the cape breton study area is also situated. the volcanic, sedimentary, and minor precambrian plutonic rocks of the cape breton study area extend into the avalon peninsula in eastern newfoundland and the caledonia highlands in new brunswick. these areas are at higher elevations and latitudes and may limit deer or geographically separate them from moose in these areas. conversely, both moose and deer occur in the more granitic and lower altitudes of the meguma terrain in which mainland moose at the control sites and in southwest nova scotia occur. while geographic and interspecific interactions do not separately explain absence of moose in the cape breton study area, in combination, soil types that favor presence of white-tailed deer and the intermediate molluscan hosts alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 91 of p. tenuis, and climatic factors related to elevation and/or coastal influences that support or favour deer, may relate to the absence of moose in the cape breton study area. while these interactions may serve to limit moose at present, and thus warrant further research, it is unlikely that these factors alone could have excluded moose historically, prior to the arrival of deer. thermoregulatory influence on moose presence is possible because nova scotia is in the southernmost range of moose with regard to heat tolerance (telfer 1984, karns 1998). moose are at risk of thermoregulatory stress while in winter coat at temperatures >5.1 °c and at >14 °c when in summer coat (karns 1998). the average maximum winter temperature does not approach 5.1 °c at any study site; however, extreme high winter temperatures, mean maximum summer temperatures, and extreme seasonally high temperatures at all sites exceed the critical temperatures. these temperature values are, however, similar at all the study sites. it is possible that both summer and extreme seasonally high temperatures could cause thermoregulatory stress at all 3 sites; however, since moose occupy the mainland control sites, it seems unlikely that a site-specific effect occurs only in southeastern cape breton. snow depth plays a critical, influencing role on moose movement during winter. snow depth >60 cm can severely impede mobility and add to seasonal energetic demands (prescott 1968, telfer 1970, kelsall and prescott 1971). the ‘highest recorded value’ for snow depth was at the cape breton study area, but it is also well above 60 cm at both mainland control sites. because average snow depth at all sites is well below 60 cm, it is unlikely that snow depth limits moose presence in the cape breton study area. road density -as road density increases, the likelihood of species extirpation also increases from the provision of access for competitors, predators, hunters, and vehicular collisions (noss 1995, forman et al. 1997, nasserden et al. 1997). moose are vulnerable to increased hunting near roads, particularly illegal hunting (lyon 1984, boer 1990), and the increased human activity associated with roads can also disrupt normal moose behavior (jalkotzky et al. 1997, gucinski et al. 2001, beazley et al. 2004, laurian et al. 2008). moose-vehicular mortality is 2-6 annually on mainland nova scotia (beazley et al. 2006). a comparison of the study and control sites demonstrated that while road density appears higher in the cape breton area than either mainland site, substantial areas with low-no road density exist within the cape breton study area, and other areas of nova scotia have higher road density with persistent localized groups of moose (snaith 2001, snaith et al. 2002, beazley et al. 2004). the cape breton study area appears to have a larger area of road density above the threshold (0.6 km/km2) than do the 2 mainland sites (fig. 5), although 11 moose were sighted on roads in 1985-2004. while it is likely that the open road corridors and higher traffic density increased the likelihood of sightings, it would not appear that the presence of roads interferes with moose movement, as only 1 sighting in the study area resulted from a vehicular mortality. moose on chebucto peninsula live in close proximity to halifax and persist to date; however, the question remains as to whether moose are able to persist in areas with relatively high road density given the direct and indirect mortality and disturbance effects. although roads do not explain the historical absence of moose in the cape breton study area, this area has higher road density than both mainland control sites, thus road density could relate to the continued absence of moose and warrants further study. geochemistry -geochemical imbalances may cause a deficiency or excess of trace elements in associated vegetation as the material is naturally liberated from the soil and bedrock through erosion and acid deposition absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 92 (frank 1998, environment canada 2003). heavy metals are highly mobile in acidic conditions (e.g., as associated with acid precipitation) particularly when the parent material has a low ph-buffering capacity (environment canada 2002). these free, naturally-occurring elements can bio-accumulate in herbivores and lead to disease (scanlon et al. 1986, frank 1998, frank et al. 2000 a, b). spatial concentrations of heavy metals (kwan 2005) suggest that there is a high occurrence of molybdenum in the cape breton study area that may warrant further investigation. lead is high at a localized site, a former economically viable mine. however, the lead formation is not likely to have wide-spread effects across the entire region. no discernable pattern of copper and moose presence was found, and cadmium data were available only for northern nova scotia. cadmium levels should be mapped for the rest of the province before conclusions are drawn as to its potential effect on moose habitat suitability; however, cadmium concentration appears high in the cape breton study area. high environmental concentration of a heavy metal does not necessarily translate into a high biological concentration, but may indicate relative availability for uptake. the trace element status of moose and white-tailed deer from various locations across nova scotia was analyzed subsequent to our review (pollock 2006, pollock and roger 2007). no moose from the specific cape breton study area were collected and measured, although deer were. in northern cape breton, fig. 5. road densities in the mainland control sites (5a. chebucto; 5b. chedabucto) and the cape breton study area (5c). road densities ≥0.6 km/km2 are indicated in black tone, whereas intervening areas of <0.6 km/km2 remain white. in figure 5c, grey-tone triangles represent moose sightings reported in wildlife incidence reports (wir) (1985-2001), and hexagons indicate moose wir post-2001. alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 93 concentration of cobalt in moose kidneys, and concentration of zinc and copper in deer were at levels considered marginal (marginally deficient or deficient with reference to values for domestic cattle (bos taurus); puls 1994) (pollock 2006). the concentration of manganese, selenium, and lead in moose livers from cape breton, and of selenium in moose from the western mainland region were also considered marginal. while concentrations of these trace elements are considered marginal in cape breton, the samples were from areas where moose populations are high. there is little evidence that clinical deficiencies of trace elements occur in moose populations in nova scotia (pollock 2006, pollock and roger 2007). however, it is possible that the health of individual animals may be impacted by marginal, deficient, or high levels of trace elements either directly or through interactions with other factors; thus, further monitoring and analyses are warranted. necropsies of 22 mainland and cape breton moose since 1998 found no gross or microscopic lesions compatible with cadmium toxicity, or cobalt, copper magnesium, selenium, and zinc deficiencies, suggesting that clinical disease associated with these trace elements has not occurred in nova scotia (beazley et al. 2006). we conclude that no evidence exists that trace element deficiencies or toxicities act to exclude moose from the cape breton study area. forest habitat assessment -model 3 consistently predicted the highest proportion of most suitable habitat (based on high probability of moose presence) at all 3 sites (table 5). this model indicated that suitable habitat (based on mainland moose occupying ecologically similar terrain) exists in the cape breton study area, with 42.6% of the area classified as class 5 (high probability). model 1 and model 2 predicted less class 5 habitat in the cape breton area, 31.6% and 39.8%, respectively. when combined with habitat in chebucto peninsula mainland control site chedabucto peninsula mainland control site cape breton study area size of study area 63,690 ha 218,312 ha 318,193 ha class 1 model 1 57.3% 85.5% 46.0% very low probability model 2 52.3% 86.4% 45.0% model 3 47.8% 85.7% 47.00% class 2 model 1 11.3% 5.2% 8.8% low probability model 2 9.6% 2.4% 5.9% model 3 4.7% 1.5% 4.4% class 3 model 1 5.3% 2.5% 6.9% medium probability model 2 6.0 % 2.1% 4.3% model 3 4.6% 1.4% 6.7% class 4 model 1 7.3% 1.9% 6.7% good probability model 2 7.3% 2.1% 5.1% model 3 7.0% 1.7% 2.9% class 5 model 1 18.7% 4.9% 31.6% high probability model 2 25% 6.9% 39.8% model 3 36.2% 9.5% 42.6% table 5. models 1-3 were applied to the cape breton study area and mainland control sites to determine the percentage of land area in each probability class for moose presence at each site. the cape breton study area had the highest percentage of land area classified as high probability, and the lowest as very low probability in all 3 models. absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 94 the class 3 (medium probability) and class 4 (good probability) categories, the total amount of suitable habitat ranged from 45.248.6% across the 3 models. comparatively, model 3 predicted 36.2% and 9.5% class 5 habitat at the chebucto and chedabucto control sites. we were not able to validate the models using statistical cross-validation due to the small sample size. previous habitat suitability modeling based on optimal habitat conditions predicted little suitable habitat on the mainland (snaith et al. 2002). such results could suggest that moose occupy sub-optimal habitat. given the lack of information concerning habitat preference of moose in nova scotia, the use of pgi offers an alternative predictive approach. habitat suitability indices of snaith et al. (2002) were unable to predict moose presence/absence, however road density alone and the combination of habitat suitability indices with road density predicted moose presence/absence; moose presence was negatively correlated with road density. the models created by this research did not contain a road component, since road occurrence was minimal in the pgi plots used to create the models, and thus could not be a variable. our models were based on those created by brannen (2004) who used a similar strategy to identify moose habitat preferences for the entire mainland with a road component. given the reported effect of roads on moose and the equivocal results of this research, further analysis of the relationship between road presence, habitat use, and moose presence/absence is warranted in the study area and elsewhere in nova scotia. non-habitat exclusion factors potential exclusionary factors identified by key informants included competition with white-tailed deer, p. tenuis, and poaching, though none of these were considered sufficient to explain exclusion. because moose and deer coexist in other parts of their range, key informants believed that the presence of deer in the study area should not prevent an established moose population. poaching, and a high social tolerance for the practice, was suggested as a potential exclusion factor. the current social attitude towards moose is believed to be one of utilization based on the acadian traditions of the region. interviewees suggested that moose would be welcome in the area as an additional source for sport hunting, but that a high social tolerance towards illegal hunting prohibits a stable population. conversely, northern cape breton has a legally and heavily hunted moose population that persists at high density. consequently, opportunistic poaching, on its own, was not considered significant enough to exclude moose. these areas differ, however, in that much of the higher elevation areas of northern cape breton are inaccessible, and a substantial portion of the area contains protected areas where hunting is prohibited or restricted as in cape breton highlands national park and the pollets cove–aspy fault wilderness area. conclusion we were not able to identify any specific cause for the continued absence of moose from the southeastern side of cape breton. none of the examined habitat elements appear to exclude moose, and sufficient, suitable habitat appears to be available assuming the forest inventory data are accurate. class 3, 4 and 5 habitat represented 154,641 ha (model 3) in the cape breton study area, an area much larger than the chebucto peninsula that maintains a small, localized group of moose. the geographical differences among the mainland control sites and the study area were not considered substantial enough to explain the lack of moose in the cape breton study area. it is possible that in combination, soil type (i.e., granitic), temperature and snow depth (i.e., that limit or exclude deer), and inter-specific interactions among white-tailed alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 95 deer, p. tenuis, molluscan hosts, and moose serve to limit moose in the cape breton study area and other areas of nova scotia. nonetheless, these factors seem insufficient to explain the historical absence of moose in the study area. future research could include further investigations of the geochemical composition of the area, the social tolerance of poaching behavior, and first nations knowledge of historical and current moose distribution. a vegetative analysis in the cape breton study area could identify whether heavy metals accumulate in preferred browse that could affect forage palatability and moose health. a health assessment of white-tailed deer in the cape breton study area could potentially identify geochemically-related health issues that may also affect moose. the habitat modeling relied heavily upon the provincial forest inventory database. this data is updated periodically, but is based primarily on the use of permanent sample plots located around the province and interpretation of aerial photographs. the information in the database should be field-verified before drawing definitive conclusions from the modeling exercise. a sampling of understory vegetation could improve the habitat model by improving its relevance to habitat suitability. as well, negatively correlated habitat features, such as road density and proximity to human population centers, should be incorporated into the modeling exercise as they often have negative impact on moose, habitat use, and habitat suitability. acknowledgements we acknowledge the support of the following, without whom this work could not have been completed: nsdnr’s wildlife division and graphic & mapping services, martin willison, hugh broders, frances mackinnon (gis technical support), dennis brannen, wade blanchard, terry goodwin, kenneth donovan, and key informants dave mccorquodale, terry power, leeanne reeves, lloyd mcdonald, trevor wilkie, bevan lock, don dodds, jenny costelo, doug archibald, and mark pulsifer. thanks to the editors of alces and anonymous reviewers. references beazley, k. f., t. v. snaith, f. mackinnon, and d. colville. 2004. road density and potential impacts on wildlife species such as american moose in mainland nova scotia. proceedings of the nova scotian institute of science 42: 339-357. _____, m. ball, l. isaac man, s. mcburney, p. wilson, and t. nette. 2006. complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada. alces 42: 89-109. benson, d. a. 1957. the moose in nova scotia. nova scotia department of lands and forests bulletin 17:1-12. _____, and g. d. dodds. 1980. the deer of nova scotia. department of lands and forests, halifax, nova scotia, canada. bingley, j. m., and g. g. richardson. 1978. regional lake sediment geochemical surveys in eastern mainland nova scotia. nova scotia department of mines and energy. open file report 371. boer, a. h. 1990. spatial distribution of moose kills in new brunswick. wildlife society bulletin 18: 431-434. bonner, f. j., p. w. finck, and r. m. graves. 1990. heavy mineral trace element analysis of till (-230 mesh), south mountain batholith. nova scotia department of mines and energy. open file report 90-006. brannen, d. c. 2004. population parameters and multivariate modeling of winter habitat for moose (alces alces) on mainland nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 96 of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8: 1309-1315. crichton, v., and p. paquet. 2000. cadmium in manitoba’s wildlife. alces 36: 205-216. davis, d., and s. browne. 2003a. the natural history of nova scotia, volume 1: topics and habitats. nimbus publishing ltd., halifax, nova scotia, canada. _____, and _____. 2003b. the natural history or nova scotia, volume 2: theme regions. nimbus publishing ltd., halifax, nova scotia, canada. denys, n. 1672. the description and natural history of the coasts of north america (acadia). translated and edited by william ganong. publications of the champlain society, toronto, 1908. dodd, a. 1805. a sketch of memorandum on the local and natural advantages of the island of cape breton. colonial office 219, vol. 124 c.b.a., vol. 9, 1815. dodds, d. g. 1963. the present status of moose (alces alces americana) in nova scotia. proceedings of the northeast wildlife conference 2: 1-40. ducreux, f. 1664. the history of canada or new france. translated and edited by robinson and conacher. publications of the champlain society, toronto, 1951. duinker, p. n., p. e. higgelke, and n. a. bookey. 1993. future habitat for moose on the aulneau peninsula, northwest ontario. pages 551-556 in 7th annual symposium on geographic information systems in forestry, environment and natural resources management, vancouver, british columbia, 15-18 february 1993. _____, _____, and s. koppikar. 1991. gis-based habitat supply modeling in northwestern ontario: moose and marten. pages 271-275 in gis’91 an international symposium on geographic information systems, vancouver, british columbia, 15-18 february 1991. dunn, c. e., s. w. adcock, and w.a. spirito. 1992a. reconnaissance biogeochemical survey, southwestern nova scotia, part 1 red spruce bark, parts of nts 200, p, 21a, b; geological survey of canada, open file 2556. _____, _____, and _____. 1992b. reconnaissance biogeochemical survey, southeastern cape breton island, nova scotia, part 1 black spruce bark, parts of nts 11f, g, j, k; geological survey of canada, open file 2558. edwards, r., and r. ritcey. 1956. the migrations of a moose herd. journal of mammalogy 37: 486-494. environment canada. 2002. envirofacts: toxic chemicals in atlantic canada – lead. (accessed april 2005). _____. 2003. atmospheric science division: acid rain faq. (accessed march 2005). _____. 2004. canadian climate normals 1971-2000. (accessed march 2005). forman, r. t. t., d. s. friedman, d. fitzhenry, j. d. martin, a. s. chen, and l. e. alexander. 1997. ecological effects of roads: toward three summary indices and an overview of north america. pages 40-54 in k. canter, editor. habitat fragmentation and infrastructure. minister of transport and public works and water management, delft, netherlands. fortier, m. 1983. the cultural landscape of 18th century louisbourg. (accessed january 2005). frank, a. 1998. “mysterious” moose disease in sweden. similarities to copper deficiency and/or molybdenosis in cattle and sheep. biochemical background of clinical signs and organ lesions. the science of the total environment 209: 17-26. alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 97 _____, r. danielsson, and b. jones. 2000a. the “mysterious” disease in swedish moose. concentrations of trace elements in liver and kidneys and clinical chemistry. comparison with experimental molybdenosis and copper deficiency in the goat. the science of the total environment 249: 107-122. _____, d. sell, r. danielsson, j. fogarty, and v. monnier. 2000b. a syndrome of molybdenosis, copper deficiency, and type 2 diabetes in the moose population of south-west sweden. the science of the total environment 249: 123-131. gillingham, m., and d. klein. 1992. latewinter activity patterns of moose (alces alces gigas) in western alaska. canadian journal of zoology 70: 293-299. gucinski, h., m. furniss, r. ziermer, and m. brookes. 2001. forest service roads: a synthesis of scientific information. united states department of agriculture, forestry service, pacific northwest research station, portland, oregon, general technical report pnw-gtr-509.1. holland, s. 1935. holland’s description of cape breton island and other documents (1865-67). public archives of nova scotia, publication no. 2. halifax, nova scotia, canada. jalkotzy, m. g., p. i. ross, and m. d. nasserden. 1997. the effects of linear developments on wildlife: a review of selected scientific literature. arc wildlife services ltd., prepared for canadian association of petroleum producers, calgary, alberta, canada. karns, p. 1998. population distribution, density and trends. pages 125-139 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d. c., usa. kelsall, j. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research supplement 1: 1-10. _____, and w. prescott. 1971. moose and deer behaviour in snow. canadian wildlife service report series number 15. canadian wildlife service, ottawa, ontario. kwan, h. 2005. an examination of the absence of established moose (alces alces) populations in southeastern cape breton island, nova scotia. m.e.s. thesis, dalhousie university, halifax, nova scotia, canada. landry, p. 1997. nicolas denys (1598-1688). biographies. (accessed march 2005). lankester, m., and w. samuel. 1998. pests, parasites and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, d.c., usa. laurian, c., c. dussault, j-p ouellet, r. courtois, m. poulin, and l. breton. 2008. behaviour of moose relative to a road network. the journal of wildlife management 72(7):1550-1557. lescarbot, m. 1609. histoire de la nouvelle france. paris. lyon, l. j. 1984. road effects and impacts on wildlife and fisheries. 7th proceedings of the forest transportation symposium, casper, wyoming, 11-13 december, 1984. mackinnon, f. 2001. analysis of moose habitat on mainland nova scotia: the development of two gis tools for moose habitat suitability and moose habitat extraction. applied geomatics research group, centre of geographic sciences, lawrencetown, nova scotia. macmichael, c. 2007. use of provincial black bear (ursus americanus) habitat associations to assess the lack of an established population in southeastern cape breton island, nova scotia, canada. absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 98 m.e.s. thesis. dalhousie university, halifax, nova scotia, canada. mercer, w., and d. kitchen. 1968. a preliminary report on the extension of moose range in the labrador peninsula. north american moose conference workshop 5: 62-81. mills, r. f. 1989. geochemical analyses of bulk stream sediment samples from northern nova scotia (parts of nts sheets 11e; 11f; 11g; 11j; 11k; and 21h). nova scotia department of mines and energy, open file report 89-007. nasserden, m., m. jalkotzy, and p. ross. 1997. the effects of linear developments on wildlife: a review of selected scientific literature. arc wildlife services. prepared for the canadian association of petroleum products, calgary, alberta. neily, p., e. quigley, l. benjamin, b. stewart, and t. duke. 2003. ecological land classification for nova scotia: volume 1 mapping nova scotia’s terrestrial ecosystems. nova scotia department of natural resources, renewable resources branch. noss, r. f. 1995. maintaining ecological integrity in representative reserve networks. world wildlife fund canada, toronto, ontario. nova scotia department of natural resources (nsdnr). 1999. forest inventory. (accessed december 2004). parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. report prepared for the nova scotia department of natural resources. peterson, w., m. lankester, and m. riggs. 1996. seasonal and annual changes in shedding of parelaphostrongylus tenuis larvae by white-tailed deer in northeastern minnesota. alces 32: 61-73. pichon, t. 1760. lettres et memoires pour server a l’histoire naturelle, civile et politique du cap breton, depuis son etablissement jusq’a la reprise de cette isle par les anglois en 1758. la haye, 1760. pimlott, d. 1953. newfoundland moose. transactions of the north american wildlife conference 18: 563-579. pollock, b. 2006. trace elements status of white-tailed deer (odocoileus virginianus) and moose (alces alces) in nova scotia. prepared for the nova scotia department of natural resources and the canadian cooperative wildlife health centre. _____, and e. roger 2007. trace element status of moose and white-tailed deer in nova scotia. alces 43: 61-77. prescott, w. h. 1968. a study of winter concentration areas and food habits of moose in nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. puls, r. 1994. mineral levels in animal health: diagnostic data. 2nd ed. clearbrook, british columbia: sherpa international, canada. pulsifer, m. d., and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. richardson, g. g., and j. m. bingley. 1980. regional lake sediment survey, southwestern nova scotia; nova scotia department of mines and energy ofr. rogers, p. j., and p. a. lombard. 1990. regional geochemical surveys conducted by the nova scotia department of mines and energy from 1957 to 1989. open file report 90-017. _____, and m. a. macdonald. 1983a. lake sediment geochemistry of the mira mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86-028, 20 maps; also geological survey of canada, open file 1269. _____, and _____. 1983b. lake sediment geochemistry of the grand narrows alces vol. 44, 2008 beazley et al. absence of moose on cape breton island 99 mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86-029, 20 maps; also geological survey of canada, open file 1270. _____, and _____. 1983c. lake sediment geochemistry of the st. peters mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86030, 21 maps; also geological survey of canada, open file 1271. _____, and _____. 1983d. lake sediment geochemistry of the framboise mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86032, 20 maps; also geological survey of canada, open file 1291. _____, and _____. 1983e. lake sediment geochemistry of the louisbourg mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86033, 20 maps; also geological survey of canada, open file 1292. _____, and _____. 1983f. lake sediment geochemistry of the glace bay mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86034, 20 maps; also geological survey of canada, open file 1293. _____, and _____. 1983g. lake sediment geochemistry of the sydney mapsheet scale 1:50 000, nova scotia department of mines and energy, open file map 86035, 20 maps; also geological survey of canada, open file 1294. scanlon, p. f., k. i. morris, a. g. clark, n. fimreite, and s. lierhagen. 1986. cadmium in moose tissues: comparison of data from maine, usa and from telemark, norway. alces 22: 303-312. schmitz, o. j., and t. d. nudds. 1994. parasite-mediated competition in deer and moose: how strong is the effect of meningeal worm on moose? ecological applications 4: 91-103. selinus, o., and a. frank. 2000. medical geology. pages 164-181 in l. moller, editor. environmental medicine. sweden. smith, t. 1801. untitled report of a summer traveling in western nova scotia. public archives of nova scotia, no. 303. snaith, t. v. 2001. the status of moose in mainland nova scotia: population viability and habitat suitability. m.e.s. thesis, dalhousie university, halifax, nova scotia, canada. _____, k. f. beazley, f. mackinnon, and p. duinker. 2002. preliminary habitat suitability analysis for moose in mainland nova scotia, canada. alces 38: 73-88. stea, r. r. 1982. pleistocene geology and till geochemistry of south central (southwestern) nova scotia, 1981 (sheet 6). nova scotia department of mines and energy, map 82-1, scale 1:100 000. _____. 1983. till geochemistry (sheet 5). nova scotia department of mines and energy, open file report 555, scale 1:100 000. _____, and p. w. finck. 1986. figure 31. till geochemistry and pebble lithology, chignecto peninsula, nova scotia (sheet 9) scale 1:100 000; quaternary geology and till geochemistry of the western part of cumberland county, nova scotia (sheet 9), by r. r. stea, p. w. finck, and d. m. wightman; geological survey of canada, paper 85-17. _____, and j. h. fowler. 1979. pleistocene geology, eastern shore region, nova scotia (sheets 1-3), scale 1:100 000; in minor and trace element variations in wisconsinon tills, eastern shore region, nova scotia. nova scotia department of mines and energy, paper 79-4. _____, and _____. 1981. pleistocene geology and till geochemistry of central nova scotia, (sheet 4), 1980. nova scotia department of mines and energy, map 81-1, scale 1:100 000. _____, and d. r. grant. 1982. pleistocene geology, southwestern nova scotia (sheet 7 and 8). nova scotia department absence of moose on cape breton island – beazley et al. alces vol. 44, 2008 100 of mines and energy, map 82-10, scale 1:100 000. telfer, e. 1967a. comparison of moose and deer winter range in nova scotia. journal of wildlife management 31: 418-425. _____. 1967b. comparison of a deer yard and a moose yard in nova scotia. canadian journal of zoology 45: 485-490. _____. 1970. winter habitat selection by moose and white-tailed deer. journal of wildlife management 3: 553-558. _____. 1972. forage yield and browse utilization on logged areas in new brunswick. canadian journal of forest research 2: 46-350. _____. 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145-182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, alberta, canada. thomas, j. e., and d. g. dodds. 1988. brainworm, parelaphostrongylus tenuis, in moose, alces alces, and white-tailed deer, odocoileus virginianus, of nova scotia. canadian field-naturalist 102(4): 639-642. turner, r. g., and r. r. stea. 1987a. till geochemistry of northern mainland nova scotia (sheet 10). nova scotia department of mines and energy. open file map 87-005, scale 1:100 000. _____, and _____. 1987b. till geochemistry of northern mainland nova scotia (sheet 11). nova scotia department of mines and energy. open file map 87-006, scale 1:100 000. _____, and _____. 1988a. till geochemistry of northern mainland nova scotia (sheet 12). nova scotia department of mines and energy, open file map 88-031, scale 1:100 000. _____, and _____. 1988b. till geochemistry of northern mainland nova scotia (sheet 13). nova scotia department of mines and energy, open file map 88-050, scale 1:100 000. underwood, e. j. 1971. trace elements in human and animal nutrition, 3rd edition. academic press, inc., new york. van ballenberghe, v. 1992. behavioural adaptations of moose to treeline habitats in subarctic alaska. alces supplement 1:193-206. whitaker, a. n. 2006. a preliminary exploration of the ecological and societal possibility of wolf recovery to nova scotia, canada, m.e.s. thesis. dalhousie university, halifax, nova scotia, canada. 4202(13-23).pdf alces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 13 moose hunting, forestry, and wolves in sweden margareta bergman1 2 1department of historical studies, s-901 87, umeå university, umeå, sweden, e-mail: margareta. bergman@formas.se; 2department of animal ecology, swedish university of agricultural sciences, s abstract: we have reviewed swedish forestry and hunting literature in order to investigate how the management of moose (alces alces) in sweden has changed during the 20th century, especially after the re-establishment of the wolf (canis lupus) in the 1980s. the focus is on the perspective of moose hunters and of the forest industry since these are the two main factors in control of the size of the swedish moose population. at about the same time as the swedish moose population was reaching half of the 19th century, wolves were relatively abundant in sweden. however, intense hunting led to their drastic decrease, so that in the beginning of the last century only a small number remained. as a result of being virtually extinct, the wolf was thus declared protected in 1965. currently, the scandinavian (i.e., the swedish and norwegian) wolf population has grown to a size of about 100 individuals. this might not sound like much in a relatively large country like sweden but in areas where hunters already have had their culling ratio for moose decreased by the forest companies to minimize forest protest to this state of affairs. there are few instances (to our knowledge) where the forest companies mercial point of view but disastrous when it comes to their relationship with the hunters. we suggest that moose management in areas with wolves should be controlled by special regulations, taking both local and national interests into account and where ownership of the hunting ground should not be the sole consideration. alces vol. 42: 13-23 (2006) key words: alces alces, canis lupus, forestry, hunting, management, moose, sweden, wolf the present distribution of gray wolves (canis lupus) in europe is tiny compared to their historical distribution. originally they were widespread and common throughout the whole of the northern hemisphere (mech 1995, wabakken et al. 2001). because the distribution of wolves followed the distribution of large presumably about the same time in history as wild ungulates were domesticated (mech moose hunters and wolves as they in many instances compete for the same prey. in addior killed in wolf territories in sweden, on 43 occasions between 1997 and 2003 (karlsson and jaxgård 2004), some hunters are hesitant to use their dogs for moose hunting in areas with wolf territories. forest companies have a strong voice in the setting of the moose hunting quotas because moose hunting in sweden is tied to the ownership of land and roughly half of the forests in sweden are owned by forest companies. the swedish management of moose in general, and in areas with wolves in particular, results in a delicate problem among not only hunters and forest companies and people who live in close proximity to the wolf, but also the state, especially the swedish environmental protection agency (sepa) moose hunting, forestry, and wolves – bergman and åkerberg alces vol. 42, 2006 14 and environmentalists who want to protect the swedish wolf. current research has shown that the attitudes toward wolves may be affected in areas experiencing increasing populations of wolves and where hunting quotas should be reduced to avoid a decline in prey populations (nilsen the swedish hunters and the forest companies by reviewing three hunting magazines and a forestry magazine we wanted to capture the current attitudes of the hunting and forestry sectors to the re-establishment of wolves in sweden. at least as far as is possible, picture of the ecological, biological, and social predictions and consequences of the re-establishment of wolves, whereas in hunting and forestry magazines a somewhat more biased picture emerges, which often can be more directly ator not the person owns land, and/or owns livestock, and/or is a hunter. to study the large forest company, we focused on one small village of sweden within an area having one of the densest populations of wolves in the country. we also give a historical background about forestry and hunting in sweden because these are the two key players concerning the management of the moose population. methods we reviewed articles dealing with moose, wolves, and forestry in the three largest hunting tion for hunting and wildlife management, published by the national swedish associapublished. also, we reviewed the forestry by the swedish forestry association. we focused on the years between 2000 and 2004 because it was in 2000 that the debate really started concerning wolves in connection with sion concerning moose hunting quotas in areas with wolves. further, we used the area and a forest company concerning the wolf, as the hunters in this area have been very vocal concerning altered hunting quotas as a result of the establishment of wolves in the area in 2001. a brief history of swedish forestry the forest industry has long been of vital exported goods from sweden in 2003, 13% were wood products (karlsson 2004). the in the beginning of the last century forest resources were alarmingly low, this mainly due to rapid human population growth demanding more land for grazing livestock and small hold leases (ekelund and hamilton 2001), and also by the expansion of the timber harvesting areas, which by the late 19th century affected pine ecosystems throughout all of sweden (axelsson and östlund 2001). however, the forestry act of 1903, stipulated that logged forests should be cultivated and replanted. this action helped turn the forest industry into the single most important industrial sector in sweden for decades. selective felling was the most common forestry technique up to the end of the 1940s. however, the second national forestry act in 1948 prohibited selective felling of mature trees and opened up the forests for clear felling, a technique which was widely used during the 1960s and 1970s. as a result, large areas were cleared of trees. alces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 15 this large scale forestry opened up the forest to extensive regeneration of trees and shrubs, and contributed to the increase of the moose population (cederlund and markgren 1987, cederlund and bergström 1996). today, large areas of the swedish boreal landscape are characterised by young, even-aged stands of pine trees (axelsson and östlund 2001). moose hunting and ownership of land (51%) is owned privately and 42% of the forest is owned by forest companies (including those owned by the state as well as by shareholding companies) (karlsson 2004). anyone who owns land, no matter its size, has the right to the national regulations concerning hunting. the forest owners may also lease out the right to others to hunt on their property and this is a very common practice concerning moose hunting in sweden. has a long tradition in sweden, dating back to 1789, when king gustav iii allowed all land owners in the country to hunt on their own property (haglund 1980). however, the resulting intense hunting meant that an already declining moose population came close to extinction and in the beginning of the 19th century there were very few moose by this situation, the swedish association for hunting and wildlife management (sahw) was founded in 1830. one of its aims was to make hunting more ethical and to see that data were gathered in order to favour and publicize knowledge to the ‘pleasure and 1938, the swedish government adopted a new law, which placed the sahw in control of the hunting and management of the swedish moose population. with the aim to increase the size of the moose population, a series of hunting restrictions suggested by the sahw were adopted by the swedish government. one of the ways to increase the moose population was by making it illegal to cull moose calves, but as the moose population started growing again this restriction was abandoned. during the 1960s hunting for calves was encouraged. this practice later proved to increase the moose population further, rather than reducing it (åkerberg 2005). at the same time as the harvesting of calves was encouraged, the harvest of adult female moose was restricted, which meant that the hunt during the late 1960s was focused on the least productive segments of the moose population (ericsson 1999). hunting restrictions in combination with modern forestry techniques were two important factors in the increase of the moose population. the harvest of moose increased from 11,318 to 32,680 between the years 1945-1960, and between1960-1980 the harvest increased a further four times (the swedish association for hunting and wildlife management 2005a). as the moose population increased, damage to trees began to be seen as a problem by the forest owners, particularly to the forest companies who had seen many decades with virtually no browsing damage to their forests. as a result, discussions and efforts were intenpopulation more actively. in 1967, completely regulated moose hunting (i.e., under licence only) was introduced in a few counties (von essen 2005) and in 1977 all of swedish hunting was completely regulated (åkerberg 2005). in 1982, the moose population peaked and 174,741 moose were culled. currently, about 300,000 hunters participate in the moose hunt each year and about 100,000 moose are harvested annually (the swedish association for hunting and wildlife management 2005a). despite the hunting regulations, the moose population increased to a level almost beyond control and the high harvesting quota was mainly a result of pressures from forest companies who feared that the level of tree moose hunting, forestry, and wolves – bergman and åkerberg alces vol. 42, 2006 16 damage would be too high. the attitude from many of the hunters at the time was that the excitement of moose hunting was almost gone 1980). the organization of swedish moose hunting the county administrative boards assign moose hunting licence quotas according to the estimated moose density for each management land coverage (ericsson 1999). the number of moose to be shot each year is generally based the hunt, aerial surveys, moose pellet counts, and by inventories of the browsing pressure (the swedish association for hunting and wildlife management 2005a). the method ing pressure by moose on pines and birches 1-4 m high. an index of browsing pressure is based on fresh browsing on top shoots, stem breakage, or bark stripping. the inventory is mainly performed on areas of 20-100,000 ha (the national board of forestry 2004a). the current moose hunting system in sweden can be described as a patchwork, are where the size and characteristics of the area have to be such that a minimum of one adult moose can be culled per year and the hunting may be permitted in these areas for one adult moose or one calf per year, the size of the area has to be a minimum of 5 ha, and size of the area has to be a minimum of 20 provide an area that is large enough and has the characteristics to allow management of its own moose population, here moose hunting without licence is allowed, the hunting season is 70 days, the size of the area is usually at least 5,000 ha, the moose population within this area has to be able to sustain a culling of 25 where, during a maximum 5 day season, an unrestricted number of calves can be culled (the county board administration västra götalands municipality 2005). wolves, moose hunting, and forestry on average, 500 wolves per year were harvested in sweden during 1827-1839. harvest of wolves and presumably numbers in the population decreased such that by 30 years later, fewer than 100 wolves per year were harvested (aronson and sand 2004). wolf hunting continued and bounties were paid out for killed wolves as late as during the mid-1960s (wabakken et al. 2001). when wolves became protected in 1966 there were10 or fewer wolves remaining in sweden (aronsince 1964 was born in northern sweden in 1978, and in 1983 at least 6 wolves were born however, it was not until the early 1990s that the scandinavian wolf population started to exceed 10 individuals. between 1991 and 1998 the average growth rate of the wolf population was 29% (wabakken et al. 2001) and by 2005 there were about 110 wolves in norway and sweden of which about 85 were http://www.naturvardsverket.se). the problem with the swedish wolf population is not the fact that they are numerous per se, the problem is rather that the wolves are concentrated in relatively small areas of the locals. a rough calculation performed by karlsson et al. (2004) estimates that sweden has a prey population that could sustain a wolf population of about 5,000 individuals. however, at those densities of wolves, there would alces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 17 be limited room for hunting of ungulates. since the swedish wolf was nearly extinct for decades, there are few swedish studies toward wolves. one exception is a study by ericsson and heberlein (2003) which showed that while swedish hunters were the strongest supporters of wolves in the 1970s, their attitudes changed after the restoration of wolves in the country, and that by 2001 the hunters were actually less supportive of wolves than was the general public. because the forest companies in sweden ing licences to the swedish moose hunters, of moose to be harvested by the hunters in a by the county administrative boards, but it is stated in the swedish hunting legislation that at least one member of this board (consisting of 11 members plus one representative from the sami community in some northern counties) should be appointed by the national board of forestry and a further 3 should represent owners of agrarian and forest land (sfs 1987:905). in 2000, forest owners were expressly encourence moose management (johansson 2000). moreover, the forest companies have other ways to put pressure on the leasing hunters. for example, there have been demands for the hunters to pay the company harvesting fees in harvesting fees depending on the gender of the shot moose (cows are usually cheaper because the companies prefer that the hunters harvest cows instead of bulls) and, if the leasing hunters cannot harvest the full quota, other hunters (swedish and/or foreign) might be brought in on the basis of external hunting tion for hunting and wildlife management, personal communication). the demand of paying the harvesting fees the hunting teams. annual leasing fees per ha are generally quite cheap in the northern parts of sweden [2-15 sek] while they increase radically as one moves further south. higher fees in the south are mainly due to more hunters or hunting teams competing for hunting opportunities. in extreme cases, the cost may be as high as 300 sek per ha but the leased areas in the south are, on the other hand, usually not as large as the areas in the north. most landowners within a certain area usually charge similar leasing fees because it is regarded as bad form to overcharge. however, in general, forest companies frequently place themselves a few percent above the regular fee hunting and wildlife management, personal 10 in sweden 4 possible 3 in norway new litters, 2004: number of wolves ~ 110 figure 1. map of scandinavia showing the location of new wolf litters in 2004. translated from swedish. source: sepa home page, http://www.naturvardsverket.se. moose hunting, forestry, and wolves – bergman and åkerberg alces vol. 42, 2006 18 communication). in addition to the leasing fee, hunting teams must pay a governmental harvesting fee of 300 sek per killed adult moose (calves are free) and the landowner seeks an additional harvest fee. one forest company charged about 3,500 sek per killed adult moose and 800 sek per killed calf (b. eriksson, stora enso, personal comunication). considering that hunting teams rarely consist of more than 15 people (usually 5 15 people), but that they can easily shoot more than 15 moose in some areas of the country, these extra harvesting fees, especially if they have to be paid in advance, can represent a considerable part of the annual expense for the hunting teams. however, it must be noted that there may be an extremely large difference between areas. it all depends on how big the competition is between the hunters in the areas in question (i.e., how much the land owner can charge for the leasing fee and how large the hunting areas are), how abundant the game is, and who the landowner is. different forest companies have different policies and the only thing consistent is that even though the various fees can represent quite a lot of money (not least for the hunting teams) overall, annual revenue – rarely more than 0.5 1%. the re-establishment of wolves in sweden presents another issue in the relationship between the hunters and the forest companies, at arise as to whether the forest companies or over the moose hunting quota. according to swedish legislation, both groups should have interest because the forestry industry is to grow commercially important trees in the amount of damage to trees by moose is desirable whereas the moose hunters wish to in order to have a meaningful hunt. a forest is the land owner who owns the moose hunting and the hunting ground and also pays for ‘there should be a better way to manage the moose population than to shoot it to pieces, one should not forget that the land owners get a large income from a well managed moose the moose population should be in balance with the available food supply in a swedish forest companies is that the level of fresh moose damage on young forests should not exceed 2% per year, corresponding to 1,000-1,800 undamaged trees (the main trunk) per ha. by 2005, the goal is that the level of fresh moose damage should be less than 2% within a minimum of 80% of the surveyed areas (steffansson 2002). however, the level predict as there are other factors than the moose density per se, such as stand density (lyly and saksa 1992, ball and dahlgren 2002) and tree species composition (danell et al. 1991) that browsing. the sahw do not agree with this tain a balance between the moose harvesting levels and the goal of maximum 2% moose damage in areas with wolves, and suggest in order to maintain moose hunting quotas that the forest owners will have to accept a higher level of moose damage in areas where there are wolves (lundvik 2002). in the three swedish hunting magazines there were hardly any articles between 20002004 in which the forest companies generally expressed a negative attitude against the presence of wolves, and the hunting manager of the forest company in åmot says in the magazine companies mission to grow food for wolves at the expense of the forest company, and at the same time, we do not say no to wolves in our alces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 19 all the three hunting magazines that express a clear resentment to how the forest companies handle the wolf issue. for example, one hunter in åmot says: ‘is there any reason at all for the moose hunters to cooperate with the land owners (i.e., a forest company) if the thank interviewed for the magazine ‘the hunting only interested in trading with wood and do the conflict in åmot panies, and the wolf is evident when looking at in south-eastern sweden. almost all land (161, 000 ha) within the åmot moose preservation district is owned by a single forest company which distributes the quota to the hunters who, in turn, lease the right to hunt there. the hunters in the area claim that they have been company would bring in their own hunters if the local hunters refused to harvest the large number of moose that the forest company is required to harvest in order to reduce the moose population and browsing damage (olsson 2003a). in 2001, the high hunting quotas came to an abrupt end when a pair of wolves established in the area had 8 pups the same year. the situation changed drastically for the hunters. instead of harvesting 22 moose as in 2,000, the following year only 3 moose were shot, and from 2002 to 2004 one adult moose (and one calf in 2004) was shot per year out to be harvested each year between 2001-2004 (the county board administration gävleborg municipality 2005). the hunters decided to refrain from shooting calves as they feared that would decrease the moose population to an unacceptably low level. this can be viewed as an example where the local hunters feel powerless with respect to decisions about the hunt in their area and use the limited means they have to express their discontent, in this case, refusing to harvest calves (3 of the 4 hunting districts in åmot are b-areas, in which the hunters can choose whether to shoot one adult moose or one calf). before the establishment of wolves in the area, moose density was 5.9 moose per 1,000 ha and a year after, in 2001, when the wolves had established in the area, the moose density was down to 0.8 moose per 1,000 ha (olsson 2003a). the hunters in åmot have suggested to the forest company that in order to have a meaningful hunt, the moose densities should be 8.45 moose per 1,000 ha, a suggestion which, according to the hunters, the forest company did not agree with. by 2003, many hunters in the area had ceased hunting as they did not think it worthwhile to hunt for only one moose (olsson 2003a). the representative of the forest company responsible for hunting issues agreed that the situation in åmot is problemhe claims that the moose population cannot be increased as the moose then would cause too much damage on young pine plantations. further, the same representative thinks that the hunters will breach the hunting agreement if they decide not to harvest moose calves. as he expresses it ‘a hunting lease is a business agreement and anyone who leases the right to hunt must try to harvest the amount of moose controlled hunting of wolves as a result of the increase in the wolf hunters, and locals in the areas having wolf territories and voices were raised demanding a hunt for wolves. however, controlled hunting of wolves in sweden is very restrictive. between 1992 and 2005, sepa has permitted controlled hunting for wolves in moose hunting, forestry, and wolves – bergman and åkerberg alces vol. 42, 2006 20 four instances, for a maximum of 5 wolves 2003, and 1 in 2005) (swedish environmental protection agency 2005). between the years of 2002 and 2005, 18 petitions were handled by the sepa, two of which came from the swedish association for hunting and wildlife management in cooperation with the federation of swedish farmers. the two petitions applied for controlled hunting of wolves in 8 wolf territories. the rationale for the petitions domesticated animals (including hunting petitions with the main argument that the swedish wolf population had not reached the population stated by the swedish government corresponding to roughly 200 animals. before controlled hunting should be allowed. also, according to the sepa interpretation of the decrease the actual number of wolves in areas having the highest concentrations of wolves rather than aiming the controlled hunting tothe wolf population. moose hunters in areas accidents, as a protest to this state of affairs. they claim that since they have to use their do not want to risk their dogs being attacked by wolves (nilsson 2004b). discussion during the 1900s, swedish moose management has been characterized by numerous rules and regulations. in hindsight it is clear that the attempts to control the moose population have failed very often. it does not seem so simple that the more facts and knowledge we have about the moose population, the more control we have over it. when reading through the four magazines it becomes clear that there their implementation in moose management. now that the wolves have entered the scene, things are further complicated by emotional arguments and any attempts to manage the moose population in wolf territories will have to take not only ecological, biological, and political factors into account but also listen to the local hunters because, in many instances, they feel like they have been neglected. if there is not better communication between local moose hunters and representatives of the forest companies concerning the wolf issue, we may be at risk of having more situations like the one in åmot. the issue concerning ownership of land is also a complicated one. should moose hunting only be tied to ownership of land or are there other options? the responsibility for management of moose hunting lies at the local and regional levels, and the goal for the management is, in general, to maintain a number and age, their reproduction, and also to strive to minimize their damage to forests (the swedish association for hunting and wildlife management 2005a). moose management under the current 2% damage goal if the wolf population continues to increase many hunters may have to accept a lower moose hunting quota unless land owners are willing to accept more browsing damage in areas with wolves. the äbin inventories performed between the years of 2000-2004 limit set by the forest companies (the national board of forestry 2004b). there is currently a suggestion for a new moose management system for sweden that has been worked out by the swedish association for hunting and wildlife management, alces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 21 the national swedish association of huntsmen, the federation of swedish farmers, and the forest industries. the suggestion is that the county administrative board divides the country into moose management areas in cooperation with hunting and landowner organizations. these areas should be, at a minimum, 50,000 ha and have a set hunting season (the swedish association for hunting and wildlife management 2005b). the idea behind the suggestion is that the administration of the hunt should be less complicated and better coordinated and that the moose hunting quota should be set to a higher degree than today for a better balance between by moose low. for example, a hunting area which has high densities of predators should have a moose hunting quota which is based on the moose population in the whole moose management area rather than the situation today where neighbouring areas may have a large reduction in the moose hunting quota, moose hunters. acknowledgements we thank j. p. ball and e. m. addison for constructive comments that greatly improved this manuscript. this work was funded by the kempe foundation via a grant to the swedish council for sustainable development. references åkerberg, s., editor. 2005. viltvård, älgar och nyheternas tryckeri, umeå, sweden. (in(in swedish). andersson, s roligare förr? svensk jakt 118:912-915. (in swedish). aronson, å., and h. sand. 2004. omom vargens utveckling i skandinavien de senaste 30 åren. skogsvilt iii: 47-53. (in swedish). axelsson, a. -l., and l. östlund. 2001. retrospective gap analysis in a swedish boreal forest landscape using historical data. forest ecology and management 147:109-122. ball, j. p., and j. dahlgren. 2002. browsing damage on pine (pinus sylvestris and p. contorta) by a migrating moose (alces alces) population in winter: relation to habitat composition and road barriers. scandinavian journal of forest research 17:427-435. bjärvall, a. 1988. lär känna vargen. thethe swedish association for hunting and wildlife management, stockholm, swe-management, stockholm, sweden. (in swedish). björklöf, s. 1994. älgen i vår historia och vardag. milano stampa, milano. (inmilano stampa, milano. (in (in swedish). cederlund, g., and r. bergström. 1996. trends in the moose-forest system in fennoscandia, with special reference to sweden. pages 265-281 in r. m. degraaf and r. i. miller, editors. conservation of faunal diversity in forested landscapes. chapman & hall, london, u.k. _____, and g. markgren. 1987. the development of the swedish moose population, 1970-1983. swedish wildlife research supplement 1:55-61. danell, k., l. edenius, and p. lundberg. 1991. herbivory and tree stand composition: moose patch use in winter. ecology 72:1350-1357. ek, b. 2003. gröna guldet blir bara massa.gröna guldet blir bara massa. skogen 9:27. (in swedish). ekelund, h., and g. hamilton. 2001. skogspolitisk historia. skogsstyrelsen, jönköping, rep. 8a. (in swedish).8a. (in swedish). ericsson, g. 1999. demographic and life history consequences of harvest in a swedish moose population. ph. d. thesis, university of agricultural sciences,sciences, umeå, sweden. _____, and t. a. heberlein. 2003. attitudesattitudes moose hunting, forestry, and wolves – bergman and åkerberg alces vol. 42, 2006 22 of hunters, locals, and the general public in sweden now that the wolves are back. biological conservation 111:149-159. haglund, b., editor. 1980. jägaren och1980. jägaren och vildnaden. svenska jägareförbundet, stockholm, sweden. (in swedish).sweden. (in swedish). johansson, s när skogsbruket tar tag i samråden. skogenskogen 11:32-35. (in swedish). karlsson, s. 2004. estate and ownership structure. pages 31-43 in j. o. loman, editor. swedish statistical yearbook of forestry 2004. volume 54. ab danagårdsvolume 54. ab danagårds _____, h. andrén, and h. sand. 2004. vadvad bestämmer antalet vargar och deras utbredning i framtiden? pages 54-57 in g. jansson, c. seiler, and h. andrén, editors. skogsvilt iii: vilt och landskap i förändring. grimsö research station,grimsö research station, swedish university of agricultural sciences, sweden. (in swedish). _____, and p. jaxgård. 2004. vargangrepp på hundar. pages 243-247 in g. jansson, c. seiler, and h. andrén, editors. skogsvilt iii: vilt och landskap i förändring. grimsögrimsö research station, swedish university of agricultural science, sweden. (in swedish). larsson, j (in swedish). lundvik, b. 2002. vargdebatt fyllde sporthall, svensk jakt 7:22. (in swedish).(in swedish). lyly, o., and t. saksa. 1992. the effect of stand density on moose damage in young pinus sylvestris stands. scandinavian journal of forest research 7:393-403. mech, l. d. 1995. the challenge and opportunity of recovering wolf populations. conservation biology 9:270-278. nilsen, e. b., t. pettersen, h. gundersen, j. m. milner, e. j. solberg, h. p. andreassen, and n. c. stenseth. 2005. moose harvesting strategies in the presence of wolves. journal of applied ecology 42:389-399. nilsson, h. 2004a. fred i älgskogen? jakt-fred i älgskogen? jakt-jaktolsson, o sociala skäl. svensk jakt 2/3: 60-62. (in swedish). _____. 2003b. för mycket älg i åmot. svenskför mycket älg i åmot. svensk jakt 2/3:63. (in swedish). steffansson, j. 2002. ungskogar med bra kvalité och med rätt trädslag! balans 1:4-5. (in swedish). (sfs) swedish code of statutes. 1987. jaktförordningen. swedish board of agriculture, stockholm, sweden.stockholm, sweden. swedish environmental protection agency. naturvardsverket.se/. (accessed 20 dec 2005). the county board administration, gävleborg municipality. mation.http://www.x.lst.se/. (accessed 20 dec 2005). the county board administration, västra götalands municipality http://www.o.lst.se/o/amnen/jakt/algthe national board of forestry, sweden. 2004a. älgbetningsinventering – äbin http://www.svo.se/minskog/templates/ page.asp?id=12188. (accessed 20 dec 2005). _____. 2004b. enkel älgbetningsinventese/minskog/templates/epfilelisting. asp?id=10193. (accessed 20 dec 2005). the swedish association for hunting and wildlife management. 2005a. http:// artpresentation/algforvaltningo.asp. (accessed 20 dec 2005). _____. 2005b. förslag till nytt älgförvaltdet.se/forslagtillnyttalgforvaltningssystem.asp. (accessed 20 dec 2005). törnström, d. 2001. även skogsbolagen hoalces vol. 42, 2006 bergman and åkerberg moose hunting, forestry, and wolves 23 _____. 2002. skogsbolagen största hotet mot von essen, h in s. åkernyheternas tryckeri, umeå, sweden. (in (in swedish). wabakken, p., o. liberg, and a. bjärvall. 2001. the recovery, distribution, and population dynamics of wolves on the scandinavian peninsula, 1978-1998. canadian journal of zoology 79:710-725. 4013.p65 alces vol. 40, 2004 moose genetics hundertmark and bowyer 103 genetics, evolution, and phylogeography of moose kris j. hundertmark1 and r. terry bowyer2 1institute of arctic biology and department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775, usa; 2department of biological sciences, idaho state university, pocatello, id 83209, usa abstract: early studies of genetic variation in moose (alces alces) indicated little variation. recent studies have indicated higher levels of variation in nuclear markers; nonetheless, genetic heterogeneity of moose is relatively low compared with other mammals. similarly, variation in mitochondrial dna of moose is limited worldwide, indicating low historic effective population size and a common ancestry for moose within the last 60,000 years. that ancestor most likely lived in central asia. moose likely exhibit low levels of heterogeneity because of population bottlenecks in the late pleistocene caused by latitudinal shifts in habitat from recurrent climate reversals. a northward movement of boreal forest associated with the end of the last ice age facilitated the northward advance of asian populations and colonization of the new world, which occurred as a single entry by relatively few moose immediately prior to the last flooding of the bering land bridge. despite suffering serial population bottlenecks historically, moose have exhibited a notable ability to adapt to a changing environment, indicating that limited neutral genetic variation may not indicate limited adaptive genetic variation. we conclude that morphological variation among moose worldwide occurred within a few thousand years and indicates that moose underwent episodes of rapid and occasionally convergent evolution. genetic change in moose populations over very short time scales (tens or hundreds of years) is possible under harvest management regimes and those changes may not be beneficial to moose in the long term. modeling exercises have demonstrated that harvest strategies can have negative consequences on neutral genetic variation as well as alleles underpinning fitness traits. biologists should consider such outcomes when evaluating management options. alces vol. 40: 103-122 (2004) key words: adaptation, alces, convergent evolution, moose, mtdna, phylogeography, pleistocene, range expansion genetics have long had a central role in biological investigations, and provide analytical tools that are applicable across a broad spectrum of investigation. for instance, genetic analysis can provide insights into such diverse investigations as evolutionary histories of species (avise et al. 1987, 2000), interactions and relationships among populations (blundell et al. 2002) or individuals (quellar et al. 1993), evaluation of the success of specific management actions (vernesi et al. 2002), population and behavioral ecology (scribner and chesser 2001), and food habits (symondson 2002). recent advances in collection and analysis of genetic data have facilitated more refined approaches to evolutionary and population genetic questions, and our understanding of moose biology has benefited as a result of those advances. evolution and taxonomy of moose (alces alces) have been reviewed previously (peterson 1955; groves and grubb 1987; geist, 1987a,b, 1998; sher 1987; lister 1993; guthrie 1995; bubenik 1998; bowyer et al. 2003) and have encompassed aspects of behavior, morphology, paleontology, and genetics, but no review has dealt specifihundertmark and bowyer moose genetics alces vol. 40, 2004 104 cally with genetics. in this review, our goal is to provide an overview of older studies while focusing on recent advances in genetics and phylogeography (see definition in appendix 1) of moose and the insights they provide. the broad scope of genetic and evolutionary investigations in species biology would make a complete review of all studies disjointed. yet, approximately half of all published studies of moose genetics have been published since the comprehensive treatise “ecology and management of the north american moose” (franzmann and schwartz 1998) was compiled; thus, we were compelled to present the most complete review possible. in an effort to present a cogent summary of all relevant studies, we have divided this review into 3 parts: (1) assessing genetic diversity, where we review the different types of markers examined in studies of moose genetics and the conclusions drawn from those studies; (2) moose evolution and phylogeography, where we examine the evolutionary descent of moose and processes that have shaped the genetic variation and structure observed today; and (3) genetic effects of harvest, which reviews a small but important body of work composed of management-based modeling that examined effects of various harvest regimes on population and genetic measures. assessing genetic diversity one potential difficulty in discussing genetic analyses is the use of specialized terminology. to avoid uncertainty and enhance understanding, we provide a brief glossary of terms used in this review (appendix 1). terms defined in the glossary are highlighted in bold in the text at their first usage. the allozyme era first reports of genetic investigations of moose were published by braend (1962, cited by gyllensten et al. 1980), nadler et al. (1967), and shubin (1969, cited by gyllensten et al. 1980), wherein those authors examined electrophoretic variation in proteins from blood serum; no variation in those genetic markers occurred in scandinavia, north america, and central russia, respectively. the first study to report genetic variability was ryman et al. (1977), who examined 1,384 moose from 3 areas of sweden for polymorphism at 23 allozyme loci. that study reported only 1 locus to be polymorphic, however, and only in 1 region. although they suspected that allele frequencies varied geographically within the 1 variable region, the difference was not statistically significant. those authors concluded that genetic drift associated with a severe population bottleneck (reduction in population size) in sweden in the 19th century was a probable cause of the observed lack of diversity. wilhelmson et al. (1978), examining variation in serum proteins, noted no differences between canadian and european moose. from that evidence, they concluded that moose populations on separate continents had not undergone significant genetic drift despite being separated for thousands of years, implying that effective population sizes of moose populations historically had been large. wilhelmson et al. (1978) also proposed that historic population bottlenecks in sweden had not been severe enough to have had an effect on genetic diversity of moose. gyllensten et al. (1980) conducted extensive screening of a transferrin locus from moose across fennoscandia and detected a single polymorphism occurring in norway, sweden, and finland. nonetheless, the polymorphism was present in only 6 of 16 populations and the uncommon allele never exceeded 6% in any population. the authors presented differences in frequency of the uncommon allele as evidence of differalces vol. 40, 2004 moose genetics hundertmark and bowyer 105 entiation of populations geographically, supporting observations of ryman et al. (1977). reliance on the occurrence of a single rare polymorphism to demonstrate population subdivision, however, is tenuous at best. those early studies and others created an impression among some biologists that certain species, including moose, possessed little genetic variation across the genome. hypotheses explaining this in evolutionary terms were proposed. selander and kaufman (1973) proposed the environmental-grain hypothesis, which stated that large, highly mobile animals exhibited less genetic variability than small, sedentary species. that hypothesis further stated that highly mobile mammals would exhibit greater homogeneity across large areas than more sedentary forms. other hypotheses proposed that r-strategists were less variable than k-strategists (harrington 1985), genetic heterogeneity was greater in species inhabiting broad arrays of habitats compared with habitat specialists (nevo 1978), or that northern cervids inhabiting boreal forests were less variable than their relatives to the south (smith et al. 1990). analyses of 23 allozyme loci in > 700 i n d i v i d u a l s r e p r e s e n t i n g 1 8 m o o s e populations in scandinavia were required to reveal extensive genetic variation in moose (ryman et al. 1980). those authors refuted the environmental-grain hypothesis, concluding that large mammals in general, and large cervids in particular, are not naturally monomorphic; previous studies of moose had examined too few loci or individuals to detect variation. nonetheless, those authors noted that genetic variability in moose was somewhat less than that observed in many other species of mammals (nevo 1978), but that genetic drift due to small historic population size was a more likely explanation than any specific evolutionary strategy. thus, genetic drift was once again proposed as being an important factor in determining the structure of genetic diversity. in another comprehensive study, chesser et al. (1982) examined 1,169 individuals from 4 regions in sweden for a single polymorphic locus and reported variation in allele frequencies among those regions and, perhaps more importantly, significant variation within 1 of those regions. detecting variation at geographic scales small enough to be considered within a single population illustrated that structure of moose populations existed at scales smaller than previously imagined, and that vagility was not inconsistent with genetic structuring. the most recent study of allozyme variation in moose reported extensive variation i n a m o o s e p o p u l a t i o n i n a l a s k a (hundertmark et al. 1992). the level of genetic diversity observed was greater than that reported elsewhere in moose and was similar to levels observed in white-tailed deer, a species known for extensive allozyme diversity (smith et al. 1984). hundertmark et al. (1992) hypothesized that lesser levels of variability described in moose from scandinavia and other regions of north america were attributable to glacial history. all other moose populations studied to that point occurred in previously glaciated terrain that was colonized by moose after retreat of pleistocene ice sheets. colonization of previously glaciated areas could have resulted in serial founder events that reduced genetic diversity (sage and wolff 1986). hundertmark et al. (1992) argued that alaska could have served as a refugium for moose in which genetic diversity could have been maintained because of a large effective population size. assessing variation at the sequence l e v e l the first investigation of highly polymorphic molecular markers in moose documented 4 alleles in a microsatellite-like hundertmark and bowyer moose genetics alces vol. 40, 2004 106 locus in 17 individuals from sweden (ellegren et al. 1991). eight genotypes were reported, which represented a heretofore unthinkable level of polymorphism. those authors investigated the inheritance of that locus in a 2-generation pedigree and determined that the alleles exhibited mendelian inheritance. that study and others like it laid the foundation for the explosion of i n t e r e s t i n p o p u l a t i o n g e n e t i c s a n d phylogenetics based on molecular markers and the polymerase chain reaction (pcr; mullis et al. 1986). the advent of pcr represented one of the truly significant advances in the history of molecular biology and genetics. the process allows in-vitro amplification of dna from miniscule amounts of starting material (theoretically as little as 1 molecule of dna) to provide sufficient quantities for analysis. no longer were researchers required to sacrifice animals to acquire sufficient quantities and types of tissues for genetic analyses because any nucleated cell held the complete genetic complement of the individual. pcr offered unsurpassed access to the genome, and researchers soon applied that to study genetics of moose. mikko and andersson (1995) conducted the first analysis of functional loci in an analysis of variation in the major histocompatibility complex (mhc) in moose from sweden and canada. the mhc is a family of genes important in immune system function, and low levels of diversity of mhc alleles have been interpreted as indicators of lost evolutionary potential and increased susceptibility to pathogens (hedrick 1994). mikko and andersson (1995) noted very low levels of mhc variation in both swedish and canadian moose. moreover, those authors documented similarity among alleles between continents and inferred the existence of a bottleneck in an ancient moose lineage prior to divergence of european and canadian lineages. mikko and andersson (1995) applied a molecular clock to dna sequence variation in the control region of mitochondrial dna (mtdna) to date the time of divergence of swedish and canadian moose, which they estimated at 165,000-350,000 years ago. the time since divergence of european moose also was analyzed by ellegren et al. ( 1 9 9 6 ) . t h e y a s s e s s e d v a r i a t i o n i n minisatellite loci of swedish moose and concluded that normal evolutionary processes could have generated the amount of variation observed within 10,000-50,000 years after a severe bottleneck. their implication, therefore, was that the estimate of divergence provided by mikko and andersson (1995) was too old by perhaps 1 order of magnitude. surprisingly, the two data sets are not inconsistent; indeed, they show similar levels of variation considering differences in evolutionary rates among marker types. the differences in estimates for date of divergence relate more to the evolutionary rate estimates used than to differences in genetic variability. microsatellite loci were first described for moose by wilson et al. (1997) and røed and midthjell (1998). broders et al. (1999) demonstrated the utility of microsatellites for assessing population structure in moose by assessing consequences of founder events in canada. heterozygosity in 3 populations founded by few individuals decreased from 14-30% compared with the source population. broders et al. (1999), in assessing variability of moose on the island of newfoundland, canada, demonstrated that 2 consecutive founder events reduced heterozygosity by 46%. although they could not discern any decrease in fitness as a result of the decrease in diversity, the authors questioned the long-term viability of those moose populations. nonetheless, levels of diversity in neutral microsatellite loci were not indicative of diversity in functional loci in moose (wilson et al. 2003). alces vol. 40, 2004 moose genetics hundertmark and bowyer 107 the future microsatellites have replaced allozymes as the most widely used molecular marker for assessing nuclear genetic diversity, and will be the choice of geneticists for the foreseeable future (bruford et al. 1996). there are problems, however, with analysis of microsatellites because the ways in which they mutate into new forms are not entirely understood (hancock 1999). in the future, a new type of analysis called single nucleotide polymorphism (snp, pronounced “snip”) may replace microsatellites for some applications (fries and durstewitz 2001, brumfield et al. 2003). this new technology allows a single nucleotide site to be queried for presence of a particular nucleotide and presence or absence can be converted to a binary code. current technology (so-called “real-time pcr” and “dna chips”) allows for fast and accurate examination of many individuals and snps, but we must await the development and characterization of marker loci before broad application of this new family of molecular markers can be considered. moose evolution and phylogeography origins of modern moose moose (alces alces) are a young species in the evolutionary scheme of large mammals. the genus alces first appears in the fossil record 2 million years ago (thouveny and bonifay 1984) and fossils attributable to a. alces are first recorded approximately 100,000 years ago (lister 1993). those dates are very recent considering that the subfamily odocoileinae, to which moose belong, diverged from other deer lineages 9-12 million years ago (miyamoto et al. 1990). paleontological evidence indicated europe as the place of origin of the genus alces (lister 1993). the genus never was diverse, with only one species present in the fossil record at any particular time. yet, the species assumed to be the precursor to a. a l c e s , t h e b r o a d f r o n t e d m o o s e ( a . latifrons) was distributed across eurasia and into northwestern north america for a time before becoming extinct in beringia at the end of the pleistocene (guthrie 1995). thus, the widespread distribution of modern moose and its immediate ancestor indicate a degree of evolutionary success despite a paucity of species diversity. up to 8 subspecies of moose are recognized worldwide (fig. 1); 4 in eurasia and 4 in north america (peterson 1955). that number is open to question, however. geist (1987a, 1998) contends that there are 2 predominant types of moose in the world: american and european, following the convention of flerov (1952). to the former type he assigns all north american moose as well as eastern asian subspecies a. a. burturlini and a. a. pfizenmayeri. he based his opinion primarily on morphology and noted similar geographic divisions in taxonomy among reindeer and caribou (rangifer tarandus) and red deer and north american elk (cervus elaphus; geist 1998). he further contended that those morphological types should correspond to subspecies designations. therefore, geist (1998) recognized a. a. alces of europe and a. a. americana in eastern asia and north america. he also suggested that a. a. a m e r i c a n u s h a s p r e c e d e n c e u n d e r nomenclatural conventions as the proper name for the east asian-north american s u b s p e c i e s . h e r e f e r r e d t o a . a . cameloides in northern china, mongolia, and southeastern russia as part of a primitive fauna native to that region and recognized that subspecies as a valid taxon although he also refers to it as an americantype moose (geist 1998:230). the 2-types hypothesis is supported to some degree by karyotype (boeskorov 1996, 1997) and some data on mtdna (mikko hundertmark and bowyer moose genetics alces vol. 40, 2004 108 fig. 1. approximate ranges of 8 subspecies of moose worldwide. a. a. a. = a. a. alces, a. a. p. = a. a. pfizenmayeri, a. a. c. = a. a. cameloides, a. a. b. = a. a. burturlini, a. a. g. = a. a. gigas, a. a. an. = a. a. andersoni, a. a. s. = a. a. shirasi, a. a. am. = a. a. americana, * = introduced population in newfoundland. and andersson 1995). most eurasian moose have a karyotype of 2n = 68, whereas north american moose have 2n = 70. that difference derives from a robertsonian translocation of 2 acrocentric chromosomes into a single metacentric chromosome or vice versa. although that chromosomal polymorphism originally was thought to separate eurasian and north american moose (groves and grubb 1987), the 2n = 70 form was recently discovered in eastern asia (boeskorov 1996, 1997). similarly, a length mutation (insertion-deletion, or indel) within the control region of mtdna originally was described as discriminating between north american and european moose (mikko and andersson 1995), but subsequent investigations documented that indel in moose from eastern asia (hundertmark et al. 2002b, udina et al. 2002). although the precise g e o g r a p h i c d i s t r i b u t i o n s o f t h o s e polymorphisms in karyotype and mtdna alces vol. 40, 2004 moose genetics hundertmark and bowyer 109 length are not well described, they seem to correspond geographically with a zone of intergradation in east-central asia, similar to that proposed for american and eurasian types of moose (flerov 1952, geist 1998). more work is needed to determine the extent of that geographic correspondence and to determine if it coincides with subspecies boundaries. moreover, the question of reproductive viability of the two chromosomal races must be addressed. indeed, boeskorov (1997) has proposed that the chromosomal races are different species and groves and grubb (1987) have identified them as “semispecies.” we caution, however, that chromosome numbers may be a poor designator of species among large mammals (bowyer et al. 2000). hundertmark et al. (2002b) tested the 2-types hypothesis by examining the distribution of genetic variance of mtdna among and within different hypothetical population structures. those authors sampled moose throughout their worldwide distribution and arranged them into either 2 groups (corresponding to the 2-types hypothesis) or 3 groups corresponding to continent of origin (asia, europe, and north america). they then examined the distribution of genetic variance within and among or between groups, predicting that the correct structure would minimize within-group variation and maximize among-group variation. percentage of total variation observed among the 3 groups was slightly greater than variation between the 2 groups (61.5% vs. 58.1%), and variation among populations within groups was minimized in the 3-group comparison (21.6% vs. 28.8% in the 2-group comparison). therefore, hundertmark et al. (2002b) concluded that mtdna data provided no support for a 2-type over a 3type hypothesis. that finding can be visualized by a phylogenetic tree constructed from haplotypes originally reported by hundertmark et al. (2002b, 2003), which shows north american moose as distinct from both asian and european forms (fig. 2). ancient bottlenecks and the mother of all moose moose worldwide exhibit little variation in a fragment of the mitochondrial cytochrome-b gene (hundertmark et al. 2002a). cytochrome b is useful for constructing mammalian phylogenies (irwin et al. 1991) and a paucity of variation in moose indicated a recent common ancestry likely due to a severe bottleneck that affected all extant lineages (hundertmark et al. 2002a), which is in agreement with the findings of mikko and andersson (1995). analysis of variation within the mitochondrial control region, which evolves at a much faster rate than cytochrome b (lopez et al. 1997), was fig. 2. unrooted phylogenetic tree of moose populations and subspecies worldwide, with the exception of a. a. pfizenmayeri, using fst a s t h e d i s t a n c e m e a s u r e [ d a t a f r o m hundertmark et al. (2002b, 2003)]. note the distinct positions of the 3 eurasian subspecies and north american moose, which do not support an hypothesis of 2 or 3 races of moose worldwide. hundertmark and bowyer moose genetics alces vol. 40, 2004 110 fig. 3. a phylogenetic tree of haplotypes of the mitochondrial control region of moose. symbols indicate region of origin, with black symbols indicating asian origin. distinct clades or phylogroups are indicated on the right. from hundertmark et al. (2002b). necessary to reveal significant levels of variation in moose and subsequently geographic patterns were revealed. control region haplotypes of moose can be divided into 3 clades, or haplogroups (fig. 3). the basal haplogroup (i.e., the group that diverges first from the base of the tree) is entirely asian, which suggests that those are the oldest moose haplotypes. the two other haplogroups are primarily european and primarily north american, although some asian haplotypes occur in both. the distribution of haplogroups on a worldwide scale illustrates the age of continental assemblages of haplotypes (fig. 4). the yakutia area contains moose from all 3 haplogroups. the russian far east contains both european and asian haplogroups, but not north american, and both europe a n d n o r t h a m e r i c a c o n t a i n o n l y 1 haplogroup each. therefore, yakutia can be identified as the area from which all extant moose lineages were derived, i.e., it is the oldest extant moose population that has been sampled. yakutia probably was the center of a single moose population during the last ice age, or at least was the location of the only population to provide descendants of modern moose. moose would have been restricted in their distribution because the cooler climate in asia at that time would have resulted in a shift of boreal forest habitat to the south. that habitat could have shifted only so far southward, because of prominent mountains running east-west, which would have formed an effective barrier to further movement to the south (hewitt 1996). during the last ice age, there were 2 periods of maximum glacial advance, termed the lower and upper pleniglacials. those episodes occurred at approximately 62,000 and 20,000 years ago, respectively (fulton et al. 1986). boreal forest habitats in asia would have shifted to the south during those cooling episodes and would have been compressed against the northern slopes of mounfig. 4. distribution of mitochondrial haplogroups worldwide. note that moose from the yakutia area have the most diverse composition and that moose from north america do not share haplogroups with moose from russian far east. alces vol. 40, 2004 moose genetics hundertmark and bowyer 111 59,000 and 14,000 years ago. when expansion times of moose are overlaid on a profile of global temperature change for the last ice age (jouzel et al. 1987), population expansion is correlated with warming trends following the pleniglacials (fig. 6). consequently, the evolution and geographic distribution of moose seems to have been affected substantially by climate change, particularly climate reversals associated with the late pleistocene and early holocene. coming to america cronin (1992) analyzed subspecific variation in north american cervids using restriction fragment length polymorphisms (rflp) of mtdna. despite analyzing 32 moose sampled from different regions of north america, he documented no variation among haplotypes from that continent. in comparison to other north american cervids, the lack of variation among subspecies of moose was interpreted as an indicafig. 5. glacial coverage during the last glacial maximum, superimposed on a map of presentday sea level. note that the bering land bridge between north america and asia would have been exposed during the glacial maximum due to lower sea levels. names of major ice sheets are provided. tain ranges. during subsequent warming intervals, moose habitat would have spread to the north, allowing moose populations to expand (guthrie 1995). unlike north america, much of eurasia was not glaciated during those periods (arkhipov et al. 1986), providing potential dispersal routes across the continent (fig. 5). the process of latitudinal shifts of range associated with episodes of climate change results in decreases of existing genetic diversity (hewitt 1996) and leaves characteristic signatures in the genome of modern moose. moose in eurasia underwent 2 distinct, recent population expansions (hundertmark et al. 2002b). any other historic population processes preceding those expansions are not detectable because low population sizes eliminated much of the genetic variation present in the pre-bottleneck populations, and hence no signature from those times exists in the present genome. by applying a molecular clock to those expansion data, hundertmark et al. (2002b) estimated that the expansions occurred approximately 160 140 120 100 80 60 40 20 0 2 0 -2 -4 -6 -8 -10 thousands of years ago te m p e ra tu re d i ff e re n c e ( c º) fig. 6. representation of mean global temperatures during the last 160,000 years relative to mean temperature in 1900. negative temperature differences indicate periods colder than today. the 2 glacial maxima of the last ice age are indicated by arrows. estimated dates of moose population expansion are indicated by dashed lines and correspond to periods of warming associated with glacial retreat and northward advance of the boreal forest in eurasia. temperature profile adapted from http://gcrio.ciesin.org/consequnces/winter96/article1-fig3.html. hundertmark and bowyer moose genetics alces vol. 40, 2004 112 ity in composition of mtdna haplotypes. as noted previously, north american subspecies are distinct from european and asian subspecies (fig. 2) and none of the asian haplotypes in the north american haplogroup were found in the russian far east. all haplotypes found in the russian far east (magadan oblast) are restricted to haplogroup 2 (fig. 3), which contains all european haplotypes. a likely colonization scenario entails closely related moose from central asia colonizing both europe and north america. lack of genetic similarity between moose in alaska and the russian far east is inconsistent with the scenario proposed by hundertmark et al. (2002b) concerning a colonizing wave of moose traveling from asia to north america through beringia. a single wave of moose colonizing north america through beringia would have left genetically similar populations on either side of the bering strait after the bering land bridge flooded. yet, ages of subfossil remains of moose from north america indicate clearly that a. alces was present in alaska prior to anywhere else on the continent (guthrie 1990, hundertmark et al. 2003), unequivocally supporting an entry through beringia. one hypothesis that accounts for this seeming paradox is that 1 or both of those populations underwent population bottlenecks shortly after the colonization of north america, and have only recently reestablished populations in those areas, leading to the presence of different genetic lineages in each. moose in the russian far east show evidence of an expansion approximately 1,200 years ago (hundertmark et al. 2002b) and continue to expand their range toward the bering sea (zheleznov 1993). similarly, alaskan moose show a surprising lack of mitochondrial diversity compared with moose elsewhere on the continent (hundertmark et al. 2003), which is tion of a recent common ancestry, consistent with colonization of the continent after the retreat of ice sheets from the last ice age (cronin 1992). similarly, no variation was detected within a fragment of cytochrome b in north american moose compared with slight variation in eurasia (hundertmark et al. 2002a). moose populations in north america were established as a result of a single entry into the continent, and that entry occurred during the population expansion of moose 14,000 years ago at the end of the last ice age, just before the closure of the bering land bridge (guthrie 1995, hundertmark et al. 2002b). a recent entry into north america is the only conclusion that is consistent with limited variation in the mtdna control region both within north america and between north america and eurasia. if moose had existed in 4 separate refugia in north america during the last ice age, as suggested by peterson (1955), or had entered north america from asia more than once, a more distant common ancestor would be indicated. the timing of the entry is supported not only by genetic data but also by the distribution of suitable moose habitat at the end of the last ice age. the cold, dry grassland habitat that prevailed in beringia for most of the last ice age was unsuitable for moose and was replaced by boreal forest only within the last 14,000 years (guthrie 1995). evidence from mtdna variation also indicates that north american moose did not originate in beringia, as some have speculated (cronin 1992, geist 1998), or recolonize the russian far east from north america (coady 1982). moose in north america are not closely related to moose on the western side of the bering strait (russian far east). if those moose were once part of the same population recently separated by the flooding of the bering land bridge, we would still expect to find similaralces vol. 40, 2004 moose genetics hundertmark and bowyer 113 indicative of a bottleneck notwithstanding the moderate levels of allozyme diversity reported for alaskan moose by hundertmark et al. (1992). diversity of mtdna can be reduced to a much greater degree by a bottleneck than diversity of nuclear dna because mtdna is 4 times more sensitive to genetic drift due to its haploid, uniparental mode of inheritance (birky et al. 1983). simultaneous bottlenecks in moose from both sides of the bering strait suggest a widespread causal factor. recent studies have found evidence of significant biotic effects of climate reversals in beringia after flooding of the bering land bridge (elias 2000, mason et al. 2001, anderson et al. 2002). those effects offer an intriguing mechanism for bottlenecks in beringian moose populations. the greatest variation in mtdna in north american moose occurs within the range of a. a. andersoni (hundertmark et al. 2003). alces a. shirasi from colorado exhibited no diversity and a. a. gigas and a. a. americana exhibited very little diversity. if the paucity of mitochondrial diversity in alaska is due to a bottleneck and recent expansion, those data would be consistent with the serial-founder-events hypothesis o f n o r t h a m e r i c a n c o l o n i z a t i o n (hundertmark et al. 1992). the pattern of colonization of north america undoubtedly was influenced by the retreating glaciers and may have had s o m e e f f e c t o n g e n e t i c s t r u c t u r e (hundertmark et al. 2003). based on the r e c o n s t r u c t i o n o f t h e r e t r e a t o f t h e laurentide ice sheet by dyke and prest (1987), we offer the following scenario of colonization. at the last glacial maximum, the cordilleran and laurentide ice sheets created an effective barrier between eastern beringia (alaska) and other parts of the continent (fig. 5). as glaciers retreated, a corridor opened on the eastern slopes of the rocky mountains allowing passage to the south. by 10,000 years ago, western canada was ice-free but central and eastern canada remained covered by the laurentide ice sheet and large proglacial lakes (fig. 7). passage to eastern canada north of the great lakes was impossible at this time and the only dispersal corridor was south of the lakes. by 8,400 years ago, moose arriving in the eastern continent could have dispersed westward north of the lakes, skirting the southern shores of the proglacial lakes to the north, and come into secondary contact with moose from the west once the proglacial lakes receded. moose in the e a s t e r n p a r t o f t h e c o n t i n e n t (a. a. americana) would have been on the end of a series of founder events, explaining their low mitochondrial variability and the presence of a contact zone between a. a. americana and the much more variable a. a. andersoni in ontario between the great lakes and hudson bay. morphological adaptation moose of the pleistocene and those that entered north america at the beginning of the holocene were significantly larger than those living today, a trait shared with other northern ruminants (guthrie 1984). guthrie (1984) proposed that the reduction in body size was an adaptation to changes in seasonal forage availability that occurred as a result of climate amelioration at the end of the last ice age. the ability of moose to respond to a rapidly changing environment belies the relatively low levels of genetic variation documented by the studies we have reviewed and demonstrates that evolutionary potential is not easily predicted solely by genetic variability but ultimately is determined by the presence of adaptive genetic variation and heritability of traits that improve fitness (lynch 1996). a general reduction in body size is not the only change to occur since the colonizat i o n o f n o r t h a m e r i c a . m o o s e i n hundertmark and bowyer moose genetics alces vol. 40, 2004 114 fig. 7. coverage of north america by the cordilleran and laurentide ice sheets and proglacial lakes at 14,000, 10,000, and 8,400 years ago (adapted from dyke and prest 1987). north america exhibit many differences in behavior and morphology. alaskan moose are perhaps the most divergent; they exhibit a degree of sociality not observed elsewhere (molvar and bowyer 1994) and have more distinctive body markings, also indicative of increased sociality (bowyer et al. 1991). molvar and bowyer (1994) suggested that moose in alaska have evolved sociality recently as a response to living in open environments. adaptation to open environments also applies to their mating system, which is harem-based. harem mating is adaptive in open environments (hirth 1977) where a male can protect a harem from competitors. moose elsewhere in north america exhibit a tending-bond system of mating, which is adaptive for forested environments. moose in alaska and siberia exhibit the largest body size of moose in north america and eurasia, respectively. the similar appearance of moose occurring on either side of the bering strait has caused some investigators to consider them the same subspecies (e.g., telfer 1984). as moose from those 2 regions are not closely related (hundertmark et al. 2002a,b), their similarity in size must result from convergent evolution. both subspecies have adapted to open, northern habitats by increasing body size. adaptation to open habitats was demonstrated with a multivariate analysis of antler size among moose inhabiting different areas and habitats in alaska (bowyer et al. 2002). those moose inhabiting open habitats (tundra) tended to have larger antlers overall than those living in boreal forest (taiga). similarly, moose occupying mountainous habitat in the southern portions of their range in north america (a. a. shirasi) and asia (a. a. cameloides) are similar; exhibiting small body and antler size (geist 1998). bubenik (1998) explained that similarity by proposing a second entry into north america by asian moose—an entry that bypassed beringia by traveling along the southern coast of alaska. neither genetic nor fossil data support that hypothesis (guthrie 1990, hundertmark et al. 2002a,b, 2003). genetic effects of management to this point we have discussed distribution of genetic and morphological diversity over space and time in the context of genetic drift and selection for locally adaptive traits. those patterns are developed over relatively long periods of time and are integral to the process of evolution. genetic alces vol. 40, 2004 moose genetics hundertmark and bowyer 115 change also can occur over short periods due to human influences, notably harvest management, and those changes may have unintended and undesirable consequences on individuals and populations (harris et al. 2002) and therefore are important to recognize. although it might be assumed that a well-designed harvest plan acts in a random fashion on genetic makeup of individuals, in reality even a managed harvest can be a highly selective force with measurable consequences in just a few generations (coltman et al. 2003). ryman et al. (1981) modeled the effect of different harvest strategies on 2 factors critical in determining the extent of genetic drift and inbreeding in populations: effective population size and generation interval, respectively. hunting strategies were defined by different probabilities of harvest for both juveniles and adult females, and resulted in stable populations with decreased generation intervals and effective sizes compared with unhunted populations. moreover, temporal changes in genetic diversity differed for different harvest strategies but always decreased. thus, ryman et al. (1981) demonstrated that improperly designed harvest regimes can affect genetic characteristics of populations and by extension may have an influence on evolutionary potential. conversely, cronin et al. (2001) detected no differences in numbers of alleles or levels of heterozygosity for 5 microsatellite loci among 3 moose populations in quebec, 1 heavily hunted, 1 lightly hunted, and 1 not hunted. another critical factor in management is the effect of harvest on genetic loci underlying characters having a direct effect on reproductive fitness, e.g., antler size in moose. controlling harvest by defining legal males according to antler size is common in management of north american elk (thelen 1991) and is a strategy employed in moose management in british columbia, canada, and alaska, usa (child 1983, schwartz et al. 1992). in an effort to evaluate genetic effects of the selective harvest system in alaska, hundertmark et al. (1993, 1998) modeled moose populations subject to harvest strategies employing different definitions of legal males. they concluded that selective harvest systems could result in allele frequency changes at loci coding for antler characteristics (hundertmark et al. 1993) and that the position of a population relative to nutritional carrying capacity of the habitat affected the rate of change in allele frequencies (hundertmark et al. 1998). those results indicated that limiting harvest to moose with large antlers could cause a genetically based decrease in antler size over time. such a reduction in adaptive genetic variation runs counter to general conservation goals (lynch 1996). a stunning example of the effect of harvest on fitness traits was recently reported by coltman et al. (2003), who documented significant genetic effects on horn size and b o d y m a s s i n b i g h o r n s h e e p ( o v i s canadensis) as a result of selective harvest of males with large horns. detecting potential changes in genetic composition of moose as they respond to various anthropogenic influences, whether related to management or to changes in the environment, is a difficult task. nonetheless, it is an important area of investigation and deserves attention. modeling exercises and studies of genetic variation have not addressed interrelationships of moose populations at fine scales. the effects of management and habitat alteration on processes involved in maintenance of connectivity of moose populations, such as malemediated gene flow via yearling dispersal are extremely important and should be described. moose are highly adaptable animals, but the intensive management of moose hundertmark and bowyer moose genetics alces vol. 40, 2004 116 populations and the environmental factors influencing their habitat (e.g., wildfire) may have unintended and significant consequences on the moose genome through a change in selective forces. proper conservation of this species requires that we recognize and avoid that possibility. acknowledgements we thank the organizers of the 5th international moose symposium for the opportunity to synthesize and present this information. we wish to acknowledge funding support from federal aid in wildlife restoration through the alaska department of fish and game as well as the institute of arctic biology, university of alaska fairbanks. s. côté and an anonymous reviewer provided constructive comments for improving the manuscript. references anderson, p. m., a. v. lozhkin, and l. b. brubaker. 2002. implications of a 24,000-yr palynological record for a younger dryas cooling and for boreal forest development in northeastern siberia. quaternary research 57:325333. arkhipov, s. a., v. g. bespaly, m. a. faustova, o. y. glushkova, l. l. isaeva, and a. a. velichko. 1986. ice sheet reconstructions. pages 269-292 in v. šibrava, d. q. bowen, and g. m. r i c h m o n d , e d i t o r s . q u a t e r n a r y glaciations in the northern hemisphere. pergamon press, oxford, u.k. avise, j. c. 2000. phylogeography: the history and formation of species. harvard university press, cambridge, massachusetts, usa. _____, j. arnold, r. m. ball, jr., e. bermingham, t. lamb, j. e. neigel, c. a. reeb, and n. c. saunders. 1987. intraspecific phylogeography: the mitochondrial dna bridge between population genetics and systematics. annual review of ecology and systematics 18:489-522. birky, c. w., t. maruyama, and p. fuerst. 1983. an approach to population and evolutionary genetic theory for genes in mitochondria and chloroplasts, and some results. genetics 103:513-527. blundell, g. m., m. ben-david, p. groves, r. t. bowyer, and e. geffen. 2002. characteristics of sex-biased dispersal and gene flow in coastal river otters: implications for natural recolonization of extirpated populations. molecular ecology 11:289-303. boeskorov, g. g. 1996. karyotype of moose (alces alces l.) from northeastern asia. proceedings of the russian academy of sciences 329:506508. (in russian). _____. 1997. chromosomal differences in moose. genetika 33:974-978. (in russian). bowyer, r. t., d. m. leslie, jr., and j. l. rachlow. 2000. dall’s and stone’s sheep. pages 491-516 in s. demarais and p. r. krausman, editors. ecology and management of large mammals in north america. prentice hall, upper saddle river, new jersey, usa. _ _ _ _ _ , j . l . r a c h l o w , v . v a n ballenberghe, and r. d. guthrie. 1991. evolution of a rump patch in moose: an hypothesis. alces 27:12-23. _____, k. m. stewart, b. m. pierce, k. j. hundertmark, and w. c. gasaway. 2002. geographical variation in antler morphology of alaskan moose: putative effects of habitat and genetics. alces 38: 155-165. _____, v. van ballenberghe, and j.g. kie. 2003. moose (alces alces). pages 931 964 in g.a. feldhamer, b. thompson, and j. chapman, editors. wild mammals of north america: biology, management, and economics. secalces vol. 40, 2004 moose genetics hundertmark and bowyer 117 ond edition. the johns hopkins university press, baltimore, maryland, usa. braend, m. 1962. studies on blood and serum groups in the elk (alces alces). new york academy of sciences 97:296305. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8:1309-1315. bruford, m. w., d. j. cheesman, t. coote, h. a. a. green, s. a. haines, c. o’ryan, and t. r. williams. 1996. microsatellites and their application to conservation genetics. pages 278-297 in t. b. smith and r. k. wayne, editors. molecular genetic approaches in conservation. oxford university press, oxford, u.k. b r u m f i e l d , r . t . , p . b e e r l i , d . a . nickerson, and s. v. edwards. 2003. t h e u t i l i t y o f s i n g l e n u c l e o t i d e polymorphisms in inferences of population history. trends in ecology and evolution 18:249-256. bubenik, a. b. 1998. evolution, taxonomy and morphophysiology. pages 77-123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. chesser, r. k., c. reuterwall, and n. ryman. 1982. genetic differentiation of scandinavian moose alces alces populations over short geographic distances. oikos 39:125-130. child, k. n. 1983. selective harvest of moose in the omineca: some preliminary results. alces 19:162-177. coady, j. w. 1982. moose. pages 902-922 in j. a. chapman and g. a. feldhamer, editors. wild mammals of north america: biology, management, and economics. johns hopkins university press, baltimore, maryland, usa. coltman, d. w., p. o’donoghue, j. t. jorgenson, j. t. hogg, c. strobeck, and m. festa-bianchet. 2003. undesirable evolutionary consequences of trophy hunting. nature 426:655-658. cronin, m. a. 1992. intraspecific variation in mitochondrial dna of north american cervids. journal of mammalogy 73:70-82. _____, j. c. patton, r. courtois, and m. crête. 2001. genetic variation of microsatellite dna in moose in québec. alces 37:175-187. dyke, a. s., and v. k. prest. 1987. late wisconsinan and holocene history of the laurentide ice sheet. geographie physique et quaternaire 41:237-263. elias, s. a. 2000. late pleistocene climates of beringia, based on analysis of fossil beetles. quaternary research 53:229-235. ellegren, h., l. andersson, and k. wallin. 1991. dna polymorphism in moose (alces alces) revealed by polynucleotide probe (tc)n. journal of heredity 82:429-431. _____, s. mi k k o, k. wa l l i n, and l. andersson. 1996. limited polymorphism at major histocompatibility complex (mhc) loci in the swedish moose a. alces. molecular ecology 5:3-9. flerov, k. k. 1952. fauna of the ussr: mammals. vol. 1 no 2. musk deer and deer. the academy of sciences of the ussr, leningrad. (english translation by the israel program for scientific translation, 1960, s. monson, jerusalem). franzmann, a. w., and c. c. schwartz, editors. 1998. ecology and management of the north american moose. smithsonian institution press, washhundertmark and bowyer moose genetics alces vol. 40, 2004 118 ington, d.c., usa. fries, r., and g. durstewitz. 2001. digitaldna signatures: snps for animal tagging. nature biotechnology 19:508. fulton, r. j., m. m. fenton, and n. w. rutter. 1986. summary of quaternary stratigraphy and history, western canada. pages 229-242 in v. šibrava, d. q. bowen, and g. m. richmond, editors. quaternary glaciations in the northern hemisphere. pergamon press, oxford, u.k. geist, v. 1987a. on the evolution and adaptations of alces. swedish wildlife research supplement 1:11-23. _____. 1987b. on speciation in ice age mammals, with special reference to cervids and caprids. canadian journal of zoology 65:1067-1084. _____. 1998. deer of the world: their evolution, behavior, and ecology. stackpole books, mechanicsburg, pennsylvania, usa. groves, c. p., and p. grubb. 1987. relationships of living deer. pages 3-59 in c. m. wemmer, editor. biology and m a n a g e m e n t o f t h e c e r v i d a e . smithsonian institution press, washington, d.c., usa. guthrie, r. d. 1984. alaska megabucks, megabulls, and megarams: the issue of pleistocene gigantism. special publications of the carnegie museum of natural history 8:482-509. _____. 1990. new dates in alaskan quaternary moose, cervalces-alces – archaeological, evolutionary and ecological implications. current research in the pleistocene 7:111-112. _____. 1995. mammalian evolution in response to the pleistocene-holocene transition and the break-up of the mammoth steppe: two case studies. acta zoologica cracoviensia 38:139-154. gyllensten, u., c. reuterwall, n. ryman, and g. stahl. 1980. geographical variation in transferrin allele frequencies in three deer species from scandinavia. hereditas 92:237-241. hancock, j. m. 1999. microsatellites and other simple sequences: genomic context and mutational mechanisms. pages 1-9 in d. b. goldstin and c. schlötterer, editors. microsatellites: evolution and applications. oxford university press, oxford, u.k. harrington, r. 1985. evolution and distribution of the cervidae. biology of deer production. the royal society of new zealand, bulletin 22:3-11. harris, r. b., w. a. wall, and f. w. allendorf. 2002. genetic consequences of hunting: what do we know and what should we do? wildlife society bulletin 30:634-643. hedrick, p. w. 1994. evolutionary genetics of the major histocompatibility complex. american naturalist 143:945964. hewitt, g. m. 1996. some genetic consequences of ice ages, and their role in divergence and speciation. biological journal of the linnaen society 58:247276. hirth, d. h. 1977. social behavior of white-tailed deer in relation to habitat. wildlife monographs 53. hundertmark, k. j., r. t. bowyer, g. f. shields, and c. c. schwartz. 2003. mitochondrial phylogeography of moose (alces alces) in north america. journal of mammalogy 84:718-728. _____, p. e. johns, and m. h. smith. 1992. genetic diversity of moose from the kenai peninsula, alaska. alces 28:15-20. _____, g. f. shields, r. t. bowyer, and c. c. schwartz. 2002a. genetic relationships deduced from cytochrome-b sequences among moose. alces 38:113122. alces vol. 40, 2004 moose genetics hundertmark and bowyer 119 _____, _____, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. schwartz. 2002b. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. molecular phylogenetics and evolution 22:375-387. _____, t. h. thelen, and r. t. bowyer. 1998. effects of population density and selective harvest on antler phenotype in simulated moose populations. alces 34:375-383. _____, _____, and c. c. schwartz. 1993. genetic and population effects of selective harvest systems in moose: a modeling approach. alces 29:225-234. irwin, d. m., t. d. kocher, and a. c. wilson. 1991. evolution of the cytochrome b gene in mammals. journal of molecular evolution. 32: 128 144. jouzel, j., c. lorius, j. r. petit, c. genthon, n. i. barkov, v. m. kotlyakov, and v. m. petrov. 1987. vostok ice core: a continuous isotope temperature record over the last climate cycle (160,000 years). nature 329:403-408. lister, a. m. 1993. evolution of mammoths and moose: the holarctic perspective. pages 178-204 in a. d. barnosky, editor. quaternary mammals of north america. cambridge university press, cambridge, u.k. lopez, j. v., m. culver, j. c. stephens, w. e. johnson, and s. j. o’brien. 1997. rates of nuclear and cytoplasmic mitochondrial dna sequence divergence in mammals. molecular biology and evolution 14:277-286. lynch, m. 1996. a quantitative-genetic perspective on conservation issues. pages 471-501 in j. c. avise and j. l. hamrick, editors. conservation geneti c s : c a s e s t u d i e s f r o m n a t u r e . chapman & hall, new york, new york, usa. mason, o. k., p. m. bowers, and d. m. hopkins. 2001. the early milankovitch thermal maximum and humans: adverse conditions for the denali complex of eastern beringia. quaternary science reviews 20:525-548. mikko, s., and l. andersson. 1995. low major histocompatibility complex class ii diversity in european and north american moose. proceedings of the national academy of sciences, usa 92:4259-4263. miyamoto, m. m., f. kraus, and o. a. ryder. 1990. phylogeny and evolution of antlered deer determined from mitochondrial dna sequences. proceedings of the national academy of sciences, usa 87:6127-6131. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75:621630. mullis, k., f. faloona, s. scharf, r. sakai, g. horn, and h. erlich. 1986. specific enzymatic amplification of dna in vitro: the polymerase chain reaction. cold spring harbor symposia in quantitative biology 51:263-273. nadler, c. f., c. e. hughes, k. e. harris, and n. w. adler. 1967. electrophoresis of the serum proteins and transferrins of alces alces (elk), rangifer tarandus (reindeer), and ovis dalli (dall sheep) from north america. comparative biochemistry and physiology 23:149157. nevo, e. 1978. genetic variation in natural populations: patterns and theory. theoretical population biology 13:121-177. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. quellar, d. c., j. e. strassmann, and c. r. hughes. 1993. microsatellites and kinship. trends in ecology and evolution 8:285-288. hundertmark and bowyer moose genetics alces vol. 40, 2004 120 røed, k. h., and l. midthjell. 1998. microsatellites in reindeer, rangifer tarandus, and their use in other cervids. molecular ecology 7:1773-1776. ryman, n., r. baccus, c. reuterwall, and m. h. smith. 1981. effective population size, generation interval, and potential loss of genetic variability in game species under different hunting regimes. oikos 36:257-266. _____, g. beckman, g. bruun-petersen, and c. reuterwall. 1977. variability of red cell enzymes and genetic implicat i o n s o f m a n a g e m e n t p o l i c i e s i n scandinavian moose (alces alces). hereditas 85:157-162. _____, c. reuterwall, k. nygren, and t. nygren. 1980. genetic variation and differentiation in scandinavian moose (alces alces): are large mammals monomorphic. evolution 34:1037-1049. sage, r. d., and j. o. wolff. 1986. pleistocene glaciations, fluctuating ranges, and low genetic variability in a large mammal (ovis dalli). evolution 40: 1092 1095. schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula, alaska. alces 28:1-13. scribner, k. t., and r. k. chesser. 2001. group-structured genetic models in analysis of the population and behavioral ecology of poikilothermic vertebrates. journal of heredity 92:180-189. selander, r. k., and d. w. kaufman. 1973. genic variability and strategies of adaptation in animals. proceedings of the national academy of sciences of the usa 70:1875-1877. sher, a. v. 1987. history and evolution of moose. swedish wildlife research supplement 1:71-97. shubin, p. n. 1969. the genetics of transferrins in the reindeer and in the european elk. genetika 5:37-41. (in russian). smith, m. h., r. baccus, h. o. hillestad, and m. n. manlove. 1984. population genetics. pages 119-128 in l. k. halls, editor. white-tailed deer: ecology and m a n a g e m e n t . s t a c k p o l e b o o k s , harrisburg, pennsylvania, usa. _____, k. t. scribner, l. h. carpenter, and r. h. garrott. 1990. genetic characteristics of colorado mule deer (odocoileus hemionus) and comparisons with other cervids. southwestern naturalist 35: 1 8. symondson, w. o. 2002. molecular identification of prey in predator diets. molecular ecology 11:627-641. telfer, e. s. 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145-182 in r. olson, f. geddes, and r. hastings, editors. northern ecology and resource management. university of alberta press, edmonton, canada. thelen, t. h. 1991. effects of harvest on antlers of simulated populations of elk. journal of wildlife management 55:243249. thouveny, n., and e. bonifay. 1984. new c h r o n o l o g i c a l d a t a o n e u r o p e a n plio-pleistocene faunas and hominid occupation sites. nature 308:355-358. udina, i. g., a. a. danilkin, and g. g. boeskorov. 2002. genetic diversity of moose (alces alces l.) in eurasia. genetika 38:951-957. (in russian). vernesi, c., e. pecchiolo, d. caramelli, r. tiedemann, e. randi, and g. bertorelle. 2002. the genetic structure of natural and reintroduced roe deer (capreolus capreolus) populations in the alps and central italy, with reference to the mitochondrial dna phylogeography of europe. molecular ecology 11:1285-1297. wilhelmson, m., r. k. juneja, and s. bengtsson. 1978. lack of polymorphism in certain blood proteins and enalces vol. 40, 2004 moose genetics hundertmark and bowyer 121 zymes of european and canadian moose. naturaliste canadien 105:445449. wilson, g. a., c. strobeck, l. wu, and j. w. coffin. 1997. characterization of microsatellite loci in caribou rangifer t a r a n d u s, a n d t h e i r u s e i n o t h e r artiodactyls. molecular ecology 6:697699. wilson, p. j., s. grewal, a. rodgers, r. rempel, j. saquet, h. hristienko, f. burrows, r. peterson, and b. n. white. 2003. genetic variation and population structure of moose (alces alces) at neutral and functional dna loci. canadian journal of zoology 81:670-683. zheleznov, n. k. 1993. historic and current distribution of moose in the northeast ussr. alces 29:213-218. hundertmark and bowyer moose genetics alces vol. 40, 2004 122 term definition adaptive genetic variation genotypic variation at loci that control traits important to fitness, such as morphology, physiology, and behavior. allozyme a gene product (protein) that is distinguished by its migratory characteristics in a gel exposed to an electric field (electrophoresis). differences among alleles (different variants of the same gene) at allozyme loci relate to amino acid composition and secondary structure of the protein. only mutations that create proteins with different migratory characteristics are detectable. control region a portion of mtdna that evolves (incorporates mutations) at a very fast rate, which makes it a valuable marker for examining intraspecific genetic variation. occasionally referred to as the d-loop. dna sequence the ultimate level of analysis of genetic material. this technique deduces the identity and order of nucleotides in a fragment of dna. mutations (nucleotide substitutions) are detectable whether or not they create different gene products. effective population size the size of a standardized population that has the same degree of genetic drift as the population being studied. an ideal population is a closed population of constant size with non-overlapping generations and no variance in reproductive success. the smaller the effective size of a population, the faster it will lose genetic diversity through drift regardless of actual population size. effective population size (ne) is almost always a fraction of the true population size (n). generation interval mean age of all parents. genetic drift changes in allele frequencies across generations due to random sampling error associated with less than infinite population size. haplotype the haploid equivalent of genotype. genetic type of an individual when haploid dna, such as mtdna, is analyzed. heritability the proportion of variance in the expression of a trait, such as antler size or body size, that is due to genetic effects (as opposed to environmental effects), i.e., the degree to which a trait can be passed on to the next generation. microsatellite segments of dna composed of sequence units varying from 2-4 nucleotides in length that are tandemly repeated. size variation (number of repeated units) defines alleles. microsatellites are mendelian in their inheritance, i.e., they are diploid, specific to a site on a chromosome, and occur in either a homozygous or heterozygous genotype. minisatellite similar to microsatellites except that the repeat units vary from 16-64 nucleotides in length and occur at many sites throughout the chromosomes, thus exhibiting more than 2 alleles per individual and creating complex genotypes of gel banding patterns that resemble bar codes. this is the technique that pioneered genetic fingerprinting. mitochondrial dna (mtdna) a circular dna molecule occurring in the mitochondrion. unlike chromosomes, which occur in pairs (diploid), mtdna occurs as a single copy (haploid) because it is inherited only from the maternal line. therefore, it is not subject to recombination and changes only via mutation. that property makes it particularly useful for tracing lineages through time. molecular clock the assumption that the average rate of mutation for a particular dna sequence is constant over evolutionary time. if a molecular clock can be assumed, the amount of genetic divergence between populations or species can be converted to time since divergence. phylogeography the study of genetic lineages in a spatial and temporal context, revealing historic population processes and evolutionary histories. rflp restriction fragment length polymorphism – a section of dna of known length is digested with restriction enzymes (endonucleases), which cleave dna at sites with specific target sequences of 4-6 nucleotides. the digested fragments are separated by length (number of nucleotides) using gel electrophoresis and haplotypes are characterized by numbers and sizes of fragments. appendix 1. glossary of specialized terms used in this review. alces37(1)_89.pdf alces37(1)_43.pdf f:\alces\vol_38\pagemaker\3812. alces vol. 38, 2002 hundertmark et al. cytb variation in moose 113 genetic relationships deduced from cytochrome-b sequences among moose kris j. hundertmark1,2, gerald f. shields2,3, r. terry bowyer2, and charles c. schwartz1,4 1alaska department of fish and game, kenai moose research center, 43961 kalifornsky beach road, suite b, soldotna, ak 99669, usa; 2institute of arctic biology and department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775, usa abstract: we studied variation in nucleotide sequences of the mitochondrial cytochrome-b gene to assess the phylogeny of moose (alces alces) in general, and the position of north american moose within that phylogeny in particular. we combined north american, asian, and european haplotypes generated for this study with 3 eurasian haplotypes obtained from genbank. no nucleotide variation occurred within moose from north america, whereas 3 haplotypes were present in european moose and 4 haplotypes in asian moose. clade structure was consistent over 6 most-parsimonious trees, with asian haplotypes composing 1 clade, and north american and european haplotypes composing a second, albeit poorly supported clade. low diversity of nucleotides in cytochrome-b indicated a recent ancestry among moose worldwide. existence of 1 north american haplotype is strong evidence of a single, recent entry into the new world via the bering land bridge, rather than multiple entries through >1 corridors. furthermore, no phylogenetic support existed for the theory of distinct lineages of european versus asian-north american moose. alces vol. 38: 113 122 (2002) key words: alces alces, cytochrome-b, genetic diversity, mitochondrial dna, moose, phylogeography moose (alces alces) arose in eurasia in the late pleistocene (lister 1993), but paleontological (guthrie 1990) and genetic (cronin 1992) evidence indicate a recent colonization of north america. such a recent colonization would result in characteristic genetic signatures in mitochondrial dna (mtdna), a haploid genome that is transmitted maternally and is informative for constructing population histories of closely related taxa (avise et al. 1987). recently colonized areas would be expected to show less genetic diversity than areas with long-established populations, particularly if the effective size of the founding population was low. moreover, haplotype composition of a recently founded population would be expected to resemble the composition of the population from which the founders originated. cronin (1992) analyzed restriction fragment length polymorphisms (rflp) of mtdna in north american cervids and reported that moose were unique because they exhibited no detectable variation. there was no comparison to eurasian moose in that study, however, to determine the degree of difference between moose inhabiting different continents. therefore, the significance of those findings is difficult to assess. 3present address: department of natural sciences, carroll college, helena, mt 59601, usa 4present address: forestry sciences lab, montana state university, bozeman, mt 59717, usa alces vol. 38, 2002 hundertmark et al. cytb variation in moose 115 cytb evolves at a moderate rate in mammals (irwin et al. 1991, lopez et al. 1997) a n d , c o n s e q u e n t l y , o f t e n i s u s e d i n phylogenetic studies of conspecific and congeneric taxa. we used our data to determine if variation within north american moose would be less than variation in eurasian moose, and if eastern and western races of moose were represented by different lineages of mtdna. methods tissue samples were solicited from moose hunters in alaska as well as biologists from across north america, europe, and asia. samples were grouped to comprise > 1 population from the range of each north american subspecies, and those populations were combined into continentlevel associations. north american subspecies and sampling locations were: a. a. gigas (n = 34) from across its range in alaska; a. a. andersoni from central north america (usa: minnesota, north dakota, isle royale in michigan; canada: western ontario and manitoba; n = 8); a. a. shirasi from colorado, usa (n = 2); and a. a. americana from new hampshire, usa and new brunswick, canada (n = 7). the colorado population originated from 3 translocations of moose from neighboring states: 12 animals (8 females) from the uinta mountains, utah, usa, in 1978, 12 animals (11 females) from grand teton national park, wyoming, usa, in 1979, and 12 animals (10 females) from jackson hole, wyoming in 1987 (duvall and schoonveld 1988). asian subspecies and sources were: a. a. burturlini (n = 10), which consisted of animals from the ola peninsula near magadan, and the omolon and chelomya rivers, russia; a. a. cameloides, represented by 1 animal housed at a zoo in harbin, china, and a sequence from genbank (accession no. ay035872); and a. a. pfizenmayeri, consisting of a sequence obtained from genbank (accession no. ay035873). the european subspecies, a. a. alces, consisted of samples collected in finland (n = 6) and sweden (n = 6), as well as a sequence of a moose from norway obtained from genbank (accession no. aj000026; randi et al. 1998). tissue samples consisted of skeletal muscle, liver, kidney, skin, blood, or hair. tissues were stored temporarily at -20°c or preserved in 100% ethanol as soon as possible after collection and were archived at -80°c. all tissue types except blood were subjected to salt extraction for isolation of genomic dna. dna from blood was extracted with chelex (walsh et al. 1991). mtdna was isolated from nuclear dna and rna from 1 moose by means of a cscl 2 density-gradient centrifugation. that sample was used to verify the mitochondrial origin of amplified sequences. we targeted the 5’ region of the cytb gene for analysis. we amplified the seq u e n c e w i t h p r i m e r s m v z 0 5 ( 5 ’ — gcaagcttgatatgaaaaaccatcgttg— 3’) and mvz04 (5’—gcagcccctcagaatgatatttgtcctc—3’) first described by smith and patton (1993). double-stranded templates were amplified in a reaction mix containing 10 mm trishcl, ph 8.3, 50 mm kcl, 1.5 mm mgcl 2 , 0.2 mm dntp, 10 mm each primer, and 0.5 units dna polymerase. cycling conditions were a 2-min soak at 94°c followed by 30 cycles of 94°c (15 sec) denaturation, 50°c (15 sec) annealing, and 72°c (45 sec) extension, followed by one extension period of 10 min at 72°c. pcr products were visualized on a 6% agarose gel with ethidium bromide staining. cleaned pcr products were cycle sequenced (both directions) with fluorescing ddntps. nucleotide composition of the final products was determined on an automated sequencer (abi 373, pe applied biosystems, foster city, ca) with standard protocols supplied by the manucytb variation in moose hundertmark et al. alces vol. 38, 2002 116 facturer. sequences were aligned with the clustal v algorithm (higgins et al. 1992) and were edited by visual examination of electropherograms with sequence navigator software (abi). we compared nucleotide sequences of individuals and identified sites at which they differed; those data served as the basis for describing individual and population-level variation. populations also were characterized by haplotype diversity (h), which is the probability that 2 randomly selected individuals would have different haplotypes (nei and kumar 2000). nucleotide diversity (π; nei and li 1979), which is the probability that 2 homologous nucleotides would differ in 2 randomly chosen individuals, and number of pairwise differences, which is the number of nucleotide substitutions observed between 2 haplotypes, also were determined. those statistics were estimated with arlequin software (schneider et al. 2000). g e n e t i c d i s t a n c e s b e t w e e n p a i r s o f populations were computed by applying the kimura 2-parameter model of sequence evolution (kimura 1980). differentiation among all populations was assessed via φ st , which incorporates differences in nucleotide and haplotype diversity within and among populations. we analyzed relationships among haplotypes with a maximum parsimony (branch-and-bound search) cladogram, and a neighbor-joining tree (saitou and nei 1987) employing 2-parameter distance estimates. those analyses were performed with the program mega version 2.0 (kumar et al. 2001). trees were rooted by a haplotype from fallow table 1. nucleotide variation within the first 403 nucleotides of the 5’ end of the mitochondrial cytochrome-b gene in moose. only variable nucleotide positions are listed, and dots represent identity with the first sequence. all haplotypes were documented in moose from this study with the exception of europe3, which was reported by randi et al. (1998) from norway. nucleotide positions were numbered according to the bovine mitochondrial genome (anderson et al. 1982), with the first nucleotide of cytochrome-b numbered 14514. distributions of haplotypes by sampling location and by subspecies also are indicated. continent subspecies nucleotide position haplotype north america 55 26 20 2 7 c t t t t t c c asia1 8 7 1 t c • • • • t • asia2 3 1 2 t c • • • • • • asia3 1 1 t c • • • c t • asia4 1 1 t c • • • c • • europe1 4 4 • • a • • • • t europe2 8 8 • • a c • • • • europe3 1 1 • • a c g • • • n o rt h a m e ri c a a s ia e u ro p e a . a . a m er ic an a a . a . a nd er so ni a . a . s hi ra si a . a . g ig as a . a . b ur tu rl in i a . a . c am el oi de s a . a . p fi ze nm ay er i a . a . a lc es 1 4 5 2 3 1 4 5 5 9 1 4 5 7 7 1 4 5 9 5 1 4 6 1 3 1 4 6 5 2 1 4 7 3 9 1 4 8 7 2 alces vol. 38, 2002 hundertmark et al. cytb variation in moose 117 deer (dama dama) obtained from genbank (accession no. x56290; irwin et al. 1991), and haplotypes derived for this study from a caribou (rangifer tarandus granti) collected in alaska (denoted rangfer1) and a reindeer (r. t. tarandus) collected in the omolon river drainage of the russian far east (denoted rangifer2). confidence in the structure of the phylogenies was assessed through 1,000 bootstrap replicates (felsenstein 1985). results we detected 8 variable sites within the 403 nucleotides of the 5’ end of cytb, defining 8 haplotypes (table 1). six of the variable sites were transitions and 2 were transversions; the transversions were restricted to european haplotypes. the overall transition:transversion ratio was 7:1 including outgroups and 3:1 for moose only. six variable sites, including 1 transversion, were synonymous substitutions at the third position of codons, resulting in no substitutions of amino acids in the gene product. in haplotype europe3, however, 1 transversion occurred at the third position of the thirty-third codon and resulted in the substitution of the amino acid phenylalanine with leucine. the remaining substitution was a synonymous, first-position transition in europe1. all new haplotypes described in this study, including 2 outgroup haplotypes, were submitted to genbank and were assigned accession numbers ay090099−ay090107. we documented an extreme degree of differentiation among continents (φ st = 0.89), with no haplotype occurring on > 1 continent (table 1). we identified 4 asian, 3 e u r o p e a n , a n d 1 n o r t h a m e r i c a n haplotypes. pairwise differences among haplotypes ranged from 1 to 7 substitutions, and associated estimates of genetic distances ranged from 0.2 to 1.8% (table 2). estimates of mean (± sd) haplotype diversity for europe (h = 0.60 ± 0.13) and asia (h = 0.56 ± 0.11) were similar, as were estimates of nucleotide diversity for haplotypes occurring within those continents (table 3). the least genetic distance between continents was the comparison between europe and north america, and the greatest was between europe and asia. europe exhibited the greatest within contitable 2. genetic distances between pairs of haplotypes for a 403-nucleotide segment of the moose cytochrome-b gene. values above the diagonal are total numbers of substitutions, and those below the diagonal are estimates of substitutions per site using kimura’s (1980) 2-parameter model. north america asia1 asia2 asia3 asia4 europe1 europe2 europe3 north america 3 2 4 3 2 2 3 asia1 0.008 1 1 2 5 5 6 asia2 0.005 0.002 2 1 4 4 5 asia3 0.010 0.002 0.005 1 6 6 7 asia4 0.008 0.005 0.002 0.002 5 5 6 europe1 0.005 0.013 0.010 0.015 0.013 2 3 europe2 0.005 0.013 0.010 0.015 0.016 0.005 1 europe3 0.007 0.015 0.013 0.018 0.015 0.007 0.002 alces vol. 38, 2002 hundertmark et al. cytb variation in moose 119 solved, each with 79 steps (consistency index = 0.91, retention index = 0.90). the strict consensus tree, in which only those clades present in all most-parsimonious trees are displayed, exhibited 2 primary clades of moose haplotypes consisting of a strictly asian clade and a european-north american clade (fig. 2a). the neighbor-joining tree exhibited an identical structure (fig. 2b), although bootstrap support for the clades was weak. discussion our data are consistent with other studies of genetic variability in cervids that indicate a relative lack of diversity in moose (wilhelmson et al. 1978, ryman et al. 1980, baccus et al. 1983, cronin 1992), although an instance of high genetic variability in moose has been reported (hundertmark et al. 1992). within the same region of cytb that we studied, kuwayama and ozawa (2000) reported 32 variable sites among 5 subspecies of north american elk and euasian red deer (cervus elaphus), and 13 variable sites within 6 subspecies of sika deer (c. nippon) restricted to the islands of japan. the maximum number of substitutions between north american and asian haplotypes in moose was 4 (all transitions). comparatively, the minimum difference between haplotypes of north american elk (c. e. canadensis = c. e. nelsoni) and asian red deer (c. e. kansuensis) was 5 substitutions, 3 of which were transversions (kuwayama and ozawa 2000). the magnitude of that difference indicated that, despite similar geographic distributions, north american elk and asian red deer have been separated longer than north american and asian populations of moose. the fossil record supports that conclusion (guthrie 1966, 1990). low levels of variability we measured in cytb within continents and small genetic distances between continents indicate a recent common ancestry for moose worldwide. also, lack of shared haplotypes between continents suggests a small number of founders or bottlenecks. if the number of founding lineages in a continent had been large, we would have expected more haplotypic diversity within continents. patterns of variation we observed in moose from asia and europe were consistent, in each instance, with founding by 1 lineage followed by divergence of 1 or 2 mutations. parsimony analysis and genetic distances indicated a closer relationship between north american and european moose than between either of those and asian moose. thus, our data provide no support for a fundamental division of moose into european and asian-north american clades. rather, the europe-north america clade was split geographically by the asia clade, indicating that phylogenetic divergence was not reflected in geographic relationships. that pattern is consistent with a scenario in which moose populations worldwide trace back to recent population expansion combined with small sizes of founding populations (hundertmark et al. 2002). absence of variation in cytb in north america is strong evidence for a single colonization characterized by a small effective size. the relatively large haplotype diversity observed in asian moose likely would have resulted in >1 haplotype in north america if >1 colonization event occurred or if the colonization wave was comprised of many moose. our data indicate that eurasian moose exhibited more diversity than moose from north america, and we find that the spatial distribution of diversity within cytb supports the idea of establishment of continental or regional populations of moose via expansion from small numbers of founders. moreover, the sharing of 1 haplotype between a. a. cameloides and a. a. burtulini indicated either recent divergence of those populations cytb variation in moose hundertmark et al. alces vol. 38, 2002 120 or the presence of female-mediated gene flow and provided no evidence of an ext r e m e t e m p o r a l s e p a r a t i o n o f a . a . cameloides from other asian subspecies. finally, our data indicate a single entry of moose into north america from asia, and do not support a fundamental division of moose into european and asian-north american clades. acknowledgements funding for this research was provided by federal aid in wildlife restoration, the boone and crockett club, and the institute of arctic biology, university of alaska f a i r b a n k s ( u a f ) . t h e c o r e d n a sequencing facility at uaf was funded by the national science foundation. we thank all the biologists that submitted tissue samples for this research. the manuscript benefited from reviews by m. a. cronin and an anonymous reviewer. references anderson, s., m. s. de bruijn, a. r. coulson, i. c. eperon, f. sanger, and i. g. young. 1982. complete sequence of the bovine mitochondrial genome. journal of molecular biology 156:683717. avise, j. c., j. arnold, r. m. ball, jr., e. bermingham, t. lamb, j. e. neigel, c. a. reeb, and n. c. saunders. 1987. intraspecific phylogeography: the mitochondrial dna bridge between population genetics and systematics. annual review of ecology and systematics 18:489-522. baccus, r., n. ryman, m. h. smith, c. reuterwall, and d. cameron. 1983. genetic variability and differentiation in large grazing mammals. journal of mammalogy 64:109-120. boeskorov, g. g. 1996. karyotype of moose (alces alces l.) from northeastern asia. proceedings of the russian academy of sciences 329:506508. (in russian). . 1997. chromosomal differences in moose. genetika 33:974-978. (in russian). cronin, m. a. 1992. intraspecific variation in mitochondrial dna of north american cervids. journal of mammalogy 73:70-82. demboski, j. r., b. k. jacobsen, and j. a. cook. 1998. implications of cytochrome-b sequence variation for biogeography and conservation of the northe r n f l y i n g s q u i r r e l ( g l a u c o m y s sabrinus) of the alexander archipelago. canadian journal of zoology 76:1771-1777. duvall, a. c., and g. s. schoonveld. 1988. colorado moose: reintroduction and management. alces 24:188-194. felsenstein, j. 1985. confidence limits on phylogenies: an approach using the bootstrap. evolution 39:783-791. flerov, k. k. 1952. fauna of ussr: mammals. volume i number. 2. musk deer and deer. the academy of sciences of the ussr, leningrad. (english translation by the israel program for scientific translation, 1960, s. monson, jerusalem). geist, v. 1998. deer of the world: their evolution, behavior, and ecology. stackpole books, mechanicsburg, pennsylvania, usa. groves, c. p., and p. grubb. 1987. relationships of living deer. pages 3-59 in c. m. wemmer, editor. biology and m a n a g e m e n t o f t h e c e r v i d a e . smithsonian institution press, washington, d.c., usa. guthrie, r. d. 1966. the extinct wapiti from alaska and yukon territory. canadian journal of zoology 44:47-57. . 1990. new dates in alaskan quaternary moose, cervalces-alces — alces vol. 38, 2002 hundertmark et al. cytb variation in moose 121 archaeological, evolutionary and ecological implications. current research in the pleistocene 7:111-112. higgins, d.g., a. j. bleasby, and r. fuchs. 1992. clustal v: improved software for multiple sequence alignment. c o m p u t e r a p p l i c a t i o n s i n t h e biosciences 8:189-191. hillis, d. m., and c. moritz. 1990. an overview of applications of molecular systematics. pages 502-515 in d. m. hillis and c. moritz, editors. molecular systematics. sinauer, sunderland, massachusetts, usa. hundertmark, k. j., p. e. johns, and m. h. smith. 1992. genetic diversity of moose from the kenai peninsula, alaska. alces 28:15-20. , g. f. shields, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. sc h w a r t z. 2002. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. molecular phylogenetics and evolution 22:375-387. irwin, d. m., t. d. kocher, and a. c. wilson. 1991. evolution of the cytochrome-b gene in mammals. journal of molecular evolution 32:128-144. kimura, m. 1980. a simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. journal of molecular evolution 16:111-120. kumar, s., k. tamura, i. b. jakobsen, and m. nei. 2001. mega2: molecular evolutionary genetic analysis software. bioinformatics 17:1244-1245. kuwayama, r., and t. ozawa. 2000. phylogenetic relationships among european red deer, wapiti, and sika deer inferred from mitochondrial dna sequences. molecular phylogenetics and evolution 15:115-123. lister, a. m. 1993. evolution of mammoths and moose: the holarctic perspective. pages 178-204 in a. d. barnosky, editor. quaternary mammals of north america. cambridge university press, cambridge, u. k. lopez, j. v., m. culver, j. c. stephens, w. e. johnson, and s. j. o’brien. 1997. rates of nuclear and cytoplasmic mitochondrial dna sequence divergence in mammals. molecular biology and evolution 14:277-286. mikko, s., and l. andersson. 1995. low major histocompatibility complex class ii diversity in european and north american moose. proceedings of the national academy of sciences of the usa 92:4259-4263. nei, m., and s. kumar. 2000. molecular evolution and phylogenetics. oxford university press, new york, new york, usa. , and w.-h. li. 1979. mathematical models for studying genetic variat i o n i n t e r m s o f r e s t r i c t i o n endonucleases. proceedings of the national academy of sciences of the usa 76:5269-5273. randi, e., n. mucci, m. pierpaoli, and e. douzery. 1998. new phylogenetic perspectives on the cervidae (artiodactyla) are provided by the mitochondrial cytochrome-b gene. proceedings of the royal society of london, series b 265:793-801. ryman, n., c. reuterwall, k. nygren, and t. nygren. 1980. genetic variation and differentiation in scandinavian moose (alces alces): are large mammals monomorphic? evolution 34:10371049. saitou, n., and m. nei. 1987. the neighborjoining method: a new method for reconstructing phylogenetic trees. molecular biology and evolution 4:406425. schneider, s., j.-m. kueffer, d. roessli, and l. excoffier. 2000. arlequin ver cytb variation in moose hundertmark et al. alces vol. 38, 2002 122 2.000: a software for population genetic analysis. genetics and biometry laboratory, university of geneva, switzerland. smith, m. f., and j. l. patton. 1993. the diversification of south american murid rodents: evidence from mitochondrial dna sequence data of the akodontine tribe. biological journal of the linnean society 50:149-177. talbot, s. l., and g. f. shields. 1996. phylogeography of brown bears (ursus arctos) of alaska and paraphyly within the ursidae. molecular phylogenetics and evolution 5:477-494. udina, i. g., a. a. danilkin, and g. g. boeskorov. 2002. genetic diversity of moose (alces alces l.) in eurasia. genetika 38:951-957. (in russian). walsh, p.s., d. a. metzger, and r. higuchi. 1991. research report: chelex 100 as a medium for simple extraction of dna for pcr-based typing from forensic material. biotechniques 10:506-513. wilhelmson, m., r. k. juneja, and s. bengtsson. 1978. lack of polymorphism in certain blood proteins and enzymes of european and canadian moose (alces alces). naturaliste canadien 105:445-449. alces 56, 2020 a journal devoted to the biology and management of moose chief editor peter j. pekins university of new hampshire submissions editor roy v. rea university of northern british columbia business editor arthur r. rodgers ontario ministry of natural resources associate editors edward m. addison ecolink science eric bergman colorado parks and wildlife vince f. j. crichton manitoba conservation michelle carstensen minnesota department of natural resources nick decesare montana fish, wildlife, and parks murray w. lankester lakehead university (retired) brian e. mclaren lakehead university steve windels voyageurs national park printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 44_front_cover v2.pdf 141 distinguished moose biologist award criteria an award was established by the north american moose conference and workshop in 1981 to honour, and bring to the public's attention, the outstanding contribution of a particular individual, individuals, and/or organizations to moose management. criteria guidelines for nominating individuals are as follows: 1. published papers on moose in a variety of refereed journals, (> alces), department documents published in their jurisdiction, and articles in popular outdoor magazines and periodicals. 2. involvement and participation in the north american moose conference and workshop. (i.e., hosting a conference, participating in workshops, committees, raising innovative ideas, and donations of time and items to auctions). 3. editing and reviewing papers submitted to alces a major consideration. 4. field experience as a manager and or researcher who has demonstrated an understanding of field management and research on moose. 5. administrative experience and attainment of a level of responsibility in overall resource management with particular emphasis on moose. academic experience in attaining a level of education and subsequent sharing of this knowledge with the public, other peers, and administrators. 7. time dedicated to moose management (i.e., # years involved). 8. personal character. a subjective index based on an individual's interaction with his/her peers and others. a person having the broadest involvement regarding the listed criteria would best qualify. criteria #1 and 2 should out-weigh those of #3-8 by a ratio of about 2:1. nominations nominations can be submitted by anyone before march 15th each year prior to the annual north american moose conference and workshop. persons submitting nominations must show (in writing) how their candidate meets the criteria and why they believe the candidate deserves the award. upon receipt of one or more nominations, a selection committee consisting of all former recipients of the distinguished moose biologist award, who choose to participate in the selection process, will review submissions and reach a decision based on a simple majority of those voting. the award when presented will be announced at the annual north american moose conference and workshop. the recipient of the dmb award is expected to attend the subsequent north american moose conference and workshop and make a special "distinguished moose biologist" presentation at the end of the conference. the local organizing committee and alces will contribute to the travel costs of the recipient (see alces website, http: //bolt.lakeheadu.ca/~alceswww/alces.html, for details). the distinguished moose biologist award should not be considered an annual award. nominations should be forwarded to: dr. arthur r. rodgers centre for northern forest ecosystem research ontario ministry of natural resources 955 oliver road thunder bay, ontario canada p7b 5e1 e-mail: art.rodgers@ontario.ca alces 45, 2009 a journal devoted to the biology and management of moose edward m. addison ecolink science vince f. j. crichton manitoba conservation murray w. lankester lakehead university (retired) brian e. mclaren lakehead university printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 kristine m. rines new hampshire fish and game edmund s. telfer canadian wildlife service richard m. p. ward yukon department of renewable resources associate editors chief editor peter j. pekins university of new hampshire submissions editor gerald w. redmond maritime college of forest technology business editor arthur r. rodgers ontario ministry of natural resources alces37(2)_245.pdf alces37(2)_329.pdf instructions for contributors to alces be used in the text for scientific names and statistical symbols. use the name-and-year system to cite published literature. cite references chronologically in the text. references – use large and small capitals for author’s last names and initials. do not use any abbreviations in the references. tables present each table on a separate page. prepare tables in the same font and font size as used in the text. titles and all parts of tables must be typed doublespaced. tables must be constructed to fit the width of the page (21.5 cm), leaving 2.5-cm margins on all sides (i.e., 16.5 cm wide). table titles must be concise. footnotes should be used to reduce the complexity of table titles and provide further details. use numerical superscripts to identify footnotes or asterisks for probabilities. use horizontal lines only to delineate the top and bottom of the table and to separate column headings from the body of the table. no vertical lines should be present in a table. table columns must be generated with tab settings or a table editor. do not use spaces (i.e., the space bar). illustrations type figure captions on a separate page. identify each illustration by printing the author’s name and the figure number on the back in soft pencil. if necessary, also indicate the orientation of the illustration on the back. each illustration (either a photograph or linedrawn figure), must be of professional graphics quality, and reduced to fit into the area of either 1 (67 mm) or 2 (138 mm) columns of text by the author(s). letters and numbers on reduced figures must remain legible and be no less than 1.5 mm high after reduction. the same size and font of lettering should be used for all figures in the manuscript. photographs must be of high contrast and printed with a matte finish. typed labels are not acceptable. the minimum resolution of electronically scanned images is 600 dpi. after revision, authors should provide the original electronic graphics files or bitmap images (preferably as tagged image file format files) in an ibm-compatible format on 9-cm (3.5-inch) diskette or cd-rom. please submit manuscripts online at: http://alcesjournal.org if a problem is encountered, please contact: roy rea, submissions editor natural resources and environmental studies institute university of northern british columbia 3333 university way prince george, british columbia canada v2n 4z9 e-mail: reav@unbc.ca telephone: (250) 960 5833 editorial policy alces invites original manuscripts describing studies of the biology and management of moose throughout their circumpolar distribution, as well as other ungulate or carnivore species that overlap their range. some manuscripts published in alces originate as papers presented at the annual north american moose conference and workshop, but works may be submitted directly to the editors at any time. reviewers judge submitted manuscripts on data originality, ideas, analyses, interpretation, accuracy, conciseness, clarity, appropriate subject matter, and on their contribution to existing knowledge. page charges current policies and charges are explained in a covering letter and invoice sent to authors with galley proofs. manuscript preparation authors should follow “manuscript guidelines for contributors to alces”, by rodgers et al. appearing in alces, vol. 34 (1): 1998 (available from the co-editors and associate editors). updates are posted on the alces web page; http://alcesjournal.org/publicdocs/manuscriptguidelines.pdf. copy – please provide an electronic copy of the manuscript in ms word to the submissions editor. this copy should maintain 2.5-cm (1-inch) margins on all pages, including tables and illustrations. double-space and leftjustify all text. except for the first page, number all pages consecutively, including tables and figure captions. revisions should be handled similarly. corresponding author do not use a title page. type the date (changed with each revision), corresponding author’s name, address, telephone, and fax numbers, singlespaced in the upper left corner of the first page. if available, the author’s electronic mail address should be provided. title – type the running head (<45 characters, including spaces) on a single line following the corresponding author information. the title (<10 words) begins left justified on the next line. type the title in upper-case bold letters. do not use abbreviations or scientific names in the title. abstract & key words following the name(s) and address(es) of the author(s), provide a one-paragraph abstract. do not use abbreviations or literature citations. type alces vol. 00: 000 000 (0000), right justified on the line following the abstract. after leaving a single blank line, provide 6-12 key words in alphabetical order. footnotes use only in tables and at the bottom of the first page to provide the present address of an author when it differs from the address at the time of the study. style accompany the first mention of a common name with its scientific name. do not use scientific names for the names of domesticated animals or cultivated plants. use système international d’unités (si) units and symbols. use digits for numbers unless the number is the first word of a sentence, in which case it is spelled out. italics should only 150 distinguished moose biologist past recipients 1991 charles c. schwartz, alaska dept. of fish and game, sol� dotna, alaska. 1990 rolf peterson, michigan �echnological university, houghton, michigan. 1989 warren b. ballard, alaska dept. of fish and game, nome, alaska. 1988 vince f. j. crichton, manitoba dept. of natural resources, winnipeg manitoba. and michel crête, ministère du loisir, de la �hasse et de la péche, service de la faune terrestre, québec, pq. 1987 w. c. (bill) gasaway, alaska dept. of fish and game, fair� banks, alaska. 1986 h. r. (tim) timmermann, ontario ministry of natural resources, �hunder bay, ontario. 1985 ralph ritcey, fish and wildlife branch, kamloops, british �olumbia. 1984 edmund telfer, �anadian wildlife service, edmonton, alberta. 1983 albert w. franzmann, alaska division of fish and game, soldotna, alaska. 1982 a. (tony) bubenik, ontario ministry of natural resources, maple, ontario. 1981 patrick d. karns, minnesota division of fish and wild� life, grand rapids, minnesota. and al elsey, ontario ministry of natural resources, �hunder bay, ontario. in 1974, prior to the establishment of the distinguished moose biologist award, the group recognized the pioneering moose research of the late laurits (larry) krefting, u.s. fish and wildlife service, with an individual award. 2007 kris j. hundertmark, university of alaska fairbanks, fairbanks, alaska. 2006 kristine m. rines, new hampshire fish and game department, new hampton, new hampshire. 2005 w. m. (bill) samuel, university of alberta, edmonton, alberta. 2004 w. eugene mercer, wildlife division, st. john's, newfoundland. 2003 arthur r. rodgers, ontario ministry of natural resources, �hunder bay, ontario. 2002 bernt-erik sæther, norwegian university of science and �echnology, �rondheim, norway. 2001 r. terry bowyer, university of alaska, fairbanks, alaska. 2000 gerry m. lynch, alberta environmental protection, edmonton, alberta. 1999 william j. peterson, minnesota department of natural re� sources, grand marais, minnesota. 1998 peter a. jordan, university of minnesota, st. paul, minnesota. 1997 margareta stéen, swedish university of agricultural sci� ences, uppsala, sweden. 1996 vic van ballenberghe, u.s. forest service, anchorage, alaska. 1995 not presented 1994 james m. peek, university of idaho, moscow, idaho. 1993 murray w. lankester, lakehead university, �hunder bay, ontario. 1992 not presented alces37(1)_147.pdf 139 editorial review committee our thanks to the following individuals who served as referees for alces volume 48. each paper was reviewed by at least 2 referees who judged its appropriateness for publication and provided editorial assistance. edward addison, ecolink science, aurora, on brad allison, ontario ministry natural resources, thunder bay, on cedric alexander, vermont fish and wildlife, st. johnsbury, vt john ball, swedish university of agricultural sciences, umea, sweden mark boyce, university of alberta, edmonton, ab james bridgland, parks canada, ingonish beach, ns hughie broders, saint mary’s university, halifax, ns vince crichton, manitoba conservation (retired), winnipeg, mb william faber, central lakes college, brainerd, mn annika felton, swedish agricultural university, alnarp, sweden james goltz, provincial veterinary laboratory, fredericton, nb alison hester, james hutton institute, aberdeen, scotland. murray lankester, lakehead university (retired), thunder bay, on gerry lynch, alberta environmental protection, edmonton, ab james maskey, jr., university of north dakota, grand forks, nd scott mcburney, university of prince edward island, charlottetown, pe peter pekins, university of new hampshire, durham, nh mark pulsifer, nova scotia department of natural resources, antigonish county, ns kris rines, nh fish and game department, new hampton, nh lisa shipley, washington state university, pullman, wa steve windels, voyageurs national park, international falls, mn alces vol. 34 (1), (1998) i jon lykke died of cancer at home in verdal, norway in the company of his family on february 1, 2008. jon was a pillar of the moose community in norway and will be dearly missed. he delivered a paper 40 years ago at the 5th north american moose conference in alaska, and subsequently attended most of the international moose symposiums where he became a lifelong friend and colleague of many moose biologists worldwide. jon completed his graduate studies in forestry at the norwegian university of life sciences and also studied internationally at the university of british columbia. he published numerous papers about moose and forest management in norway, provided countless lectures about moose and forest management in norway, and was an important contributor to alces. he was also the chief editor of two books: “moose and moose hunting in norway” (1986) and “the forest” (1988). his recent paper in 2005 was a historical compilation and analysis of moose in memoriam jon lykke 1935 – 2008 management techniques and harvest data that included data collected by both he and his father leif. they were the managing directors and foresters of værdalsbruket from 1931-2002, jon the final 31 years. værdalsbruket is one of the largest private properties in norway consisting of 900 km2 of forests, mountains, marshland, rivers, and lakes. it is unlikely that any comparable long-term records of integrated moose-forest management exist. to produce a culminating paper after retirement is tremendous testimony to his passion for moose and forest management in norway. jon’s constant involvement with the international moose community was both admirable and sincerely respected, and his passion and generosity toward his colleagues make him a great role model for international cooperative efforts in moose research and management. alces provides this tribute in an international spirit of cooperation, in recognition and sincere appreciation of our honored colleague jon lykke. 152 editorial review committee our thanks to the following individuals who served as referees for al�es volume 45. each paper was reviewed by at least 2 referees who judged its appropriateness for publica� tion and provided editorial assistance. edward addison ecolink science, aurora, on �edric alexander vermont fish and wildlife, st. johnsbury, v� warren ballard �exas �ech university, lubbock, �x karen beazley dalhousie university, halifax, ns james bridgland parks �anada, ingonish beach, ns hughie broders saint mary's university, halifax, ns andy edwards 1854 �reaty authority, duluth, mn william faber �entral lakes �ollege, brainerd, mn shawn haskell vermont fish and wildlife, st. johnsbury, v� mary hindelang michigan �echnological university, houghton, mi dexter hodder university of northern british �olumbia, prince george, b� steve kilpatrick wy game and fish, jackson, wy murray lankester lakehead university (retired), �hunder bay, on gerry lynch alberta environmentalprotection (retired), edmonton, ab alice mc�ulley university of northern british �olumbia, b� peter pekins university of new hampshire, durham, nh bill peterson minnesota department of natural resources (retired), grand marais, mn roy rea university of northern british �olumbia, prince george, b� william samuel university of alberta (retired), edmonton, ab david scarpitti mass. division of fisheries and wildlife, westborough, ma peggy smith lakehead university, �hunder bay, on donald �homas �anadian wildlife service (retired), edmonton, ab steven weber new hampshire fish and game department, �oncord, nh don young alaska department of fish and game, ak alces37(1)_201.pdf 133 previous meeting sites of the north american moose conference and workshop and international moose symposia 1963 st. paul, minnesota 1964 st. paul, minnesota 1966 winnipeg, manitoba 1967 edmonton, alberta 1968 kenai, alaska 1970 kamloops, british columbia 1971 saskatoon, saskatchewan 8th 1972 thunder bay, ontario 9th 1973 québec city, québec, in conjunction with the 1st international moose symposium 10th 1974 duluth, minnesota 11th 1975 winnipeg, manitoba 12th 1976 st. john’s, newfoundland 13th 1977 jasper, alberta 14th 1978 halifax, nova scotia 15th 1979 soldotna kenai, alaska 16th 1980 prince albert, saskatchewan 17th 1981 thunder bay, ontario 18th 1982 whitehorse, yukon territory 19th 1983 prince george, british columbia 20th 1984 québec city, québec 1984 uppsala, sweden, 2nd international moose symposium 21st 1985 jackson hole, wyoming 22nd 1986 fredericton, new brunswick 23rd 1987 duluth, minnesota 24th 1988 winnipeg, manitoba 25th 1989 st. john’s, newfoundland 26th 1990 regina and ft. qu’apelle, saskatchewan 1990 syktyvkar, russia, 3rd international moose symposium 27th 1991 anchorage and denali national park, alaska 28th 1992 algonquin park, ontario 29th 1993 bretton woods, new hampshire 30th 1994 idaho falls, idaho 31st 1995 fundy national park, new brunswick 32nd 1996 banff national park, alberta 33rd 1997 fairbanks, alaska, in conjunction with the 4th international moose symposium 34th 1998 québec city, québec 35th 1999 grand portage, minnesota 36th 2000 whitehorse, yukon territory 37th 2001 carrabassett valley, maine 38th 2002 hafjell, norway, in conjunction with the 5th international moose symposium 134 39th 2003 jackson hole, wyoming 40th 2004 corner brook, newfoundland and labrador 41st 2005 whitefish, montana 42nd 2006 baddeck, nova scotia 43rd 2007 prince george, british columbia 2008 yakutsk, russia, 6th international moose symposium 44th 2009 pocatello, idaho 45th 2010 international falls, minnesota 46th 2011 jackson hole, wyoming future meetings 2012 bialowieza, poland, 7th international moose symposium 47th 2013 new hampshire alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 125 how moose select forested habitat in gros morne national park, newfoundland brian. e. mclaren1, s. taylor2 and s. h. luke1 1lakehead university, faculty of forestry and the forest environment, 955 oliver road, thunder bay, on, canada p7b 5e1; 2gros morne national park, p.o. box 130, rocky harbour, nl, canada a0k 4n0. abstract: current parks canada policy does not allow moose (alces alces) to be hunted in national parks in newfoundland and labrador; combined with the extirpation of wolves (canis lupus), this policy creates a situation where introduced moose (a. a. americana) are relatively predator-free in gros morne national park. forested areas of this park are frequently disturbed by defoliating insects resulting in extensive young conifer forest; increasingly, more areas are identified as failing to regenerate to normal tree densities or “not sufficiently restocked” (nsr). we used data from gps-collared moose that occupy areas of the park where limited timber cutting is allowed for domestic purposes and a very detailed and current forest inventory exists; such areas are still dominated by insect and wind disturbance, including a large designation of nsr forest. we hoped to determine whether moose are found preferentially in disturbed forest versus other landscape patches during summer or winter, during day or night, and under certain temperature conditions. variability in habitat availability and habitat use by moose appears to preclude forest management options directed at specific habitat types. alces vol. 45: 125-135 (2009) key words: alces alces, absence of predators, gros morne national park, moose, newfoundland, overabundance, population dynamics, resource selection function. when moose (alces alces) face fewer predators they can reach higher densities (peterson et al. 2003). moose (a. a. americana) were first introduced to central newfoundland, canada in 1878 with the release of a male and female from nova scotia (pimlott 1953). a second release of 2 males and 2 females from new brunswick into western newfoundland followed in 1904. gray wolves (canis lupus), their only potential predator, were extirpated in 1932 (pimlott 1959). therefore, for most of their occupation of the island province, newfoundland moose were preyed on only by human hunters, with black bear (ursus americana) predating only calves. consequently, their density averages about tenfold higher than in other parts of their range in north america (crête and daigle 1999). newfoundland moose experience large population fluctuations even where heavily hunted in central areas of the island, an effect that the limited functional response of human predation and delayed density dependence in hunter kill have been implicated (ferguson and messier 1996). relationships with their food source during extreme population peaks were first described by bergerud and manuel (1968); declines in population following such peaks identified by mercer and mclaren (2002) suggest that food availability is often the only limiting factor for moose in newfoundland. moose are not hunted in the national parks in newfoundland, creating a situation where populations are predator-free in 1,805 km2 of western newfoundland (gros morne national park) and 404 km2 of eastern newfoundland (terra nova national park). the resulting extreme moose demographics in both parks (mclaren et al. 2000), the fact that moose are not native to newfoundland, and the potential for their high densities to alter natural ecosystem processes (mclaren habitat selection in gros morne mclaren et al. alces vol. 45, 2009 126 et al. 2004), add up to policy and management challenges for parks canada (corbett 1995). in canada, the response to hyperabundant species in protecting the ecological integrity within and by means of national parks follows a definition that very explicitly requires relating the reasons for hyperabundance to human impacts (pca 2000). examples of acceptable reasons to control large herbivores include artificial introductions and the loss of key predator species like carnivores, both of which have occurred in newfoundland. however, to institute any control program, parks canada also demands confirmation that the reasons for hyperabundance are well understood and the program is conducted under an adaptive management framework where the original assumptions are subject to review. moreover, control is better justified when “ecological integrity” is compromised by the large herbivore; according to the canada national parks act, ecological integrity means “…a condition that is determined to be characteristic of its natural region and likely to persist, including…rates of change and supporting processes.” it is in this context that we embarked on a review of how moose exploit areas of forest disturbance in gros morne national park (gmnp). it has been hypothesized that the gros morne moose population expanded as its range expanded into new habitat created by insect, wind, and timber cutting disturbances in gmnp (connor et al. 2000). we argue that forecasts for forest regeneration and moose habitat in gmnp depend on how much the cause for high moose densities can be ascribed to forest disturbance, and how confident we are that moose generally select and occupy disturbed areas. therefore, we matched the forest inventory and disturbance database to information from a series of collars with global positioning system (gps) capability placed on adult female moose in 1997-1998 (mclaren et al. 2000). while the sample size is small (n = 4), we focus on a subset of collared moose that occupy areas of the park where timber cutting is allowed for domestic purposes (home and boat construction and fuelwood) and, therefore, a very detailed and current forest inventory exists. the sample size and inventory data quality are sufficient to determine whether moose are found preferentially in disturbed forest versus other landscape patches during summer or winter, during day or night, and under certain temperature conditions. we explored these expected differences against a null hypothesis that selection within the home range varies among individual moose to an extent that general prediction about their use of landscape patches is not possible. study area gros morne national park is situated on the gulf of st. lawrence in newfoundland, encompassing parts of the northern peninsula, long range barrens, and western newfoundland forest ecoregions (damman 1983). moose likely invaded the area now protected by gmnp in 1925 and became common by the 1950s, with modest population increases until the 1970s. surveys in the late 1970s indicated that after gmnp became established (and human hunting was excluded), moose population density increased much more rapidly (gmnp, unpublished data). increases first occurred in upland areas, but by the 1980s moose density increased throughout the park (connor et al. 2000). in a 1998 survey, approximately 8,000 animals occurred in <1,000 km2 of potential habitat (gmnp unpublished data). some survey units in 2007 and 2008 had densities >15 moose/km2 in lowland areas, where the average density remains ~ 4 moose/km2. gros morne national park includes an allowance for domestic timber cutting in 12 “cutblocks” (193 km2), but excludes 6 adjacent community enclaves (140 km2). the cutblocks used by moose in this study were located in the coastal plain sub-region of the northern peninsula ecoregion, east of the flat coastal areas of the gulf and including the western slopes of the long range mountains. at elevations alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 127 <425 m, these areas experience a cool boreal climate with a relatively long growing season of 110-150 days (damman 1983). forests consist mostly of balsam fir (abies balsamea), with some spruce (picea spp.) and a mix of pioneer (mostly betula spp.) and tolerant (mostly acer spp. and sorbus spp.) hardwoods; a more detailed description is found in connor et al. (2000). insect outbreaks, primarily of spruce budworm (choristoneura fumiferana) in 1977 and hemlock looper (lambdina fiscellaria) in 1969, 1988 and 1996 affected a large area of the forest: 7,550 ha in total (2,800 ha in the cutblocks) with individual areas of canopy disturbance up to 49 ha in extent (gmnp unpublished data). a study just east of gmnp on the northern peninsula concluded that insects caused defoliation and death of the forest canopy in >60% of the landscape, plus an additional canopy break-up and gap regeneration in >10% of the forest (mccarthy and weetman 2007). in contrast, domestic timber cutting entails manual tree removal from small patches (maximum 2.2 ha); to an extent within the higher areas of this range, ‘high-grading’ may occur where large, dominant trees are preferentially cut leaving smaller trees in place. at the local level then, cutting, wind disturbance, and insect outbreaks create very similar forest structures. the disturbance types also frequently occur in combination; for example, cutting has occurred in 340 ha of defoliated forest following insect outbreaks in the cutblocks, and windfall is frequently associated with defoliation. timber cutting, on the other hand, is a relatively minor contributor to new disturbances in gmnp, amounting to approximately 20 ha per year, equivalent to just over 1% of the total forested areas within the cutblocks since records began in the mid-1990s. methods forest inventory a new, detailed forest inventory for gmnp was completed in 1997. forest stand information was delineated using data on colour, 1:12,500 scale aerial photographs taken in 1995. this information included age, height, and crown density estimates, and the approximate species composition of each forested stand >0.5 ha; the disturbance type and year (if known) were tagged to disturbed forest stands identified on these photographs. if disturbed forest was interpreted as failing to regenerate to normal tree densities at the time of interpretation, based on absence of crown closure, stands in this category were tagged “not sufficiently restocked” (nsr). in addition, non-forest vegetation types such as barren, bog, scrub, residential areas and water were mapped. in 2004, black-and-white, 1:10,000 scale aerial photographs were acquired to update and classify more accurately the forest disturbances limited to the cutblocks. forest stands that were labelled in the 1995 inventory as disturbed, nsr, or regenerating forest were re-evaluated to determine their status. if regeneration had partly or completely failed, they were labelled (or re-labelled) nsr; stands with sufficient regeneration were labelled with the appropriate regenerating forest label from the original age of disturbance. approximately half of these stands were labelled in the “0-20 year” age category and half in the “20-40 year” category. new disturbances, such as recent cutting or insect outbreaks (~ 1996-2004) were also delineated in 2004. moose locations we used location data collected from 4 adult, female moose collared and monitored 26 june 1997-13 october 1998. collars were set to record differentially correctable gps locations every third hour (lotek engineering inc., newmarket, ontario, canada). details of the collaring and of accuracy testing of the location data are found in mclaren et al. (2000). only locations with 3d accuracy and differential correction were used in this study; these locations had a 95% accuracy habitat selection in gros morne mclaren et al. alces vol. 45, 2009 128 of <25 m. a night location dataset was created approximately equal in size to a daytime location dataset for each moose by first choosing from the 3d locations recorded closest to 0300 and 1500 hr each monitoring day. the night location datasets were supplemented in 13-21% of cases with locations recorded at approximately 0000 hr when no 3d location occurred at 0300 hr, and the daytime datasets were supplemented similarly in 18-25% of cases with locations recorded at approximately 1200 hr. the resulting datasets covered 9197% of possible night locations and 96-99% of possible day locations during a 351-402 day monitoring period, depending on the moose. the monitoring periods were further divided into summer and winter seasons following a method developed by vander wal (2005). for each moose, cumulative distance moved was calculated in arcview version 9 (esri, redlands, california) beginning with the first 3d location on 1 january 1998, and ending with either 26 june 1998 (2 cases) or 13 october 1998 (2 cases). the time series were completed with data beginning on 26 june 1997 or 13 october 1997 and ending 31 december 1997. winter was defined as the period when slope of cumulative distance over cumulative time was less than the annual mean slope; summer was when movement within the home range was faster. each time series was closest matched to a logistic curve, using regression with statistical software (spss, version 16). summer fell between 8-28 april and 17 september-18 october, ranging from 158-186 days in length, depending on the moose. for 3 of the 4 moose, 2 nearly complete summer seasons (1997 and 1998) could be quantified in the gps database. the fourth moose was sufficiently monitored only during the 1998 summer season. data analysis summer (1997 and 1998) and winter (1997-1998) home ranges were calculated using 100% minimum convex polygons. size of seasonal ranges was compared across all 4 moose by t-test. areas outside the cutblocks and large areas of open water were then excluded from each of the seasonal ranges. the remaining area was divided into a) young nsr forest ≤ 20 years since disturbance, b) older nsr forest, c) young (regenerating) conifer forest ≤ 20 years since disturbance, d) older conifer forest, e) mixed forest and deciduous forest of all ages, f) scrub forest, and g) barren (non-forested) areas. the first 2 categories were defined according to information in both the original forest inventory (1995) and the update (2004); disturbances that were labelled nsr in either database (or both databases) were re-categorized as “young” or “older” nsr forest. recognizing that the moose locations were recorded during 1997-1998, we classed “young” nsr based either on a date in the inventory indicating ≤20 years since disturbance by 1997, or on a forest age class in the 1995 inventory of “0-20 years.” areas of nsr first identified in the 2004 update were counted as nsr habitat only if they were classed as “0-20 years” of age (or older) at that time, because most insect outbreaks occurred before 1997; new areas of nsr in the update were not counted as nsr areas in the habitat analysis if they were assigned to a date after 1998. the reason 2 of the habitat categories refer to “young” forest is that forage for moose in newfoundland has long been estimated to be highest in regenerating balsam fir forest at 8-10 years of age (parker and morton 1978), where the amount of browsing by moose increases with the fraction of balsam fir among trees <3 m in height (thompson 1988). thus, “young” forest is a special and desired habitat for foraging by moose, while older forest may serve largely as cover, but not foraging habitat. habitat selection was tested for all 7 of the categories in 4 separate resource selection functions (rsf), one for each moose. rsfs are statistical tools that describe the relative alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 129 probability of occurrence of animals based on their response to their habitat (boyce 2006). in this study, each observed moose location was linked to 10 random locations within a 700-m radius circle, an area that encompassed approximately 50% of the distances between 2 successive locations during any season for all moose. thus, the observed moose locations were considered “selected” among random areas within 700 m, a distance chosen to assume that moose selected the habitat class recorded by the gps collar as their location among random locations to which they would have been capable of moving, but were not found. in other words, locations are tracked relative to the patches of habitat immediately available to a moose at any given time. the rsf typically distinguishes among observed locations and available habitats using log-linear modelling, for this study constructed in spss (version 16). comparisons among resulting rsf models were made using a combination of the significance of change in deviance in an analysis of deviance table (manley et al. 2002) and the akaike’s information criterion (aic, burnham and anderson 2002). relative habitat selection was calculated, setting the probability of selection of young nsr forest at 1.000. reporting relative rather than absolute habitat selection probabilities follows the recommendation of arthur et al. (1996), for situations when availability of habitats is not constant and comparisons among habitats may be affected by the choice of which habitats to include. variation in habitat selection among individual moose was compared to 1) variation in habitat selection between summer and winter and between day and night in each season, 2) variation in habitat selection when the collar recorded cold (≤0º c) versus warm (>0º c) winter temperatures, 3) variation in habitat selection when the collar recorded cool (≤8º c) versus warm (>8º c) summer temperatures, and 4) variation in habitat selection between summers (1997 and 1998) only when warm (>8º c) collar temperatures were recorded (limited by data availability). the temperature thresholds were chosen as close as possible to the temperatures at which heat stress may begin for moose, −5º c in winter and 15º c in summer (renecker and hudson 1986), while also dividing the 3d locations approximately evenly among cold, warm, and cool categories. young nsr forest was considered the best choice for a standard against which to compare habitat selection in the other 6 categories, both because it serves as a theoretical target for foraging by moose, and because it provides the highest actual likelihood among the other young forest categories that it was visited by moose due to its nsr designation and the likelihood that regeneration failure is linked to moose overabundance. results seasonal range sizes varied from 1351,692 ha (table 1), and were larger (p = 0.002) in winter (1,200 ± 257 ha, n = 4) than in summer (419 ± 101 ha, n = 7). range composition also varied considerably among individual moose (table 1); 3-86% of a seasonal range occurred outside a cutblock. in 3 of 4 cases, the proportion of range outside cutblocks was greater in winter than in summer. for moose 15 and 16, old and young nsr forest made up more than one-third of all range area within the cutblocks, whereas scrub forest filled a dominant proportion of every range for moose 22 and 25. by contrast, non-forested (barren) areas comprised anywhere from 1-33%, mixed and deciduous forest 4-33%, and older conifer forest 2-32% of ranges. habitat selection was significant and varied significantly for all tested factors (table 2). variation in habitat selection among individual moose was significant and larger than variation in habitat selection by season, by temperature, by time of day, or by year for the summer season. habitat selection in gros morne mclaren et al. alces vol. 45, 2009 130 ta bl e 1. c ha ra ct er is tic s of s ea so na l r an ge s fo r a du lt fe m al e m oo se c ap tu re d (m oo se 1 5, 1 6, 2 2 an d 25 ) i n g ro s m or ne n at io na l p ar k (g m n p) . m on ito ri ng o cc ur re d fr om co lla rs s et to r ec or d g lo ba l p os iti on in g sy st em ( g ps ) lo ca tio ns e ve ry th ir d ho ur ; r ec or ds b eg an o n 26 j un e 19 97 a nd w er e ce ns or ed a t t he e nd o f th e se co nd s um m er , 26 s ep te m be r 19 98 . r an ge s iz e an d co m po si tio n w er e ca lc ul at ed f ro m m in im um c on ve x po ly go ns ( m c ps ) en co m pa ss in g 10 0% o f 3d d if fe re nt ia lly c or re ct ed lo ca tio ns . s am pl e si ze s in s ub se qu en t a na ly se s ar e sh ow n as th e nu m be r o f l oc at io ns , d iv id ed in to c at eg or ie s of d ay a nd n ig ht , a nd c oo l o r c ol d v er su s w ar m -t em pe ra tu re lo ca tio ns . a d as he d lin e in di ca te s a ce ns or ed c at eg or y in c om pa ra tiv e an al ys es d ue to lo w s am pl e si ze . alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 131 ta bl e 2. a na ly si s of d ev ia nc e ta bl e fo r m od el s of h ab ita t u se . n is th e nu m be r o f c ol la re d, a du lt fe m al e m oo se u si ng c ut bl oc ks . r es ou rc e se le ct io n fu nc tio ns (r sf s) w er e co ns tr uc te d fr om o ne lo glin ea r m od el f or e ac h m oo se f ro m n lo ca tio ns a nd 1 0n a dd iti on al r an do m lo ca tio ns . m od el s ig ni fic an ce w as c al cu la te d by c om pa ri so n of th e di ff er en ce in d ev ia nc e (∆ d ) be tw ee n su cc es si ve m od el s w ith th e χ2 d is tr ib ut io n fo r df d eg re es o f fr ee do m . a ka ik e' s in fo rm at io n c ri te ri a (a ic s) a re s ho w n fo r ea ch m od el s te p. habitat selection in gros morne mclaren et al. alces vol. 45, 2009 132 young nsr forest was the most selected among all habitat types, although less selected by moose 22 (table 3). among winter, daytime locations for moose 15, young nsr forest was selected more than any other habitat type. the same pattern, although not significant, held at all times except on summer nights when scrub forest was slightly, but not significantly, more selected than young nsr forest. for moose 25 the same pattern held for winter day-time locations, with young nsr forest selected more often than 4 of the 6 other habitat types. in contrast, moose 22 selected only barren areas significantly less than young nsr forest, while mixed forest and scrub forest were generally more selected than young nsr forest. moose 25 had the largest shift in selection between winter and summer; both older conifer forest and scrub forest were less selected than young nsr forest on winter days, but more selected on summer days. in contrast to young nsr forest, older nsr forest did not appear to be selected. there was a significant difference in winter habitat selection during cold versus warm periods for all moose. conclusions about these differences were difficult, due to the higher variation in selection among individual moose (table 2). for 3 moose (16, 22, and 25), mixed and deciduous forest was either the most selected habitat type or the second most selected throughout winter, next to young nsf forest; however, moose 15 selected significantly less mixed and deciduous forest and significantly more older coniferous forest in colder periods during winter, relative to young nsf forest (table 4). on the other hand, coniferous forest was significantly less selected in colder periods by moose 22 and moose 25, relative to young nsr forest. during warmer periods in winter, selection among moose was more constant when conifer forest was less selected than young nsr forest. there was a smaller, but significant difference in summer habitat selection during cool versus warm periods for the 3 moose with sufficient records for this comparison. as in winter, moose 15 selected significantly less mixed and deciduous forest than young nsr forest during the cooler period; the same pattern held for moose 25 for the summer period. moose 25 selected scrub forest significantly more than nsr forest on warm summer days, but significantly less scrub forest was selected on cool summer days. table 3. relative selection probabilities when comparing habitat use by season and time of day. relative selection was assessed using 1.000 as the probability of selection for young (≤20 year-old) nsr forest. probabilities were derived from parameters in rsfs constructed for individual moose for which habitat selection varied significantly (table 2a). resource selection varied for the other factors, season and time of day, for moose 15 and 22, but not for moose 25. for these conclusions, ∆ d is compared to the χ2 distribution for 6 degrees of freedom. asterisks indicate cases where probabilities are derived from significant parameter estimates (p <0.05). alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 133 the difference in summer habitat selection between 1997 and 1998 was significant for only 1 (moose 16) of the 3 moose with a long enough monitoring record to allow comparison. in this case, the least selected habitat in 1997 was mixed and deciduous forest, while in 1998 it was either older conifer forest or barren areas. except for scrub forest on warm summer days, young nsr forest was the most selected habitat type throughout the summer in both years, while older nsr forest was among the least selected habitat types. discussion for gmnp we conclude that selection within the home range varies among individual moose to an extent that general prediction about use of landscape patches by even 4 moose, let alone the population, is not possible. this conclusion is not different from that made recently by gillingham and parker (2008) for northern british columbia, and it may be a general caution about the interpretation of rsfs, as well as guidance for managers in the management of moose we also garnered additional information from the gps-collared moose that differed from conclusions made from vhf-collared moose in the same area (mclaren et al. 2000). for example, migrations did not occur from cutblocks during the early months of summer among the 4 moose reported in this study. even though local movement rates increase in summer months, moose residing in the cutblocks may be less prone to long-distance migrations if food can be found in young forest, even in young nsr forest. if this condition describes the current situation, it likely also explains smaller summer home ranges, while ranges sufficiently larger to include older coniferous forest in colder periods, for example, were likely required during winter. the observation of higher daily movement rates in summer versus winter, the only uniform pattern among the moose in this study, is consistent with vander wal’s (2005) conclusion that winter is a period of energy conservation and limited daily movement. variation in snow conditions in the coastal plain subregion of newfoundland is probably the source for variation in habitat selection by moose in winter. except for warmer summer days, young nsr forest appears generally the most selected habitat type in the cutblocks. table 4. relative selection probabilities comparing habitat use on cold and warm days in winter (∆ d1), cool and warm days in summer (∆ d2), and for warm summer days, 2 monitoring seasons in 1997 and 1998 (∆ d3). other definitions and calculations are as in table 3. habitat selection in gros morne mclaren et al. alces vol. 45, 2009 134 however, overall high variability in habitat selection during winter and summer suggests that conclusions drawn to assist moose or moose habitat management in gmnp will not be general ones. one possibile generalization is that in 20 years, when young nsr forest converts to older nsr forest, foraging opportunities for moose may be depleted. even with modestly high densities, moose habitat in newfoundland demonstrably declines over a few decades (mercer and mclaren 2002). however, this study is unable to show that moose are food limited in gmnp. variability in habitat availability and habitat use by moose appears to preclude forest management options directed at specific habitat types. concern with high-density moose populations in national parks arises from the possibility that forests will not regenerate naturally. a study in terra nova national park detailed that even after moose are removed from the ecosystem, alteration of forest dynamics by hyperabundant moose can persist at least 2 decades (mclaren et al. 2009). only circumstantial evidence of altered forest composition in gmnp was found by connor et al. (2000). however, in extreme cases in gmnp, persistent invasive plants, including coltsfoot (tussilago farfara) and canada thistle (cirsium arvense), as well as native grasses (e.g., calamagrostis canadensis), become widespread problems as a result of moose browsing and trampling in disturbed areas (rose and hermanutz 2004). we recommend continued monitoring in gmnp to evaluate impacts by moose on forested habitats relative to forest regeneration and composition. acknowledgements we are grateful to the inland fish and wildlife division of the government of newfoundland and labrador for contributing to the effort to resolve questions of moose overabundance in gmnp. we are also very grateful to carson wentzell, gmnp resource conservation, who leads the forest inventory in the cutblocks and is responsible for the detailed aerial photo updates that allowed this study to be undertaken. the original fieldwork involving gps collars was coordinated under the direction of shane mahoney (government of newfoundland and labrador), and douglas anions and christopher mccarthy (parks canada), and with the assistance of pilots baxter slade (newfoundland helicopters) and hughie day (canadian helicopters). references arthur, s. m., b. f. j. manly, l.l. mcdonald, and g. w. garner. 1996. assessing habitat selection when availability changes. ecology 77: 215-227. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir-white birch forests in central newfoundland. journal of wildlife management 32: 729-746. boyce, m. s. 2006. scale for resource selection functions. diversity & distributions 12: 269-276. burnham, k. p., and d. r. anderson. 2002. model selection and multi-model inference: a practical information-theoretic approach. springer-verlag, new york, new york, usa. connor, k. j., w. b. ballard, t. dilworth, s. mahoney, and d. anions. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 36: 111-132. corbett, g. n. 1995. review of the history and present status of moose in the national parks of the atlantic region: management implications? alces 31: 255-268. crête, m., and c. daigle. 1999. management of indigenous north american deer at the end of the 20th century in relation to large predators and primary production. acta veterinaria hungarica 47: 1-16. damman, a. w. h. 1983. an ecological subdivision of the island of newfoundland. pages 163-205 in g. r. south (ed.) biogeography and ecology of the island alces vol. 45, 2009 mclaren et al. habitat selection in gros morne 135 of newfoundland. w. junk publishers, boston, ma, usa. ferguson, s. h., and f. messier. 1996. can human predation of moose cause population cycles? alces 32: 149-161. gillingham, m. p., and k. l. parker. 2008. the importance of individual variation in defining habitat selection by moose in northern british columbia. alces 44: 7-20. manly, b. f. j., l. l. mcdonald, d. l. thomas, t. l. mcdonald, and w. p. erickson. 2002. resource selection by animals, second edition. kluwer academic publishers, dordrecht, the netherlands. mccarthy, j. w., and g. weetman. 2007. stand structure and development of an insect-mediated boreal forest landscape. forest ecological management 241: 101-114. mclaren, b., l., hermanutz, j. gosse, b. collet, and c. kasimos. 2009. broadleaf competition interferes with balsam fir regeneration following experimental removal of moose. forest ecological management 257: 1395-1404. ______, b. a. roberts, n. djan-chékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 45-59. ______, c. mccarthy, and s. p. mahoney. 2000. extreme moose migrations in gros morne national park, newfoundland. alces 36: 217-232. mercer, w. e., and b. e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose. alces 38: 123141. parker, g. r., and l. d. morton. 1978. the estimation of winter forage and its use by moose on clearcuts in northcentral newfoundland. journal of range management 31: 300-304. pca (parks canada agency). 2000. managing hyperabundant species. pages 5-12 and 5-13 in panel of the ecological integrity of canada’s national parks. unimpaired for future generations? protecting ecological integrity with canada’s national parks. volume ii. setting a new direction for canada’s national parks. ottawa, ontario, canada. peterson, r. o., j. a. vucetich, r. e. page, and a. chouinard. 2003. temporal and spatial aspects of predator-prey dynamics. alces 39: 215-232. pimlott, d. h. 1959. reproduction and productivity of newfoundland moose. journal of wildlife management 23: 381-401. ______. 1953. newfoundland moose. transactions north american wildlife conference 18: 563-581. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditure and thermoregulatory response of moose. canadian journal of zoology 64: 322-327. rose, m., and l. hermanutz. 2004. are boreal ecosystems susceptible to alien plant invasion? evidence from protected areas. oecologia 139: 467-4776. thompson, i. d. 1988. moose damage to precommercially thinned balsam fir stands in newfoundland. alces 24: 56-61. vander wal, e. j. 2005. core areas of habitat use: the influence of spatial scale of analysis on interpreting summer habitat selection by moose (alces alces). m. sc. thesis, lakehead university, thunder bay, ontario, canada. 4012bb.p65 alces vol. 40, 2004 potvin and courtois moose in clear-cuts 61 winter presence of moose in clear-cut black spruce landscapes: related to spatial pattern or to vegetation? françois potvin and réhaume courtois ministère des ressources naturelles et de la faune, direction de la recherche sur la faune, 930, chemin ste-foy, 4e étage, québec, pq, canada g1s 2l4 abstract: winter aerial surveys of moose (alces alces) were completed on 14 landscapes (10– 256 km2 ) formed of aggregated black spruce (picea mariana) clear-cuts logged 3–9 years ago in southcentral québec. moose were present in 8 landscapes (11 yards) and had a mean density of 0.20 moose/10 km2, which was 50% of the density observed in the same hunting zone with a similar forest composition. based on previous work, effects of variability in hunting pressure and time since cutting were assumed not to influence distribution and abundance of moose. browse density did not increase with age of cuts. moose density was not related to the size of the clear-cut landscapes or the proportion of residual forest (18–40%) within each landscape (p = 0.14). moose yards were not located close to uncut forest surrounding the landscapes and did not have a greater proportion of residual forest than clear-cut landscapes. moose yards had a denser shrub layer and more browse available than random sites selected in the same landscapes. the presence of moose in large clearcut black spruce landscapes is related to vegetation characteristics and not the spatial pattern of the forest. the authors propose two strategies to maintain moose populations and moose hunting activity in this type of forest after harvesting. alces vol. 40: 61-70 (2004) key words: alces alces, black spruce, cover, food, habitat management, picea mariana résumé: nous avons réalisé l’inventaire aérien hivernal de l’orignal (alces alces) dans 14 paysages formés de grandes coupes totales agglomérées (10–256 km2) effectuées au cours des 3 à 9 dernières années au centre-sud du québec. la forêt d’origine était dominée par l’épinette noire (picea mariana). la présence de l’orignal a été confirmée dans 8 paysages (11 ravages), pour une densité moyenne de 0,20 orignal/10 km2, soit 50% de celle retrouvée dans cette zone de chasse ayant le même type de forêt. d’après un travail antérieur, nous assumons que la variation de la pression de chasse et le nombre d’années après coupe n’ont pas influencé la distribution et l’abondance de l’orignal. la taille des paysages de coupe totale ou la proportion de forêt résiduelle (18–40%) à l’intérieur de ceux-ci n’a pas influencé la densité (p = 0,14). les ravages d’orignaux n’étaient pas situés à proximité de la forêt intacte autour des paysages de coupe et ne contenaient pas davantage de forêt résiduelle que les paysages de coupe. cependant, les ravages d’orignaux avaient une strate arbustive plus dense et une disponibilité de brout plus grande que des sites aléatoires choisis dans les mêmes paysages. la présence de l’orignal dans des paysages formés de grandes coupes totales en pessière noire est davantage reliée aux caractéristiques de la végétation qu’à la configuration spatiale de la mosaïque forestière. nous proposons deux stratégies pour favoriser après coupe le maintien de l’orignal et de l’activité de chasse dans ce type de forêt. alces vol 40: 61-70 (2004) mots clés: alces alces, aménagement forestier, couvert, épinette noire, nourriture, picea mariana moose in clear-cuts – potvin and courtois alces vol. 40, 2004 62 although forest logging increases the amount of browse for moose (alces alces), large clear-cuts can have detrimental effects in the short term due to a lack of protective cover (girard and joyal 1984, courtois and beaumont 2002, courtois et al. 2002). clear-cutting, the prevalent harvesting regime in the boreal forest in canada, remains a controversial issue (bliss 2000). provinces have adopted regulations to make this technique more acceptable ecologically and socially. for example, over the past 15 years, regulations in québec have limited the maximum size of clear-cuts to 250 ha (up to 1995) or 150 ha (since 1996). uncut forest strips (60–100 m wide) are left between 2 adjacent clear-cut patches. riparian buffer strips, 20-m wide along lakes and on each side of permanent streams, are also mandatory and can be used to separate adjacent cuts, but their width must then be increased to 60 or 100 m. aggregating cutover patches, a common management strategy, has resulted in clear-cut dominated landscapes that may exceed tens and even hundreds of km2. large clear-cut landscapes are more frequent in black spruce (picea mariana) boreal forests, characterized by a more uniform forest mosaic, than in balsam fir (abies balsamea) or mixed forests. a previous study in coniferous and mixed boreal forest in south-western québec has shown that moose seldom use recent clearcuts, except in areas providing dense high regeneration and where logging is more of a partial type (courtois and beaumont 2002, courtois et al. 2002). because clear-cut landscapes are larger in black spruce forests, we suspect that the short-term impact of logging might be more severe in this type of forest than where that previous study took place. traditionally, higher moose densities have been associated with mixed stands (crête 1988) or areas disturbed by forest fires, insect outbreaks, or logging some 15-40 years ago (crête 1977, peek 1998). rarely has moose abundance been related to the pattern of the forest mosaic, except for the edge between food and cover (thompson and stewart 1998, courtois et al. 2002). the original black spruce forest being a poor habitat for moose (brassard et al. 1974, girard and joyal 1984), recent cutovers in that type of forest must consequently be of very low value for moose. in such clear-cut landscapes, the pattern of the residual forest, such as buffer strips, might play a greater role than vegetation for the persistence of moose in recent cuts. from 1998 to 2000, we surveyed moose in very large black spruce clear-cuts in central québec. we hypothesized that clear-cut landscapes constitute poor habitats for moose. to verify this hypothesis, we predicted that (1) moose would be at a lower density in recently clear-cut landscapes, especially the largest ones, than in similar uncut black spruce forests. we also hypothesized that (2) moose yards would be located close to uncut forest surrounding the clear-cut landscapes or in those parts of the landscapes having a higher proportion of residual forest, and that (3) moose yards would be located in parts of the clear-cut landscapes where the shrub layer is more dense and browse more abundant. study area aerial surveys were conducted in forest management unit 25-03 (49°02'–50°00' n, 72°31'–74°17' w), located north-west of lac saint-jean, québec (fig. 1). we selected all clear-cut landscapes 10 km2 in this unit, within the black spruce–moss ecological zone (grondin 1996). these 14 landscapes had been clear-cut between 1991 and 1997 and ranged in size between 10 and 256 km2 (fig. 2). the proportion of residual forest (uncut) within landscapes ranged between 18 and 40% of the productive area (land area minus wetlands/bogs), with a ≥ alces vol. 40, 2004 potvin and courtois moose in clear-cuts 63 fig. 1. location of the 14 aerial survey blocks for moose in forest management unit 25-03 (gray shade) in québec. each block corresponds to a clear-cut landscape. mean value of 30%. the bulk of residual forest was made up of riparian and nonriparian buffer strips, but also included noncommercial forest patches (forests too young or at low density) and commercial forests in inaccessible areas (slope 40%) in some landscapes. black spruce was the dominant tree species, with balsam fir, white birch (betula papyrifera), trembling aspen (populus tremuloides), and jack pine (pinus banksiana) being locally present. the study area is part of hunting zone 18 west, which supports a density of 0.95 moose/10 km2 (dussault 1999). in the northern portion of forest management unit 25-03 (black spruce-moss ecological zone), moose density is much lower than in the southern portion (balsam fir–white birch zone) (0.37 vs. 1.26 moose/10 km2). methods aerial surveys and habitat map aerial surveys were conducted in january of winters 1998 (1 landscape), 1999 (4), and 2000 (9). we used a bell 206-b helicopter flying at an altitude of 110 m and a speed of 160 km/hr (courtois 1991). a navigator and 2 observers were on board. north-south survey lines were spaced 600 m along longitude transects (0.5 min). a 1km distance was added to the beginning and end of each transect to delineate the survey blocks (fig. 3). moose sign was noted on 1:20,000 (1998) or 1:50,000 (1999, 2000) topographical maps showing recent clearcut patches. a spatial database was built to produce the aerial survey maps and to analyse the data. the base map was the québec 1:20,000 topographical map in numeric format. it included watercourses, wetlands/ bogs, roads, and elevation contour lines. the delineation of recent cuts was obtained from the industrial forest company abitibiconsolidated inc. we used arcview 3.2 software and spatial analyst extension (esri 1996) to manage the database and to analyse the aerial survey data. the analysis of aerial surveys was done at 2 scales: clear-cut landscapes and moose yards. at the first scale, moose density was computed inside each landscape (excluding the 1-km buffers). we used the pearson correlation coefficient to test if moose density was related to the size of the landscape or to the proportion of residual forest within 0 50 100 150 200 250 300 1 2 3 4 5 12 13 14 15 17 18 19 20 b landscape t o t a l a r e a ( km 2 ) 0 10 20 30 40 50 r e s id u a l f o r e s t ( % ) . total area residual forest fig. 2. total area of the 14 studied clear-cut landscapes and proportion of residual forest (uncut) within each one, expressed as the proportion of the productive area (land area minus wetlands/bogs). ≥ moose in clear-cuts – potvin and courtois alces vol. 40, 2004 64 each landscape. we also verified if moose yards tended to be located close to the uncut forest surrounding the landscapes. for that purpose, we generated 500-m buffers inside each clear-cut landscape to evaluate the proportion of the surveyed area occupied by each distance class to surrounding uncut forest. using a test, the actual number of moose yards by distance class was then compared with the theoretical number that corresponded to a distribution proportional to the area surveyed in each class. at the scale of the yards, we generated a 1-km2 circle around the center point of moose tracks to approximate the area used by moose. using fresh moose tracks to delineate yards was not possible because the track network was too small and sometimes consisted of a single linear track. to evaluate habitat selection of moose in cutlandscapes, we computed the chessonmanly selection index (manly et al. 1993) for each yard for 4 habitat classes (water, wetlands/bogs, recent cuts, and uncut forest): where u i = proportion of habitat class i inside the 1-km2 circle and a i = proportion of habitat i inside the clear-cut landscape. the friedman test was applied to this index in order to test the influence of habitat class on moose yard selection (yard = subject, habitat class = treatment). vegetation survey time since cutting influences vegetation in cutovers. to account for that variation, we measured habitat variables at 19 sites randomly selected within clear-cut patches of the studied landscapes. we also compared vegetation in moose yards (n = 11) with random sites (n = 11) selected within clear-cuts in the same landscapes. cm index i = u i / a i ∑ u i / a i measurements were made at 6 sampling stations for each random site or yard, systematically distributed (50 × 50 m spacing). vertical crown closure of the tree (> 4 m) and shrub (1.5-4 m) layers was measured by the presence/absence of canopy over 10 points spaced on a line at a 3-m interval at each station (adapted from bunnell and vales 1990). we used a 2× metric prism and dbh measurements to evaluate the basal area and the number of trees > 9 cm d b h ( 1 p o i n t s a m p l e p e r s t a t i o n ) (grosenbaugh 1952). the number of stems in the shrub layer (1–9 cm dbh) was counted inside a 25-m2 circular plot. the lateral cover was measured on a 2-m profile board at a 15-m distance (2 readings in opposite directions per station) (nudds 1977). the median height of forest regeneration was estimated visually inside a 15-m radius circle at each station. browse availability was evaluated by counting stems having at least 1 twig (10 cm long) between 50 and 300 cm from the ground in three 1m radius circular plots per station. black spruce was not included as browse because it is seldom used by moose. to compare each variable among cuts of different ages, we used a 1-factor anova. the same test was also applied between moose yards and random sites selected in the same landscapes. fig. 3. aerial survey plan for a clear-cut landscape. north-south transects are spaced 0.5 min of longitude (600 m) apart. χ2 alces vol. 40, 2004 potvin and courtois moose in clear-cuts 65 results moose densities snow conditions were very good in all aerial surveys, with 40-80 cm of snow depth, 5-20 cm of fresh dry snow, and absence of crust. wind was light or absent and a clear sky enabled easy detection of moose tracks in these open landscapes. moose tracks were detected in 8 landscapes, for a total of 11 yards and 22 moose (4 adult males, 10 adult females, and 8 calves). the overall density for the 14 landscapes (1,089 km2) was 0.20 moose/10 km2. there was no significant relationship between moose density by clear-cut landscape and the size of the landscape (r = -0.31, p = 0.14) or the proportion of residual forest within each landscape (r = 0.32, p = 0.14) (fig. 4). moose yards the number of moose yards by 500-m distance classes from the uncut forest inside each landscape had a similar distribution to the proportion of the landscape's area by distance classes ( = 0.02, p = 0.99) (fig. 5). the 14 clear-cut landscapes contained 4.2% water, 7.8% wetlands/bogs, 61.8% recent cuts, and 26.1% uncut forest. the proportions of these cover types were different between moose yards and landscapes, with moose yards containing less water (0.9%) and wetlands/bogs (3.8%) (friedman test, p < 0.01) (fig. 6). when testing only for recent cuts and uncut forest, there was no difference between moose yards and landscapes (p = 0.76). vegetation vegetation was quite similar among cuts of different ages, with only white birch stems in the shrub layer being more numerous in 8–9 year-old clear-cuts than in younger ones (p = 0.04) (table 1). moose yards had quite a different vegetation than the clear-cut landscapes where they were r = -0.31 p = 0.14 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0 50 100 150 200 250 size of the landscape (k m 2) m o o s e / 1 0 k m 2 r = 0.32 p = 0.14 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 15 20 25 30 35 40 residual forest (%) m o o s e / 1 0 k m 2 fig. 4. relationship between moose density in clear-cut landscapes and the size of the landscape or the proportion of residual forest within each landscape. located (table 2). moose yards had greater vertical cover for the shrub layer (37 vs. 18%, p < 0.01), greater lateral cover (69 vs. 53%, p < 0.01), a more abundant shrub layer (2,600 vs. 1,300 stems/ha, p < 0.01), especially balsam fir (p = 0.02), and 3 times more browse available (23,200 vs. 7,100 stems/ha, p < 0.01). conversely, black spruce was less abundant in the tree layer (basal area and stems/ha, p = 0.02) and shrub layer (p = 0.01) of moose yards. the height of the regeneration was the same in moose yards and in random sites (2.2 m). discussion black spruce forests are poor habitats for moose, as opposed to mixed or deciduous stands (brassard et al. 1974, girard and joyal 1984). the general density in this type of forest in the hunting zone was only 0.37 2 2 moose in clear-cuts – potvin and courtois alces vol. 40, 2004 66 0 1 2 3 4 5 0-500 5001000 10001500 15002000 20002500 25003000 30003500 35004000 4000+ distance (m ) m o o s e y a r d s . 0 5 10 15 20 25 30 35 40 l a n d s c a p e a r e a ( % ) . moose yards landscape area fig. 5. distribution of the number of moose yards in relation to the distance from the uncut forest surrounding each clear-cut landscape and proportion of the landscape area by 500-m distance classes. moose/10 km2, a low value (dussault 1999). as expected from our first prediction, moose density was even lower inside our 14 clearcut landscapes (0.20 moose/10 km2). however, such density still represents about 50% of the density before cut. in a previous study, courtois and beaumont (2002) measured a smaller decrease in moose density (23–30%) in 2 blocks that had been partially clear-cut (29–43% of their total area). in our study, the treatment was more severe because surveys were restricted to clearcut landscapes instead of larger blocks containing a proportion of uncut forest. in the early 1980s, girard and joyal (1984) conducted aerial surveys in an area located at the same latitude as our study area. they reported densities of 0.36 moose/10 km2 in 60-km2 plots located in clear-cuts, as opposed to 0.50 moose/10 km2 in uncut forest. all these results indicate that moose can persist in landscapes that have been recently logged (< 10 years) and where large clear-cuts (= 250 ha) are aggregated. moose densities in such landscapes are much lower though than in uncut forest. this confirms part of our first prediction. on the other hand, moose density was not higher in the smaller clear-cut landscapes (fig. 4). while 0 10 20 30 40 50 60 70 water bogs cuts forest % o f t h e a r e a . landscapes moose yards 5 landscapes smaller than 100 km2 had the highest densities ( ≥ 0.4 moose/10 km2), 5 other landscapes in the same area class had no moose. in ontario, increase in hunting pressure in logged areas had a negative effect on moose densities (eason et al. 1981, eason 1989, rempel et al. 1997). conversely, in southwestern québec, the accessibility offered by new logging roads had a minor impact on harvest rates in blocks recently clear-cut (courtois and beaumont 1999). we did not measure hunting pressure in our study but bertrand and potvin (2002) analysed moose harvest statistics in those same 14 clear-cut landscapes and in the entire management unit 25-03 from 1981 to 2000. a similar trend between both data sets suggests that the effect of hunting on moose fig. 6. cover types of 14 clear-cut landscapes and of 11 moose yards located in the same landscapes. alces vol. 40, 2004 potvin and courtois moose in clear-cuts 67 table 1. vegetation of clear-cut black spruce landscapes according to the time since cutting. ( x se) abundance was not different in the clearcut landscapes than outside. contrary to our second prediction, spatial analysis was unable to explain the presence or the location of moose yards inside clear-cut landscapes. moose yards were not located close to the uncut forest surrounding the landscapes and did not have a higher proportion of residual forest. forest strips between clear-cut patches or riparian buffer strips along streams are not attractive to moose in winter probably because of their small size (60–100 m width) and low browse production (black spruce forest). furthermore, as suggested by courtois et al. (2002), cover does not appear to be a major component of habitat for moose even in late winter, at least in regions where snow depth is usually < 90 cm. in the clear-cut landscapes in general, browse availability and density of the shrub layer were still poor 3–9 years after logging. in moose yards, the vegetation was quite different from that of random sites within the same landscapes in that shrubs and small trees were more abundant, especially balsam fir, and there was greater vertical cover, lateral cover, and more browse. before conducting the ground survey in recently logged areas, we did not expect to find this type of vegetation, which is more characteristic of balsam fir or mixed 3–4 years 5–7 years 8–9 years (n = 10) (n = 6) (n = 3) crown closure (%) trees (> 4 m) 2 ± 1 2 ± 1 7 ± 3 0.91 shrubs (1.5-4 m) 7 ± 1 13 ± 3 24 ± 6 0.25 tree layer (> 9 cm dbh) basal area (m2 / ha) black spruce 0.3 ± 0.2 0.9 ± 0.3 0.1 ± 0.1 0.97 balsam fir 0.0 ± 0.0 0.0 ± 0.0 0.3 ± 0.2 1 white birch 0.1 ± 0.1 0.3 ± 0.2 0.7 ± 0.4 0.96 total 0.4 ± 0.2 1.4 ± 0.4 1.0 ± 0.6 0.99 stems / ha black spruce 32 ± 18 87 ± 35 9 ± 9 0.86 balsam fir 0 ± 0 0 ± 0 18 ± 15 1 white birch 3 ± 2 13 ± 7 18 ± 11 0.85 total 35 ± 18 120 ± 37 44 ± 25 0.92 shrub layer (1–9 cm dbh) black spruce 500 ± 110 520 ± 170 170 ± 97 0.91 balsam fir 93 ± 35 0 ± 0 470 ± 160 0.14 white birch 7 ± 7 89 ± 48 390 ± 170 0.04 trembling aspen 7 ± 7 480 ± 200 13 ± 13 0.69 total 610 ± 110 1,200 ± 240 1,300 ± 340 0.22 lateral cover (%) 42 ± 2 52 ± 4 57 ± 4 0.11 height of regeneration (m) 1.8 ± 0.1 2.0 ± 0.2 1.8 ± 0.2 0.22 available browse (stems / ha) 6,200 ± 750 10,700 ± 1,800 11,900 ± 1,900 0.07 variable p moose in clear-cuts – potvin and courtois alces vol. 40, 2004 68 table 2. vegetation of moose yards located in clear-cut black spruce landscapes and of random sites selected within clear-cuts of the same landscapes. ( x se) moose yards random sites (n = 11) (n = 11) crown closure (%) trees (> 4 m) 4 ± 1 5 ± 2 0.83 shrubs (1.5-4 m) 37 ± 3 18 ± 3 < 0.01 tree layer (>9 cm dbh) basal area (m 2 / ha) black spruce 0.1 ± 0.1 0.7 ± 0.2 0.02 balsam fir 0.1 ± 0.0 0.1 ± 0.1 0.56 white birch 0.5 ± 0.2 0.4 ± 0.2 0.91 total 0.6 ± 0.2 1.4 ± 0.4 0.1 stems / ha black spruce 10 ± 6 71 ± 25 0.02 balsam fir 8 ± 5 8 ± 7 0.97 white birch 16 ± 7 14 ± 6 0.82 total 33 ± 12 100 ± 27 0.02 shrub layer (1–9 cm dbh) black spruce 240 ± 60 590 ± 130 0.01 balsam fir 550 ± 160 140 ± 60 0.02 white birch 410 ± 120 150 ± 70 0.07 trembling aspen 850 ± 280 250 ± 110 0.05 total 2,600 ± 400 1,300 ± 200 < 0.01 lateral cover (%) 69 ± 3 53 ± 2 < 0.01 height of regeneration (m) 2.2 ± 0.1 2.2 ± 0.1 0.8 available browse (stems / ha) 23,200 ± 2,400 7,100 ± 1,300 < 0.01 variable p forests than of black spruce stands, at such northern latitudes. in conformity with our third prediction, this confirms that food production is the main factor driving habitat selection by moose, as suggested by crête (1977), peek (1998), and courtois et al. (2002). management implications in black spruce forest, it takes more than 10 years for browse availability to become attractive to moose in recent clearcuts. in our study, moose density was very low and in most cutovers, even the oldest ones, and food production was much lower in control sites than in moose yards. these observations confirm our hypothesis that clear-cut black spruce landscapes are poor moose habitats. in balsam fir or mixed forests, 10-year cutovers are much more productive (potvin et al. 2004). special measures are therefore needed in the planning process if the goal is to maintain moose populations and moose hunting activity in this type of forest. as proposed by courtois et al. (2002), there can be 2 strategies: (1) identify stands that have greater value for moose and use silvicultural techniques that will maintain their characteristics, or (2) leave a higher proportion of residual forest (50–60%) by applying a dispersed patch cutting strategy. since mixed alces vol. 40, 2004 potvin and courtois moose in clear-cuts 69 and deciduous stands are rare within the black spruce–moss ecological zone, the first strategy might be implemented as a basic step where moose production is a moderate priority issue. the second strategy is more aimed at maintaining moose hunting activity. although moose density in our study was not related to the proportion of residual forests, moose hunters have a negative perception of forest harvesting systems (courtois et al. 2001). they prefer landscapes where residual forests are dominant over those where clear-cut patches are aggregated and narrow uncut strips are the sole residual forest. the second strategy might be suited to areas where moose production is a higher priority (outfitters areas, areas devoted to forest–wildlife integrated management). in other areas, the goal might be to decrease moose numbers. for example, the conservation of woodland caribou (rangifer tarandus) may be at risk because of predation, if a large moose population exists that can sustain high wolf numbers (courtois 2003). in this case, caribou habitat management involves preserving large forest blocks (> 250 km2) interconnected with a corridor network. in order to diminish moose and predator densities, clearcuts outside preserved areas should aim at regenerating black spruce and very few deciduous species. in this context, stands that may be suitable to moose (e.g., mixed forests) might even be transformed towards coniferous stands. acknowledgements we would like to thank charles faucher (deceased), who conducted most of the spatial analyses, wildlife technicians of saguenay-lac-saint-jean region at société de la faune et des parcs du québec, who took part in the aerial surveys, and normand bertrand, laurier breton, and claude paquet, who implemented the vegetation survey. special thanks to abitibi-consolidated inc. for providing the maps of the recent cuts. references bertrand, n., and f. potvin. 2002. utilisation par la faune de la forêt résiduelle dans de grandes aires de coupe: synthèse d’une étude de trois ans réalisée au saguenay-lac-saint-jean. québec ministère des ressources naturelles, de la faune et des parcs, report def0231. bliss, c. j. 2000. public perceptions on clearcutting. journal of forestry 98(12):4-9. brassard, j. m., é. audy, m. crête, and p. grenier. 1974. distribution and winter habitat of moose in québec. naturaliste canadien 101:67-80. bunnell, f. l., and d. j. vales. 1990. comparison of methods for estimating forest overstory cover: differences among techniques. canadian journal of forest research 20:101-107. courtois, r. 1991. normes régissant les t r a v a u x d ’ i n v e n t a i r e s a é r i e n s d e l’orignal. québec ministère du loisir, de la chasse et de la pêche, report 1907. _____. 2003. la conservation du caribou forestier dans un contexte de perte d’habitat et de fragmentation du milieu. ph.d. thesis, université du québec à rimouski, rimouski, québec, canada. _____, and a. beaumont. 1999. the influence of accessibility on moose hunting in northwestern québec. alces 35:41-50. _____, and _____. 2002. a preliminary assessment on the influence of habitat composition and structure on moose density in clearcuts of north-western québec. alces 38: 167-176. _____, c. dussault, f. potvin, and g. daigle. 2002. habitat selection by moose in clear-cuts – potvin and courtois alces vol. 40, 2004 70 nudds, t. d. 1977. quantifying the vegetation structure of wildlife cover. wildlife society bulletin 5:113-117. peek, j. m. 1998. habitat relationships. pages 351-373 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. potvin, f., l. breton, and r. courtois. 2004. response of beaver, moose and snowshoe hare to clearcutting in québec boreal forest: a reassessment 10 years after cut. canadian journal of forest research 35: 151 160. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. thompson, i. d., and r. w. stewart. 1998. management of moose habitat. pages 377-401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. moose (alces alces) in clear-cut landscapes. alces 38: 177-192. _____, j. p. ouellet, and a. bugnet. 2001. moose hunters’ perception of forest harvesting. alces 37:19-33. crête, m. 1977. importance de la coupe forestière sur l’habitat hivernal de l’orignal dans le sud-ouest du québec. canadian journal of forest research 7:241-257. _____. 1988. forestry practices in quebec and ontario in relation to moose population dynamics. forestry chronicle 64:246-250. dussault, c. 1999. inventaire aérien de l’orignal (alces alces) dans la zone de chasse 18 ouest à l’hiver 1998. québec ministère de l’environnement et de la f a u n e , d i r e c t i o n r é g i o n n a l e d u saguenay-lac-saint-jean. eason, g. 1989. moose response to hunting and 1 km2 block cutting. alces 25:6374. _____, e. thomas, r. jerrard, and k. oswald. 1981. moose hunting closure in a recently logged area. alces 17:111125. esri. 1996. arcview gis. environmental systems research institute incorporated, redlands, california, usa. girard, f., and r. joyal. 1984. l’impact des coupes à blanc mécanisées sur l’orignal dans le nord-ouest du québec. alces 20:3-25. grondin, p. 1996. écologie forestière. pages 133-279 in manuel de foresterie. presses de l’université laval, québec, québec, canada. grosenbaugh, l. r. 1952. plotless timber estimates new, fast, easy. journal of forestry 50(1):32-37. manly, b., l. mcdonald, and d. thomas. 1993. resource selection by animals. statistical design and analysis for field studies. chapman and hall, london, u.k. 4209(55-64).pdf alces vol. 42, 2006 clough et al. abnormal incisor breakage in moose 55 elemental composition of incisors in nova scotia moose: evaluation of a population with abnormal incisor breakage michael clough1,2, marcos zentilli1, hugh g. broders3, andtony nette4 1department of earth sciences, dalhousie university, halifax, ns, canada b3h 4j1; 3department of biology, st. mary’s university, halifax, ns, canada b3h 3c3, e-mail: hugh.broders@smu.ca; 4nova scotia department of natural resources, wildlife division, kentville, ns, canada b4n 4e5 abstract: this study compared the concentrations of major and trace elements in the enamel of incisors from moose (alces alces andersoni) in cape breton highlands, where the incidence of incisor tooth breakage was believed to be unusually high, and moose in southwest nova scotia (a. a. americana) where there was no evidence of breakage. our goal was to determine which elements, if any, might be related to the incisor breakage in moose from cape breton highlands. there was a positive relationship between age and frequency of incisor breakage, and most moose had a broken i2 incisor by 4 years of age in the cape breton highlands. we analyzed i2 incisors for 51 trace elements with inductively coupled plasma-mass spectrometry. concentrations of 8 elements, including barium, beryllium, cadmium, cobalt, lead, tin, strontium, and yttrium, were lower (p < 0.05) in incisors from cape breton highlands; gallium had a higher concentration. reduced intake of barium, beryllium, and alces vol. 42: 55-64 (2006) key words: alces alces, barium, beryllium, cadmium, cobalt, gallium, incisor, inductively coupled plasma-mass spectrometry, lead, moose, nova scotia, strontium, teeth, trace elements, yttrium the role that major and trace elements play in the metabolism of organisms has been the focus of intense, well-documented research (underwood 1977, prasad 1993, underwood and suttle 1999, bogden and klevay 2000). disease may result from either a toxic or a (maisironi 2000). teeth are good indicators of the abundance of many elements (brown et al. 2004, kang et al. 2004, dolphin et al. 2005) due to the crystal structure of enamel (sharaway and yeager 1991, simmelink 1994). major and trace elements are incorporated within the hydroxyapatite crystal framework during the mineralization period (sharaway and yeager 1991, simmelink 1994) and, due to the semipermeable nature of hydroxyapatite, small ions and molecules are able to pass through the enamel framework preceding eruption (cutress 1983). therefore, to an extent, teeth remain in chemical equilibrium with the oral environment (zimmerman 1976, driessens 1982). enced by the concentration of major and trace elements (zimmerman 1976), which is affected by various factors such as normal wear, food composition, and regional geochemistry (cutress 1983). therefore teeth are an excellent bio-indicator of local environmental conditions (lee et al. 1999, lochner et al. 1999, gdula-argasinska et al. 2004). this is especially true for non-migratory ruminant 2present address: department of biology, st. mary's university, halifax, ns, canada b3h 3c3 abnormal incisor breakage in moose clough et al. alces vol. 42, 2006 56 herbivores such as moose (alces alces) that obtain their entire dietary mineral intake from distinct regional geological localities (cederlund and okarma 1988, lepitch and gilbert 1989, hundertmark 1998). observations of physical and behavioral anomalies suggest that the moose population is stressed in the cape breton highlands (a. a. andersoni) in nova scotia, canada. wildlife authorities have observed osteophagia (roger and nette 2002), an increased incidence of bark stripping, and tooth breakage (fig. 1) in cape breton highlands moose where densities are high and heavy browsing of preferred vegetation is evident. distinct browse lines 2–3 meters off the ground extend several kilometers through the forest (basquill and thompson 1997). the only previously reported case of incisor breakage was in a moose (a. a. gigas) population on the seward peninsula, alaska. smith (1992) documented ‘incisiform breakage’ over a 2-year period (1988-1990), which closely resembles the breakage observed in cape breton highlands moose. the breakage observed in the cape breton highlands is markedly different from the tooth wear frequently reported in other ungulate species (hewison et al. 1999, loe et al. 2003). it also differs from the incisiform wear reported in kenai peninsula moose by peterson et al. (1982), and the unusual wear described by young and marty (1986) in a manitoban moose population. the distinctive incisor breakage begins as a brown stained crack on the tooth surface (fig. 1) that is a precursor to breakage. where breakage has occurred, the tooth is subsequently rounded down and stained brown. this rounded, staining characteristic is important because it indicates that breakage occurred during the lifetime and not at, or after, death. the purpose of this study was to determine whether concentrations of major and trace elements in incisors might explain breakage. to do this we compared trace element concentrations of teeth from cape breton highlands moose to those of a control population on mainland nova scotia (a. a. americana) without evidence of abnormal incisor breakage. study areas the highlands region is located in northern cape breton island, in northeastern nova scotia, and covers an area of approximately 3,900 km2. the area is a characteristic boreal region (pulsifer and nette 1995) with a peak elevation of approximately 535 m above sea level (canada: national and historic parks branch 1970). approximately 1,600 mm of precipitation is recorded annually, with > 400 cm falling as snow, resulting in snow pack that lasts 181-212 days (phillips 1990). the control area was the tobeatic wilderness area, which is a characteristic acadian forest region (farrier et al. 1991). the area receives approximately 1,400 mm of precipitation annually, with 150-200 cm falling as snow, resulting in a snow pack that lasts 110-140 days (phillips 1990). methods in the cape breton highlands, incisor samples were obtained from hunters during the 2001 hunting season from management zones immediately north and south of cape 1 2 3 fig 1. lower mandible from a cape breton highlands moose. numbers on teeth represent breakage scores. note the stained, polished, and rounded surfaces of the fractured teeth indicating breakage occurred during lifetime of the animal (photo by vince power, 2001). alces vol. 42, 2006 clough et al. abnormal incisor breakage in moose 57 breton highlands national park (fig. 2). in the tobeatic wilderness area in southwestern nova scotia, incisors were obtained by the nova scotia department of natural resources (nsdnr) from accidental kills, illegal kills, and live extraction when collaring animals as part of ongoing studies (fig. 2). based on the work of smith (1992), nette and power (nsdnr 1999) devised a ing the degree of incisor breakage used in this study: slight – indicates < 30% of tooth material missing from the incisal surface; moderate – indicates 30-60% of tooth material missing from the incisal surface; and severe – indicates > 60% of tooth material missing from the incisal surface. individual i1s were used to determine age from cementum annuli measurements (matson 1981). the concentrations of 37 elements and the 14 rare earth elements (ree) were determined from the i2s. the ree comprising the lanthanide series were summed and considered as one element (brown et al. 2000). the root was separated from the crown with a buhler isomet low speed saw with a diamond tip blade. crowns were crushed with a percussion cutter and enamel fragments were retrieved tipped paintbrush. enamel was reduced to a with ethanol as a lubricant. an ultrasound bath agate mortar pestle between samples to prevent cross-contamination. enamel was placed in drying dishes and dried at room temperature for 3-5 hours, recovered, and placed into labeled vials. samples were sent to geo labs laboratory (sudbury, ontario, canada) and analyzed for element content using inductively coupled plasma-mass spectrometry (icp-ms). two-sample t-tests were used to compare element concentrations of the two populations. regression analysis was used to assess the relationship between age and breakage score (sokal and rohlf 1995) using 46 i2s collected during the 2001 hunting season. age/breakage data compiled between 1996-2002 (nsdnr 2002) were analyzed using chi-square analysis of breakage between the i1s and i2s. statistiresults there was a positive relationship between incisor breakage and age in cape breton 1.28 + 0.158age, p < 0.05, r2 i1 had a higher frequency of breakage than the i2 ( 2 p frequency of breakage of i1 was 2-3x that of i2 at > 2.5 years old and 2 out of 3 moose had a broken i1 after 4.5 years. breakage of i1 in moose > 3.5 years old was > 50% and more than doubled from 3.5 to 5.5 years old. breakage of i2 was < 30% in moose < 6.5 years old (table 1). relative to the control population in southwest nova scotia, cape breton highlands moose had lower (p < 0.05) concentrations of barium (ba), beryllium (be), cadmium (cd), cobalt (co), lead (pb), strontium (sr), tin (sn), and yttrium (y), and higher concentration of gallium (ga) in the enamel of their teeth (table 2). largest absolute differences occurred in fig. 2: moose teeth sample collection areas in nova scotia, canada. note: no samples were taken from cape breton highlands national park indicated by the lined inset area. abnormal incisor breakage in moose clough et al. alces vol. 42, 2006 58 cd, co, pb, sn, sr, and y. although not different, relatively large absolute differences were found in chromium (cr), iron (fe), and vanadium (v) (table 2). discussion breakage type and trends measured in cape breton highland moose were similar to results from alaskan moose that indicated that breakage severity increased with age, and breakage severity was greater in i1s than i2s (smith 1992). there were no differences in elemental concentrations of 40 teeth collected from the seward peninsula and 20 teeth from a control population near galena, alaska (smith 1992). the results for their microbeam analysis were not presented, therefore, comparisons between nova scotia and alaska moose were in mineral requirements, diagnosis of problems (underwood and suttle 1999). however, the lower concentrations of ba, be, co, sr, and y within the cape breton highlands population are consistent with previous studies that concluded that a reduction of these elements may result in higher enamel solubility and compromised strength that lead to dental disease (bibby and losee 1970, curzon 1983a, b, c, d). barium, be, co, and sr depletion result in depressed growth, even evidence of hypoplasia within erythrogenic tissue and bone marrow (underwood 1971, curzon 1983a, d). strontium may also that results in retardation of apatite dissolution (curzon 1983d). trations of cd, ga, pb, and sn may have on enamel strength and solubility are unclear (curzon 1983c, stack 1983a, b), although sn may inhibit the dissolution of enamel via acid neutralization within the oral environment (curzon 1983c). knowledge of essential vitamins and minerals within the diet of moose is limited (schwartz and renecker 1998). frank et al. (2004) examined elemental concentrations from various tissues of nova scotia moose and found that tobeatic moose displayed unusually high levels of cd within kidney tissue as compared to cape breton highlands moose; cd concentration in enamel was also high in tobeatic moose (table 2). they also found that both tobeatic and cape breton highland moose had lower co concentrations than swedish and alaskan moose. these data and the low concentration value of co in the cape breton highlands population (table 2) suggest that nova scotia populations though not different, it is important to note that the relative concentrations of cr, fe, and v were low, and magnesium (mg) high in the cape breton highlands population (table 2). with the exception of mg, these results were not surprising because geochemical position within the periodic table (albarede 2003). elements of the same group would be expected to have similar, relative availability within the environment. cadmium, fe, and v are all grouped in the transition metals of the periodic table with cd, co, and y that age incisor 1 incisor 2 n breakage frequency (%) n breakage frequency (%) 1.5 210 2 (1.0%) 214 1 (0.5%) 2.5 282 33 (11.7%) 287 8 (2.8%) 3.5 299 98 (32.8%) 308 24 (7.8%) 4.5 228 122 (53.5%) 232 49 (21.1%) 5.5 118 86 (72.9%) 120 32 (26.7%) 218 148 (67.9%) 220 71 (32.3%) table 1. sample size, age, and frequency of incisor breakage of harvested moose from the cape breton highlands, 1999-2002 (nsdnr 2002). alces vol. 42, 2006 clough et al. abnormal incisor breakage in moose 59 table 2. element concentrations in tooth enamel from 2 moose populations in nova scotia, canada. concentrations are expressed in ppm (except ca, which is expressed as weight percent). bold numelement cape breton highlands, n mainland (control) population, n mean se range mean se range ag 0.055 0.015 0.011 0.294 0.063 0.012 0.038 0.089 al 65.3 18 12.9 329 72.5 8.4 59.9 95.8 as 4.842 0.22 3.870 4.212 5.685 0.58 4.71 7.37 ba 125.7 8.6 72.6 235 180.5 6.7 171 – 200 be 0.015 0.002 0 0.034 0.029 0.01 0 0.045 bi 0.015 0.003 0.005 0.058 0.024 0.006 0.017 0.041 ca (wt%) 34.8 0.4 30.8 38.5 35.2 0.7 33.9 37.2 cd 0.114 0.032 0.028 0.64 0.363 0.12 0.201 0.717 co 0.017 0.015 0 0.297 0.824 0.31 0 1.15 cr 0.082 0.024 0 0.323 0.18 0.065 0.096 0.37 cs 0.005 0.001 0.02 0.019 0.008 0.001 0.006 0.001 cu 17.4 2.6 5.4 48.5 24.3 2.2 18 28.7 fe 143.7 44 41.7 914 308 111 176 – 639 ga 0.207 0.007 0.144 0.267 0.169 0.009 0.144 0.182 hf 0.011 0.004 0 0.072 0.017 0.014 0.001 0.060 li 1.031 0.17 0.270 3.16 1.216 0.21 0.902 1.81 mg 3429 155 2640 4720 2767.5 86 2590 – 2960 mn 56.8 8.8 14.4 168 34.5 11 14.7 – 65 mo 0.055 0.012 0.015 0.242 0.061 0.006 0.047 0.076 nb 0.02 0.005 0.009 0.104 0.02 0.002 0.015 0.023 pb 0.805 0.077 0.330 1.88 1.935 0.49 1.06 3.30 rb 0.607 0.054 0.287 1.25 0.847 0.053 0.689 0.926 sb 0.027 0.007 0.004 0.141 0.026 0.009 0.007 0.048 se 1.349 0.029 1.115 1.74 1.425 0.027 1.37 1.48 si 173.2 26 62.2 552 152 9.3 137 – 179 sn 0.084 0.018 0.01 0.322 0.343 0.16 0.151 0.744 sr 268.3 13 192 415 518.5 22 474 – 564 ta 0.001 0.000 0 0.008 0.001 0.000 0 0.001 th 0.004 0.001 0 0.018 0.009 0.003 0.004 0.018 ti 88.6 4.7 70 153 84.6 6.3 70.9 97.9 tl 0.008 0.001 0.005 0.015 0.008 0.001 0.006 0.011 u 0.003 0.001 0.001 0.015 0.006 0.001 0.004 0.008 v 0.387 0.045 0.168 0.982 0.597 0.098 0.353 0.824 w 7.61 1.7 0.99 20.8 4.16 1.4 1.88 8.39 y 0.062 0.009 0.03 0.19 0.133 0.02 0.097 0.169 abnormal incisor breakage in moose clough et al. alces vol. 42, 2006 60 had lower concentration in the cape breton highlands population. however, the alkaline – earth metal mg has a higher concentration in the cape breton highlands population than the tobeatic population, unlike the other alkaline earth metals (ba, be, and sr). the reason for this is unclear, however it may be a result of isobaric interference during the icp-ms analysis, which is a common issue with biological applications (taylor et al. 2005). lying cause for the incisor breakage observed in the cape breton highlands moose population. osteophagia is generally considered to occur within ruminant species as a result of rette 1985, denton et al. 1986). bark stripping by moose is generally considered to be a sign of starvation (renecker and schwartz 1998), and heavily browsed plants generally contain less nutritive value for moose (schwartz and renecker 1998). it is important for moose to consume a wide variety of plant species, or (ohlson and staaland 2001). whether tooth breakage is a result of osteophagia or bark stripping, or even a combination of the two, is unclear. however, these phenomena have been well documented in other ungulate species (bowyer 1983, barrette 1985, denton et al. 1986, duthie and skinner 1986, kierdorf 1994, baker et al. 1997, yokoyama et al. 2001, ueda et al. 2002) with no reports of breakage similar to that observed in cape breton highlands moose. it was interesting that teeth from the control area were consistently richer in tin by one differences in bedrock geology. southwestern economic) deposits of tin and the glacial deposits and soils are locally enriched with this and other elements common in granitoid intrusive rocks. this correlation of enamel geochemistry and bedrock geology may have valuable forensic usefulness. continued use of a broken incisor results in the incisal surface becoming rounded (fig.1) and this probably reduces foraging likely has negative implications for the overall health and vitality of individual moose. to on population dynamics, it would be necessary to compare the mortality rate of older age classes and population age structure of several moose populations. we suggest that research on incisor breakage in the cape breton highlands moose continue and be supplemented with more control samples from other populations. acknowledgements we would like to thank vince power (nsdnr) for supplying hunting data, frequency and breakage data, and photos of moose mandibles. alexander (sandy) grist trained mc in sample preparation techniques in the fission track research laboratory of dalhousie university. during its early stages tions by prof. milton graves, phil casey element cape breton highlands, n mainland (control) population, n mean se range mean se range zn 54.7 3.3 37.9 85.9 64.6 5.6 53.6 80.3 zr 0.456 0.19 0.027 3.35 0.865 0.8 0.027 0.084 ree 0.301 0.1 0.076 2.196 0.899 0.41 0.271 0.711 table 2 (continued). element concentrations in tooth enamel from 2 moose populations in nova scotia, canada. concentrations are expressed in ppm (except ca, which is expressed as weight percent). alces vol. 42, 2006 clough et al. abnormal incisor breakage in moose 61 and bsc. honours thesis by vanessa walsh. we acknowledge funding from the nsdnr wildlife division and the natural sciences and engineering research council of canada (nserc) through discovery grants to mz and hgb. we are thankful for the suggestions of two anonymous reviewers, which considerably improved the manuscript. references albarede, f. 2003. geochemistry: an introduction. cambridge university press, cambridge, u.k. baker, w. l., j. a. munroe, and a. e. hessl. 1997. the effects of elk on aspen in the winter range in rocky mountain national park. ecography 20: 155-165. barrette, c. 1985. antler eating and antler growth in wild axis deer. mammalia 49: 492-499. basquill, s., and r. thompson. 1997. moose (alces alces) browse availability and utilization in cape breton highlands national park. parks canada report 010. parks canada, atlantic region, halifax, nova scotia, canada. bibby, b. g., and f. l. losee. 1970. environmental factors and dental disease. eastman dental center, rochester, new york, usa. bogden, j. d., and l. m. klevay, editors. 2000. clinical nutrition of the essential trace elements and minerals. the guide for health professionals. humana press, totowa, new jersey, usa. bowyer, r. t. 1983. osteophagia and antler breakage among roosevelt elk. california fish and game 69: 84-88. brown, c. j., s. r. n. chenery, b. smith, c. mason, a. tomkins, g. j. roberts, l. sserunjogi, and j. v. tiberindwa. 2004. element content of teeth implications for disease and nutritional status. archives of oral biology 49: 705-717. brown, t. l., h. e. lemay, and b. e. bursten. 2000. chemistry: the central science. eighth edition. prentice hall incorporated, upper saddle river, new jersey, usa. canada national and historic parks branch. 1970. cape breton highlands national park provisional master plan. queen’s printer, ottawa, ontario, canada. cederlund, g. n., and h. okarma. 1988. home range and habitat use of adult female moose. journal of wildlife management 52: 341-343. curzon, m. e. j. 1983a. barium. pages 305-310 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. _____. 1983b. other essential trace elements. pages 267-282 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. _____. 1983c. other nonessential trace elements. pages 339-356 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. _____. 1983d. strontium. pages 283-304 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. cutress, t. w. 1983. teeth, calculus, and bone. pages 33-106 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. denton, d. a., j. r. blair-west, m. j. mckinley, and j. f. nelson. 1986. physiological analysis of bone appetite (osteophagia). bioessays 4. dolphin, a. e., a. h. goodman, and d. d. amarasiriwardena. 2005. variation in elemental intensities among teeth and between preand postnatal regions of abnormal incisor breakage in moose clough et al. alces vol. 42, 2006 62 enamel. american journal of physical anthropology 128: 878-888. driessens, f. c. m., editor. 1982. mineral aspects of dentistry. volume 10. karger, basel, switzerland. duthie, a. g., and j. d. skinner. 1986. osteophagia in the cape porcupine hystrix africaeaustralis. south african journal of zoology 21: 316-318. farrier, r., c. drysdale, and g. kenney. 1991. kejimkujik national park seaside adjunct. resource description and analysis. parks canada, atlantic region, maitland bridge, nova scotia, canada. frank, a., j. mcpartlin, and r. danielsson. 2004. nova scotia moose mystery a moose sickness related to cobaltand total environment 318: 89-100. gdula-argasinska, j., j. appleton, k. sawicka-kapusta, and b. spence. 2004. further investigation of the heavy metal content of the teeth of the bank vole as an exposure indicator of environmental pollution in poland. environmental pollution 131: 71-79. hewison, a. j. m., j. p. vincent, j. m. angibault, d. delorme, g. van laere, and j. m. gaillard. 1999. tests of estimation of age from tooth wear on roe deer of known age: variation within and among populations. canadian journal of zoology 77: 58-67. hundertmark, k. j. 1998. home range, dispersal and migration. pages 303-336 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. kang, d., d. amarasiriwardena, and a. h. goodman. 2004. application of laser ablation-inductively coupled plasma-mass spectrometry (la-icp-ms) to investigate trace metal spatial distributions in human tooth enamel and dentine growth layers and pulp. analytical and bioanalytical chemistry 378: 1608-1615. kierdorf, u. 1994. fork formation and other signs of osteophagia on long bone swallowed by a red deer stag (cervus elaphus). international journal of osteoarchaeology 3: 37-40. lee, k. m., j.appleton, m. cooke, f. keenan, and k. sawicka-kapusta. 1999. use of laser ablation inductively coupled plasma mass spectrometry to provide element chimica acta 395: 179-185. lepitch, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53: 880-885. lochner, f., j. appleton, f. keenan, and m. cooke of human deciduous teeth by laser ablation-inductively coupled plasma-mass spectrometry. analytica chimica acta 401: 299-306. loe, l. e., a. mysterud, r. langvatn, and n. c. stenseth. 2003. decelerating and sex-dependent tooth wear in norwegian red deer. oecologia 135: 346-353. maisironi, r. 2000. the epidemiology of in j. d. bogden and l. m. klevay, editors. clinical nutrition of the essential trace elements and minerals. the guide for health professionals. humana press, tetowa, new jersey, usa. matson, g. m. 1981. workbook for cementum analysis. matson’s, milltown, montana, usa. (nsdnr) nova scotia department of natural resources. 1999. unpublished reports and tables. nova scotia department of natural resources wildlife division, kentville, nova scotia, canada. _____. 2002. unpublished reports and tables. nova scotia department of natural resources wildlife division, kentville, nova scotia, canada. ohlson, m., and h. staaland. 2001. mineral alces vol. 42, 2006 clough et al. abnormal incisor breakage in moose 63 for moose. oikos 94: 442-454. peterson, r. o., j. m. scheidler, and p. w. stephens. 1982. selected skeletal morphology and pathology of moose from the kenai peninsula, alaska, and isle royale, michigan. canadian journal of zoology 60: 2812-2817. phillips, d. w. 1990. the climates of canada. environment canada, ottawa, ontario, canada. prasad, a. s., editor. 1993. essential and toxic trace elements in human health and disease: an update. wiley-liss incorporated, new york, new york, usa. pulsifer, m. d., and t. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behavior. pages 403-440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. roger, e., and t. nette. 2002. osteophagia among moose of cape breton highlands. moose call 14. schwartz, c. c., and l. a. renecker. 1998. nutrition and energetics. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. sharaway, m., and j. a. yeager. 1991. enamel. pages 49-105 in s. n. bhaskar, editor. orban’s oral histology and embryology. mosby year book, st. louis, missouri, usa. simmelink, j. w. 1994. histology of enamel. pages 228-241 in j. k. avery, editor. oral development and histology. thieme medical publishers incorporated, new york, new york, usa. smith, t. e. 1992. incidence of incisiform tooth breakage among moose from the seward peninsula, alaska, usa. alces supplement 1: 207-212. sokal, r. r., and f. j. rohlf. 1995. biometry. the principles and practice of statistics in biological research. third edition. w.h. freeman and company, new york, new york, usa. stack, m. v. 1983a. cadmium. pages 387400 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, massachusetts, usa. _____. 1983b. lead. pages 357-286 in m. e. j. curzon and t. w. cutress, editors. trace elements and dental disease. john wright. psg, littleton, masschusetts, usa. taylor, a., s. branch, d. halls, m. patriarca, and m. white. 2005. atomic spectrometry update. clinical and biological materials, foods and beverages. journal of analytical atomic spectrometry 20: 323-369. ueda, h., s. takatsuki, and y. takahashi. 2002. bark stripping of hinoki cypress by sika deer in relation to snow cover and food availability on mt takahara, central japan. ecological research 17: 545-551. underwood, e. j. 1971. trace elements in animal and human nutrition. third edition. academic press incorporated, new york, new york, usa. _____. 1977. trace elements in animal and human nutrition. fourth edition. academic press incorporated, new york, new york, usa. _____, and n. f. suttle. 1999. the mineral nutrition of livestock. third edition. cabi publishing, new york, new york, usa. yokoyama, n., i. maeji, t. ueda, m. ando, and e. shibata. 2001. impact of bark stripping by sika deer, cervus nippon, on subalpine coniferous forests in central japan. forest ecology and management 140: 93-99. abnormal incisor breakage in moose clough et al. alces vol. 42, 2006 64 young, w. g., and t. m. marty. 1986. wear and microwear on the teeth of a moose (alces alces) population in manitoba, canada. canadian journal of zoology 64: 2467-2479. zimmerman, s. 1976. physicochemical properties of enamel and dentine. pages 112-133 in e. p. lazzari, editor. dental biochemistry. lea & feiger, philadelphia, pennsylvania, usa. f:\alces\vol_39\p65\3901.pdf alces vol. 39, 2003 spears et al. moose bone marrow fat 273 bone marrow fat content from moose in northeastern minnesota, 1972-2000 brian l. spears1, william j. peterson2, and warren b. ballard1 1department of range, wildlife, and fisheries management, box 42125, texas tech university, lubbock, texas 79409, usa; 2minnesota department of natural resources, section of wildlife, p.o. box 115, grand marais, mn 55604, usa abstract: percent fat in femur bone marrow has been used as an indicator of animal condition at time of death. however, femur bone marrow is not always available for collection. we used linear regression to examine relationships among marrow fat values for long bones (i.e., femur, tibia, mandible, humerus, radius, tarsal and carpal bones) of moose (alces alces) from northeastern minnesota during 1972-2000. linear regressions for bone marrow fat in each set of bones (paired with femur) in calves and adults were significant and highly correlated (r2 = 0.83-0.99). linear regressions for femur bone marrow fat for yearling moose were significant and highly correlated for tibia, humerus and radius bones (r2 = 0.86-0.93), and less so for tarsal bones (r2 = 0.63). bone marrow fat deposition appeared first in proximal and distal bones and was mobilized last in distal bones. calves had higher femur fat in fall and early winter than late winter and spring. month, season, and year had no significant effect on femur marrow fat percent for yearlings or adults. percent femur marrow fat was lower in accidentally killed calves than accidentally killed yearlings or adults. adults killed by disease had lower percent femur fat than those killed by accident or wolves (canis lupus). amount of adult male femur fat was loosely correlated to a winter severity index for the previous winter. our results suggest that fat deposition and mobilization were similar to that found in other studies and that bone marrow fat content may be a good indicator of relative moose health within a population. alces vol. 39: 273-285 (2003) key words: alces alces, bone marrow, death, deposition, fat, femur, minnesota, moose, wolves bone marrow fat percentages have been widely used as an index of ungulate condition at time of death (cheatum 1949, baker and leuth 1966, neiland 1970, franzmann and arneson 1976, peterson 1977, peterson and bailey 1984, ballard et al. 1987, mech et al. 1995). marrow fat is mobilized after other body fat reserves are depleted or exhausted (cheatum 1949, smith and jones 1961). reduction in bone marrow fat may therefore be a good indicator of decreased fitness of an individual. however, the percentage of marrow fat depletion indicating a stressed animal has been debated (mech and delgiudice 1985, ballard 1995). moreover, what constitutes a healthy animal may be relative to other members of the population each year (ballard 1995), and a high marrow fat content would not necessarily indicate an animal in good condition (mech and delgiudice 1985, ballard et al. 1987). despite discrepancies in conclusions regarding stage of individual health, bone marrow fat depletion often remains the only indicator of individual condition at time of death (ballard 1995). fat levels from the femur have typically been used to draw conclusions regarding the condition of individual moose at time of death (cheatum 1949, franzmann and arneson 1976, peterson et al. 1982, ballard et al. 1987, hayes et al. 1991, ballard 1995, moose bone marrow fat spears et al. alces vol. 39, 2003 274 mech et al. 1995). however, predators often consume proximal leg bones (ballard et al. 1981, peterson et al. 1982), and frost wedging may expose femur marrow to the air, rendering the sample useless (peterson et al. 1982). researchers must often settle for collection of other bones for analysis, and the correlation of marrow fat in these bones with that of the typically used femur is desirable. while investigating natural and humancaused mortality of moose (alces alces) in minnesota during 1972-2000, amount of marrow fat was estimated in femur, tibia, mandible, humerus, radius, tarsal, and carpal bones of individual moose. we examined correlations between femur marrow fat and marrow fat in the above bones for moose divided into 3 age classes to determine their usefulness as indices of health as related to the femur marrow fat standard. we also present data on the annual cycle of bone marrow fat deposition and mobilization in moose, and examine differences in amount of femur bone marrow fat among seasons, years, cause of death, and age class. study area and methods the study was conducted from 19722000 in cook and lake counties in northeastern minnesota. the majority of moose examined were within 65 km of poplar lake, located in superior national forest, cook county, minnesota. a description of the study area was provided by nelson and mech (1981). we sexed moose by examining sex organs or by the presence or absence of antlers or pedicles. age of calves and yearlings was determined by tooth eruption and replacement. dental cementum was used to determine adult ages. a femur, tibia, humerus, radius, tarsal, carpal, and mandible was collected from each moose carcass if possible, usually within several hours after death. bone samples from each moose were stored frozen for up to a few months before being processed. bones were broken with a hammer. marrow samples were removed by hand from the central portion of the marrow tube. marrow fat content was determined by the oven drying technique described by neiland (1970). we separated bone marrow fat data by age class of individual due to possible differences of marrow fat mobilization at different stages of life (ballard et al. 1981). moose within 1 year of birth were classified as calves, 1-2 years of age as yearlings, and >2 years of age as adults. all moose were assumed born 1 june (schwartz 1998). we classified causes of death as automobile or train collision, natural accidental death, hunter-kill, killed intentionally as a danger to humans, diseased (brainworm [parelaphostrongylus tenuis] confirmed or suspected, or heavy tick load with heavy hair loss), wolf (canis lupus)-kill, or unknown. marrow fat content means are given + sd. fat was analyzed by month and season of death. month of death was lumped into 1 of 4 seasons: 1 = decemberfebruary, 2 = march-may, 3 = june-august and 4 = september-november. we calculated regression lines and pearson's product-moment correlations for each comparison between an individual's femur and other bones examined within each age class. we assumed the simple linear model: y = b + mx + ε, where y = marrow fat value of bone being correlated, x = marrow fat value of femur, b = the value of y when x = 0, m = best-fit slope of comparison line, and ε = residual error unaccounted for by the model. f-tests were calculated to determine whether models of similar bone pair regressions differed significantly among age classes (graybill 1976). f-statistics were calculated to examine differences in femur marrow fat content alces vol. 39, 2003 spears et al. moose bone marrow fat 275 among age classes. f-statistics were also calculated to examine effects of month, season, and cause of death on bone marrow fat content in femurs within each age class. where f-tests were significant, student's ttest was used to determine differences in femur marrow fat means among age classes, causes of death, and seasons. following ballard et al. (1981) and davis et al. (1987), we used paired t-tests to examine differences in marrow fat content between femurs and all other bones examined, as well as humerus-tibia,humerus-carpal, and proximal-distal bone pairs. welch's approximate t values and associated degrees of freedom were used in cases where variances between student's t-test groups were not homogeneous (zar 1999). a winter severity index (wsi) was calculated for each winter from weather data collected at poplar lake in cook county, minnesota (minnesota department of natural resources, section of wildlife, grand marais, minnesota). the wsi was defined as: sum of number of days <-17.8 degrees celsius + number of days with >38.1 cm of snow. f-tests were calculated to examine relationships between femur marrow fat content and wsi the preceding winter. all computer regressions, t-tests, and f-test calculations were completed using statistica 2000 (statsoft, tulsa, ok). results one-hundred four adults (37m, 67f), 47 yearlings (26m, 21f) and 45 calves (26m, 19f) were examined. calf marrows were obtained for all months except july, and marrows from only the femur and mandible were collected from october death samples. yearling samples were collected in all months except february, april, and december. adult marrows were collected in every month except march. linear regressions for bone marrow fat in each set of bones (paired with femur) in calves were highly correlated and significant (table 1). linear regressions for marrow fat in yearling moose femurs were significant and highly correlated for tibia, humerus, and radius bones. regressions in marrow fat between femur and mandible, table 1. regression equations for bone marrow fat percentage comparisons among bones in calf moose (<1 year old) from northeastern minnesota, 1972-2000. bone pair x (sd) n pairs regression equation r 2 p femur48.2 (24.7) 30 t = 4.8820 + 0.9976 (f) 0.94 <0.001 tibia 53.0 (26.3) femur50.0 (24.8) 35 m = 14.574 + 0.70435 (f) 0.91 <0.001 mandible 50.0 (19.6) femur48.2 (24.7) 30 h = 0.49474 + 1.0243 (f) 0.99 <0.001 humerus 49.9 (25.6) femur48.2 (24.7) 30 r = 0.2615 + 1.0579 (f) 0.96 <0.001 radius 50.8 (27.2) femur48.2 (24.7) 30 t = 15.012 + 0.80347 (f) 0.90 <0.001 tarsal 53.8 (22.0) femur48.2 (24.7) 30 c = 23.639 + 0.69442 (f) 0.85 <0.001 carpal 57.1 (20.2) moose bone marrow fat spears et al. alces vol. 39, 2003 276 tarsal, and carpal bones were significant but less correlated (table 2). for adults, linear regressions for marrow fat in each set of bones were significant and highly correlated (table 3; figs. 1-6). regression models (i.e., either slopes, intercepts, or both) among adults, yearlings, and calves differed for all bone pair comparisons except for the yearling-calf comparison in the femur-humerus bone pair (table 4). mean marrow fat in calves was lower in the femur than tarsal, humerus, carpal, and tibia bones and lower in the humerus than carpal (table 5). mean marrow fat in yearlings was lower in the femur than tarsal, humerus, and carpal bones, higher in the femur than mandible, and higher in the humerus than carpal (table 6). mean marrow fat in adults was lower in the femur than tarsal and carpal bones, higher in the femur than mandible, and lower in the humerus than carpal (table 7). femur marrow fat did not differ among moose killed by vehicle, accident, hunters, or unknown causes in calves (f 2,29 = 0.195; p = 0.824), yearlings (f 3,35 = 1.081; p = 0.370), or adults (f 3,58 = 2.372; p = 0.080). these mortality categories were therefore combined as accidental deaths. no relationship existed between age in years and femur fat in accidentally killed males (f 1,61 = 0.042; p = 0.838). femur marrow fat in accidentally killed females increased with age (y = 72.202 + 1.702(age); r2 = 0.112; f 1,67 = 8.429; p = 0.005). femur marrow fat was lower in accidentally killed calves than accidentally killed yearling or adult moose (fig. 7). femur marrow fat and sample size in each age class by month is shown in figure 8. femur marrow fat in calf moose ( x = 52.500 + 24.035; range = 5-88) differed by month (f 10,31 = 4.117; p = 0.001) and season (f 3,38 = 5.020; p = 0.005), but not year (f 20,21 = 1.342; p = 0.254). femur fat content differed among months (f 10,31 = 4.117; p = 0.001) and seasons in calves killed accidentally. calves had higher femur fat in seasons 3-4 ( x = 64.250 + 19.461; n = 16) than 1-2 ( x = 42.269 + table 2. regression equations for bone marrow fat percentage comparisons among bones in yearling moose (1-2 years old) from northeastern minnesota, 1972-2000. bone pair x (sd) n pairs regression equation r2 p femur78.5 (11.1) 28 t = 16.739 + 0.815 (f) 0.81 <0.001 tibia 80.7 (11.1) femur78.7 (10.5) 31 m = 44.191 + 0.265 (f) 0.42 0.028 mandible 65.0 (6.7) femur78.9 (11.1) 27 h = 0.9519 + 0.899 (f) 0.93 <0.001 humerus 80.4 (10.7) femur78.9 (11.1) 27 r = 11.156 + 0.8759 (f) 0.86 <0.001 radius 80.2 (11.3) femur78.9 (11.1) 27 t = 49.722 + 0 .445 (f) 0.63 <0.001 tarsal 85.0 (7.9) femur78.8(11.1) 27 c = 68.058 + 0.24492 (f) 0.53 0.004 carpal 87.4 (5.1) alces vol. 39, 2003 spears et al. moose bone marrow fat 277 table 3. regression equations for bone marrow fat percentage comparisons among bones in adult moose (>2 years old) from northeastern minnesota, 1972-2000. bone pair x (sd) n pairs regression equation r2 p femur71.0 (26.3) 62 t = 5.2776 + 0.94322 (f) 0.96 <0.01 tibia 72.3 (25.9) femur71.3 (26.2) 63 m = 19.277 + 0.60666 (f) 0.86 <0.001 mandible 78.4 (19.7) femur70.2 (27.1) 70 h = 0.98791 + 0.99693 (f) 0.99 <0.01 humerus 61.6 (18.8) femur70.7 (26.4) 61 r = 4.3726 +0.94837 (f) 0.97 <0.01 radius 71.4 (25.8) femur71.3 (26.2) 63 t = 32.581 + 0.64246 (f) 0.85 <0.001 tarsal 78.4 (19.7) femur71.3 (26.2) 63 c = 36.891 + 0.59844 (f) 0.83 <0.001 carpal 79.6 (19.0) humerus = 0.92751 + 0.99875(femur) femur h um er us 0 20 40 60 80 100 120 0 20 40 60 80 100 120 fig. 1. relationship between percent marrow fat in the femur and humerus for adult moose from northeastern minnesota, 1972-2000. tarsal = 32.539 + 0.64380(femur) femur t ar sa l 0 20 40 60 80 100 0 20 40 60 80 100 120 fig. 2. relationship between percent marrow fat in the femur and tarsal for adult moose from northeastern minnesota, 1972-2000. mandible = 18.476 + 0.60800(femur) femur m an di bl e 0 20 40 60 80 100 120 0 20 40 60 80 100 120 fig. 3. relationship between percent marrow fat in the femur and mandible for adult moose from northeastern minnesota, 1972-2000. tibia = 6.4169 + 0.93866(femur) femur t ib ia -10 10 30 50 70 90 110 0 20 40 60 80 100 120 fig. 4. relationship between percent marrow fat in the femur and tibia for adult moose from northeastern minnesota, 1972-2000. moose bone marrow fat spears et al. alces vol. 39, 2003 278 24.029; n = 26) (t 40 = -2.664; p = 0.011). calves killed by accident ( x = 60.563 + 18.371; n = 32) had more femur fat than those killed by disease ( x = 24.222 + 22.174; n = 9) (t 39 = 5.013; p <0.001). only one calf killed by wolves was examined. femur marrow fat for this individual was 49%. femur marrow fat among all yearling moose killed ( x = 77.732 + 11.194 range = 46-92) did not differ by month (f 8,32 = 1.453; p = 0.213), season (f 3,37 = 0.634; p = 0.598), or year (f 14,26 = 0.898; p = 0.571). femur fat did not differ among months (f 7,30 = 1.627; p = 0.166), seasons (f 3,34 = 0.598; p = 0.620), or years (f 14,23 = 0.825; p = 0.638) for yearlings accidentally killed. only one yearling killed by wolves and one killed by disease were examined. femur marrow fat for yearling moose killed accidentally averaged 78% + 11.4 (n = 38). femur marrow fat content was 68% for the wolf-killed yearling and 73% for the yearling dying of disease. for adults, number of moose examined radius = 3.3440 + .96917(femur) femur r ad iu s -10 10 30 50 70 90 110 0 20 40 60 80 100 120 fig. 5. relationship between percent marrow fat in the femur and radius for adult moose from northeastern minnesota, 1972-2000. carpal = 29.937 + 0.68225(femur) femur c ar pa l -10 10 30 50 70 90 110 0 20 40 60 80 100 120 fig. 6. relationship between percent marrow fat in the femur and carpal for adult moose from northeastern minnesota, 1972-2000. table 4. f-test degrees of freedom and f-values for tests of equality of models in moose bone pairs from northeastern minnesota, 1972-2000. all tests evaluated at alpha = 0.05. adults vs. yearlings adults vs. calves yearlings vs. calves bone pair df f df f df f femur2, 85 5.50 2, 88 138.60 2, 53 0.041 humerus femur2, 97 592.45 2, 101 741.30 2, 62 11.81 mandible femur2, 208 2864.60 2, 114 999.70 2, 148 2381.40 tibia femur2, 86 94.3 2, 89 134.86 2, 53 321.37 carpal femur2, 86 83.55 2, 89 111.48 2, 53 159.23 tarsal femur2, 84 13.57 2, 87 141.73 2, 53 5.14 radius 1failed to reject h o : models tested are equal. alces vol. 39, 2003 spears et al. moose bone marrow fat 279 table 5. means, standard deviations, differences between, and t-tests for mean differences of fat content in calf moose (<1 year old) bones from northeastern minnesota, 1972-2000. bone pair n pairs x sd diff sd diff t p femur30 48.233 24.713 tarsal 53.767 21.970 -5.533 10.582 -2.864 0.008 femur35 49.971 24.823 mandible 49.771 19.629 0.200 11.552 0.102 0.919 femur30 48.233 24.713 humerus 49.900 25.578 -1.667 3.717 -2.456 0.020 femur30 48.233 24.713 radius 50.767 27.211 -2.533 7.678 -1.808 0.081 femur30 48.233 24.713 carpal 57.133 20.190 -8.900 13.044 -3.737 <0.001 femur30 48.233 24.713 tibia 53.000 26.265 -4.767 9.058 -2.882 0.007 humerus30 49.900 25.578 radius 50.767 27.211 -0.867 7.776 -0.610 0.546 humerus30 49.900 25.578 carpal 57.133 20.190 -7.233 13.723 -2.887 0.007 table 6. means, standard deviations, differences between, and t-tests for mean differences of fat content in yearling moose (1-2 years old) bones from northeastern minnesota, 1972-2000. bone pair n pairs x sd diff sd diff t p femur27 78.851 11.128 tarsal 84.963 7.862 -6.111 8.657 -3.668 0.001 femur31 78.677 10.537 mandible 65.032 6.681 13.645 9.841 7.720 <0.001 femur27 78.852 11.128 humerus 80.444 10.714 -1.593 3.983 -2.078 0.048 femur27 78.852 11.128 radius 80.222 11.274 -1.370 5.832 -1.221 0.233 femur27 78.852 11.128 carpal 87.370 5.123 -8.519 9.456 -4.681 <0.001 femur28 78.500 11.077 tibia 80.679 11.049 -2.179 6.700 -1.721 0.097 humerus27 80.444 10.714 radius 80.222 11.274 0.222 5.199 0.222 0.826 humerus27 80.444 10.714 carpal 87.370 5.122 -6.926 9.583 -3.755 <0.001 moose bone marrow fat spears et al. alces vol. 39, 2003 280 and mean femur marrow fat percentages by season, month, and cause of death are given in table 8. femur marrow fat percent in adult moose ( x = 69.477 + 26.202; range = 8-95) did not differ among months (f 10,75 = 1.47; p = 0.167), seasons (f 3,82 = 2.355; p = 0.078), or years (f 24,61 = 1.43; p = 0.133). femur fat did not differ among months (f 9,52 = 1.325; p = 0.247), seasons (f 3,58 = 0.737; p = 0.534), or years (f 21,40 = 1.746; p = 0.064) for adults dying from accidental causes. femur marrow fat was higher in adult moose dying accidentally ( x = 79.210 + 16.861; n = 62) than those dying from disease ( x = 45.380 + 28.409; n = 21) (t25 = 5.158; p <0.001). femur marrow fat did not statistically differ between adult moose dying from accidental causes or those killed by wolves ( x = 37.00 + 42.673; n = 3) (t2 = 1.707; p = 0.228), or between those killed by wolves or disease (t 22 = 0.453; p = 0.655) (fig. 9). mean adult male femur fat within years was loosely related to yearly wsi (fig. 10). mean yearly adult female femur fat content was not related to yearly wsi (f 1,18 = 0.3692; p = 0.551). discussion the relationships among moose femur bone marrow fat and other examined moose table 7. means, standard deviations, differences between, and t-tests for mean differences of fat content in adult moose (>2 years old) bones from northeastern minnesota, 1972-2000. bone pair n pairs x sd diff sd diff t p femur63 71.254 26.175 tarsal 78.413 19.727 -7.159 13.860 -4.100 <0.001 femur70 70.157 27.103 mandible 61.600 18.787 8.557 14.017 5.108 <0.001 femur62 71.016 26.320 humerus 71.855 26.522 -0.839 3.521 -1.875 0.066 femur61 70.721 26.435 radius 71.443 25.821 -0.721 6.330 -0.890 0.377 femur63 71.254 26.175 carpal 79.603 18.966 -8.349 14.917 -4.443 <0.001 femur62 71.016 26.320 tibia 72.290 25.895 -1.274 7.446 -1.347 0.183 humerus62 71.790 26.472 radius 71.758 25.729 0.032 6.724 0.038 0.970 humerus63 72.143 26.406 carpal 79.603 18.966 -7.460 15.169 -3.904 <0.001 age class f em ur m ar ro w f at p er ce nt 52 58 64 70 76 82 88 calf (n = 32) yearling (n = 38) adult (n = 62) ± 95% ci ± se mean f2,129 = 15.857; p <0.001 fig. 7. femur marrow fat percent and age class in moose kills classified as accidental from northeastern minnesota, 1972-2000. alces vol. 39, 2003 spears et al. moose bone marrow fat 281 table 8. mean femur bone marrow fat percentages of adult moose (>2 years old) by season, month, and cause of death from northeastern minnesota, 1972-2000 (standard deviations in parentheses; sample sizes appear below marrow fat percentages). mortality all accidental disease wolf kill season dec-feb 56.55 (33.99) 79.78 (16.40) 35.44 (31.05) 47.00 (55.15) 20 9 9 2 mar-may 76.83 (19.67) 83.60 (11.84) 43.00 (0.00) 6 5 1 jun-aug 73.76 (22.73) 81.56 (15.47) 43.83 (25.79) 33 27 7 sep-nov 71.04 (22.62) 74.37 (20.65) 65.60 (23.19) 17.00 (0.00) 27 19 5 1 month jan 46.71 (26.20) 76.20 (20.52) 15.00 (9.90) 8.00 (0.00) 8 5 2 1 feb 55.20 (24.50) 88.00 (0.00) 46.20 (25.47) 6 1 5 mar apr 43.00 (0.00) 43.00 (0.00) 1 1 m a y 83.60 (11.84) 83.60 (11.84) 5 5 jun 71.36 (24.45) 78.33 (18.26) 29.50 (3.54) 14 12 2 jul 75.50 (21.60) 80.44 (15.81) 31.00 (0.00) 10 9 1 a u g 79.00 (23.01) 89.67 (2.25) 57.67 (32.89) 9 6 3 sep 82.44 (11.61) 84.57 (6.92) 78.33 (18.90) 9 7 2 oct 64.08 (26.55) 66.45 (23.97) 85.00 (0.00) 17.00 (0.00) 131 11 1 1 nov 64.08 (26.55) 83.33 (6.43) 46.50 (14.85) 5 3 2 dec 66.00 (35.06) 85.25 (11.07) 33.67 (42.74) 86.00 (0.00) 8 4 3 1 1includes 2 unknown causes of death. moose bone marrow fat spears et al. alces vol. 39, 2003 282 bones were significant and highly correlated for calves, yearlings, and adults. our study corroborates findings by others that bone marrow fat in the mandible and leg bones can be just as useful as that of the femur in determining relative health of moose individuals (snider 1980, ballard et al. 1981). unlike davis et al.'s (1987) findings in caribou (rangifer tarandus), we found significant differences among regression lines of adults, yearlings, and calves in all bone pair comparisons except for the yearling-calf comparison in the femur-humerus bone pair (table 4). this is in agreement with ballard et al. (1981), suggesting that calves, yearlings, and adults may deposit and mobilize bone marrow fat differently. our study corroborated previous findings that bone marrow fat was deposited first and mobilized last in distal bones in ungulates (cheatum 1949, peterson et al. 1982, and ballard 1995). femur and humerus (proximal) fat was lower in all age classes than fat in tarsal and carpal (distal) bones, respectively, suggesting that fat is deposited in distal bone marrow first in calves, and depleted first from proximal bone marrow in older moose. however, femur fat did not differ from tibia fat in yearlings or adults, and humerus fat did not differ from radius fat in any age class. peterson et al. (1982) indicated that while marrow fat withdrawal was sequential from proximal to distal bones in many ungulates, this pattern was not as marked in moose. overall, calf femur marrow content increased after birth (1 june) and peaked in november, a trend also observed in moose in southcentral alaska (ballard and whitman 1987). femur fat significantly decreased following the first winter, and increased as moose became yearlings. yearling femur marrow fat remained relatively constant throughout the year. although not statistically different, average femur marrow fat values seemed to increase in the spring and fig. 9. femur marrow fat percent and cause of death of adult moose from northeastern minnesota, 1972-2000. cause of death f em ur m ar ro w f at p er ce nt -20 0 20 40 60 80 100 accident (n = 62) diseased (n = 21) wolf kill (n = 3) ± 95% ci ± se mean fig. 10. yearly average adult male femur fat percent vs. the preceding winter's winter severity index for moose from northeastern minnesota, 1972-1999. winter severity index f em ur f at p er ce nt 40 50 60 70 80 90 100 60 100 140 180 220 260 f1,13 = 3.365; p = 0.090 y = 78.321-0.092*x+eps r2 = 0.206 fig. 8. femur bone marrow fat percent by month from moose in northeastern minnesota, 19722000. sample sizes are given in table below graph. alces vol. 39, 2003 spears et al. moose bone marrow fat 283 summer months while declining throughout the winter months for adult moose, corresponding to the timing of increased and decreased availability of high-quality forage (schwartz and renecker 1998). this trend was also observed in moose in alaska (franzmann and arneson 1976). previous authors suggested that bone marrow fat content can be used as an index of relative health for individuals within a population (cheatum 1949, ballard 1995). assuming marrow fat values (as an index of animal condition) had no bearing on moose killed by accident, we can use marrow fat content in these random animals as an index for animals of normal health (or "healthy") within the population. calf and adult moose we examined killed by disease had lower femur marrow fat than that of healthy individuals, supporting the use of femur marrow fat as a relative health index. we were only able to examine 1 calf and 1 yearling killed by wolves, and were therefore unable to statistically compare femur marrow fat content in calves and yearlings killed by accident or disease and wolves. using femur fat as an index of health for adults, we did not find statistical evidence that wolves hunting moose were singling out sick individuals from this population. our results were similar to franzmann and arneson (1976) and ballard et al. (1987), which found no difference in marrow fat values between wolf-killed and accidentally-killed adult female moose in alaska. however, our low sample size of wolf-killed adult moose examined here (n = 3) may not be a good indication of moosewolf interactions in this population. several studies suggested universal femur marrow fat content ranges that indicate animals that are healthy or are in poor condition. bischoff (1954) classified most mule deer (odocoileus hemionus) individuals with 80-100% bone marrow fat as in fair or poor condition. mech et al. (1995) suggested that caribou individuals with <7087% femur fat content had depleted muscle or fat reserves, and were in marginal condition. franzmann and arneson (1976) and peterson and bailey (1984) suggested that moose with <20% femur marrow fat were dying of starvation. mean femur fat for normal representative moose calves we examined was 64% before their first winter and 42% after. femur marrow fat content did not statistically change throughout the year for yearling or adult moose we examined and used as normal representatives. mean values for these animals were 77% and 79%, respectively. given that these animals were representative of normal individuals, our results are in contrast with suggestions by previous authors that these animals may be in an abnormal sub-healthy state. ballard (1995) indicated that a declining trend in bone marrow fat content by late winter and early spring was common for many northern ungulate populations and relatively low values might be considered normal for many populations. alternatively, the relative condition of this moose population as a whole may be diminished due to factors we did not examine, such as a possible lack of highly nutritious forage or other stressors. marrow fat content has been observed at or near 100% in caribou, moose, and deer individuals (bischoff 1954, peterson et al. 1982, davis et al. 1987, this study). however, a baseline marrow fat content for healthy animals has not been established against which to judge individuals from different populations. a universal baseline for using marrow fat content in moose as an overall index of health may only be determined through examination of total body fat depletion patterns in animals at different states of health (ballard 1995). acknowledgements data for this paper were collected while author peterson was employed as area moose bone marrow fat spears et al. alces vol. 39, 2003 284 wildlife manager at grand marais, mn by the minnesota department of natural resources, section of wildlife. he appreciates having been permitted the freedom of action to pursue this and other studies that were not covered by formal research proposals. we thank rick fields, tim webb, and dave ingebrigtsen who helped collect data, and pam coy who processed marrows in early years of the study. special thanks are due peterson's wife, dale peterson, who gave up many nights, weekends, and holidays to help examine moose and collect samples in addition to putting up with having her oven used for drying marrows during the later years of the study. this is texas tech university, college of agricultural sciences and natural resources publication t-9-977. references baker, m. f., and f. x. leuth. 1966. mandibular cavity tissue as a possible indicator of condition of deer. proceeding of the annual southeast game and fish commissioners 20:69-74. ballard, w. b. 1995. bone marrow fat as an indicator of ungulate condition-how good is it? alces 31:105-109. , c. l. gardner, j. h. westlund, and s. m. miller. 1981. use of mandible versus longbone to evaluate percent marrow fat in moose and caribou. alces 17:147-164. , and j. s. whitman. 1987. marrow fat dynamics in moose calves. journal of wildlife management 51:66-69. , , and c. l. gardner. 1987. ecology of an exploited wolf population in south-central alaska. wildlife monographs 98. bischoff, a. i. 1954. limitations of the bone marrow technique in determining malnutrition in deer. proceedings of the western association of state game and fish commissioners 34:205-210. cheatum, e. l. 1949. bone marrow as an index of malnutrition in deer. new york state conservation 3:19-22. davis, j. l., p. valkenburg, and d. j. reed. 1987. correlations and depletion patterns of marrow fat in caribou bones. journal of wildlife management 51:365371. franzmann, a. w., and p. d. arneson. 1976. marrow fat in alaskan moose femurs in relation to mortality factors. journal of wildlife management 40:336339. graybill, f. a. 1976. theory and application of the linear model. duxbury press, north scituate, massachusetts, usa. hayes, r. d., a. m. bayer, and d. g. larson. 1991. population dynamics and prey relationships of an exploited and recovering wolf population in the southern yukon. yukon fish and wildlife branch final report tr-91-1. whitehorse, yukon, canada. mech, l. d., and g. d. delgiudice. 1985. limitations of the marrow fat technique as an indicator of body condition. wildlife society bulletin 13:204-206. , t. j. meier, j. w. burch, and l. g. adams. 1995. patterns of prey selection by wolves in denali national park, alaska. pages 231-249 in l. s. carbyn, s. fritts, and d. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, university of alberta, edmonton, alberta, canada. neiland, k. a. 1970. weight of dried marrow as an indicator of fat in caribou femurs. journal of wildlife management 34:904-907. nelson, m. e., and l. d. mech. 1981. deer social organization and wolf predation in northeastern minnesota. wildlife monographs 77. peterson, r. o. 1977. wolf ecology and alces vol. 39, 2003 spears et al. moose bone marrow fat 285 prey relationships on isle royale. u.s. national park service science monograph serial 11. , d. l. allen, and j. m. dietz. 1982. depletion of bone marrow fat in moose and a correction for dehydration. journal of wildlife management 46:547-551. , and t. n. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88. schwartz, c. c. 1998. reproduction, natality and growth. pages141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d. c., usa. , and l. a. renecker. 1998. nutrition and energetics. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d. c., usa. smith, h. a., and t. c. jones. 1961. veterinary pathology. lea and febiger, philadelphia, pennsylvania, usa. snider, j. b. 1980. an evaluation of mandibular marrow fat as an indicator of condition in moose. proceedings of the north american moose conference and workshop 16:37-50. zar, j. h. 1999. biostatistical analysis. fourth edition. simon and schuster, upper saddle river, new jersey, usa. 4008.p65 alces vol. 40, 2004 edwards et al. ne minnesota moose management 23 northeastern minnesota moose management a case study in cooperation andrew j. edwards1, mike schrage2, and mark lenarz3 11854 authority, 4428 haines road, duluth, mn 55811, usa; 2fond du lac resource management division, 1720 big lake road, cloquet, mn 55720, usa; 3minnesota department of natural resources, 1201 east highway 2, grand rapids, mn 55744, usa abstract: this paper provides an overview of moose management in northeastern minnesota with an emphasis on relationships between the state and tribal entities that share management responsibility. specific topics discussed include settlement of treaty rights issues, harvest allocation and strategies, and the evolving state-tribal partnerships that have been created during the past 15 years. brief updates on the status of moose in minnesota, population monitoring efforts, population goals, and the future direction of management are provided. alces vol. 40: 23-31 (2004) key words: alces, allocation, management, minnesota, moose, partnerships, population, treaties, tribal moose (alces alces) are common in the northeastern portion of minnesota. a small population is also found in the northwestern portion of the state. previously much larger, the northwest population was hunted from 1971 to 1997 when the season was closed due to an unexplained, precipitous decline in that herd. however, moose numbers in northeastern minnesota have remained relatively stable and are sufficient to support hunting by both state-licensed and tribal hunters under several jurisdictions. historical superior national forest records for northeastern minnesota indicate that a dramatic increase in moose numbers occurred during the late 1920s but numbers plummeted in the mid-1930s, and remained low until the midto late 1960s (peek et al. 1976). population estimates, conducted from aerial surveys since 1960, suggest that the population gradually began to increase through the 1970s and 1980s to a peak of 6,900 in 1988 and then dropped sharply to 3,700 by 1990. recent surveys indicate that the population is relatively stable and at approximately 4,000 animals. human interactions with moose predate european settlement of the region. local bands of native americans relied heavily upon moose for subsistence. following settlement of the region, establishment of reservations, and the creation of the state of minnesota, moose management in its various forms fell on the shoulders of the state of minnesota. however, during the past two decades, various changes to the management process have occurred. this paper covers the structure and history of current moose management in northeastern minnesota with a focus on the structure of tribal management and cooperation between state and tribal management authorities. historical information on the development of the tribal management agencies is presented for background. management of moose within reservation boundaries will not be discussed in this paper. ne minnesota moose management – edwards et al. alces vol. 40, 2004 24 historical background the main population of moose in minnesota inhabits the forested northeastern region. much of the area is public land under state, county, and federal ownership. private individuals, paper companies, and native american chippewa bands control other land. habitat management is generally the responsibility of the landholder, though there are cases where other jurisdictions may have say in management activities that require review prior to issuing any necessary permits. outside of tribal reservation lands, harvest/population management primarily falls under the jurisdiction of the minnesota department of natural resources (dnr). prior to european settlement of the region, the local native american bands led a somewhat nomadic lifestyle, frequently moving their base camps from one area to another to take advantage of seasonally abundant foods. moose were a staple large game animal taken by the bands. as settlement of the region occurred, changes to the region, its wildlife, and its inhabitants took place. by the mid-1800s increasing pressure to settle the region led to numerous treaties between the u.s. government and the chippewa bands in the great lakes region. some of these treaties established reservation boundaries, while others ceded large portions of land to the u.s. government in exchange for various types of payments. on 30 september 1854, 10 of the lake superior chippewa bands signed a treaty ceding much of present-day northeastern minnesota to the united states government (fig. 1). article 11 of the treaty held in part, “and such of them as reside in the territory hereby ceded, shall have the right to hunt and fish therein, until otherwise ordered by the president” (glifwc 1992). in the years preceding and following the 1854 treaty, several additional treaties ceding lands and/or establishing reservations were signed by various bands in the great lakes region, some of which contained similar language retaining off-reservation rights to hunt and fish. of the 10 bands that signed the 1854 treaty, only the bois forte (two reservations, the main one just west of the ceded territory and a second smaller reservation within the ceded territory), grand portage, and fond du lac bands eventually ended up residing in the ceded territory (fig. 1). following this period of treaty signing, increased settlement of the region occurred and minnesota was established as the 32nd state of the union on 11 may 1858. the issue of treaty-reserved rights offreservation did not become a major issue in minnesota until late 1984. tribal hunting and fishing within reservation boundaries falls under the jurisdiction of individual bands and except for some federal regulations, little outside input enters into how resources are managed. the subject of who had the right to regulate tribal hunting and fishing off-reservation had not been addressed by the courts. all of that changed in december 1984. a grand portage band member legally hunting moose under tribal jurisdiction wounded a moose on what he believed to be the grand portage reservation. while tracking the wounded moose the hunter eventually realized that he was off the reservation and rapidly running out of light to pursue the moose. upon returning to the reservation, the band member talked with the tribal game warden and tribal council and told them that he intended to resume tracking the moose the following morning. the tribal council offered their support of the hunter in whatever might come from his actions. the band member also contacted the local state dnr game warden that evening and told him of his intentions. the following morning the hunter resumed tracking the moose but never was able to find it. alces vol. 40, 2004 edwards et al. ne minnesota moose management 25 fig. 1. territory ceded by the lake superior chippewa to the u.s. government in a 30 september 1854 treaty. reservation locations of the bois forte, grand portage, and fond du lac bands, which retain hunting and fishing rights within the 1854 ceded territory, are also shown. when the hunter returned to his truck the dnr warden was waiting for him. after discussing the events with the dnr warden, and taking a closer look at a map of reservation boundaries, it was determined that the band member was actually about a half-mile outside the reservation when he wounded the moose. at that time the hunter was issued a state citation for hunting moose out of season. as a result, in 1985, the grand portage band and 2 of its members filed a civil action in u.s. district court claiming that the state of minnesota had no jurisdiction over band members exercising their treaty reserved rights to hunt and fish in lands ceded under the 1854 treaty. in may 1987 the parties involved in the lawsuit requested that the court remove the matter from its trial calendar to give the parties time to negotiate an out-of-court settlement. the bois forte and fond du lac bands joined in the negotiations. in february 1988, after lengthy negotiations, the end result of the lawsuit was an out-of-court settlement between the 3 bands and the state of minnesota. the triband authority, an inter-tribal natural resource agency governed by the duly elected officials of the 3 bands, was formed to regulate the exercise of off-reservation treaty rights. under the terms of the agreement the bands agreed to forbear, or limit, the exercise of certain treaty rights in exchange for an annual monetary payment by the state of minnesota. included in the language of the agreement were provisions that outlined establishment of seasons, methods of take, and in some cases harvest limits. one of the terms of the agreement established that the moose season to be held by the bands would run concurrently with the state of minnesota’s season. this agreement marked the first time that the bands had any real say in how management of moose outside of reservation boundaries was handled. in 1989, the situation changed yet again, when the fond du lac band left the triband authority to pursue their own settlement with the state by utilizing a provision in the agreement that allowed any party to withdraw with a 1-year notice. the triband authority then became the 1854 authority, governed by the reservation councils of the bois forte and grand portage bands. this resulted in 3 separate management “agencies” that were responsible for moose management in the 1854 ceded territory. since leaving the tri-band authority the fond du lac band has continued to hunt moose in the 1854 ceded territory under seasons, regulations, and desired harvest levels established by the band. discussions with the state are ongoing and ultimately it is likely that a formal agreement will be reached clarifying their role in sharing the resource. both bands represented by the authority might also be involved in these discussions, depending on the proposed ne minnesota moose management – edwards et al. alces vol. 40, 2004 26 final terms of the agreement. under its current structure, the 1854 authority has 3 divisions: administrative, enforcement, and the biological services division. the fond du lac band also created their own resource management division and developed a conservation code regulating off-reservation hunting and fishing by their members. there is often a question of whether or not the various treaties, signed so long ago, still are legitimate law in our modern society. to answer that we will step back and examine several relevant legal rulings. first and foremost, as set forth in article vi of the united states constitution “…and all treaties made, or which shall be made, under the authority of the united states, shall be the supreme law of the land;…” (woods institute 2002). that would seem to set the question at rest. however, there are several specific court cases that have upheld the treaty rights issues. in a 1942 ruling, the u.s. supreme court ruled that since a treaty takes precedence over state law, indians with tribal rights cannot be required to buy a state license to exercise those rights (glifwc 1992). in 1969, a federal judge ruled that the state of oregon could only regulate tribal rights when “reasonable and necessary for conservation,” that state regulations should not discriminate against indians, and must be the least restrictive means (glifwc 1992). regionally, in 1983, the u.s. court of appeals for the 7th circuit reaffirmed that treaty rights to hunt, fish, and gather on ceded lands were reserved and protected through a series of treaties (glifwc 1992). in 1994, a u.s. district court decision upheld the mille lacs band of ojibwe’s treaty rights in the 1837 ceded territory (glifwc 2002). a 1996 ruling also found that the fond du lac band retained their treaty reserved rights to hunt, fish, and gather in the 1854 ceded territory (glifwc 2002). by default, this was also a verification of bois forte and grand portage claims to the same rights since they also signed the same treaty. most recently, in 1999, the u.s. supreme court upheld the rights of the mille lacs band and other signatory bands within the 1837 ceded territory (glifwc 2002). based on these decisions, the question of whether or not bands retain their treaty-reserved rights in the midwest has been essentially laid to rest. that said, there are still questions about how tribal management of rights and the underlying resources are to fit in with state management. questions will continue to evolve. overview of cooperation as mentioned previously, prior to 1984, no tribal off-reservation management of tribal hunting, fishing, and gathering in the 1854 ceded territory existed. that changed in 1988 with the out of court settlement and the formation of the tri-band authority, now the 1854 authority. initially, most of the cooperation between the bands and the state of minnesota was limited to enforcement and a single harvest report that followed the close of hunting and trapping seasons. with the development of the fond du lac resource management division in 1993 and the 1854 authority’s biological services division in 1994, a new era of cooperation began. at that time both the authority and fond du lac began building their respective natural resources staff to become more involved in off-reservation natural resources management. initial involvement began with attendance of tribal staff at season setting meetings and discussing state and tribal harvest as a combined effort. in 1995, the 1854 authority and fond du lac began contributing both funding and occasional manpower to the annual moose survey with the stipulation that there be actual involvement of the bands in the procalces vol. 40, 2004 edwards et al. ne minnesota moose management 27 ess. in 1997, the band personnel started participating in the planning for the survey and were included in setting seasons and quotas for state moose hunters. in 1998, fond du lac and authority personnel were included as permanent members of the survey crew and began working with the state to investigate the potential of expanding hunting opportunities through new zones. currently, the bands work with the state in both funding and staffing the annual survey and in evaluating annual harvest and setting season structure and quotas. in addition tribal biologists work with state wildlife managers to fund and implement habitat improvement projects. in 2002, a new partnership between the 1854 authority, the fond du lac band, the minnesota department of natural resources, and the u.s. geological survey was formed. the new partnership launched a 5-year study of the northeastern minnesota moose herd with an emphasis on causes and rates of nonhunting mortality. since 2002, 84 moose have been fitted with radio-collars as part of this study. areas of cooperation moose survey — surveys of the moose population in northeastern minnesota began in 1960 (lenarz 1998). survey protocols have changed slightly over the years, but have followed the same general methodology since 1984, with very consistent conditions since 1997. the purpose of the survey is to estimate numbers and age/ sex ratios for setting harvest levels and to investigate potential new hunting zones. currently, we use a stratified random block protocol. about 10,878 km2 of northeastern minnesota are considered to be moose range, almost all of which lies within the 1854 ceded territory (fig. 2). survey plots (ranging from 70 to 847 km2) are chosen randomly from 3 different strata (expected low, medium, and high winter density) in proportion to the total area of the moose fig. 2. northeast minnesota moose range in relation to the 1854 ceded territory. moose range is broken into aerial moose survey plots stratified according to expected winter moose density (low < 0.2 moose/km2, medium 0.2 – 0.6 moose/km2, high > 0.6 moose/km2). about 30 survey plots are flown each year. range that falls within each category. tribal and state staff jointly review plot stratification every 5 years. the survey itself is flown by two 3-person crews in cessna 185s. parallel transects 0.54 km apart are flown, with a circular resurvey area of 5.2 km2 being flown following the initial run. each moose located during the survey is sexed and located with the plane’s gps. although survey timing has varied considerably over the years, we now start the first possible working day in january. currently the bands contribute from 33 to 50% of the annual survey costs, in addition to providing 2 full time crew members. moose population estimates from 1985 to 2003 have ranged from 3,500 to 9,000 animals (fig. 3). ninety percent confidence intervals on the estimates have also varied from a low of 23% in 2002 to a high of 126% in 1995. conditions for the last several years have been similar, as has survey protocol, enabling us to make better comparisons among years. currently it appears that the herd is relatively stable around 4,000 animals. ne minnesota moose management – edwards et al. alces vol. 40, 2004 28 season setting and framework – state licensed hunters – the first modern harvest season was held in 1971 and was continued on an every-other-year basis until 1991. in 1991 the season was closed for 1 year following a large winter die-off mainly attributed to tick (dermacentor albipictus) associated mortality. at that point, it was also decided to go to a once-in-a-lifetime hunt, meaning that once you had received a license you were no longer eligible for future hunts, regardless of your success. in 1994 the decision was made to switch to annual hunts and permit levels were dropped to roughly half of their previous, everyother-year levels. prospective hunters are required to apply in parties of 2 – 4 hunters for a lottery process. demand for the licenses remains high, with odds of receiving a permit around 1 in 20 overall. annual hunts have been held since 1994, though the state cancelled its 2000 season due to budget constraints. season dates are set by law, with a current regulation that opens the season for 16 days on the saturday nearest october 1. each party receiving a permit is allowed to harvest 1 animal of any sex or age from the 1 zone for which their permit is valid. season setting and framework 1854 authority — one of the terms of the agreement established between the bands and the state of minnesota in 1988 was that the moose season to be held by the bands would run concurrently with the state season. at the time that the agreement was signed, the state of minnesota was holding a moose season in the northeastern portion of the ceded territory every-other-year. the agreement set a quota for tribal offreservation moose harvest at 60 animals every 2 years. this provision stated that in the event that the state went to an annual hunt, the band quota would be 30 moose per year. the tribal quota could increase if the state ever increased their permit levels, and fig. 3. estimated moose numbers (bars) and 90% confidence intervals (lines) from aerial surveys in northeastern minnesota, 1985 – 2003. would be done so on a proportional basis. the annual base permit level upon which the 1854 authority quota is based is 264 permits (1/2 of the 528 permits offered in 1987, the year prior to the agreement). in recent seasons, the state has been issuing roughly 200 permits per year, so it will likely be some time before an increase in the authority’s harvest quota can be discussed. in 1994, the state of minnesota did go to an annual hunt and, as a result, the bands’ quota moved to 30 moose annually. in 1989, the fond du lac band pulled out of the agreement with the other two bands and the state of minnesota. since that time the fond du lac band has continued to hunt moose in the ceded territory, but has yet to reach a final agreement with the other parties that would clarify their role in sharing the resource. currently, the 1854 authority must manage its moose season based on the current agreement, meaning an annual quota of 30 moose that may be taken during the same open season dates as the state. interest in moose hunting by 1854 authority licensed hunters is fairly high, so annual permits are distributed through a lottery process. band members must apply 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 19 85 19 87 19 89 19 91 19 93 19 95 19 97 19 99 20 01 20 03 year e st im at ed m oo se alces vol. 40, 2004 edwards et al. ne minnesota moose management 29 in parties of 3 or 4 for a single permit that is valid for 1 moose of any sex or age. if that party is chosen to receive a permit through the drawing, they are issued a permit for that year. the following year they are ineligible for the initial drawing. in the event that not enough applicants fill the drawing in a given year, the parties that were allowed to hunt the previous year are entered into a second chance drawing to fill any leftover permits. generally, band members retain the same party from year to year and most parties receive a permit every other year. parties are allowed to hunt within any open state hunting zone. however, most parties tend to hunt in the same areas from year to year, allowing us to make some predictions as to where the 1854 authority moose harvest is likely to occur. permit numbers are based on average success rate from the previous 3-year period. for example, the average success rate the last 3 years has been 61%, so the authority issued 49 permits in 2003. included in the annual permit numbers are 4 subsistence permits which are distributed to the 1854 authority main office, and each of the 3 bois forte and grand portage reservations without competition, allowing them to harvest a moose to provide meat for tribal elders. in essence, this means that in 2003, the authority issued 49 permits in hopes of obtaining a moose harvest of 30 animals. of those 49 permits 45 were available to band members through the lottery process. there are currently no fees for band members to obtain a license, and all duly enrolled bois forte and grand portage band members are eligible to apply for a permit provided they have an 1854 authority identification card, are not restricted from hunting due to court imposed sanctions, and meet hunter safety certificate requirements. season setting and framework fond du lac — the fond du lac band opens their moose season on the same day as the state and the 1854 authority. however, they generally run for an additional 8 weeks. currently fond du lac has 10 moose hunting zones that combine 2 – 5 state zones. permits are allocated through a drawing. the number of permits available for each zone varies depending on estimated moose numbers and past success rates for the zone. band members must apply for permits in parties of 3 or 4 hunters, indicating their preferred zones on their application. a $20 non-refundable fee is required to enter the drawing. as parties are drawn, they are assigned a permit for a given zone, following the preference indicated on their application. all parties receiving a permit are required to provide a $50 deposit that is returned if they properly register a moose or return their unused tag following the season. proceeds from the drawing are used to offset costs incurred in the annual moose survey. although the number of permits varies from year to year, recently it has been around 70 permits annually. the fond du lac band also has treaty reserved rights to hunt in the 1837 ceded territory that lies to the immediate south of the 1854 ceded territory. moose populations in the 1837 ceded territory are very low. although there is a tribal quota of 5 moose available for the 1837 ceded territory and 5 permits are made available annually, there has been little interest in the permits and no harvest thus far. harvest allocation — currently, the state of minnesota issues permits to their hunters based on a harvest goal of 5% of the moose population in each zone. estimates of each zone’s population are conducted annually, based on the results of the annual aerial survey. local wildlife managers and band biologists review the estimates and adjust them accordingly if their experience indicates that the estimates are biased high or low. at that point, the state designates 5% of the moose as harvestable. the ne minnesota moose management – edwards et al. alces vol. 40, 2004 30 biologists from the bands are then asked to give their best estimate as to how many moose tribal hunters will harvest from each state zone for the upcoming season. for example, if the population estimate for zone 1 is 200 moose, state considers the total harvestable surplus as 10 moose. if total tribal harvest is expected to be 5 moose, the remaining harvestable surplus for state hunters is 5 moose. the state then looks at their average success rates for the previous 3 seasons in that zone and issues permits based on that success rate to attempt to harvest the remaining surplus. in this example, assuming that the success rate by state hunters in zone 1 has averaged 50% in the past, the state would designate 10 permits for that zone. this process is a little tricky due to some complicating factors. success rates by state hunters can vary dramatically from year to year in any one zone. another factor is the difference in permit zoning by the 3 bands and the state. state hunters are only allowed to hunt in the state permit zone for which they are drawn. fond du lac hunters are currently required to hunt within the fond du lac moose zone for which they are drawn. however the fond du lac zones encompass 2 5 state zones. hunters receiving a permit from the 1854 authority may hunt in any state zones. therefore, it is currently difficult to predict exactly what the maximum tribal harvest from any one state zone will be. however, based on knowledge of moose numbers, access, permit numbers, tribal harvest histories, and past success rates, tribal biologists can be reasonably close in their predictions of where harvest will occur. the other complicating factor occurs when predicted tribal harvest from a state zone may exceed the 5% harvest goal of the state. in that situation, the state generally will offer at least 1 permit to state hunters. in cases where this is proposed, the permits are fig. 4. total (state and tribal) northeast minnesota moose harvests by state licensed hunters, 1854 authority licensed hunters, and fond du lac licensed hunters, 1971 2003. offered only if there is consensus among all involved that doing so is not likely to result in any long-term damage to the local population. despite the complications in predicting harvest and the involvement of three management agencies, we have been fairly successful in keeping harvest around 5%. since 1997, at the point at which we feel the herd has stabilized at about 4,000 animals (fig. 3), total hunting mortality has been approximately 200 moose per year, or 5% (fig. 4). summary the development of trust and cooperation between the state of minnesota and the native american bands did not happen overnight. it required time, commitment to reaching a workable arrangement, and in some cases litigation or threat of litigation. cooperative management of moose between the state of minnesota and the 3 bands retaining off-reservation harvest rights to moose in the 1854 ceded territory continues to evolve. great strides have been made since 1988 when the initial agreement was signed between the state and the bands. increasing levels of trust on both sides, as well as the involvement of all parties in monitoring and managing the resource has led to a good working situation. personalities of involved personnel also play a critical role and, fortunately, the current 0 50 100 150 200 250 300 350 400 450 500 19 71 19 73 19 75 19 77 19 79 19 81 19 83 19 85 19 87 19 89 19 91 19 93 19 95 19 97 19 99 20 01 20 03 year h ar ve st ed m oo se fdl 1854 state alces vol. 40, 2004 edwards et al. ne minnesota moose management 31 suite of players involved at the local level work well together. of course, it is unlikely that all parties will agree on all issues over time. future changes in the moose herd, changes in tribal and state allocation, or other unforeseen issues may invite conflict. however, given the current working relationships between the natural resources staff of the bands and the state of minnesota there is a better chance that the parties involved will be able to reach workable arrangements without further litigation. in short, we have elevated co-existence to cooperation much to everyone’s benefit. acknowledgements the authors would like to thank several anonymous reviewers for their insightful reviews of this manuscript. references (glifwc) great lakes indian fish and wildlife commission. 1992. a guide to understanding chippewa treaty rights, minnesota edition. glifwc, po box 9, odanah, wisconsin 54861, usa. _____. 2002. a guide to understanding chippewa treaty rights. 2003 edition. glifwc, po box 9, odanah, wisconsin 54861, usa. lenarz, m. s. 1998. precision and bias of aerial moose surveys in northeastern minnesota. alces 34:117-124. peek, j. m., d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. woods institute. 2002. the constitution of the united states with index and the declaration of independence. woods institute, 2231 california street, n.w., washington, d.c. 20008, usa. f:\alces\vol_39\p65\3924.pdf alces vol. 39, 2003 persson – visibility of moose pellets 233 seasonal and habitat differences in visibility of moose pellets inga-lill persson department of animal ecology, swedish university of agricultural sciences, se-90183 umeå, sweden abstract: counts of faecal pellet groups have been widely used to estimate population densities and trends of large ungulates like moose (alces alces). the visibility of pellet groups affects the accuracy of estimates, decreases with time, and varies among habitat types. i investigated the impact of season and habitat type on how, over time, visibility of moose pellets decreased along a forest productivity gradient in boreal forests of northeastern sweden. visibility decreased at the fastest rate during the transition from spring to summer due to concealment by new vegetation. visibility also varied significantly among habitat types and was correlated with vegetative litter production. after one winter of exposure, more than 95% of all pellet groups were visible independent of habitat type, but thereafter visibility decreased fast in more productive habitats. the results demonstrated that if study plots are cleared in late autumn after the vegetation period and then visited as soon as possible after snowmelt, pellet counts can be used to estimate population trends and habitat use of moose in winter without bias caused by differences in visibility within different habitat types. also, the correlation with litter production suggests that if a sightability correction factor is developed, pellet counts could be used to estimate habitat use and population distribution during the vegetation period and with longer periods between plot visits. alces vol. 39: 233-241 (2003) key words: alces alces, habitat, management, moose, pellet counts, productivity, season, visibility counts of faecal pellet groups have been widely used to estimate population densities and trends of large ungulates ( w a l l m o e t a l . 1 9 6 2 , n e f f 1 9 6 8 , timmermann 1974, harestad and bunnell 1987). it is a less expensive technique than p o p u l a t i o n e s t i m a t e s f r o m a i r c r a f t (härkönen and heikkilä 1999) and might be more precise in forests with dense overstory vegetation (jordan et al. 1993), although there are methodological problems (neff 1968). the visibility of pellet groups is an important factor affecting the estimates from pellet counts (wallmo et al. 1962, lehmkuhl et al. 1994), but is rarely attributed much importance (harestad and bunnell 1987, aulak and babinska-werka 1990). visibility decreases with time as a result of concealment by vegetation and decay processes. the commonly used “clearanceplot” method is based on faecal accumulation in previously cleared plots. correcting for pellet groups which have become invisible between consecutive plot visits is critical for the reliability of the method (massei et al. 1998). how fast visibility decreases depends to a great extent on habitat type (smith 1968, lavsund 1975, harestad and bunnell 1987, lehmkuhl et al. 1994, massei et al. 1998). visibility decreases faster in moister habitats with more vegetation than in dry habitats with scarce vegetation (lavsund 1975, harestad and bunnell 1987, lehmkuhl et al. 1994). faster concealment in more visibility of moose pellets – persson alces vol. 39, 2003 234 vegetated habitats is suggested to be a primary factor contributing to observed differences in visibility of pellet groups (harestad and bunnell 1987), and vegetative litter is especially important (lavsund 1975). the decrease of visibility also varies by season and decreases faster during the growing season and in autumn than in winter (aulak and babinska-werka 1990, massei et al. 1998). there is a general lack of information on the fate of moose pellets, especially in areas where moose densities are high and where pellet counts could be an alternative to helicopter surveys and other methods of population estimation. my study objective was to estimate seasonal and habitat differences in visibility of moose pellets (i.e., the percentage of the original surface area of each pellet group still visible). more specifically, i tested how fast visibility decreased in spring, summer, and autumn respectively, if visibility was correlated to habitat productivity, and if some plants and substrates (i.e., the ground cover the pellet groups were lying on) contributed more than others to a decrease in visibility. study area the study was done in the middle boreal zone (ahti et al. 1968) of coastal northern sweden (fig. 1), at 8 sites (table 1) situated 50 90 km north and north-west of umeå (63°50' n, 20°18' e). the length of the vegetation period (average day temperature > 5° c) was 120 150 days with onset between 10 and 20 may. yearly precipitation was 600 700 mm (raab and vedin 1996), and precipitation during the vegetation period was 300 350 mm (nilsson 1996). snow covered the ground approximately from 20 25 october to 5 15 may (raab and vedin 1996). all sites used in this investigation were young forest stands of scots pine (pinus sylvestris), interspersed with various deciduous trees; mainly b i r c h e s ( b e t u l a p u b e s c e n s a n d b . p e n d u l a ) , b u t a l s o r o w a n ( s o r b u s aucuparia), aspen (populus tremula), and willows (salix spp.). raspberry (rubus idaeus), wavy hair grass (deschampsia f l e x u o s a ) , b l u e b e r r y ( v a c c i n i u m myrtillus), lingonberry (v. vitis-idaea), heather (calluna vulgaris), and fireweed (epilobium angustifolium) were common. to prevent bias caused by differences in soil moisture among sites, the sites were selected to have as little slope as possible. methods the study sites (table 1) were selected to represent a forest productivity gradient, covering the range of forest types in the region (hägglund and lundmark 1987, fridman et al. 2001). site productivity was estimated as site index (estimated top height at 100 years) for scots pine using methods developed for young forest stands (lindgren et al. 1994, elfving and kiviste 1997). site index is a common measurement of habitat productivity in forestry in sweden (persson fig. 1. map of sweden showing the study area. # umeå # stockholm n ew s study area alces vol. 39, 2003 persson – visibility of moose pellets 235 table 1. study sites ranked by increasing site index for scots pine (mean top height in meters at 100 years). litter production (g dry mass per m2 and year; all treatment plots per exclosure pooled) estimated 2001-02 (persson 2003), mean age of trees (years), geographic location (wgs84), altitude (m above sea level), and major tree species present. site site litter mean geographic altitude tree index age location species1 lögdåberget 12.9 18.08 16 64o 00' n, 18 o 45' e 300 b, p, po, (s, sa) skatan 14.7 15.85 9 64 o 13' n, 19 o 09' e 265 b, p, (po, sa) djupsjöbrännan 24.3 15.83 9 64 o 06' n, 19 o 12' e 250 b, p, sa åtmyrberget 24.8 55.02 9 64 o 12' n, 19 o 17' e 305 b, p, s, sa, (po) selsberget 26.3 11.40 7 64 o 15' n, 19 o 16' e 175 b, p, s, (po, sa) mörtsjöstavaren 26.4 56.38 7 64 o 22' n, 20 o 07' e 280 b, p, s, sa ralberget 27.3 27.15 9 64 o 13' n, 20 o 42' e 250 b, p, s, sa, (po) rönnäs 27.9 12.01 9 64 o 02' n, 20 o 40' e 62 b, p, (po) 1 b = betula spp., p = pinus sylvestris, po = populus tremula, s = sorbus aucuparia, and sa = salix spp. tree species occurring sparsely are in brackets. 2003). however, site index is developed for conifers, and studies have shown that conifers and hardwoods have fundamentally different soil-plant interactions (ollinger et al. 2002). litter production might be a better biological measurement of habitat productivity (persson 2003). thus, i also tested for relationships between the visibility of pellet groups and vegetation litter production in each plot. the study of pellet visibility was done in connection with an experimental study, where browsing, defecation and urination of different levels of moose population density were simulated in 8 exclosures (i.e., the study sites) along a forest productivity gradient (persson 2003). to simulate defecation, 39 pellet groups of 0.8 litres were laid out at each study site on 4 occasions (may 1999, october 1999, june 2000, and august 2000). the number of pellet groups was estimated based on defecation rates of moose (persson et al. 2000). the pellet groups were deposited within circles of 35 cm in diameter and marked with a plastic stick in the centre. the moose pellets were collected from a nearby moose farm. pellet groups showing clear evidence of decomposition were not collected. the animals were using mainly natural habitats and had free access to natural food (nyberg and persson 2002). it was thus an experimental study with pellet groups of equal size and origin. twelve randomly selected pellet groups from each of the 4 age classes of pellets (i.e., pellet groups laid out may 1999, october 1999, june 2000, and august 2000) were investigated at each site; 48 pellet groups per site and 384 for the whole study (one pellet group was not found in spring and some others were not found in summer and autumn; the total n was therefore 383 in spring, 376 in summer, and 371 in autumn). the reason for the missing pellet groups was that some markers were accidentally kicked down by fieldworkers during the experimental treatment. the same pellet groups were investigated in spring (21 31 may), in summer (16 visibility of moose pellets – persson alces vol. 39, 2003 236 20 july), and after the vegetation period (22 31 october) in 2001. at each visit, visibility (from standing position) was estimated to the nearest 5%. all pellet groups which became invisible disappeared due to concealment by vegetation, not as a result of decay. no signs of other wildlife having disturbed the pellet groups by scratching in search for insects or nematodes were found. the relationships between visibility and site index, as well as litter production, were checked. percentage of the different plants which covered the pellet groups either by growing over them or by litterfall was visually estimated to the nearest 5%, and classified as lichens, mosses, grasses, forbs, dwarf shrubs, raspberry, ferns/horsetails, and litter. because there were few values for some plant groups, site differences were only tested for the most common groups. ground cover was recorded as lichens, mosses, grass, or bare ground. data were analysed statistically using the sas system for windows (version 8.2, 2001). the kruskal wallis test was used to test for differences in visibility among study sites, plants concealing the pellet groups, and substrate types (siegel and castellan 1988). correlations between visibility, site index, and litter production were checked using the pearson correlation coefficient (fry 1999) or (if the assumptions for a parametric test failed) the spearman correlation coefficient (siegel and castellan 1988). when significant correlations were found, linear regression models were developed. the significance level was set at α = 0.05. results the mean visibility (the percentage of the original surface area of each pellet group visible) decreased with time (fig. 2). the largest decrease in visibility was 10% per month, and occurred during the transition from spring to summer after one winter (9 months of exposure). thereafter, visibility decreased at a slower rate. the mean visibility differed significantly among study sites for all seasons: (spring: kruskal wallis test, χ2 = 100.92, p < 0.001; summer: χ2 = 174.25, p < 0.001; and autumn: χ2 = 149.60, p < 0.001; df = 7 in all tests, fig. 3). no correlations were found between the mean visibility and site index for all pellet groups pooled or for spring, summer, and autumn respectively for pellet groups laid out in 1999 and 2000 combined (spearman r, p > 0.16 for all tests) or pellet groups laid out in 2000 (p > 0.07 for all tests). however, there were significant negative correlations between mean visibility and litter production for pellet groups laid out in 1999 and 2000 (fig. 4a), which could be described by linear regressions (table 2). for pellet groups laid out in 2000 there was a trend, although not significant, that the mean visibility of all pellet groups pooled decreased with litter production (fig. 4b). significant negative correlations were also found between litter production and visibility of pellet groups investigated in spring, summer, and autumn, respectively (table 2). the relative proportions of plant groups concealing pellet groups differed signifi 10 15 20 25 30 0 20 40 60 80 100 % v is ib ili ty months of exposure fig. 2. the change in percent visibility (mean and standard deviation) with time of moose faecal pellets, all study sites pooled. alces vol. 39, 2003 persson – visibility of moose pellets 237 cantly among study sites for all seasons (table 3, fig. 5). grasses contributed most to the concealment in all seasons, and at all study sites except lögdåberget and skatan. mosses were also important in all seasons and at all study sites, whereas lichens only were important at lögdåberget and skatan. åtmyrberget and mörtsjöstavaren had the richest field vegetation, and in addition to grass, raspberry, forbs, ferns, and horsetails covered the pellet groups at those sites in summer, whereas litter was important in autumn. the visibility of pellet groups differed significantly among substrate types for all seasons: (spring: kruskal wallis test, χ2 = 73.61, p < 0.001; summer: χ2 = 128.44, p < 0.001; and autumn: χ2 = 114.44, p < 0.001; df = 2 in all tests, fig. 6). only 4 pellet groups were on barren ground, and because low sample size makes statistical tests unreliable, these groups were omitted from the analysis. visibility was highest on lichen substrate and lowest on grass. the difference among substrate types was fig. 3. the mean visibility of moose faecal pellets at different study sites ranked according to site index, investigated in spring (a), summer (b), and autumn (c). a) b) c) fig. 4. the negative correlation between mean visibility (%) and habitat productivity estimated as litter production (g dry mass per m2 and year): (a) all faecal pellet groups and investigations pooled for pellet groups laid out in 1999 and 2000, exposed for an average of 18 months; and (b) all pellet groups and investigations pooled for pellet groups laid out in 2000, exposed for an average of 12 months. a) b) 8 10 12 14 16 18 20 22 24 26 0 20 40 60 80 100 3a åtmyrbegetmörtsjöstavaren skatan lögdåberget lögdåberget skatan djupsjöbrännan åtmyrberget selsberget mörtsjöstavaren ralberget rönnäs spring % v is ib ili ty months of exposure 10 12 14 16 18 20 22 24 26 28 0 20 40 60 80 100 months of exposure 3b mörtsjöstavaren åtmyrberget skatan lögdåberget summer % v is ib ili yy 12 14 16 18 20 22 24 26 28 30 0 20 40 60 80 100 3c months of exposureåtmyrberget and mörtsjöstavaren skatan lögdåberget autumn % v is ib ili ty 1.5 2.0 2.5 3.0 3.5 4.0 2.0 2.5 3.0 3.5 4.0 4.5 ln (% visibility) = 4.27 0.90 (ln litter) r square = 0.60, p = 0.025 l n m ea n v is ib ili ty ( % ) lnlitter (g per m2 and year) 1.5 2.0 2.5 3.0 3.5 4.0 2.5 3.0 3.5 4.0 4.5 ln (% visibility) = 4.29 0.63 (ln litter) r square = 0.49, p = 0.054 l n m ea n v is ib ili ty ( % ) lnlitter (g per m2 and year) visibility of moose pellets – persson alces vol. 39, 2003 238 higher in summer and autumn than in spring: in summer and autumn, the mean visibility was 13 16 times higher for pellet groups lying on lichen than on grass, whereas visibility was 3 times higher in spring. visibility was 4.5 5 times higher for pellet groups lying on moss than on grass in summer and autumn, and only 1.7 times higher in spring. discussion my study demonstrated that visibility of moose pellets decreased considerably with time. as a result of concealment by new vegetation, the largest decrease occurred in the transition from spring to summer after one winter of exposure. there were large habitat differences in how fast visibility decreased. i found no correlations between visibility and site index, whereas visibility was negatively correlated with vegetative litter production. litter production could therefore be used to predict how fast the visibility of pellet groups decreases. composition of the ground cover vegetation was an important factor affecting the visibility of pellet groups. visibility was highest on lichen-rich sites without grass (i.e., lögdåberget and skatan) and lowest on sites with rich field vegetation of grass, forbs, ferns, horsetails, and raspberry (i.e., åtmyrberget and mörtsjöstavaren). pellet groups disappeared as a result of table 2. statistics from the regression models (f, p, and r2) describing correlations between mean visibility (%) of faecal pellet groups and vegetative litter production (g dry mass per m2 and year) for pellet groups laid out in 1999 and 2000 combined, as well as in 2000. the regressions are estimated as ln visibility = ln litter for all pellet groups and investigations pooled for each site (“total”), as well as for all pellet groups at each site of the investigations in spring, summer, and autumn 2001, respectively. 1999 and 2000 2000 total f = 8.84, p = 0.025, r2 = 0.60 f = 5.69, p = 0.054, r2 = 0.49 spring f = 9.34, p = 0.022, r2 = 0.61 f = 8.01, p = 0.030, r2 = 0.57 summer f = 12.08, p = 0.013, r2 = 0.67 f = 8.88, p = 0.023, r2 = 0.60 autumn f = 13.41, p = 0.011, r2 = 0.69 f = 8.00, p = 0.030, r2 = 0.57 table 3. results of kruskal wallis test for statistical differences in the relative contribution of various plant groups concealing moose pellet groups among study sites. χ2 and p-values are presented in the table, df = 7 for all tests. forbs, shrubs, raspberry, ferns, horsetails, and litter (in summer) contributed less to the concealment and were omitted in the tests. vegetation spring (χ2, p) summer (χ2, p) autumn (χ2, p) lichens 140.13, < 0.001 165.31, < 0.001 192.11, < 0.001 mosses 14.13, = 0.049 26.51, < 0.001 37.52, = 0.005 grass 139.80, < 0.001 139.80, < 0.001 192.11, < 0.001 litter 161.97, < 0.001 207.59, < 0.001 alces vol. 39, 2003 persson – visibility of moose pellets 239 concealment by vegetation rather than decay processes. studies from vancouver island, british columbia (harestad and bunnell 1987), and the olympic peninsula, 1 2 3 4 5 6 7 8 0 20 40 60 80 100 5a rönnäs ralberget mörtsjöstavaren selsberget åtmyrberget djupsjöbrännan skatan lögdåberget spring r el at iv e % c o ve r lichens mosses grass litter fig. 5. the relative contribution (mean and standard deviation) of the most important vegetation types concealing moose faecal pellet groups at different study sites ranked after site index; investigated in spring (a), summer (b), and autumn (c). forbs, shrubs, raspberry, ferns, horsetails, and litter (in summer) contributed less to the concealment and were omitted in the figures. lichens mosses grass 0 20 40 60 80 100 % v is ib ili ty substrate type spring summer autumn fig. 6. the visibility (mean and standard deviation) of moose faecal pellets lying on different substrate types, all study sites pooled. a) c) 1 2 3 4 5 6 7 8 0 20 40 60 80 100 5b rönnäs ralberget mörtsjöstavaren selsberget åtmyrberget djupsjöbrännan skatan lögdåberget summer r el at iv e % c o ve r lichens mosses grass 1 2 3 4 5 6 7 8 0 20 40 60 80 100 5c rönnäs ralberget mörtsjöstavaren selsberget åtmyrberget djupsjöbrännan skatan lögdåberget autumn r el at iv e % c o ve r lichens mosses grass litter b) washington (lehmkuhl et al. 1994), have reported the opposite. however, their studies were done in areas with milder climate and higher annual precipitation. the relatively cold and dry climate in my study area was probably the main explanation for lower decay rates. vegetation concealment likely results in higher moisture which increases decomposition rate and decreases visibility in more moist and vegetated habitats (lehmkuhl et al. 1994). it is likely that the sites with the richest field vegetation also offered the moistest conditions at ground level. reliable methods to estimate population density, habitat use, and distribution of moose are important for wildlife and forest management (härkönen and heikkilä 1999). unless pellet counts are done immediately after snow melt before green up, one needs to know how fast visibility decreases, and habitat differences must be regarded. after one winter of exposure, more than 95% of the deposited pellet groups were visible (i.e., visibility > 0 from standing position) independent of habitat type in the study area. population trends and habitat use by moose in winter can thus be estimated without biases caused by visibility differences in boreal forest habitats with a relatively dry and cold climate. because of the fast visibility of moose pellets – persson alces vol. 39, 2003 240 decline in visibility after the first winter, it is important to clear the study plots in late autumn after the vegetation period and then visit them as soon as possible after snowmelt. however, moose have different food preferences in summer and winter (cederlund et al. 1980, bergström and hjeljord 1987), and habitat use and distribution of moose in the vegetation period is also interesting to reveal for moose managers. the correlations between visibility and litter production suggest that visibility can be estimated as a function of habitat productivity. more studies of the relationship between visibility and habitat productivity should be done to establish a sightability correction factor for various habitat types. the pellet count method could then be used to estimate habitat use and population distribution in the landscape during the vegetation period and with longer periods between plot visits. acknowledgements thanks to christer johansson for allowing me to collect moose pellets at his moose farm in bjurholm, and the forest companies assi domän and holmen skog where the study sites were located. i am also grateful to professor kjell danell and professor roger bergström for help with the planning of the study as well as valuable discussions and comments concerning the manuscript. the study was financed by the swedish council for forestry and agricultural research and the swedish environmental protection agency (grants to kjell danell and roger bergström). references ahti, t., l. hämet-ahti, and j. jalas. 1968. vegetation zones and their sections in northwestern europe. annales botanici fennici 5:169-211. aulak, w., and j. babinska-werka. 1990. estimation of roe deer density based on the abundance and rate of disappearance of their faeces from the forest. acta theriologica 35:111-120. bergström, r., and o. hjeljord. 1987. moose and vegetation interaction in northwestern europe and poland. swedish wildlife research supplement 1:213-228. c e d e r l u n d , g . , h . l j u n g q u i s t , g . markgren, and f. stålfelt. 1980. foods of moose and roe deer at grimsö in central sweden results of rumen content analyses. swedish wildlife research supplement 11:169-247. elfving, b., and a. kiviste. 1997. construction of site index equations for pinus sylvestris l. using permanent plot data in sweden. forest ecology and management 98:125-134. fridman, j., g. kempe, p. nilsson, h. toet, and b. westerlund. 2001. forestry statistics 2001. official statistics of sweden. swedish university of agricultural sciences, umeå, sweden. (in swedish with english summary). fry, j. c. 1999. biological data analysis: a practical approach. oxford university press, oxford, uk. hägglund, b., and j-e. lundmark. 1987. h a n d l e d n i n g i b o n i t e r i n g m e d skogshögskolans boniteringssystem. d e l 2 , d i a g r a m o c h t a b e l l e r . skogsstyrelsen, jönköping, sweden. (in swedish). harestad, a. s., and f. l. bunnell. 1987. persistence of black-tailed deer fecal pellets in coastal habitats. journal of wildlife management 51:33-37. härkönen, s., and r. heikkilä. 1999. use of pellet group counts in determining density and habitat use of moose alces alces in finland. wildlife biology 5:233239. jordan, p. a., r. o. peterson, p. campbell, and b. mclaren. 1993. comparison of pellet counts and aerial counts for estimating density of moose at isle royale: alces vol. 39, 2003 persson – visibility of moose pellets 241 a progress report. alces 29:267-278. lavsund, s. 1975. investigations on pellet groups. research notes 23. department of vertebrate ecology, royal college of forestry, stockholm, sweden. (in swedish with english summary). lehmkuhl, j. f., c. a. hansen, and k. sloan. 1994. elk pellet-group decomposition and detectability in coastal forests of washington. journal of wildlife management 58:664-669. lindgren, d., c. c. ying, b. elfving, and k. lindgren. 1994. site index variation with latitude and altitude in iufro pinus contorta provenance experiments in western canada and northern sweden. scandinavian journal of forest research 9:270-274. massei, g., p. bacon, and p. genov. 1998. fallow deer and wild boar pellet group disappearance in a mediterranean area. j o u r n a l o f w i l d l i f e m a n a g e m e n t 62:1086-1094. neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. journal of wildlife management 32:597-614. nilsson, n.-e. 1996. the forest. national atlas of sweden. bokförlaget bra böcker, höganäs, sweden. nyberg, å., and i.-l. persson. 2002. habitat differences of coprophilous fungi on moose dung. mycological research 106:1360-1366. ollinger, s. v., m. l. smith, m. e. martin, r. a. hallett, c. l. goodale, and j. d. aber. 2002. regional variation in foliar chemistry and n cycling among forests of diverse history and composition. ecology 82:339-355. persson, i.-l. 2003. moose population density and habitat productivity as drivers of ecosystem processes in northern boreal forests. ph.d. thesis, swedish university of agricultural sciences, umeå, sweden. , k. danell, and r. bergström. 2000. disturbance by large herbivores in boreal forests with special reference to moose. annales zoologici fennici 37:251-263. raab, b., and h. vedin. 1996. climate, lakes and rivers. national atlas of sweden. bokförlaget bra böcker, höganäs, sweden. sas institute. 2001. sas system for windows. version 8.2. sas institute incorporated, cary, north carolina, usa. siegel, s., and j. n. castellan. 1988. nonparametric statistics for the behavioural sciences. second edition. mcgraw-hill book company, singapore. smith, r.h. 1968. a comparison of several sizes of circular plots for estimating deer pellet-group density. journal of wildlife management 32:585-591. timmermann, h. r. 1974. moose inventory methods: a review. naturaliste canadien 101:615-629. wallmo, o. c., a. w. jackson, t. l. hailey, and r. l. carlisle. 1962. influence of rain on the count of deer pellet groups. journal of wildlife management 26:50-55. alces 48, 2012 a journal devoted to the biology and management of moose edward m. addison ecolink science vince f. j. crichton manitoba conservation murray w. lankester lakehead university (retired) printed at lakehead university thunder bay, ontario, canada (called proceedings of the north american moose conference from 1972 through 1980) issn 0835-5851 brian e. mclaren lakehead university gerald w. redmond maritime college of forest technology kristine m. rines new hampshire fish and game associate editors chief editor peter j. pekins university of new hampshire submissions editor roy v. rea university of northern british columbia business editor arthur r. rodgers ontario ministry of natural resources alces36_17.pdf 101 moose population dynamics during 20 years of declining harvest in british columbia gerald kuzyk1, ian hatter2, shelley marshall3, chris procter4, becky cadsand5, daniel lirette6, heidi schindler7, michael bridger8, patrick stent1, andrew walker9, and michael klaczek3 1ministry of forests, lands and natural resource operations and rural development, 205 industrial road g, cranbrook, british columbia v1c 7g5, canada; 2nature wise consulting, 308 uganda avenue, victoria, british columbia v9a 5x7, canada; 3ministry of forests, lands and natural resource operations and rural development, 2000 s ospika boulevard, prince george, british columbia v2n 4w5, canada; 4ministry of forests, lands and natural resource operations and rural development, 1259 dalhousie drive, kamloops, british columbia v2c 5z5, canada; 5ministry of forests, lands and natural resource operations and rural development, suite 400 640 borland street, williams lake, british columbia v2g 4t1, canada; 6ministry of forests, lands and natural resource operations and rural development, 300 s cariboo highway, 100 mile house, british columbia v0k 2e0, canada; 7ministry of forests, lands and natural resource operations and rural development, 3726 alfred avenue, smithers, british columbia v0j 2n0, canada; 8ministry of forests, lands and natural resource operations and rural development, suite 400 10003 110th avenue, fort st. john, british columbia v1j 6m7, canada; 9ministry of forests, lands and natural resource operations and rural development, 102 industrial place, penticton british columbia v2a 7c8, canada. abstract: licenced harvest of moose (alces alces) in british columbia, canada declined by approximately half over the 20-year period from 1996–2015. to better understand changes in moose populations coinciding with this period of declining harvest, we modelled population dynamics within 31 game management zones (gmzs). we used aerial survey data (180 density and 159 composition surveys) combined with licensed harvest to develop 4 competing statistical models to assess population dynamics based on constant parameters and temporal trends in calf:cow ratios at 6 months, juvenile survival from 6–18 months, or cow survival. the models indicated that moose populations declined (λ < 1) in 7 gmzs (23%) from 1996–2005 and in 22 gmzs (71%) from 2006–2015. over the 20-year period, the best model was fit with declining trends in calf:cow ratios in 8 gmzs, declining juvenile survival in 6 gmzs, and declining cow survival in 8 gmzs. population growth rate was slightly reduced in those gmzs where licenced antlerless (cow and calf) hunting occurred but was not considered the primary factor causing population decline. total licenced bull harvest influenced bull:cow ratios that were significantly lower in 2006–2015 (x = 37:100) than 1995–2005 (x = 48:100); significant predictive relationships existed between harvest rates and bull:cow ratios. provincial moose numbers and harvest were highly correlated (r = 0.81) suggesting that declining harvest was a reaction to declining population trends. we found that the provincial moose population increased 6% from 1996–2005, subsequently declined 32% from 2006–2015, and declined 29% overall during the 20-year study period. alces vol. 54: 101–119 (2018) key words: alces alces, bull:cow ratio, calf survival, cow survival, harvest, juvenile recruitment moose (alces alces) remain an important hunted species throughout their circumpolar range (telfer 1984, kelsall 1987, karns 2007, boman et al. 2011), but concern exists about declining populations in parts of north america (lenarz et al. 2009, decesare bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 102 et al. 2014, timmermann and rodgers 2017). moose populations are vulnerable to factors such as predation and hunting, both of which can be influenced by human-induced landscape change (murray et al. 2006, brown 2011). moose co-exist with predators throughout most of their range in canada and alaska, and in areas where predators are lightly hunted, moose density is typically low (<400 moose/1000 km2), and male (hereafter bull)—only hunting may be the only viable harvest option (gasaway et al. 1992, boertje et al. 1996). in these systems moose populations may fluctuate over time and decline in the absence of hunting (gasaway et al. 1983). a recent assessment of the provincial moose population and licenced harvest trends in british columbia (bc) documented that licenced harvest declined by approximately half over a 20-year period from 1995–2014 as regional population declines of ~50–70% occurred in the central interior (kuzyk 2016). these declines occurred concurrently with a mountain pine beetle (dendroctonus ponderosae) outbreak and associated salvage logging and road building, which presumably facilitated predator and hunter access to moose (ritchie 2008). in 2013, the province initiated research to determine factors leading to the population declines in the central interior. because the declines occurred over a relatively short period of 20 years, it was assumed that female moose (hereafter cow) survival was the most influential population parameter in the decline (kuzyk and heard 2014). these declines continue to create concern among stakeholders and first nations who have requested more refined information about provincial moose population and harvest trends to help inform management decisions (gorley 2016). to provide a consistent and objective assessment of relationships between moose population dynamics and declining harvest levels (kuzyk 2016), we followed an approach similar to that of hatter (1999) who assessed moose population trends in relation to harvest strategies in 19 game management zones (gmzs) in northern and central bc from 1994–1996. gmzs generally share similar ecological characteristics and hunter harvest patterns providing a suitable geographic unit for managing moose hunting. all gmzs are amalgamations of 1–13 wildlife management units (wmus) which is the geographic area used to collect hunting information. we expanded hatter’s (1999) approach and included all gmzs (n = 31) with licenced moose hunting seasons and survey data over a 20-year period (1996–2015). our objectives were to: 1) evaluate competing models of moose population dynamics in each gmz to determine which parameters (i.e., calf:cow ratio at 6 months, juvenile survival from 6–18 months, or cow survival) best predicted moose population trend, 2) assess the density, composition, and trend of moose populations based on the best model over the 20-year period within each gmz, and 3) evaluate the influence of licenced moose hunting based on the best model within each gmz. study area we assessed moose population dynamics and licenced harvest in 31 gmzs where moose hunting is authorized in bc (fig. 1). these areas are ecologically diverse (meidinger and pojar 1991) with moose occupying landscapes of differing topography and vegetation including northern boreal forests, dry interior forests of the central plateau, and some mountainous habitats (eastman and ritcey 1987, kuzyk et al. 2016). a mountain pine beetle outbreak beginning in the early 1990s occurred over much of the central interior of the province leaving large amounts of dead pine, which alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 103 led to increased salvage logging and road building (alfaro et al. 2015). the increased number of roads and cutblocks were thought to facilitate hunter and predator access to moose (ritchie 2008, kuzyk and heard 2014). moose co-exist with wolves (canis lupus), grizzly bears (ursus arctos), and black bears (u. americanus) throughout most of their bc range and overlap with cougars (felis concolor) in the central and southern areas (spalding and lesowski 1971, mowat et al. 2013, kuzyk and hatter 2014). a diversity of ungulates including mule deer (odocoileus hemionus), whitetailed deer (o. virginianus), elk (cervus elaphus), bison (bison bison), and caribou (rangifer tarandus) also occur here (shackleton 1999). licenced hunting seasons for bulls occurred in all gmzs having a range of season dates between august 15 and november 30. bull hunting was regulated with general open seasons with or without antler restrictions, limited-entry seasons (i.e., hunters must draw an appropriate authorization) with no antler restrictions, or a combination of general open and limited entry seasons. hunting seasons for cows and calves were mostly limited-entry hunts in 7 gmzs between october 1 and december 10 over most of the 20-year period; general open seasons for calves existed in a few select areas. methods moose density and composition two winter aerial survey methods were used to collect reliable population information following provincial standards (risc 2002). fig. 1. game management zones (n = 31) with licenced moose hunting from 1996-2015 in british columbia, canada. bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 104 density surveys collected a combination of population size, density, and composition (i.e., bull:cow and calf:cow ratios) information, whereas composition surveys gathered only bull:cow and calf:cow ratios. density surveys were typically conducted in 5–7 consecutive days in december–march using stratified random blocks that could be remeasured to detect population trends 5–7 years later (gasaway et al. 1986). certain surveys were modified to include habitat based stratification (heard et al. 2008), and distance sampling surveys were used in more open habitats (peters et al. 2014); all survey types produced comparable density estimates. a sightability correction factor developed in central bc (quayle et al. 2001) was used to account for detection probability. these surveys followed established standards for accuracy and precision (90% ci) with allowable error from ±15–25% of the estimated population size (risc 2002). composition surveys were conducted over 1–3 days in early winter (december– january) prior to typical antler drop. these surveys provided bull:cow and calf:cow ratios which were used as a general index of population trend to gauge progress towards harvest management objectives. frequency of surveys varied among gmzs due to population objectives, and financial and logistical constraints. there were a total of 180 density surveys (gmz range = 1–15) and 159 composition surveys (gmz range = 1–24) used in our analysis. over the 20-year period, the average number of density and composition surveys per gmz was 5.8 (min = 1, max = 15) and 5.1 (min = 0, max = 24), respectively. licenced harvest licenced resident harvest was estimated annually from 1996–2015 with a provincial hunter survey generating data from mail-out questionnaires sent to a random sample of moose hunters. these estimates (95% ci) were produced from an annual average of 15,477 questionnaires with an average response rate of 68%. reporting of non resident licenced harvest was mandatory and obtained from guide declarations. we used both resident and non-resident harvest in the models, but only used resident kill per unit effort (kpue; resident kill/100 resident hunter days) as an indicator of population trend. we recognized that success rates differed between resident and non-resident hunters, and that the overwhelming majority of moose hunters were residents. unlicenced hunter harvest, which we defined as moose legally harvested by first nations and moose harvested illegally, was not quantified as an annual harvest statistic due to the limited availability of these data. unlicenced harvest mortality was incorporated in estimates of annual non-hunting survival rates derived from radio-collared moose (kuzyk et al. 2016), and was assumed as a constant proportion of annual mortality. modelling approach our models provided a reasonable trade-off between what can be measured practically by biologists and what is needed to help predict moose population trends and responses to harvest management strategies (hatter 1999). the intent was to utilize a statistically rigorous and objective modelling approach while maintaining a relatively easy-to implement procedure for constructing population models from multiple types of population data. because we wanted to make comparisons among gmzs, it was important that the standardized approach and assumptions were consistent among models. we used models built in microsoft excel® and model fitting was accomplished using excel’s built-in optimizer function, solver. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 105 we constructed discrete-time, stage structured population models for each of the 31 gmzs. our model and model-fitting approach followed white and lubow (2002), where the focus was on building a series of simple candidate models and selecting the most parsimonious model (i.e., the model that best fits the data with the fewest number of estimated parameters) using akaike information criteria (aic) (burnham and anderson 2002). an exception was that we included resident kpue as a trend index. although kpue was available for all gmzs and was consistently collected over the 20-year study period, we did not use kpue in gmzs 3b, 3c, 8a, and 8b because the trends in kpue were inconsistent with the trends from periodic population estimates. for all other gmzs, we fit the kpue index to the model using procedures described by hatter (1998) and haddon (2001). the data for each gmz typically included estimates of absolute abundance in winter, which were extrapolated from density surveys, winter age and sex ratios (i.e., bull:cow and calf:cow ratios), kpue, and the annual harvest (bulls, cows, and calves). the key parameters estimated by the model included: r, the recruitment rate or the calf:cow ratio when calves were ~6 months old which we define as the calf:cow ratio at 6 months; sj, the calf survival rate from 6–18 months which we define as juvenile survival from 6–18 months; and sf, the annual cow (>18 months) survival rate excluding licenced hunting which we define as cow survival. we used cow survival rates estimated from radio-collared cows from 5 study areas, and juvenile survival rates from 2 of these areas (kuzyk et al. 2016). survival rates were available for a 2-year period in 5 study areas corresponding to gmzs 7ob, 7oc, 6c, 5d, and 3c; we used the documented survival rate from the corresponding study area for these gmzs each year. for the remaining gmzs, we used the average of the 5 study areas for each year. the juvenile survival rates were obtained from gmz 7ob and 7oc in 2016. the estimates were based on the change in calf:cow ratios between early winter and late winter surveys, and included an adjustment for adult cow survival between these 2 time periods (skalski et al. 2005). all field-based estimates except harvest level included the standard error (se). we fixed 3 model parameters at values reported in the literature: wounding loss equaled 15% of licenced hunting (gasaway et al. 1983, boer and keppie 1988, fryxell et al. 1988), calf sex ratio was 50:50 (ballard et al. 1991, boer 1992), and the natural adult bull survival rate was 96% of the cow survival rate (peterson 1977). we built 4 competing models for each gmz (n = 124) and considered each as a potential hypothesis driving population dynamics. model 1 held all 3 parameters constant; i.e., calf recruitment, juvenile survival from 6–18 months, and cow survival. the other 3 models included a single parameter as a linear trend while holding the other 2 constant (table 1). for each model we calculated the corrected aicc for a small sample in order to determine the best model for each gmz and aic weight (wi) (burnham and anderson 2002). changes in moose population dynamics over 4 time periods (1996–2015, 1996–2005, 2006–2015, 2011–2015) were evaluated. we chose 1996–2015 as a long-term (20 year) overview as it corresponded to the earlier moose population assessment (kuzyk 2016) and when provincial survey methods became standardized. the period of 1996–2005 corresponded to when the provincial harvest was relatively stable, and the 2006–2015 period corresponded to when annual harvests declined by 37%. the 2011–2015 period corresponded to when additional harvest restrictions were imposed for numerous bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 106 gmzs due to increasing concerns about declining moose densities (kuzyk and heard 2014, kuzyk 2016). annual population estimates were summed across all gmzs to provide the provincial estimates from 1996–2015. we used the best fit model estimates of post-hunt females (cows and calves) to assess the influence of cow and calf hunting on moose population growth. we calculated both the potential rate of population change without female hunting (λp) and the actual rate of change with female hunting (λh). a λp value >1 indicated potential growth and thus some opportunity for cow harvest, and a value <1 implied that licenced cow harvest had some negative influence on population growth. we used the modelled estimates of the post-hunt adult ratios to assess the influence of bull hunting on bull:cow ratios. it is unknown what adult sex ratio ensures that cows are synchronously bred to avoid impeding population growth (timmermann 1992). ministry policy is to maintain post-hunt, bull:cow ratios ≥ 30 bulls:100 cows in areas with densities > 200 moose/1000 km2, and 50 bulls:100 cows in areas with <200 moose/1000 km2 (bc moe 2010). results best models model 1 was the best model (i.e., lowest aicc) in 7 (23%) gmzs, model 2 in 12 (39%) gmzs, model 3 in 2 (6%) gmzs, and model 4 in 10 (32%) gmzs (table 2). model 1 had substantial support (i.e., δaicc <2) in 12 gmzs (39%), model 2 in 16 gmzs (52%), model 3 in 6 gmzs (19%), and model 4 in 14 gmzs (45%). models 2 and 4 had very strong support in 10 gmzs, each with a relative aic weight of 1 in 5 gmzs. twenty gmzs had a single top model with strong support (i.e., δaicc <2), 5 gmzs had 2 models, and 6 gmzs had 3 models (table 2). population dynamics we calculated the average (1996–2015) moose density, population composition, key population parameters, and rate of change by gmz for the best fitting models, and the average rate of population change (λ) in 1996– 2005, 2006–2015, and 2011–2015 (table 2). the rate of population change varied spatially between 1996–2005 and 2006–2015 (fig. 2). population density in 1996–2015 varied from 6 (gmz 4wb) to 1078 (gmz 7ob), averaging 334 moose/1000 km2 (sd = 271) across table 1. sequence of models fit to moose population survey and harvest data for using parameters of calf:cow ratios at 6 months (r), juvenile survival from 6–18 months (sj), or cow survival (sf) in each of 31 game management zones in british columbia, canada. model calf:cow ratio (r) juvenile survival (sj) cow survival (sf) model structure 1 constant constant constant rt = r sj,t = sj sf,t = sf 2 linear trend constant constant rt = rintercept + rslope · (yeart – year0) sj,t = sj sf,t = sf 3 constant linear trend constant rt = r sj,t = sj,intercept + sj,slope · (yeart – year0) sf,t = sf 4 constant constant linear trend rt = r sj,t = sj sf,t = sf,intercept + sf,slope · (yeart – year0) alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 107 (c on tin ue d ) ta bl e 2. t he a ve ra ge m oo se d en si ty , p op ul at io n co m po si tio n, k ey p op ul at io n pa ra m et er e st im at es , a nd r at e of c ha ng e ba se d on th e be st m od el s (l ow es t a ic c) in 3 1 g am e m an ag em en t z on es fr om 1 99 6– 20 15 in b ri tis h c ol um bi a, c an ad a. t he a ve ra ge ra te o f c ha ng e is s ho w n fo r 1 99 6– 20 15 , 1 99 6– 20 05 , 2 00 6– 20 15 , an d 20 11 –2 01 5; th os e in g ra y ha ve λ <1 . g m z m od el s1 ,2 w ith δ a ic c < 2 a ic w ei gh t (w i)3 d en si ty se x/ a ge r at io s k ey p op ul at io n pa ra m et er s4 r at e of c ha ng e (λ )5 /1 00 0 km 2 b :1 00 c c a: 10 0 c r s j s f 19 96 –2 01 5 19 96 –2 00 5 20 06 –2 01 5 20 11 –2 01 5 3b 1, 2 , 4 0. 43 24 2 18 34 34 89 89 1. 03 1. 04 1. 03 1. 05 3c 1, 3 0. 57 20 9 29 37 37 68 89 1. 01 1. 01 1. 02 1. 02 3d 2 1. 00 16 1 69 47 15 –7 9 81 81 0. 97 1. 03 0. 91 0. 87 4e a 2 1. 00 20 4 46 30 6– 53 90 92 1. 02 1. 07 0. 97 0. 94 4e b 2 1. 00 12 8 40 27 7– 48 90 91 1. 02 1. 07 0. 97 0. 94 4e c 1 0. 60 94 60 27 27 86 86 0. 96 0. 95 0. 97 0. 98 4w a 2, 1 0. 61 85 55 49 23 –7 5 90 90 1. 10 1. 17 1. 04 1. 02 4w b 4 0. 99 6 49 30 30 81 –9 0 81 –9 9 0. 99 1. 01 0. 97 0. 95 5a 2, 1 0. 55 26 4 38 28 19 –3 6 90 90 0. 99 0. 99 0. 99 0. 98 5b 1, 2 , 4 0. 41 53 5 28 34 34 84 84 0. 98 0. 98 0. 98 0. 99 5c 4, 2 , 1 0. 47 35 2 34 40 40 79 –8 4 79 –8 4 0. 98 0. 96 1. 00 1. 01 5d 2 1. 00 22 4 38 35 25 –4 5 84 84 0. 99 1. 01 0. 96 0. 95 6b 2 1. 00 92 56 46 33 –6 0 79 79 0. 98 1. 01 0. 95 0. 93 6c 4 1. 00 52 6 35 37 37 77 –8 9 77 –8 9 0. 98 1. 02 0. 95 0. 92 6d 4 0. 96 36 1 47 46 46 25 86 –9 7 0. 97 0. 99 0. 95 0. 93 6e 4 1. 00 30 5 58 36 36 81 –9 0 81 –9 1 1. 01 1. 03 0. 99 0. 98 6f 2, 4 , 1 0. 51 51 1 56 31 27 –3 5 69 90 1. 00 1. 01 1. 00 0. 99 7o a 2 0. 68 22 5 43 25 6– 45 90 94 0. 98 1. 00 0. 96 0. 95 7o b 4 1. 00 10 78 35 32 32 64 81 –9 6 0. 97 1. 00 0. 94 0. 92 7o c 4, 3 0. 57 95 4 43 25 25 59 85 –9 9 0. 98 1. 00 0. 95 0. 94 7o d 3, 4 , 2 0. 46 56 8 49 35 35 50 –8 6 86 0. 98 1. 00 0. 96 0. 95 7o e 4 0. 99 21 5 49 25 25 79 –9 0 79 –9 9 1. 00 1. 04 0. 96 0. 93 7p a 4 1. 00 57 4 42 30 30 74 –9 0 74 –9 6 0. 98 1. 05 0. 92 0. 88 7p b 4 1. 00 57 2 32 34 34 73 –8 8 73 –9 3 0. 97 1. 03 0. 92 0. 88 7p c 2 0. 79 90 3 28 17 13 –2 2 89 89 0. 96 0. 98 0. 95 0. 94 7p d 2, 3 0. 71 31 6 49 25 15 –3 4 87 87 0. 98 1. 01 0. 95 0. 93 bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 108 all gmzs. the adult bull:cow ratio in 1996– 2015 varied from 18:100 (gmz 3b) to 69:100 (gmz 3d), averaging 42:100 (sd = 12) across all gmzs. bull:cow ratios were significantly lower (paired t test: t = 4.48, df = 30, p < 0.001) in 2006–2015 (x = 37:100, sd = 12.3) than in 1996–2005 (x = 48:100, sd = 15.3). calf:cow ratios varied from 17:100 (gmz 7pc) to 49:100 (gmz 4wa), averaging 34:100 (sd = 8.0) across all gmzs. across the 20-year period, calf:cow ratios at 6 months declined in 12 gmzs (39%), juvenile survival from 6–18 months declined in 9 gmzs (29%), and cow survival declined in 10 gmzs (32%). no gmz experienced an increase in calf:cow ratio, juvenile survival, or annual survival of cows. the annual rate of population change (λ) varied from 0.96 (gmz 7pc) to 1.10 (gmz 4wa) from 1996– 2015; λ was ≥1 in 24 gmzs (77%) from 1996–2005, but only ≥9 in gmzs (29%) from 2006–2015 (χ2 = 12.7, df = 1, p < 0.001). only 8 gmzs (3b, 3c, 4wa, 5c, 8a, 8b, 8c, and 8d) had an average λ > 1 in the most recent 5-year period (2011–2015). the highest modelled population estimate across the province occurred in 2002 (258,532) and the lowest in 2015 (169,752) (fig. 3). percent change within the 4-time periods was: -29% from 1996–2015 (λ = 0.98), 6.4% from 1996–2005 (λ = 1.01), -32% from 2006–2015 (λ = 0.96), and -20% from 2011–2015 (λ = 0.95). there was a strong correlation between the decline in the population estimate and licenced harvest (r = 0.81, p <0.001). licenced hunting the average annual licenced harvest during 1996–2015 varied from 18 (gmz 8c) to 1490 moose per gmz (gmz 7ob; table 3). harvest composition was predominantly bulls in all gmzs, with the antlerless harvest >10% of the total harvest in 8 gmzs: 3b (12%), 3c (15%), 3d (10%), 4wb (31%), 7oa (32%), ta bl e 2. (c on tin ue d ) g m z m od el s1 ,2 w ith δ a ic c < 2 a ic w ei gh t (w i)3 d en si ty se x/ a ge r at io s k ey p op ul at io n pa ra m et er s4 r at e of c ha ng e (λ )5 /1 00 0 km 2 b /1 00 c c a/ 10 0 c r s j s f 19 96 –2 01 5 19 96 –2 00 5 20 06 –2 01 5 20 11 –2 01 5 7p e 2, 1 , 3 0. 38 17 3 57 29 22 -3 6 86 86 0. 98 1. 00 0. 96 0. 95 8a 1 0. 87 59 38 43 43 90 92 1. 09 1. 07 1. 13 1. 12 8b 1 0. 62 20 8 35 35 35 87 87 1. 01 0. 99 1. 02 1. 02 8c 3 0. 74 10 4 42 48 48 26 -9 0 95 1. 07 1. 12 1. 03 1. 02 8d 1 0. 64 91 30 45 45 83 83 1. 02 1. 02 1. 02 1. 02 1 m od el 1 . c al f: co w ra tio a t 6 m on th s (r ), ju ve ni le (s j) a nd c ow s ur vi va l ( s f ) r at es a re c on st an t. m od el 2 . l in ea r t re nd in c al f: co w ra tio a t 6 m on th s. j uv en ile a nd c ow s ur vi va l r at es a re c on st an t. m od el 3 . l in ea r t re nd in ju ve ni le s ur vi va l r at e. c al f: co w ra tio a t 6 m on th s an d co w s ur vi va l r at es a re c on st an t. m od el 4 . l in ea r t re nd in c ow s ur vi va l r at e. c al f: co w ra tio a t 6 m on th s an d ju ve ni le s ur vi va l r at es a re c on st an t. 2 m od el s lis te d st ar tin g w ith lo w es t δ a ic c v al ue . 3 a ic re la tiv e w ei gh t o f b es t f itt in g m od el . 4 r = r ec ru itm en t r at e (c al ve s: 10 0 co w s) , s j = % c al f s ur vi va l r at e (6 18 m th s) , s f = % c ow s ur vi va l r at e (> 18 m th s) . 5 r at e of c ha ng e (λ ) b y tim e pe ri od . λ < 1 .0 a re h ig hl ig ht ed in g re y. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 109 7ob (41%), 7oc (32%), and 7od (27%). the harvest rate of bulls relative to the population (bull harvest/pre-hunt population) averaged 5% (range = 2–12%), and the proportional bull harvest rate (bull harvest/pre-hunt bulls) averaged 17% (range = 5–36%). in comparison, the proportional cow (range = 1–3% of cows) and calf harvest rates (range = 1–9% of calves) were low. antlerless hunting generally had minimal influence on population growth rate. during 1996–2015, the average λp for gmzs 3b, 3d, 7ob, and 7oc was 1.00; λh was slightly lower in these and other units (0.97 in 7ob and 0.99 in 3b, 3c, and 7oc). when the 1996–2005 growth rate estimates were compared to those in 2006–2015, both λp and λh were > 1 in 1996–2005, and < 1 in fig. 2. population rate of change (λ) in 31 game management zones having licenced moose hunting seasons from 1996–2005 (a.) and 2006–2015 (b.); λ values < 1 are highlighted in burgundy and values ≥1 in gray. (footnote: λ for gmz 8b was 0.993 from 1996–2005 and 1.024 from 2006–2015. gmz 5c was 0.961 from 1996–2005 and 1.000 from 2006–2015). (a) (b) bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 110 2006–2015, with the exception of gmz 3b where λp = 1.00 and λh = 0.99. antlerless harvest may have reduced population growth rate somewhat during both time periods in certain gmzs, but was not a primary influence in the overall decline of cows during 2006–2015. modelled adult sex ratios averaged 43 bulls:100 cows during 1996–2015, 49 bulls:100 cows from 1996–2005, 37 bulls:100 cows from 2006–2015, and 35 bulls:100 cows from 2011–2015. adult sex ratios were <30 bulls:100 cows in 4 gmzs from 1996–2015, 3 gmzs from 1996–2005, 11 gmzs from 2006–2015, and 13 gmzs from 2011–2015 (χ2 [1996–2005, 2006–2015] = 4.5, p = 0.034). bull harvest rate (r2 = 0.82, p <0.001; fig. 4) and moose harvest rate (r2 = 0.70, p < 0.001; fig. 5) were strong predictors of the bull:cow ratio (i.e., # post-hunt bulls:100 cows). a ratio of 30 bulls:100 cows was achieved at a population harvest rate of 6% and bull harvest rate of 23%. discussion we found that moose populations were declining (λ < 1) in 23% of the gmzs from 1996–2005, and in most (71%) gmzs from 2006–2015. while cow survival was declining in certain gmzs, variation (decline) in the calf:cow ratio at 6 months and juvenile survival from 6–18 months was extensive and likely added to population decline in areas where cow survival was constant (gaillard et al. 1998). the variation and decline measured in both parameters were consistent with previous studies indicating that their temporal variation is more common than in cow survival (gaillard et al. 2000, eberhardt 2002). further, they may be more influential in local population change (monteith et al. 2014), especially in variable seasonal environments (raithel et al. 2007, hurley et al. 2014). others conclude that population growth rate is most influenced by calf and juvenile survival in predator-limited populations, especially in summer (ballard et al. 1991) due to high calf predation by wolves, grizzly bears, and black bears (gasaway et al. 1992, boertje et al. 1996, hayes et al. 2003), and presumably certain regions of bc are similar. nutritionallycompromised cows also experience reduced pregnancy rate and calf survival (murray et al. 2006, schwartz 2007), both of which negatively influence recruitment and population growth rate. predation is considered the major factor influencing cow survival in parts of canada and alaska (hauge and keith 1981, mytton fig. 3. trend in modelled population size of moose (solid circles) and the licensed hunter harvest (open circles) from 1996–2015 in british columbia, canada. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 111 (c on tin ue d ) ta bl e 3. s um m ar y of a ve ra ge h ar ve st p ar am et er s ba se d on th e be st m od el s (l ow es t a ic c) , a nd a ss es sm en t o f h un tin g in fl ue nc e (λ p, λ h , a nd b ul l:c ow ra tio ) i n 31 g am e m an ag em en t z on es fr om 1 99 6– 20 15 in b ri tis h c ol um bi a, c an ad a. g m z m od el s1 ,2 w ith δ a ic c < 2 a ic w ei gh t (w i)3 to ta l h ar ve st % h ar ve st c om po si tio n % h ar ve st r at e r at e of c ha ng e4 b ul ls :1 00 c ow s5 b ul ls c ow s c al ve s b ul ls c ow s c al ve s to ta l λ p λ h 19 96 –2 01 5 19 96 –2 00 5 20 06 –2 01 5 20 11 –2 01 5 3b 1, 2 , 4 0. 43 19 1 88 9 3 36 1 1 7 1. 03 1. 02 18 18 19 23 3c 1, 3 0. 57 13 7 85 12 3 26 1 1 6 1. 03 1. 02 29 36 21 21 3d 2 1. 00 97 89 8 2 11 1 0 4 1. 00 0. 99 69 87 51 45 4e a 2 1. 00 10 4 98 1 0 18 0 0 5 1. 05 1. 05 46 63 29 21 4e b 2 1. 00 38 97 1 2 16 0 1 4 1. 03 1. 03 40 44 35 28 4e c 1 0. 60 24 95 5 0 9 0 0 3 0. 97 0. 97 60 72 47 46 4w a 2, 1 0. 61 23 10 0 0 0 14 0 0 4 1. 11 1. 11 55 65 44 43 4w b 4 0. 99 75 69 24 7 18 3 3 7 1. 04 1. 00 49 67 31 29 5a 2, 1 0. 55 13 7 93 7 0 24 1 0 6 1. 02 1. 02 38 54 23 21 5b 1, 2 , 4 0. 41 43 3 99 1 0 25 0 0 6 0. 98 0. 98 28 29 27 26 5c 4, 2 , 1 0. 47 51 9 99 0 1 22 0 0 5 0. 98 0. 98 34 33 36 37 5d 2 1. 00 16 0 10 0 0 0 18 0 0 5 0. 99 0. 99 38 37 38 36 6b 2 1. 00 11 0 99 1 0 11 0 0 3 0. 98 0. 97 56 53 60 60 6c 4 1. 00 10 01 97 2 1 21 0 0 5 0. 99 0. 99 35 36 35 34 6d 4 0. 96 86 99 1 0 6 0 0 1 0. 98 0. 98 47 54 40 38 6e 4 1. 00 38 0 10 0 0 0 10 0 0 3 1. 02 1. 02 58 63 52 52 6f 2, 4 , 1 0. 51 19 1 10 0 0 0 5 0 0 2 1. 00 1. 00 56 56 56 56 7o a 2 0. 68 26 1 68 21 11 24 3 6 9 1. 06 1. 02 43 64 22 18 7o b 4 1. 00 14 90 58 15 26 17 2 9 7 1. 00 0. 97 35 37 33 32 7o c 4, 3 0. 57 80 9 68 15 17 11 1 5 4 1. 00 0. 99 43 53 34 32 7o d 3, 4 , 2 0. 46 42 4 73 16 11 10 1 2 4 0. 99 0. 98 49 50 48 46 7o e 4 0. 99 10 1 10 0 0 0 10 0 0 3 1. 00 1. 00 49 53 45 46 7p a 4 1. 00 37 2 10 0 0 0 12 0 0 3 0. 98 0. 98 42 40 43 42 7p b 4 1. 00 79 8 95 1 3 22 0 1 6 0. 98 0. 97 32 32 31 30 7p c 2 0. 79 36 1 10 0 0 0 14 0 0 3 0. 96 0. 96 28 27 29 28 7p d 2, 3 0. 71 17 2 10 0 0 0 6 0 0 2 0. 97 0. 97 49 45 53 51 7p e 2, 1 , 3 0. 38 18 4 99 0 0 6 0 0 2 0. 98 0. 98 57 56 57 56 bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 112 and keith 1981, larsen et al. 1989, ballard et al. 1991, gasaway et al. 1992). given that a third of gmzs had declining cow survival, particularly in remote areas of northern bc, these areas probably reflect a natural predator-prey system with minimal human influence (ballard et al. 1991). in such systems, the main cause of cow mortality is generally wolf predation (hauge and keith 1981, ballard et al. 1991) with less predation by grizzly bears (boertje et al. 1988). a current study with radio-collared cows in central bc is evaluating a landscape change hypothesis that assumes cow survival has a greater proportional effect on population growth rates than calf survival (kuzyk and heard 2014). in that study, annual cow survival rate ranges from 86–92% (kuzyk et al. 2016) and is within the expected range of stable populations (bangs et al. 1989, ballard et al. 1991, bertram and vivion 2002), and exceeds rates measured in the northwest territories (85%; stenhouse et al. 1995) and northern alberta (75–77%; hauge and keith 1981). the primary causes of mortality (as of kuzyk et al. 2016) were predation (43%), health-related (28%), and unlicenced hunting (16%). we found proportionally steeper population declines in the last 5 years (2011–2015) when 45% of populations were in >20% decline compared to 29% of populations declining ≤20% previously. these accelerated declines are consistent with other studies in which ungulate species globally have declined (≤23%) in the past 40 years (di marco et al. 2014). for example, recent declines have been reported for mule deer (lendrum et al. 2013) and pronghorn antelope (antilocapra americana) in north america (christie et al. 2015), ungulates in africa (western et al. 2009), and saiga antelope (saiga tatarica) among other species in the soviet union (milner-gulland et al. 2001). the level of decline in our study was likely influenced by the rate of early calf ta bl e 3. (c on tin ue d ) g m z m od el s1 ,2 w ith δ a ic c < 2 a ic w ei gh t (w i)3 to ta l h ar ve st % h ar ve st c om po si tio n % h ar ve st r at e r at e of c ha ng e4 b ul ls :1 00 c ow s5 b ul ls c ow s c al ve s b ul ls c ow s c al ve s to ta l λ p λ h 19 96 –2 01 5 19 96 –2 00 5 20 06 –2 01 5 20 11 –2 01 5 8a 1 0. 87 38 99 0 1 40 0 0 12 1. 11 1. 11 38 53 24 25 8b 1 0. 62 86 97 3 0 26 0 0 6 1. 03 1. 02 35 47 24 25 8c 3 0. 74 18 97 3 0 17 0 0 4 1. 09 1. 09 42 56 28 22 8d 1 0. 64 41 99 1 1 28 0 0 6 1. 02 1. 02 30 33 27 28 1 m od el 1 . c al f: co w ra tio a t 6 m on th s (r ), ju ve ni le (s j) a nd c ow s ur vi va l ( s f ) r at es a re c on st an t. m od el 2 . l in ea r t re nd in c al f: co w ra tio a t 6 m on th s. j uv en ile a nd c ow s ur vi va l r at es a re c on st an t. m od el 3 . l in ea r t re nd in ju ve ni le s ur vi va l r at e. c al f: co w ra tio a t 6 m on th s an d co w s ur vi va l r at es a re c on st an t. m od el 4 . l in ea r t re nd in c ow s ur vi va l r at e. c al f: co w ra tio a t 6 m on th s an d ju ve ni le s ur vi va l r at es a re c on st an t. 2 m od el s lis te d st ar tin g w ith lo w es t δ a ic c. 3 a ic re la tiv e w ei gh t o f b es t f itt in g m od el . 4 λ p = p ot en tia l a nn ua l g ro w th ra te o f f em al es w ith ou t h un tin g, λ h = a nn ua l g ro w th ra te o f f em al es w ith h un tin g. λ < 1 is h ig hl ig ht ed in g re y. 5 b ul l:c ow ra tio s by ti m e pe ri od . a du lt se x ra tio s < 30 b ul ls :1 00 c ow s ar e hi gh lig ht ed in g re y. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 113 survival which is directly related to the level of predation (larsen et al. 1989, bertram and vivion 2002, patterson et al. 2013). presumably, stochastic environmental events during the 20-year period added to the variability of reproductive rates and juvenile and adult survival rates (sæther 1997, gaillard et al. 2000), and that decline in individual gmzs reflected the combined influence of these parameters and local anthropogenic disturbances (brown 2011). a major challenge for wildlife managers is maintaining sustainable harvests of ungulates in systems with multiple factors influencing population growth (fryxell et al. 2014), which is especially relevant when populations are declining (palazy et al. 2012). moose in bc have high cultural and economic importance, and harvest management objectives are set to ensure that first nations requirements are addressed, while maintaining diverse opportunities for licenced hunters (bc flnro 2015). licenced antlerless harvest rates were low during the 20-year period, averaging ~1.5% for cows and 3% for calves. these rates were insufficient to initiate population declines, although may have minor additive effect on fig. 4. predictions of the post-hunt, bull:cow ratio based on the modelled bull harvest rate from 31 game management zones in british columbia, canada; = −y eˆ 61.595 x0.031 where ŷ is the post-hunt bull:100 cows ratio, and x is the bull harvest rate (bull harvest/pre-hunt bull population). fig. 5. predictions of the post-hunt, bull:cow ratio based on the modelled population harvest rate from 31 game management zones in british columbia, canada; = −y eˆ 67.339 x0.145 where ŷ is the posthunt bull:100 cows ratio, and x is the population harvest rate (bull harvest/pre-hunt population of bulls, cows, and calves). bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 114 populations in decline. our results are suggestive of areas in alaska where predator density has been reduced or moose populations are near their nutritional limitations, such that harvest mortality is partially compensatory to natural mortality (boertje et al. 2007, boertje et al. 2009). a sustainable harvest in either system can include a combination of bulls, cows, and calves. however, in areas where predation remains high, hunter harvest is often considered additive and restricted to bull-only harvest to avoid further population decline (van ballenberghe and dart 1982, gasaway et al. 1992, boertje et al. 1996). although a combination of harvest strategies may be required to adapt to variable factors influencing ungulate population growth (fulton and huntermark 2004, stedman et al. 2004), an ongoing challenge is the lack of adequate data to implement and monitor diverse harvest strategies (bunnefeld et al. 2011), especially with high hunter demand concurrent with declining populations. ministry policy in bc is to maintain post-hunt, bull:cow ratios ≥30 bulls:100 cows in densities >200 moose/1000 km2, and 50 bulls:100 cows in densities <200 moose/1000 km2 (bc moe 2010). although adult sex ratios were > 30 bulls:100 cows in most gmzs, bull:cow ratios were significantly lower during 2006–2015 (x = 37:100) than 1996–2005 (x = 48:100); declining recruitment may have been a contributing factor. we found significant predictive relationships between harvest rates and bull:cow ratios. although the population harvest rate (bull harvest/pre-hunt population) accounted for 68% of the variance in bull:cow ratios for the time period 2006–2015, it was a poor predictor of adult sex ratios in the other periods. in contrast, the bull harvest rate (bull harvest/pre-hunt bulls) was a good predictor of bull:cow ratios in all time periods. in 2006–2015 we found that average population harvest rate of 6% and bull harvest rate of 24% resulted in post-hunt, adult sex ratios of 30 bulls:100 cows. in areas with limited survey and monitoring data, our modelling suggests that a maximum population harvest rate of 5% and maximum bull harvest rate of 20% would be sustainable for most populations in bc. these recommendations are similar to those in other northern systems where moose are limited by predation (hayes et al. 2003). importantly, calculation of sustainable bull harvest rates must account for bull selectivity in the unlicensed harvest and trends in calf recruitment. our assessment provides a more rigorous and refined determination of moose population and licenced harvest trends in bc than previously available at the provincial and regional scales (hatter 1999, kuzyk 2016). for example, we found that the provincial moose population declined by 32% (λ = 0.96) from 2006–2015, which more closely aligns with views expressed by many first nations and other stakeholders (gorley 2016, kuzyk 2016). further, we determined that the provincial moose population trend and licenced harvest were highly correlated, suggesting that declining harvest was a reaction to the declining population, a conclusion similar to those in certain western united states where harvest and populations have declined in the past 15 years (decesare et al. 2014, nadeau et al. 2017, timmermann and rodgers 2017). despite our efforts to improve understanding of moose population trends by using a statistical model-fitting approach (white and lubow 2002), we acknowledge that limited survey data in certain gmzs could lead to uncertainty in interpretation and extrapolation of our model results; e.g., 4 gmzs had only a single population estimate and 2 gmzs had only a single year of alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 115 sex/age ratio data. therefore, we used kpue as a trend index to assist the models in determining population growth rates, especially in certain northern gmzs with limited data. while we acknowledge the limitations of kpue as a trend index (crichton 1993, bowyer et al. 1999, hatter 2001, decesare et al. 2016), particularly in areas with increasing road access, we found similar trends in survey density estimates and kpue in many gmzs. another data concern was that the first nation’s harvest (leblanc et al. 2011) was unknown and not quantifiable as an annual harvest statistic. although field (i.e., radio-telemetry) estimates of cow survival rates have included unlicenced harvest (kuzyk et al. 2016), sample sizes and distribution of study areas on the provincial landscape were likely insufficient to adequately estimate losses at the provincial scale or account for temporal changes in harvest rate. the amount of unlicenced harvest of bulls and calves is also unknown at the provincial scale, and likely varied spatially and temporally during the 20-year study period. a system of reliably estimating first nations harvest would benefit provincial and regional moose population modelling by providing a more complete representation of harvest statistics. and, as with most studies, increasing the number and frequency of moose density estimates and composition surveys would also provide more reliable data and population estimates. despite these limitations, we were able to assess moose population dynamics and licenced harvest trends over a 20-year period suitable for management purposes. we used science-based and repeatable methods to provide an assessment of provincial and gmz moose populations, their temporal trends, and predictive relationships useful for developing management strategies. our approaches and ability to detect declining population trends will be beneficial when compiling future, broadscale ungulate population trend assessments. acknowledgements we thank d. heard, m. gillingham, d. dupont, and v. harriman for their review comments that improved this manuscript, and m. anderson for producing the figures. references alfaro, r. i., l. van akker, and b. hawkes. 2015. characteristics of forest legacies following two mountain pine beetle outbreaks in british columbia, canada. canadian journal of forest research 45: 1387–1396. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. bangs, e. e., t. n. bailey, and m. f. portner. 1989. survival rates of adult female moose on the kenai peninsula, alaska. journal of wildlife management 53: 557–563. bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior alaska. journal of wildlife management 66: 747–756. boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1 (1992): 1–10. _____, and d. m. keppie. 1988. modelling a hunted moose population in new brunswick. alces 24: 201–217. boertje, r. d., w. c. gasaway, d. v. grangaard, and d. g. kelleyhouse. 1988. predation on moose and caribou by radio-collared grizzly bears in east central alaska. canadian journal of zoology 66: 2492–2499. _____, m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314–327. bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 116 _____, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvest. journal of wildlife management 71: 1494–1506. _____, p. valkenburg, and m. mcnay. 1996. increases in moose, caribou, and wolves following wolf control. journal of wildlife management 60: 474–489. boman m., l. mattsson, g. ericsson, and b. kristrom. 2011. moose hunting values in sweden now and two decades ago: the swedish hunters revisited. environmental and resource economics 50: 515–30. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999. moose on kalgin island: are density-dependent processes related to harvest? alces 35: 73–89. british columbia ministry of environment (bc moe). 2010. moose harvest management procedure manual. fish and wildlife branch, victoria, british columbia, canada. british columbia ministry of forests, lands and natural resource operations (bc flnro). 2015. provincial framework for moose management in british columbia. fish and wildlife branch, victoria, british columbia, canada. brown, g. s. 2011. patterns and causes of demographic variation in a harvested moose population: evidence for the effects of climate and density-dependent drivers. journal of animal ecology 80: 1288–1298. bunnefeld, n. e. hoshino, and e. j. milner-gulland. 2011. management strategy evaluation: a powerful tool for conservation? trends in ecology and evolution 26: 441–447. burnham, k. p., and d. s. r. anderson. 2002. model selection and multimodel inference: a practical informationtheoretic approach, 2nd edition. springerverlag, new york, new york, usa. christie, k. s., w. f. jensen, j. h. schmidt, and m. s. boyce. 2015. long-term changes in pronghorn abundance index linked to climate and oil development in north dakota. biological conservation 192: 445–453. crichton, v. 1993. hunter effort and observations – the potential for monitoring trends of moose populations – a review. alces 29: 181–185. decescare, n. j., j. r. newby, v. j. boccadori, t. chilton-rad, t. t. thier, d. waltee, k. podruzny, and j. a. gude. 2016. calibrating minimum counts and catch-per-unit-effort as indices of moose population trend. wildlife society bulletin 40: 537–47. _____, t. d. smucker, r. a. garrot, and j. a. gude. 2014. moose status in montana. alces 50: 35–51. dimarco, m., m. l. boitani, d. mallon, m. hoffmann, a. iacucci, e. meijaard, p. visconti, j. schipper, and c. rondinini. 2014. a retrospective evaluation of the global decline of carnivores and ungulates. conservation biology 28: 1109–1118. eastmam, d., and r. ritcey. 1987. moose habitat relationships and management in british columbia. swedish wildlife research supplement 1: 101–117. eberhardt, l. l. 2002. a paradigm for population analysis of long-lived vertebrates. ecology 83: 2841–2854. fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52: 14–21. _____, a. r. e. sinclair, and g. caughley. 2014. wildlife ecology, conservation, and management. john wiley and sons, new york, new york, usa. fulton, d. c., and k. hundertmark. 2004. assessing the effects of a selective harvest system on moose hunters’ behaviors, beliefs, and satisfaction. human dimensions of wildlife 9: 1–16. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 117 gaillard, j.-m., m. festa-bianchet, and n. g. yoccoz. 1998. population dynamics of large herbivores: variable recruitment with constant adult survival. trends in ecology and evolution 13: 58–63. _____, _____, _____, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual review of ecology and systematics 31: 367–393. gasaway, w. c., r. d. boertje, d. grangaard, d. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. _____, s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. institute of arctic biology, university of alaska, fairbanks, alaska, usa. _____, r. o. stephenson, j. l. davis, p. e. k. shepard, and o. e. burris. 1983. interrelationships of wolves, prey and man in interior alaska. wildlife monographs 84. gorley, a. 2016. a strategy to help restore moose populations in british columbia. prepared for the ministry of forests lands and natural resource operations. fish and wildlife branch, victoria, british columbia, canada. haddon, m. 2001. modelling and quantitative methods in fisheries. chapman & hall/ crc, london, united kingdom. hatter, i. w. 1998. a bayesian approach to moose population assessment and harvest decisions. alces 34: 47–58. _____. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. _____. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces 37: 71–77. hauge, t. m., and l. b. keith. 1981. dynamics of moose populations in northeastern alberta. journal of wildlife management 45: 573–597. hayes, r. d., r. farnell, r. m. p. ward, j. carey, m. dehn, g. w. kuzyk, a. m. baer, c. l. gardner, and m. o’donoghue. 2003. experimental reduction of wolves in the yukon: ungulate responses and management implications. wildlife monographs 152. heard, d. c., a. b. d. walker, j. b. ayotte, and g. s. watts. 2008. using gis to modify a stratified random block survey design for moose. alces 44: 111–116. hurley, m. a., m. hebblewhite, j. m. gaillard, s. dray, k. a. taylor, w. k. smith, p. zager, and c. bonenfant. 2014. functional analysis of normalized difference vegetation index curves reveals overwinter mule deer survival is driven by both spring and autumn phenology. philosophical transactions of the royal society b 369 (1643): 20130196. karns, p. d. 2007. population distribution, density and trends. pages 125–139 in a. w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. kelsall, j. p. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research supplement 1: 1–10. kuzyk, g. w. 2016. provincial population and harvest estimates of moose in british columbia. alces 52: 1–11. _____, and i. w. hatter. 2014. using ungulate biomass to estimate abundance of wolves in british columbia. wildlife society bulletin doi: 10.1002/wsb.475. _____, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. bc moose population dynamics – kuzyk et al. alces vol. 54, 2018 118 british columbia ministry of forest, lands and natural resource operations. victoria, british columbia, canada. _____, s. marshall, m. klaczek, c. procter, b. cadsand, h. schindler, and m. gillingham. 2016. determining factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife working report no. wr-123. progress report: february 2012 – 30 april 2016. british columbia ministry of forest, lands and natural resource operations, victoria, british columbia, canada. larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rate of moose mortality in southwest yukon. journal of wildlife management 53: 548–557. leblanc, j. e., b. e. mclaren, c. pereira, m. bell, and s. atlookan. 2011. first nations moose hunt in ontario: a community’s perspectives and reflections. alces 47: 163–174. lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–10. lendrum, p. e., c. r. anderson jr., k. l. monteith, j. a. jenks, and r. t. bowyer. 2013. migrating mule deer: effects of anthropogenically altered landscapes. plos one 8: e64548. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. british columbia ministry of forests, special report series number 6. british columbia ministry of forests, victoria, british columbia, canada. milner-gulland, e. j., m. v. kholodova, a. bekenov, o. m. bukreeva, i. a. grachev, l. amgalan, and a. a. lushchekina. 2001. dramatic declines in saiga antelope populations. oryx 35: 340–345. monteith k. l., v. c. bleich, t. r. stephenson, b. m. pierce, m. m. conner, j. g. kie, and r. t. bowyer. 2014. life-history characteristics of mule deer: effects of nutrition in a variable environment. wildlife monographs 186. mowat, g., d. c. heard, and c. j. schwartz. 2013. predicting grizzly bear density in western north america. plos one 8: e82757. doi: 10.1371/journal. pone.0082757 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate change influences on a declining moose population. wildlife monographs 166. mytton, w. r., and l. b. keith. 1981. dynamics of moose populations near rochester, alberta, 1975–1978. canadian field-naturalist 95: 39–49. nadeau, m. s., n. j. decesare, d. g. brimeyer, e. j. bergman, r. b. harris, k. r. hersey, k. k. huebner, p. e. matthews, and t. p. thomas. 2017. status and trends of moose populations and hunting opportunity in the western united states. alces 53: 99–112. palazy, l., c. bonenfant, j. m. gaillard, and f. courchamp. 2012. rarity, trophy hunting and ungulates. animal conservation 15: 4–11. patterson, b. r., j. f. benson, k. r. middel, k. j. mills, a. silver, and m. e. obbard. 2013. moose calf mortality in central ontario, canada. journal of wildlife management 77: 832–841. peters, w., m. hebblewhite, k. g. smith, s. m. webb, n. webb, m. russe, c. stambaugh, and r. b. anderson. 2014. contrasting aerial moose survey population estimate methods and evaluating sightability in west-central alberta, canada. wildlife society bulletin 38: 639–649. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. national park service scientific monograph series, no. 11. alces vol. 54, 2018 kuzyk et al. – bc moose population dynamics 119 quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43–54. raithel, j. d., m. k. kauffman, and d. h. pletscher. 2007. impact of spatial and temporal variation in calf survival on the growth of elk populations. journal of wildlife management 71: 795–803. resources information standards committee (risc). 2002. aerial-based inventory methods for selected ungulates: bison, mountain goat, mountain sheep, moose, elk, deer and caribou. standards for components of british columbia’s biodiversity no. 32, version 2.0. british columbia ministry of sustainable resource management, victoria, british columbia, canada. ritchie, c. 2008. management and challenges of the mountain pine beetle infestation in british columbia. alces 44: 127–135. saether, b. e. 1997. environmental stochasticity and population dynamics of large herbivores: a search for mechanisms. trends in ecology and evolution 12: 143–149. schwartz, c. c. 2007. reproduction, natality, and growth. pages 141–171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. shackleton, d. 1999. hoofed mammals of british columbia. royal british columbia museum handbook. university of british columbia press, vancouver, canada. skalski, j. r., k. e. ryding, and j. j. millspaugh. 2005. wildlife demography: analysis of sex, age and count data. elsevier academic press, cambridge, massachusetts, usa. spalding, d. j., and j. lesowski. 1971. winter food of the cougar in south central british columbia. journal of wildlife management 35: 378–381. stedman, r., d. r. diefenbach, c. b. swope, j. c. finley, a. e. luloff, h. c. zinn, g. j. san julian, and g. a. wang. 2004. integrating wildlife and human-dimensions research methods to study hunters. journal of wildlife management 68: 762–773. stenhouse, g. b., p. b. latour, l. kutny, n. maclean, and g. glover. 1995. productivity, survival, and movements of female moose in a low-density population, northwest territories, canada. arctic 48: 57–62. telfer, e. s. 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145–182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. timmermann, h. r. 1992. moose sociobiology and implications for harvest. alces 28: 59–77. _____, and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. van ballenberge, v., and j. dart. 1982. harvest yields from moose populations subject to wolf and bear predation. alces 18: 258–275. western, d., s. russell, and i. cuthill. 2009. the status of wildlife in protected areas compared to non protected areas of kenya. plos one 4: e6140. white, g. c., and b. c. lubow. 2002. fitting population models to multiple sources of observed data. journal of wildlife management 66: 300–309. _hlk524694781 pone.0163249.ref001 alces36_69.pdf 4301.pdf alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 13 modeling the impact of moose and wolf management on persistence of woodland caribou réhaume courtois1,2 and jean-pierre ouellet3 1ministère des ressources naturelles et de la faune, direction du développement de la faune, 880, ch. ste-foy, 2e étage, québec, pq, canada g1s 4x4; 2université laval, département de foresterie et de géomatique, cité universitaire, ste-foy, pq, canada g1k 7p4; 3jean-pierre ouellet, université du québec à rimouski, département de biologie, centre d’études nordiques, 300, allée des ursulines, rimouski, pq, canada g5l 3a1 abstract: limiting factors of caribou (rangifer tarandus) populations vary regionally. in tundra environments, this species appears to be regulated by food, either because wolves (canis lupus) are absent or because migration of caribou allows escape from predation during part of the year. in the boreal forest, the main limiting factors are hunting and predation but because of low caribou densities, no regulation mechanism seems to exist between caribou and wolves. moose (alces alces) is the primary prey species of wolves and consequently, if moose abundance increases, wolves should also increase, independently of the caribou population. thus, caribou could experience high predation rates and be eliminated in high wolf densities. here we attempted to identify the necessary conditions to maintain caribou numbers in the presence of moose. to do so, we built a deterministic model that simulated the relationship between a caribou population regulated by food competition and limited by predation, a moose population regulated by predation, and a wolf population, the abundance of which is determined by moose abundance. at current hunting rates for caribou and moose in the boreal forest, and in the absence of wolf trapping, the model predicted that the caribou population would be extirpated in approximately 100 years. wolf trapping was not adequate to conserve the caribou population unless very intensive control was undertaken. in the absence of trapping, cessation of caribou hunting allowed a 3-fold increase in caribou numbers over the long term, if the moose population remained low. according to our model, the best management measure for caribou consisted of maintaining a low moose density through appropriate population and habitat management strategies, which prevented expansion of the wolf population and limited predation on caribou. alces vol. 43: 13-27 (2007) key words: caribou, hunting, interactions, moose, predation, simulation, trapping, wolf the hypothesis of exploitation ecosystems predicts that the number of trophic levels and population regulating factors depend on the primary productivity of an ecosystem (oksanen et al. 1981, oksanen and oksanen 2000). in very poor environments (e.g., high arctic and deserts, where productivity is < 40 g/m2/year), ecosystems consist of vegetation communities that are regulated by competition for resources. in poor environments (e.g., tundra: 40 – 700 g/m2/year), plants are regulated by herbivores. in contrast, more productive environments are made up of three trophic levels that are regulated from the top. here, herbivores are regulated by carnivores that have only a limited impact on vegetation (oksanen 1988), so both carnivores and plants because regulating factors of caribou (rangifer tarandus) vary between ecotypes, the exploitation ecosystem model seems appropriate for prediction of caribou population changes. in tundra environments, low productivity and low carrying capacity conbiological communities. predators are even absent on certain arctic islands (klein 1968, ouellet et al. 1996) and in such cases, caribou impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 14 populations are regulated by competition for food and population changes can be described by a logistic model (caughley 1977). net population growth rate varies with density: and then growth rate gradually slows down as populations approach carrying capacity. in such environments, the carrying capacity seems to be on the order of 60 – 100 caribou per 100 km2, as observed on coats island in hudson bay where caribou have occurred since at least the beginning of the 20th century (ouellet et al. 1996). massive mortalities from starvation sometimes take place overwinter due to food overexploitation or density-independent factors, such as climatic conditions preventing access to food (e.g., ice crust; klein 1968, reimers 1982). in continental tundra, caribou populations also seem to be regulated by food. caribou may undertake very long migrations that allow escape from predation most of the year, partly because wolf (canis lupus) packs are because wolf packs cannot move away from a den site when raising pups during summer (bergerud 1996). in tundra environments, prepredation rate does not increase as a function of caribou abundance. caribou numbers can therefore increase and, at high density, they can overexploit food available during summer (bergerud 1996, crête and doucet 1998) or winter (ouellet et al. 1994, 1996, 1997). as resources available per animal diminish with increasing population, a decrease in birth rate and an increase in adult and calf mortality are noted. densities averaging 60 – 110 caribou per 100 km2 have been observed in this environment (messier et al. 1988, seip 1991). using the annual increase in lichen biomass, which is the primary food source for caribou (gauthier et al. 1989), along with losses caused by animal trampling, arsenault et al. (1997) have estimated the carrying capacity of caribou in tundra at approximately 20 animals per 100 km2. in the boreal forest, the carrying capacity for caribou is not precisely known. based on lichen biomass, crête and manseau (1996) estimated that carrying capacity should be at least that observed in tundra, because alternative food sources (such as leaves, twigs, and deciduous shrubs) are abundant and the climate is milder in that environment. for example, in east-central québec, carrying capacity based solely on terrestrial lichen has recently been estimated at 4.1 – 7.7 caribou per 100 km2 (courtois 2003). despite a relatively high potential carrying capacity, woodland caribou populations experience very low densities of between 1 and 3 individuals per 100 km2 (seip 1991, courtois 2003) and most are declining in north america (mallory and hillis 1998). those observations suggest that woodland caribou populations for winter food. moreover, as caribou food habits are much less restrictive in summer, the main limiting factors seem to be hunting and predation (stuart-smith et al. 1997, rettie and messier 1998). caribou densities in the boreal forest (1 – 3 per 100 km2; seip 1991, courtois et al. 2003) are typically incapable of supporting wolf populations. wolves in the boreal forest therefore depend on moose (alces alces), a larger and more abundant ungulate that generally lives at densities of between 10 and 20 individuals per 100 km2 (messier 1985). due to the absence of a regulating mechanism between caribou and wolves in the boreal forest (seip 1991), an increase in moose abundance should provoke a wolf population increase, independent of caribou abundance. in such a situation, more frequent encounters between wolf and caribou should be expected leading to an increased predation rate on caribou that could decline, eventually down to extirpation (seip 1991). caribou abundance decreases when there are more than 0.60 – 0.65 wolves per 100 km2 (bergerud and elliot 1986, bergalces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 15 erud 1996) and caribou populations increase when wolf numbers are controlled (boertje et al. 1996, hayes et al. 2003). the necessity for caribou to adopt avoidance strategies of both predators and other ungulates in order to survive is now recoget al. 1984; bergerud 1985, 1996; seip 1991, 1992; cumming et al. 1996; stuart-smith et al. 1997; rettie and messier 1998). however, the mechanisms underlying caribou, moose, and wolf interactions still remain obscure. in this study, we built a deterministic model that mimicked the relationships between moose, wolves, and caribou and then simulated how variation in hunting of caribou and caribou abundance. model results should help to determine which management strategies are most suitable to maintain equilibrium between these three species in order to help maintain caribou, a threatened species in the boreal forest. methods the model the model simulates changes in caribou, moose, and wolf numbers in a 1,000-km2 study area of the québec boreal forest where the three species live in sympatry. according to the model, in the absence of wolves, the moose and caribou populations are regulated by food competition and follow a logistic equation (nt+1 = nt r 1 nt/kcc , where nt and nt+1 = numbers at time t and time t+1, respectively, r = maximum population growth rate, and kcc = food carrying capacity; caughley 1977). in the presence of wolves, the moose population is regulated by predation in accordance with the predation model of messier (1994). wolf numbers are determined by moose numbers, as predicted by the michaelis-menten hyperbolic equation of messier (1994). in the combined model, where the three species interact, wolves have access to caribou and carry out density-independent survival of calves and adults, as predicted by the bergerud and elliot (1986) model. we completed the model by adding management parameters, which allowed changing the moose and caribou hunting and wolf trapping mortalities that were additive to natural mortality. additional parameters allowed stochastic variation of moose and caribou productivity due to uncontrolled environmental conditions (crête and courtois 1997). the model (appendix i) was elaborated using stella 4.0 software (isee systems, lebanon, nh, usa). in summary, our combined model is of predation or when predators are controlled, caribou and moose populations are regulated by competition for food (messier 1994, crête and manseau 1996); (2) in the presence of wolves, the moose population is regulated by wolf predation (messier 1994); (3) wolf numbers are determined by moose abundance (messier 1994), but there is no dependance of wolves on caribou numbers (seip 1991); (4) caribou predation increases non-linearly with wolf abundance (bergerud and elliot 1986); and (5) there is no immigration or emigration in the system or these two opposite processes are equal. caribou population parameters due to the paucity of woodland caribou population dynamics data, maximum growth rate and food carrying capacity were taken from published data for barren-ground caribou. maximum population growth rate of caribou was based on observations of a population reintroduced on southampton island in hudson bay (ouellet et al. 1996, 1997). in 1991, this population was estimated at 13,700 1,580 adults, with an annual growth rate (r) of 0.245 since the introduction of 38 one-yearold individuals in 1967 (nt = n0 e rt, hence 13,700 = 38 er 24; where nt = number of caribou at time t, n0 = number of caribou in 1967, and t = time in years). no decline in growth impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 16 rate was observed over the 24 years, so this value of r should be close to the maximum attainable by a caribou population that is subject to negligible hunting (< 1% per year) and lacks wolf predation. maximum growth rate is comparable to that reported for the caribou herd introduced on st. matthew island in the bering sea (17 – 29%; klein 1968). carrying capacity based on availability of terrestrial lichens was estimated at 20 caribou per 100 km2 (crête and manseau 1996, arsenault et al. 1997). caribou can eat various items (arboreal lichens, deciduous leaves, sedges, forbs, etc.), but these were not included in the carrying capacity estimates to avoid overestimation since winter diet is largely dominated by terrestrial lichens (gauthier et al. 1989) and this season is the most likely to be limiting caribou (klein 1968). initial winter density was set at 1.63 individuals per 100 km2, with 16% calves in the population, as noted in a recent aerial survey (courtois et al. 2003). annual losses due to predation and hunting (adults only) were initially established at 3% and 8%, respectively, as noted in central québec (courtois 2003). we considered that using empirical data would allow more realistic simulations for the québec boreal forest. moose population parameters maximum annual growth rate of the moose population was estimated to be 25%, based on observations in newfoundland in the absence of predation (fryxell et al. 1988). similar rates of increase were reported in south-central québec (23-24%, laurian et al. 2000). other parameters were taken from studies carried out in east-central québec where densities are approximately 3.0 moose per 100 km2 (gingras et al.1989). the habitat carrying capacity was estimated at 84 moose per 100 km2, based on annual production of deciduous twigs available during winter (courtois et al. 1993). although this estimate may appear low, it is four times higher than the minimum density required to maintain viable wolf populations (20 moose per 100 km2; messier 1985). natural mortality rate was set at 9.2% per year (of which 4.5% was due to causes other than predation) while annual hunting mortality was estimated to be 9.0% (courtois et al. 1994b). as for caribou, these estimates were the only ones available for the québec boreal forest. additional simulations involved annual stochastic variation in birth rate in order to mimic productivity changes (0.56 – 1.00) that were caused by variations in snowfall and summer temperature (crête et courtois 1997). it would have been possible to use stochastic variability in all parameters but we preferred using actual estimates to allow realistic simulations for the study site. besides, anthropogenic mortality does not vary substantially from year to year because hunting regulation remains stable in the area. finally, interpretation would have many parameters. wolf population estimation our model considers wolf population size to be determined by moose population size, according to the predation model developed by messier (1994, 1995) from 27 north american studies. this author demonstrated that it was possible to predict wolf abundance as a function of moose density (numerical response: number of wolves/1,000 km2 = 58.73 [number of moose/km2 – 0.031]/[0.76 + number of moose/km2], r2 = 0.62). the number of moose exponentially as a function of prey density then decelerates and starts declining at 0.65 moose/km2 (functional response: number of moose killed per wolf per 100 days = 3.3 number of moose per km2/[0.46 + number of moose per km2], r2 = 0.53; messier 1994). in this model, the total response is the product of numerical and functional responses. following recommendations of messier (1994), a correction factor of 0.71 was applied to winter predation rates to obtain annual rates and values for 100 days were converted to alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 17 annual values. the impact of wolves on the caribou population was estimated using the bergerud and elliot (1986) density-independent model that was based on 17 north american studies. we thus predicted annual recruitment rate (percentage calves during winter = e(3.340-0.127 wolves/1,000 km2), r2 = 0.69) and annual mortality rate of adults (percentage adult mortality = 4,766 + (0.699 wolves per 1,000 km2)1,275, r2 = 0.73). simulations parameters and initial and calculated variables are presented in appendix ii. preliminary simulations were performed on caribou and moose populations separately illustrated the trajectories of populations that petition for food. in a second trial, we used the moose-wolf model of messier (1994) to illustrate the reaction of wolves in the presence of moose only. after these preliminary trials, we ran the combined model using the three species’ interactions to investigate the outcome of a limited number of realistic management scenarios over a 100-year span. four deterministic scenarios were conducted. in trajectories under management strategies that had prevailed in central québec until autumn 2000 (i.e., moose and caribou hunting at 9% and 8%, respectively, but no wolf trapping). then we successively simulated effects of (2) a caribou hunting ban; (3) intensive wolf trapping (at 30%; larivière et al. 2000), and (4) moose hunting increased to 15% (in order to allow only a light population increase) without wolf trapping. finally (5), in order to experiment with the effects of changes not controlled by harvest regulation, we included negative effects of stochastic environmental variations on productivity (crête and courtois 1997) of moose (from 0% – 40%) and caribou (20%), on scenarios 1 and 2. results simple models for caribou, moose, and wolf populations in the absence of hunting and predation, caribou reached their carrying capacity (k) in 53 years, with numbers increasing from 16 to 200 individuals in the 1,000-km2 site. half of the carrying capacity (optimal density 0.5k) was reached in 10 years, allowing a maximum sustainable yield of 12 caribou per year in a stable population of 100 individuals. adding a natural mortality of 3% resulted in 0.5 k being reached in 12 years, with a maximum sustainable yield of 9 individuals per year. when an additional anthropogenic harvest of 8% was included, the population reached 0.5 k after 28 years and did not exceed 112 caribou after 100 years. in the absence of hunting and predation, the 30 initial moose reached their carrying capacity (840 individuals) in 62 years. the maximum sustainable yield (52 moose) was attained with a population of 420 moose (optimal density: 42 moose per 100 km2) in 13 years. including a natural mortality of 9% resulted in 0.5 k being reached in 25 years, but the total population did not exceed 537 moose after 100 years. after adding another 9% hunting mortality, the population comprised only 234 moose (0.28 k) after 100 years. according to the predation model (messier 1994), the wolf population was sustained by the moose population and also regulated moose numbers. both wolf abundance and predation rate increased with moose abundance. without hunting, the moose population stabilized at 650 individuals (65 moose per 100 km2). at this plateau, predator density reached 2.58 wolves per 100 km2, as predicted by the numerical response equation of messier (1994). impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 18 combined model incorporating interactions between moose, wolves and caribou scenario 1: moderate moose (9%) and caribou (8%) hunting but no wolf trapping.— when moose and caribou were exploited at moderate rates and in the absence of wolf trapping, moose abundance increased from 30 to 110 individuals in 55 years and then remained stable (fig. 1). simultaneously, wolves increased from 0 at the start of the run to 5 individuals after 30 years only to stabilize afterwards. moose predation rate reached 8% per year, which stabilized the moose-wolf system. caribou abundance increased from 16 to 40 individuals in 14 years and then gradually declined to quasi-extirpation in 100 years (2 caribou remaining). the caribou decline initiated when moose and wolf reached 76 and 32 individuals (7.6 and 0.32 individuals per 100 km2), respectively. scenario 2: ban on caribou hunting with 9% moose hunting and no wolf trapping.— the moose and wolf population trajectories, since the relevant population parameters remained the same (fig. 2). the caribou hunting ban allowed the caribou population to increase and reach 94 individuals after 22 years; however, despite this increase, the caribou population started to decline as soon as moose numbers attained 95, thus supporting 4.5 wolves (9.5 moose and 0.45 wolves per 100 km2). caribou were not extirpated and the population stabilized at approximately 64 individuals (6.4 per 100 km2). scenario 3: intensive wolf trapping (30%) with 9% moose and no caribou hunting.— introducing intensive trapping of wolves (30%) to parameters of the previous simulation did not markedly improve the situation for caribou. wolf abundance increased steadily despite exploitation, in parallel with the increase in moose abundance (fig. 3). abundance after 100 years was 5 wolves and 141 moose (0.5 and 14.7 per 100 km2, respectively). the caribou population reached 109 individuals after 23 years, but then declined and stabilized at (75 individuals 7.5 per 100 km2). scenario 4: increased moose hunting (15%) with no wolf and caribou harvesting.— intensifying moose harvest to 15% but discontinuing wolf trapping maintained the moose population at a relatively low level of 61 individuals (6.1 moose per100 km2), which directly reduced the wolf population to less than 2.2 individuals (< 0.22 wolves per 100 km2; fig. 4). this management strategy was very favourable for caribou and the population reached 146 individuals after 30 years and then stabilized at 141 caribou (14.1 per 100 km2). scenario 5: stochastic environmental variations.— with random annual variation in birth rates, each simulation followed a new trajectory (fig. 5). stochastic variation 0 20 40 60 80 100 120 0 10 20 30 40 50 60 70 80 90 100 time (years) n o .i n d iv id u a l s moose caribou wolf fig. 1. moose, wolf, and caribou population trajectories with moderate moose (9% per year) and caribou (8% per year) hunting but no wolf trapping (scenario 1). 0 20 40 60 80 100 120 0 10 20 30 40 50 60 70 80 90 100 time (years) n o .i n d iv id u a ls moose caribou wolf fig. 2. moose, wolf, and caribou population trajectories with moderate moose hunting (9% per year) but no caribou hunting or wolf trapping (scenario 2). alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 19 tions in both moose and wolf numbers, which always remained lower than in previous simulations. these variations had a positive impact on caribou numbers. in the presence of moderate moose harvesting (9%) but no wolf trapping or caribou hunting (previous scenario 2), the caribou population reached 125-135 individuals and stabilized at this level (fig. 5a). if caribou hunting was also introduced into this model (previous scenario 1), the caribou population stabilized at approximately 50 – 60 individuals (fig. 5b). when environmental stochasticity was allowed to increase the variation in caribou birth rate in addition to the moose birth rate, the caribou population declined to approximately 35 individuals (fig. 5c). equilibrium points of the combined models theoretical equilibrium points of the simple moose-wolf model were 65.0 moose and 2.58 wolves per 100 km2. in the moosewolf-caribou model, equilibrium points were much lower than these values. in simulations without stochastic environmental variation, equilibrium points were approximately 7 – 10 0 20 40 60 80 100 120 140 160 0 10 20 30 40 50 60 70 80 90 100 time (years) n o .i n d iv id u a l s moose caribou wolf fig. 3. moose, wolf, and caribou population trajectories with moderate moose hunting (9% per year) and intensive wolf trapping (30% per year) but no caribou hunting (scenario 3). 0 20 40 60 80 100 120 140 160 0 10 20 30 40 50 60 70 80 90 100 time (years) n o .i n d iv id u a l s moose caribou wolf fig. 4. moose, wolf, and caribou population trajectories with moose hunting set at 15% per year but no caribou hunting or wolf trapping (scenario 4). a) 0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 time (years) n o .i n d iv id u a ls _ b) 0 20 40 60 80 100 120 140 0 20 40 60 80 100 time (years) n o .i n d iv id u a ls _ c) 0 20 40 60 80 100 120 140 0 20 40 60 80 100 time (years) n o .i n d iv id u a ls _ moose caribou wolf fig. 5. typical moose, wolf, and caribou population trajectories according to different management strategies and by including random variations in moose (all simulations) and caribou birth rates (last simulation) (scenario 5). (a) moderate moose hunting (9% per year), with no caribou hunting or wolf trapping; moose birth rate varying up to 40% a year; (b) moderate moose (9% per year) and caribou (8% per year) hunting with no wolf trapping; moose birth rate varying randomly (0 – 40%); (c) moderate moose (9% per year) and caribou (8% per year) hunting with no wolf trapping; both moose and caribou birth rates varying randomly, between 0 and 40% and 0 and 20%, respectively. impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 20 moose, 0.3 – 0.5 wolves, and 3 – 14 caribou per 100 km2. in the last simulation, which appears a plausible situation in nature, populations stabilized at 6 – 8 moose, 0.2 – 0.3 wolves, and 3.5 – 4 .0 caribou per 100 km2. discussion usefulness of the combined model in caribou, moose, and wolf management our moose-wolf-caribou model suggests interesting alternatives for the management of these three species. seip (1991, 1992) proposed that predation could eliminate woodland caribou whenever wolves are sustained by another species because there is no retroactive mechanism that decreases the impact of wolves when the caribou population declines. our simulations suggest that certain management strategies could help maintain caribou numbers by limiting wolf expansion. wolf control could be considered (bergerud and elliot 1986, seip 1991), but this strategy would only have a minor impact on caribou, unless a very intensive control is performed. weclaw and hudson (2004) obtained similar results using another model based on responses of wolves, moose, caribou, lichens, and vascular plants to various natural and anthropogenic factors. a 30% harvest rate exerted by trappers is not high birth rate, wolf populations can increase quickly if moose abundance is high. cessation of caribou hunting seems more of the caribou population in the long term if the moose (and wolf) population does not the habitat carrying capacity for moose is relatively low. moose densities are generally higher in disturbed habitats and young forests rich in deciduous browse than in mature spruce forests (timmermann and mcnicol 1988). moose expansion could therefore be controlled by limiting forest exploitation in sites used by caribou or by promoting spruce wolves, our simulations suggest that caribou can be maintained in natural ecosystems, at least in the absence of human disturbance as suggested by weclaw and hudson (2004). the best strategy, however, would consist of maintaining low moose densities through population and habitat management. low moose densities imply low wolf densities and, therefore, low predation rates on caribou. increasing the harvest of moose would be the easiest and most convenient management measure to adopt. moose hunting is highly popular. with a 3to 4-week hunting season, the harvest rate could reach 15% if all segments of the population (both sexes of adults and calves) were targeted without setting a limit on the number of hunting permits issued (courtois et al. 1994a). in such a situation, our simulation suggests that moose density would stabilize at about than 7 – 8 individuals per 100 km2 and wolf density at less than 0.2 – 0.3 individuals per 100 km2. annual caribou recruitment would then be approximately 19 – 20% and adult predation rate would be around 7 – 8%. a population increase in caribou to 3 – 14 individuals per 100 km2 could thus be seen in several decades. even if conservative caribou hunting was maintained, a doubling or tripling of the population could be observed over the long term, depending on the magnitude of variations due to environmental factors. this scenario supports the growing evidence that moose management may be an integrated aspect of caribou management, as indicated by controlled experiments in the moose hunting rates higher than 15% in order to maintain a stable or slightly increasing population. management that encourages declining moose populations will lead to lower hunting success rate, which in the long term may be a less popular decision among hunters. our simulations seem optimistic in comparison with actual woodland caribou densities (1 – 3 individuals per 100 km2, seip 1991, courtois 2003); however, historical densities alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 21 were much higher. brassard (1967) measured a density of 26.2 caribou per 100 km2 in a 25,723 km2 study site in the boreal forest of the québec north shore. in labrador, bergerud (1967) estimated the 1958 density of the mealy mountain herd at 7.9 individuals per 100 km2 over 30,303 km2. both these densities rapidly declined following intensive hunting during the 1960s and 1970s following the invention of snowmobiles that greatly facilitated hunter access. similarly, caribou density was approximately 4 – 5 individuals per 100 km2 in 1992 in the charlevoix herd, north of québec city, in the absence of hunting and restricted forest exploitation (sebbane 2003). caribou declines followed anthropogenic and natural habitat disturbance in the absence of hunting, presumably due to increased predation. controlled experiments (boertje et al. 1996, hayes et al. 2003) and other simulation models (weclaw and hudson 2004) also suggest that caribou densities can attain 10 individuals per 100 km2, depending on the extent of wolf predation and hunting-related mortality. effects of increase in caribou biomass we intentionally developed a simple model based on empirical data from the quéorder to avoid including variables that were depended exclusively on moose density. this ence of ungulates is necessary to sustain wolf populations (messier 1994, 1995) and also caribou density was too low to constitute a white-tailed deer, could be an important prey for wolves during winter, but this deer is not found where caribou live in the northern boreal forest. in sites used by woodland caribou, for wolves. although wolves certainly consume beavers in summer, which presumably increases pup survival, beavers are not readily available in winter because beaver movements are mostly restricted to beneath the ice where nevertheless, we tested the effect of a possible caribou-wolf retroaction mechanism. we the carrying capacity of wolves depended on the total biomass of available ungulates (fuller 1989) and we considered that 1 caribou was equivalent to 0.29 moose, based on respective masses. as a result, the wolf population increased by 2% compared to scenario 2 (harvest rates: 9%, 0%, and 0% for moose, caribou, and wolves, respectively), which created population declines of 14% in moose and 6% in caribou. assuming that wolves make hunting decisions based solely on prey biomass, not risk associated with their capture (which is lower for caribou than moose), the impact of an increase in wolves would probably be greater on moose than caribou. seip (1991) and bergerud (1996) have suggested that caribou could be extirpated from an area in the presence of high moose densities due to increased predation on caribou from increased wolf densities. in contrast, hayes et al. (2003) have suggested that reduced moose densities would lead to caribou extirpation in areas where wolves, moose, and caribou use similar habitats in summer because in such a case caribou lose the advantage of their spacing strategy to avoid predation. in the model including both moose and caribou biomass to support wolves, increasing the moose harvest rate to 22% led to extirpation of this species in 100 years, while caribou increased to 156 individuals. despite a 10-fold increase, the caribou biomass (4.5 moose-equivalent per 100 km2 a viable wolf population that subsequently declined to 1 animal after 100 years, suggesting that an abundant primary prey (moose or deer) is required to maintain wolves in the boreal forest. our model implies that natural and anthropogenic mortality are additive. this assumption is probably valid for ungulates, impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 22 but wolf trapping could be partly compensatory (fuller 1989). harvesting wolves would indeed leave more food available for each surviving wolf, which could contribute to a wolf population increase. we did not have any empirical evidence to quantify this possibility, however. moreover, if trapping is extensive, the reduced wolf packs should be biomass per animal. limits of the model simulation results are strongly dependent on which combinations of parameters are used and on the accuracy of the parameters included. caribou and moose densities, food carrying capacity, and hunting mortalities were based on empirical data from the québec boreal forest (courtois 2003), but wolf densities and their related impact on caribou numbers could be greater than those predicted by our model. larivière et al. (2000) used the number of wolf howls heard by hunters to predict wolf density. for the port-cartier – sept-îles wildlife reserve, which is located in the main distribution area of woodland caribou, these authors estimated a density of 0.85 wolves per 100 km2. using fuller’s model (1989), 0.48 wolves per 100 km2 would be obtained. the model we used (messier 1994) predicted 0.22 wolves per 100 km2 for a similar moose abundance (6 individuals per 100 km2), and maximum density of 0.38 wolves per 100 km2 at carrying capacity (84 moose per 100 km2). only an inventory could provide accurate wolf numbers in the study area. however, if the wolf densities predicted by larivière et al. (2000) and fuller (1989) are correct, more restrictive measures would be needed to conserve caribou in the presence of moose and wolves. similarly, the population growth rate for both by certain environmental variables, such as snow conditions (schaefer and messier 1991). therefore, simulations that do not include the impact of such stochastic limiting factors are likely to yield optimistic results. bergerud and elliot (1986) estimated that caribou abundance should decline when wolf density exceeds 0.65 individuals per 100 km2. in our simulations, caribou numbers started declining as soon as the wolf population exceeded 0.45 individuals per 100 km2. caribou declines always followed an expansion phase, which means that food density-dependent mechanisms intervened when caribou numbers increased and that these mechanisms were additive to predation effects. the carrying capacity that we selected (20 caribou per 100 km2, arsenault et al. 1997) could also be too high based on recent estimates (4.1 – 7.7 individuals per 100 km2, courtois 2003), and recent studies suggest that caribou avoid human-related infrastructures leading to loss of available habitat (see weclaw and hudson 2004 for a review). black bear (ursus americanus) are not abundant in prime caribou habitat, but this species can exert an important additional density-independent pressure on caribou in some areas (ballard 1994), particularly the southern region of the boreal forest. therefore, the model may be inappropriate in areas with bears, cougars, or other predators. taken management decisions. model improvements improvements could be made to our model. for example, the wolf population could be considered as a reservoir, rather than as a parameter linked to the moose population. this change would allow consideration of other variables that affect wolves, such as the presence of other prey species or demographic uncontrolled variables and measurement errors were responsible for approximately 40% of the variance in wolf density between studies consulted by messier (1994). the regulation mechanism between moose and wolves in our alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 23 model, however, seems more appropriate to mimic natural predator-prey dynamics than a priori parameters. furthermore, instead of considering ungulate populations as homogeneous reservoirs, we could categorize populations by age and only if proportions of males and females vary substantially. according to sensitivity analyses performed by fancy et al. (1994), moose and caribou population growth rates are not highly sensitive to age structure, but rather depend on recruitment rate, which is strongly and their survival. more importantly, a spatially explicit model would be particularly useful to investigate the possibility of excluding wolves from areas used by caribou, for example through habitat management that renders areas unsuitable for moose. such a model could be based on the movement rate and direction of wolves could help to determine the minimal size of a protected area for caribou that would permit avoidance of most encounters between wolves and caribou. acknowledgements we wish to thank pierre etcheverry for helpful insights during elaboration of the model and selection of the parameters. we also address special thanks to pierre drapeau, robert hudson, françois potvin, jim schaefer, and two anonymous reviewers for their helpthe ministère des ressources naturelles et de la faune, the association des manufacturiers de bois de sciage du québec, the fondation de la faune du québec, abitibi-consolidated inc., and kruger (scierie-manic). references arsenault, d., n. villeneuve, c. boismenu, y. leblanc, and j. deshaye. 1997. estimating lichen biomass and caribou grazing on the winter grounds of northern québec: an application of fire history and landsat data. journal of applied ecology 34:65-78. ballard, w. b. 1994. effects of black bear predation on caribou – a review. alces 30:25-35. banfield, a. w. f. 1974. les mammifères du canada. les presses de l’université laval, québec, québec, canada. bergerud, a. t. 1967. management of labrador caribou. journal of wildlife management 31:621-635. _____. 1985. antipredator strategies of caribou: dispersion along shorelines. canadian journal of zoology 63:1324-1329. _____. 1996. evolving perspectives on caribou population dynamics, have we got it right yet? rangifer, special issue 9:95-116. _____, and j. p. elliot. 1986. dynamics of caribou and wolves in northern british columbia. canadian journal of zoology 64:1515-1529. _____, r. d. jaminchuk, and d. r. carruthers. 1984. the buffalo of the north: woodland caribou at calving. animal behavior 39:360-368. boertje, r. d., p. valkenburg, and m. e. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal of wildlife management 60:474-489. brassard, j. m. 1967. inventaire aérien des ongulés sauvages de la côte-nord et identification des aires d’hivernement en fonction des formes du relief et de la végétation. service de la faune. québec, québec, canada. caughley, g. 1977. analysis of vertebrate populations. john wiley and sons, london, u.k. courtois, r. 2003. la conservation du caribou forestier dans un contexte de perte d’habitat et de fragmentation du milieu. impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 24 ph.d. thesis. université du québec à rimouski, rimouski, québec, canada. _____, m. crête, and f. barnard. 1993. productivité et dynamique d’une population d’orignaux du sud de la taïga québécoise. ministère du loisir, de la chasse et de la pêche. québec, québec, canada. _____, a. gingras, c. dussault, l. breton, and j.-p. ouellet. 2003. an aerial survey technique for the forest-dwelling ecotype of woodland caribou, rangifer tarandus caribou. canadian field-naturalist 117:546-554. _____,y. leblanc, j. maltais, and h. crépeau. 1994a. québec moose aerial surveys: methods to estimate population characteristics and improved sampling srategies. alces 30:159-171. _____, d. sigouin, j. -p. ouellet, a. beaumont, and m. crête. 1994b. mortalité naturelle et d’origine anthropique de l’orignal au québec. ministère de l’environnement et de la faune. québec, québec, canada. crête, m., and r. courtois. 1997. limiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive forests. journal of zoology 242:765-781. _____, and j. doucet. 1998. persistent suppression in dwarf birch after release from heavy summer browsing by caribou. arctic and alpine research 30:126-132. _____, and m. manseau. 1996. natural regulation of cervidae along a 1000 km latitudinal gradient: change in trophic dominance. evolutionary ecology 10:51-62. cumming, h. g., d. b. beange, and g. lavoie. 1996. habitat partitioning between woodland caribou and moose in ontario: the potential role of shared predation risk. rangifer, special issue 9:81-94. fancy, s. g., k. r. whitten, and d. e. russell. 1994. demography of the porcupine caribou herd, 1983-1992. canadian journal of zoology. 72:840-846. fryxell, j. m., w. e. mercer, and r. b gellatel. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52:14-21. fuller, t. k. 1989. population dynamics of wolves in north-central minnesota. wildlife monographs 105. gauthier, l., r. nault, and m. crête. 1989. variations saisonnières du régime alimentaire des caribous du troupeau de la rivière george, québec nordique. naturaliste canadien 116:101-112. gingras, a., r. audy, and r. courtois. 1989. inventaire aérien de l’orignal dans la zone de chasse 19 à l’hiver 1987-88. ministère du loisir, de la chasse et de la pêche. sept-îles, québec, canada. hayes, r. d., r. farnel, r. m. p. ward, j. carey, m. dehn, g. w. kuzyk, a. m. baer, c. l. gardner, and m. o’donoghue. 2003. experimental reduction of wolves in the yukon: ungulate responses and management implications. wildlife monographs 152. klein, d. r. 1968. the introduction, increase, and crash of reindeer on st. mattew island. journal of wildlife management 32:350-367. larivière, s., h. jolicoeur, and m. crête. 2000. status and conservation of the gray wolf (canis lupus) in wildlife reserves of québec. biological conservation 94:143-151. laurian, c., j. -p. ouellet, r. courtois, l. breton, and s. st-onge. 2000. the effects of intensive harvesting on moose reproduction. journal of applied ecology 37:515-531. mallory, f. f., and t. l. hillis. 1998. demographic characteristics of circumpolar caribou populations: ecotypes, ecological constraints/releases, and population dynamics. rangifer, special issue 10:49-60. messier, f. 1985. social organization, spaalces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 25 tial distribution and population density of wolves in relation to moose density. canadian journal of zoology 63:10681077. _____. 1994. ungulate population models with predation: a case study with north american moose. ecology 75:478-488. _____. 1995. trophic interactions in two northern wolf-ungulate systems. wildlife research 22:131-146. _____, j. huot, d. le henaff, and s. luttich. 1988. demography of the george river caribou herd: evidence of population regulation by forage exploitation and range expansion. arctic 41:279-287. oksanen, l. 1988. ecosystem organization: mutualism and cybernetics or plain darwinian struggle for existence? the american naturalist 131:424-444. _____, and t. oksanen. 2000. the logic and realism of the hypothesis of exploitation ecosystem. the american naturalist 155:703-723. _____, d. stephen, j. a. fretwell, and n. pekka. 1981. exploitation ecosystems in gradients of primary productivity. the american naturalist 118:240-261. ouellet, j.-p., s. boutin, and d. c. heard. 1994. responses to simulated grazing and browsing of vegetation available to caribou in the arctic. canadian journal of zoology 72:1426-1435. _____, d. c. heard, s. boutin, and r. mulders. 1997. a comparison of body condition and reproduction of caribou on two predator-free arctic islands. canadian journal of zoology 75:11-17. _____, _____, and r. mulder. 1996. population ecology of caribou populations without predators: southampton and coats island herds. rangifer, special issue 9:17-26. reimers, e. 1982. winter mortality and population trends of reindeer on svalbard, norway. arctic and alpine research 14:295-300. rettie, w. j., and f. messier. 1991. the implications of environmental variability on caribou demography: theoretical consideration. rangifer, special issue 7:53-59. _____, and _____. 1998. dynamics of woodland caribou populations at the southern limit of their range in saskatchewan. canadian journal of zoology 76:251-259. sebbane, a., r. courtois, s. st-onge, l. breton, and p. -é. lafleur. 2003. trente ans après sa réintroduction, quel est l’avenir du caribou de charlevoix? naturaliste canadien 127:55-62. seip, d. r. 1991. predation and caribou populations. rangifer, special issue 7:46-52. _____. 1992. factors limiting woodland caribou populations and their interrelationships with wolves and moose in southeastern british columbia. canadian journal of zoology 70:1494-1503. stuart-smith, a. k., j. a. corey, s. boutin, d. h. hebert, and a. b. rippin. 1997. woodland caribou relative to landscape patterns in northeastern alberta. journal of wildlife management 61:622-633. timmermann, h. r., and j. g mcnicol. 1988. moose habitat needs. the forestry chronicle 64:238-245. weclaw, p., and r. j. hudson. 2004. simulation of conservation and management of woodland caribou. ecological modeling 177:75-94. impact of moose and wolves on caribou – courtois and ouellet alces vol. 43, 2007 26 appendix i stella model incorporating moose, wolf, and caribou interactions in the boreal forest. rectangles represent reservoirs (populations), double arrows identify sources (births) and sinks tions tested the impact of changes in caribou and moose hunting rates (caribouhuntingrate and moosehuntingrate, respectively), wolf trapping rate (wolftrappingrate) and changes in caribou and moose productivity due to random variations in environmental conditions (caribourandomenvcdn and mooserandomenvcdn) on caribou and moose reproductive rates. other parameters and equations were obtained from the literature (see methods). minprodcaribou and minprodmoose indicate the minimum productivity of the species under the worst environmental conditions. other acronyms are self explanatory. alces vol. 43, 2007 courtois and ouellet impact of moose and wolves on caribou 27 appendix ii parameters and variables used in the models, values employed in the simulations, and data sources. parameter or variable values employed data source moose habitat carrying capacity 840 in the study area (84 per 100 km2) courtois et al. (1993) population 30 in the study area at time 0; calculated by the model afterwards gingras et al. (1989) maximum growth rate 25% per year fryxell et al. (1988) annual predation rate calculated from messier’s predation model (depends on moose density) messier (1994) annual hunting rate 9% (simulations 1, 2, 3, 5) or 15% (simulation 4) courtois et al. (1994b) annual rate of mortality due to other causes 4.50% courtois et al. (1994b) annual variability of recruitment 0% (simulations 1 to 4); random variation between 0 and 40% (simulation 5) crête and courtois (1997) wolf population calculated from messier’s michaelismenten hyperbolic equation (depends on moose population) messier (1994) trapping rate 0% (simulations 1, 2, 4, 5) or 30% (simulation 3) larivière et al. (2000) caribou habitat carrying capacity 200 in the study area (20 per 100 km2) arsenault et al. (1997) population 163 in the study area at time 0; calculated by the model afterwards courtois et al. (2003) maximum growth rate 24.5% per year estimated from ouellet et al. (1996, 1997) data annual recruitment rate calculated from bergerud and elliot’s model (depends on wolf density) bergerud and elliot (1986) annual rate of adult natural mortality calculated from bergerud and elliot’s model (depends on wolf density) bergerud and elliot (1986) annual hunting rate 9% (simulations 1, 2, 3, 5) or 15% (simulation 4) courtois et al. (2003) annual variability of recruitment 0% (simulations 1 to 4, 5a and 5b); random variation between 0 and 20% (simulation 5c) arbitrary f:\alces\vol_38\pagemaker\3818. alces vol. 38, 2002 courtois et al. habitat selection in cut areas 177 habitat selection by moose (alces alces) in clear-cut landscapes réhaume courtois1, christian dussault2, françois potvin1, and gaétan daigle3 1 société de la faune et des parcs du québec, direction de la recherche sur la faune, 675 boulevard rené-lévesque est, 11e étage, boîte 92, québec, pq, canada g1r 5v7; 2 université laval, département de biologie, cité universitaire, québec, pq, canada g1k 7p4; 3 université laval, service de consultation statistique, cité universitaire, québec, pq, canada g1k 7p4 abstract: habitat selection by moose was studied over 4 years in two large sectors subject to intensive forest harvesting using a two-scale approach. at the coarser scale, i.e. location of the home range within the landscape, habitat selection did not appear to be influenced by the presence of clear-cuts. in one sector, moose preferred mature mixed stands, young coniferous, and mature coniferous stands. in the second sector, the highest preference was noted for cut areas and mature deciduous stands. moose home ranges were located in areas with higher edge and interspersion among habitat patches. home range size for females was positively related to the proportion of cuts, but movements were not. habitat selection was more pronounced at the finer scale (animal locations within home range) and did not differ between sectors. mixed stands were preferred in all seasons. mature conifer stands were preferred in summer and in early winter while young conifer stands were preferred in late winter. clear-cuts were avoided except in early winter. moose were located in areas closer to edge between food and cover stands than were random locations, especially in late winter. a marked decrease in movements also was noted in late winter. this study shows differences in habitat selection pattern between the coarser and finer scales. for example, clear-cuts did not seem to markedly influence home range location at a coarser scale, and adaptations to minimize their impact seemed to operate at a finer scale. coarser scale habitat selection was probably linked to a trade-off between predator avoidance and browse availability, whereas seasonal changes suggest behavioural adaptations of moose to maximize energy gain and counteract predation and other adverse environmental conditions at the finer scale. alces vol. 38: 177-192 (2002) key words: alces alces, clear-cuts, cover, food, forest management, habitat selection, mortality resumé: la sélection d’habitat par l’orignal a été étudiée durant quatre ans dans deux grands secteurs comportant d’importantes coupes forestières. deux échelles d’analyse ont été retenues. à l’échelle brute, correspondant à l’emplacement du domaine vital dans le paysage, la sélection d’habitat ne paraissait pas influencée par la présence de coupes totales. dans l’un des secteurs, l’orignal préférait les peuplements mélangés matures ainsi que les résineux jeunes et mûrs. dans le deuxième secteur, la préférence la plus grande était notée pour les coupes forestières, les feuillus matures et les mélangés matures. les domaines vitaux de l’orignal étaient localisés dans des sites comportant plus de bordure et plus d’entremêlement entre les parcelles d’habitat. chez les femelles, la superficie du domaine vital était positivement liée à l’importance des coupes, mais les déplacements ne l’étaient pas. la sélection d’habitat était plus prononcée à l’échelle plus fine (localisation des animaux à l’intérieur du domaine vital) et elle ne différait pas entre les deux secteurs. les peuplements mélangés étaient préférés en toutes saisons. les résineux matures l’étaient en été et en début d’hiver alors que les jeunes résineux étaient préférés en fin d’hiver. les coupes étaient évitées, sauf en début d’hiver. les localisations d’orignaux étaient situées plus près des bordures entre des peuplements d’alimentation et de couvert que ne l’étaient des localisations aléatoires. une diminution marquée des déplacements était aussi notée en fin d’hiver. cette étude montre que les préférences d’habitat diffèrent entre les échelles brute et fine. ainsi, les coupes forestières ne habitat selection in cut areas courtois et al. alces vol. 38, 2002 178 johnson (1980) suggested that habitat selection, i.e. the decision to choose a specific habitat, is a hierarchical process, with decisions being made at different spatial and temporal scales. for example, an animal can choose to establish its home range (hr) in an area dominated by dense conifer stands while preferring rare deciduous stands within its hr to fulfil nutritional requirements. at a coarser scale, animals are expected to select habitats that reduce main limiting factors such as predation (rettie and messier 2000). when animals have successfully overcome the more important limiting factors at the coarser scale, selection could diverge at a finer scale to meet more specific needs, for example adequate shelter during the calving period, food intake in summer, and protection against unfavourable environmental conditions in late winter. moose (alces alces) thrive in young boreal forests originating from perturbations (timmermann and mcnicol 1988). the highest densities are recorded in mixed stands (brassard et al. 1974, joyal 1987, crête 1988) and in areas disturbed by forest fires, insect outbreaks, forest harvesting, or windfall (krefting 1974, peek et al. 1976, timmermann and mcnicol 1988, loranger et al. 1991). such stands provide abundant deciduous shrubs that constitute the main food source for moose (crête 1989). the benefits of forest harvesting for moose, however, are restricted to the 15-40 year period following disturbance (cowan et al. 1950, crête 1977, franzmann and schwartz 1985). over the short-term, clearcutting leads to landscapes dominated by large openings separated by small strips of uncut forest. such recently clear-cut landscapes seem much less favourable to moose than older cuts since previous studies have indicated low moose densities in these areas (eason et al. 1981, girard and joyal 1984, eason 1989). increased moose mortality caused by hunting and predation following forest harvesting were suggested to explain these low densities (eason et al. 1981, girard and joyal 1984, dalton 1989, eason 1989) and as shown by rempel et al. (1997). consequently, cutovers seem less favourable for moose survival. whether or not moose avoid these areas in order to increase their survival and at which scale such an avoidance could occur is unknown. in this study, we used radio-telemetry to evaluate habitat selection by moose in an area subject to intensive forest harvesting during the last 10-15 years. we used a twoscale approach to test (1) if moose avoid landscapes with clear-cuts, and (2) if they avoid clear-cuts and edges within their hr. at a coarser scale (hr within the landscape), we expected moose to select habitats that are believed to decrease their vulnerability to predation and hunting (eason et al. 1981, girard and joyal 1984, dalton semblaient pas influencer grandement l’emplacement des domaines vitaux à l’échelle brute, et les adaptations visant à diminuer leur impact semblaient opérer à fine échelle. à l’échelle brute, la sélection d’habitat semblait résulter d’un compromis entre l’évitement des prédateurs et la recherche de nourriture alors qu’à l’échelle fine, les changements saisonniers suggéraient des adaptations comportementales visant à maximiser les gains énergétiques et à contrer les effets de la prédation et des conditions environnementales adverses. alces vol. 38: 177-192 (2002) mots clés: alces alces, aménagement forestier, coupes totales, couvert, mortalité, nourriture, sélection d’habitat alces vol. 38, 2002 courtois et al. habitat selection in cut areas 179 1989, eason 1989). at that scale moose s h o u l d e x h i b i t l o w p r e f e r e n c e (disproportional use, hall et al. 1997) for areas dominated by clear-cuts, and should increase their hr size and movements in cut areas as suitable habitats are more scattered. at a finer scale (within hr), we expected that habitat use would be oriented toward maximizing energy gains and would vary seasonally depending on the severity of environmental conditions. habitat requirements are more easily fulfilled during the growing season and in early winter when high quality food is abundant and snow (< 60 cm) does not restrict movement. however, during the growing season, females must protect their calves from predation by wolves (canis lupus) and black bears (ursus americanus) (ballard 1992). consequently, in summer and early winter, moose should then look for stands offering both ample food supply to maximize nutritional intake (schwartz 1992), and cover to keep predation risk at a moderate level (edenius 1992), such a trade-off being more pronounced in females than males. moose are sensitive to heat stress in late winter and their movements are hampered by snow depth and hardness (timmermann and mcnicol 1988). during this period, moose should reduce their daily movements, prefer cover to food supply, and be located closer to edges between food and cover than in other seasons (peek et al. 1992). methods study area the study was conducted in a 3,200 km2 area located in northwestern québec, canada (fig. 1). the study area is a typical boreal forest dominated by conifer stands that were intensively harvested (22% of the fig. 1. location of the study area in northwestern québec (78° 40’ w, 47° 50’ n). dots indicate sites of capture of marked animals. habitat selection in cut areas courtois et al. alces vol. 38, 2002 180 area) between mid-1980 and 1994. at the end of the study, 31% conifer, 13% mixed, and 8% deciduous stands covered the area. other habitat classes available include stands currently unproductive for forestry (12%, mostly alder, alnus rugosa), conifer stands a f f e c t e d b y t h e s p r u c e b u d w o r m (choristoneura fumiferana) outbreak (5%) during the 1980s, and water bodies or lakes (9%). dominant trees are black spruce (picea mariana), balsam fir (abies balsamea), jack pine (pinus banksiana), paper birch (betula papyrifera), and trembling aspen (populus tremuloides). clearcutting without protecting advanced regeneration was practiced in the study area 7 to 11 years (94% of all cuts) before the study. in areas cut between 1992 and 1994, fellerbunchers and skidders circulated in trails spaced 10-15 m apart to protect advanced regeneration. the terrain is gently rolling with hills rarely exceeding 350 m above sea level. temperature (mean ± se (n years)), as measured at the nearby belleterre meteorological station, was -16.2°c ± 0.6 (29) in january, and 17.3°c ± 0.2 (29) in july. annual precipitation was 1013 mm ± 23 (24) including 291 mm ± 12 (26) that fell as snow. maximum snow depth (66 cm ± 7 (17)) occurs in february and did not exceed 90 cm during the study. wolf and black bear are found in the study area at approximately 1 and 14 individuals / 100 km2 respectively (lamontagne et al. 1999, larivière et al. 2000). capture and telemetry sixty-five moose (1991: 29; 1992: 16; and 1993: 20) were equipped with radio transmitters. track networks in the snow were identified by aerial surveys, numbered and randomly selected to determine the collaring sequence. males (14), females (37), and calves (14) were collared according to their proportion in the population of the hunting zone in which the study was conducted (paré and courtois 1990). all animals present in a chosen track network were collared provided they corresponded to the desired category. moose were immobilized with succinylcholine chloride (van ballenberghe 1989, delvaux et al. 1999), at a dosage of 5-6 mg / 100 kg of visually estimated body weight. moose were equipped with lmrt-4 vhf transmitters (lotek engineering, newmarket, ontario) and ear tagged. to estimate habitat use and movements, moose were located from an aircraft (helicopter: 96%; fixed wing: 4%) in 3 seasons: summer and fall (15 april to 31 october), early winter (mid-december to late january: snow depth < 60 cm), and late winter (mid-february-end of march: snow depth > 60 cm). each year, marked animals were located 3-4 times per season. transmitters included motionless sensors that allowed the identification of dead individuals. for dead animals, the cause of death was determined by inspection of the carcass and its surrounding area. predation was presumed when the presence of wolves was obvious around the carcass. animals killed by hunters were reported to mandatory registration stations. home range estimates the minimum convex polygon (eddy 1977) was retained to estimate hr size due to its widespread use in the literature (aebischer et al. 1993). post-capture outliers (distance > 10 km in < 3 days postcapture) were discarded to eliminate their effect on hr size, leaving 2,160 valid locations (1991: 297 locations; 1992: 626; 1993: 732; and 1994: 505). most animals were followed for 2-3 years, yielding 32.6 locations per animal (se = 2.1; n = 65 moose). the plot of multi-annual hr size against the number of locations per animal indicated an inflection point at 20 locations per animal, as noted by courtois et al. (1998a). for moose alces vol. 38, 2002 courtois et al. habitat selection in cut areas 181 collared at 0.5-year-old, only locations collected after they were 18 months old were used in analyses to avoid the influence of dispersing animals (labonté et al. 1998) and to prevent pseudo-replication with the data of their mothers, which were also monitored. consequently, analyses were conducted only for the 47 moose (11 males and 36 females > 1.5 year-old) having > 20 locations. multi-annual hr sizes were calculated in arcview gis 3.1 (esri 1996) using the animal movement extension developed by hooge and eichenlaub (1997). habitat composition we used the term “habitat classes” to designate forest stands and unproductive areas identified on forest maps. habitat composition was obtained from digitized 1:20,000 forest maps produced by interpretation of 1:15,000 aerial photographs (mer 1984). the minimum size of stands on these maps is usually 8 ha. this information was used to produce a 10 m × 10 m raster map imported into arcview 3.1 and managed with the spatial analyst extension (esri 1996). to maximize the statistical power of comparisons (alldredge and ratti 1992) and to avoid having missing values in the availability data sets (aebischer et al. 1993), habitats were grouped into eight classes according to their age (young: 20-60 years; mature: >60 years) and composition (y_dec, m_dec, y_mix, m_mix, y_con, m_con, respectively, young and mature deciduous, mixed and coniferous stands; cuts: clearcuts < 11 years; other: water (mainly) and unproductive stands). all cuts were grouped because those harvested between 1992 and 1994 were not available to all animals. such grouping of habitat classes provides good correspondence between the forest maps and ground survey validations (dussault et al. 2001). it also reflects the basic requirements of moose in terms of food (browse, leaves, and aquatics) and cover (heat stress, snow) (thompson and stewart 1998). cuts can potentially provide summer and winter food as well as y_dec and m_dec which could also offer summer cover against heat; other was expected to give access to aquatics; y_mix and m_mix could offer both food and cover all year long while y_con and m_con were supposed to provide summer and winter cover. one habitat map per year (1 april to 31 march) was produced to take into account the chronology of cutting operations. habitat composition of a given area (i.e., study area, hr) was computed as the percentage of that area covered by each habitat class. habitat structure habitat structure at the coaser scale was estimated with 7 fragstats landscape pattern indices (mcgarigal and marks 1995) computed with the arcview extension patch analyst (elkie et al. 1999). we selected two indices to measure edge (the length of the interface) between habitat patches (ed: edge density, perimeter of all habitat patches per unit area [m/ha]; cwed: contrast-weighted edge density [m/ha], edge between deciduous or mixed vs. coniferous stands [weights = 1; other edges = 0]). one metric (cad: core area density) was selected to estimate the density of core areas per 100 ha. core area was defined as the interior of the patch exclusive of a band 100 m wide immediately inside the patch boundary. the size of the buffer was based on observations indicating that moose overused the 100 m zone on each side of food-cover edges (dussault 2002). two indices measured the diversity of the landscape (sdi: shannon’s diversity index, amount of information per habitat patch [without units]; iji: interspersion and juxtaposition index, measure the extent to which patch types are equally adjacent to each other [percent]). two other indices were selected to quantify the shape and size of stands (msi: mean shape index, average habitat selection in cut areas courtois et al. alces vol. 38, 2002 182 perimeter-to-area ratio [without units]; tcai: total core area index, relative importance (%) of the core areas). finally, an index of edge preference was calculated as the mean distance (m) between moose locations and the nearest edge between deciduous or mixed vs. coniferous stand. data analysis habitat components were ranked in order of preference, rather than classified as preferred or avoided, because preference and avoidance depend markedly upon the array of components assumed to be available (johnson 1980). in all analyses, the level of rejection of the null hypothesis was set at α= 0.05. conclusions drawn from habitat selection studies are critically dependent upon t h e d e l i n e a t i o n o f a v a i l a b l e h a b i t a t (aebischer et al. 1993, mcclean et al. 1998, wilson et al. 1998). to ensure that habitats deemed available were really accessible to each animal, we delimited two sectors (rapide-sept, 1,065 km2, in the south eastern part of the study area; rouyn-noranda, 1,184 km2, in the north western) using 100% convex polygons joining all the locations of moose frequenting each sector (fig. 1). each of these sectors were 13-15 times the size of the mean hr size, which seems sufficient to provide a precise estimate of availability outside a given hr. at the same time such an area remains potentially accessible to moose. for example, dispersal of yearlings is usually 10-30 km and some animals can travel up to 100 km before settling down (labonté et al. 1998). preference analyses (percent used/percent available) were conducted using the animal as the sample unit (aebischer et al. 1993). zero values in percent use were replaced by a small value (0.01%) about one order of magnitude smaller than the minimum encountered in the data set. analyses were conducted at two scales to identify: (1) preferences revealed by the multiannual hr location within the landscape (coarser scale); and (2) seasonal preferences within the hr (finer scale). at the coarser scale, habitat within the hr was defined as use while availability corresponded to habitat composition of the sector where each animal was collared. to test the influence of landscape pattern indices, each moose hr was paired with 30 areas of identical size and shape (random hr) randomly positioned (x and y coordinates of the centroid) and oriented (0-360 degrees of rotation from the observed orientation) in the sector frequented by that animal (wilson et al. 1998, potvin et al. 2001, dussault 2002). the use of random hr facilitated the comparison of landscape pattern indices because it eliminates the influence of size and shape on landscape pattern indices of hr under comparison (mcgarigal and marks 1995). preference at the coarser scale was studied using the 1994 habitat map. however, this will overestimate moose preference towards cuts since some stands that were frequented before cutting will appear to be cut in the analysis. any conclusion that cutovers are avoided will therefore be conservative. at the finer scale, habitat composition in a 100-m buffer zone around each moose location was compared to that of the total hr of that animal. the buffer minimized the influence of spatial imprecision of habitat map and telemetry locations (rettie and mcloughlin 1999). the size of the buffer was relatively small because locations were precise (20-100 m; potvin 1998) due to the use of a helicopter that permitted seeing most animals (69%) during telemetry locations. habitat use was estimated using the annual forest map corresponding to each location. data collected during a given season (summer, early winter, and late winter) and year were pooled per animal to avoid temporal pseudo-replication (thus givalces vol. 38, 2002 courtois et al. habitat selection in cut areas 183 ing one record per animal per season of each year). edge preference (distance [m] from an edge) was evaluated by comparing moose locations with paired random locations (one per moose location) within the hr of each animal. paired locations were retained to assure that habitat characteristics at the random locations were accessible to the animal under consideration. preference analyses (aebischer et al. 1993) were preceeded by multivariate analyses of variance (proc glm, sas institute 1989) to evaluate the effect of independent factors (year, sector, sex, season, and their interactions) on habitat preference for all habitat classes taken simultaneously. in these analyses, percentage use and availability were transformed into log-ratios using other habitats as the denominator to render habitat variables independent (aebischer et al. 1993). multivariate normality of the residuals was assessed with mardia skewness and kurtosis (program multnorm.sas, available from sas institute). potential outliers were identified with the program outlier.sas (m. friendly, http:/ / w w w . m a t h . y o r k u . c a / s c s / s s s g / outlier.html). predicted preference indices were estimated by a posteriori univariate analysis (proc mixed, sas institute 1989) followed by a bootstrap procedure (available at gaetan.daigle@mat.ulaval.ca). the bootstrap technique with 500 samples was used in order to approximate the exact distribution of the estimation procedure, conditional on the observed data (efron and tibshirani 1993). bootstrap technique allows estimating the statistic of interest with many subsamples drawn from the original data with replacement. the associated distribution of the estimates is an approximation of its true distribution, and thus can be used in statistical analyses. univariate analyses (proc glm) were used to test the effect of independent factors on hr size, female productivity, and distance between successive locations. we used the same factors as in the multivariate analyses but time interval between two successive locations was added in analyses involving distance. the normality and homoscedasticity of the residuals were evaluated respectively with the shapiro-wilk’s test and the plot of the residuals against predicted values. paired t-tests were used to compare (1) landscape pattern indices between moose hr and the mean value of the 30 paired random hrs, and (2) the distance from an edge between moose and random locations. this last analysis was followed by an univariate analysis to assess the influence of independent factors on that variable. to test the influence of clear-cuts on moose mortality, animals were partitioned into 2 groups according to the proportion of cuts within their hr (few: < 35%; many: > 35%; based on the mean proportion found in hr). proc lifetest (sas institute 1989) was then used to compare equality of survival curves over these two groups using the non-parametric log-rank test. results habitat preference at the coarser scale at the coarser scale, two outliers were removed to respect the multinormality of the residuals. no transformation of the data permitted reaching multinormality with these two outliers. we detected a sector effect in habitat preference (f[ 7,35 ] = 9.25, p <0.0001) but no sex (f[ 7,35 ] = 0.31, p = 0.9461) or sector*sex effect (f[ 7,35 ] = 1.25, p = 0.3009). consequently, preferences were estimated by sector. important differences were noted between the two sectors (fig. 2). in the rapide-sept sector, the highest preference was noted for mature mixed stands followed by mature coniferous and young coniferous. the rank order for all habitat classes was: m_mix > m_con = y_con > m_dec = cuts = y_mix = other > y_dec. in habitat selection in cut areas courtois et al. alces vol. 38, 2002 184 the rouyn-noranda sector, the most preferred habitat classes were clear-cuts followed by mature deciduous, and all other habitat classes (cuts > m_dec > m_mix = other = m_con = y_con = y_dec = y_mix). six of the seven habitat pattern indices differed between moose and random hr (table 1). moose hr had higher total edge and number of core areas per surface area, higher interspersion and diversity (iji and sdi), and lower perimeter-to-area ratio, and core areas size than random hr within the landscape. the difference between moose and random hr was particularly important for the number of core areas (7.1 vs. 5.1 core areas/ha) and total edge (81 vs. 69 m/ ha). other significant differences were low (<9%). habitat preference at the finer scale habitat preference within the hr varied among seasons (f[ 14,566 ] = 2.24, p < 0.0059) without any sector (f[ 7,37 ] = 1.80, p < 0.1156), sex (f[ 7,37 ] = 0.50, p < 0.8266) or year (f[ 21,813 ] = 1.43, p < 0.0960) influence, and no significant interactions. preference analyses indicated that mature mixed and coniferous stands, and mature deciduous stands, were the most preferred habitat classes in summer and fall (m_mix = m_con > m_dec > y_con = other > cuts = y_dec = y_mix). in early winter, moose preferred mature mixed stands, and cuts and mature coniferous (m_mix > cuts = m_con > m_dec > y_con = other = y_dec = y_mix). in late winter, only m_mix and y_con were preferred over other habitat classes (m_mix = y_con > m_con = y_mix = y_dec = other = m_dec = cuts). globally, the highest preferences were noted for mature mixed in all seasons (fig. 3). the major seasonal changes were an important increase in the preference for cuts in early winter, and y_con in late winter. young and mature deciduous, and other habitats were more preferred in summer-fall than in the two other seasons. moose were located closer to edge between deciduous or mixed stands and coniferous stands (food-cover edge) than random locations within the hr (mean difference = 52 ± 16 m; n = 297; t = 4.63, p < 0.001; fig. 4). changes were noted among seasons (f[ 2,296 ] = 9.39, p = 0.0001) but not between sex (f[ 1,296 ] = 0.55, p = 0.4615), sectors (f[ 1,296 ] = 1.91, p = 0.1736), year fig. 2. habitat preference (mean ± se) by moose at the coarse scale in the two sectors of the study area in northwestern québec, 1991-1994. preference indices (% used/% available) were standardized to sum to unity. 0.00 0.05 0.10 0.15 0.20 0.25 y_dec m_dec y_mix m_mix y_con m_con cuts other p re fe re n c e i n d e x rapide-sept rouyn-noranda alces vol. 38, 2002 courtois et al. habitat selection in cut areas 185 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 y_dec m_dec y_mix m_mix y_con m_con cuts other p re fe re n c e i n d e x summer-fall early winter late winter (f[ 3,296 ] = 2.31, p = 0.0767), or any interactions among factors. the difference was greater in late winter (80.9 m) than in the two other seasons (30.0 and 33.5 m). moose were closer to an edge in late winter (254.2 ± 22.9 m, n = 118 moose-season) than in summer-fall (279.4 ± 14.5, 120) or in early winter (282.1 ± 17.9, 124). effects of clear-cuts on space use, mortality and productivity mean multi-annual hr size was related to sex (87.7 km2 ± 20.3 and 73.7 ± 10.9 for males and females respectively; f[ 1,44 ] = 4.68, p = 0.037) and the interaction between sex and the proportion of cuts within the hr (f[ 1,44 ] = 9.23, p = 0.004); hr size increased with the proportion of cuts for females (r = 0.44, p = 0.009, n = 34) (fig. 5) but not for males. however, the proportion of cuts within hr, or sex, sector, and interactions among these variables did not influence the mean distance travelled between two successive locations (f[ 4,44 ] = 1.38, p = 0.2573). movements depended on season (f[ 2,346 ] = 32.59, p < 0.001), interaction table 1. mean value ± se of seven landscape pattern indices in moose (n = 47) and random (n = 30 per moose) home ranges in northwestern québec, 1991-1994. paired landscape pattern indices moose hr random hr student’s t p ed: total edge (m/ha) 80.6 ± 1.5 68.8 ± 0.8 7.40 <0.0001 cwed: food-cover edge (m/ha) 16.3 ± 0.7 16.5 ± 0.5 -2.27 0.7874 cad: n core areas beyond 100 m/100 ha 7.1 ± 0.3 5.1 ± 0.1 7.17 <0.0001 msi: perimeter to area ratio (without units) 1.84± 0.01 1.92 ± 0.01 -6.19 <0.0001 tcai: importance (%) of core areas beyond 100 m 88.1 ± 0.2 89.6 ± 0.1 -6.80 <0.0001 ji: interspersion of habitat patches (%) 80.0 ± 1.2 75.5 ± 0.7 3.81 <0.0004 sdi: shannon’s diversity index (without units) 1.75 ± 0.03 1.61 ± 0.02 4.94 <0.0001 fig. 3. seasonal changes in habitat preference (mean ± se) within the home range by moose in northwestern québec, 1991-1994. preference indices (% used/% available) were standardized to sum to unity. habitat selection in cut areas courtois et al. alces vol. 38, 2002 186 between year and season (f[ 6,346 ] = 2.48, p= 0.024), sector*year (f[ 2,346 ] = 5.06, p= 0.0069), and time interval between locations (f[ 1,146 ] = 3.99, p= 0.047). mean distance travelled by moose was lower in late winter during all the years of the study (fig. 6). twenty-three collared moose died during the study from hunting (17), predation (4), and other causes (2). survival rate of adults was similar for both sexes (χ2 = 1.84, df = 1, p = 0.1746; males: 83.2% ± 5.8 [12,364 collar-days]; and females: 86.9% ± 3.3 [33,340 collar-days]). similarly, no significant difference was found between survival for moose with many or few cuts in their hr (χ2 = 0.50, df = 1, p = 0.4812). female productivity depended on the sector (f[ 1,34 ] = 5.29, p = 0.0284) but not on the proportion of cuts within the hr (f[ 1,34 ] = 0.56, p = 0.4615), and no interaction was detected. discussion johnson (1980) suggested a four-order scale to explain habitat selection by an animal: the geographic range of the species, the hr, the specific habitats within the hr, and the micro-habitat (e.g., feeding sites). further, rettie and messier (2000) suggested that differences in selection pattern among the scales could reflect the importance of limiting factors with animals escaping the most important limiting factors at the coarser scales. habitat selection at the coarser and finer scales habitat selection patterns of moose differed between the coarser and finer scales in an area subject to intensive forest harvesting, suggesting that moose resource selection at the level of hr and within hr may reside in different domains (sensu wiens 1989). however, our results do not support the hypothesis that moose avoid clear-cuts at the coarser scale. in the rapide-sept sector, moose preferred forested habitats over clear-cuts, but the reverse trend was noted in the rouyn-noranda sector. females increased hr size, but not movements, in the presence of cuts. differences in habitat preference befig. 4. distance (mean ± se) between moose locations or random locations and the proximal limit of the nearest edge between deciduous or mixed stands vs. coniferous stands in northwestern québec, 1991-1994. fig. 6. seasonal changes in distance (mean ± se) travelled by moose between two successive locations in northwestern québec, 1991-1994. 0 50 100 150 200 250 300 350 400 450 summer-fall early winter late winter d is ta n c e f ro m a f o o d -c o v e r e d g e ( m ) moose random r² = 0.19 0 50 100 150 200 250 0 10 20 30 40 50 60 % of hr cut h r s iz e ( k m ²) 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 1991 1992 1993 1994 d is ta n c e t ra v e ll e d ( k m ) summer-fall early winter late winter fig. 5. influence of the proportion of cuts within the home range (hr) on female moose hr size in northwestern québec, 1991-1994. alces vol. 38, 2002 courtois et al. habitat selection in cut areas 187 tween sectors could not be related to different availability since both sectors comprised similar proportions of clear-cuts (24.4 vs. 23.6%) of similar age (≤ 11-year-old). the use of the 1994 habitat map and multiannual hr could have over-represented the proportion of clear-cuts within moose hr, but the same situation prevailed for both sectors. moreover, annual hrs were examined and the general trend was a propensity to stay in the same area despite the presence of clear-cuts. only 3 of the 47 animals seemed to gradually shift hr following the progression of the cuts. preference at the finer scale was more pronounced than at the landscape scale. in both sectors, a marked preference was noted for mature over young stands, principally mixed and coniferous ones. in early winter, however, there was an increase in preference for cuts. mixed stands and regenerated clear-cuts could provide significant browse availability (courtois et al. 1998b). during the summer-fall period, there was an increased preference for young and mature deciduous, and for other habitat class. deciduous trees can provide food and thermal cover while unproductive areas and water give access to aquatics. preference for cover slightly increased in late winter. according to thompson and euler (1987), in cut areas, moose preferably choose young cutovers in early winter before moving to older ones, and finally leave clear-cuts to use undisturbed areas in late winter. during that period, between midmarch and mid-april, they also reduce their movements (courtois and crête 1988) probably because snow accumulation > 65 cm and crust impede movements (peek et al. 1976). also, moose are in poorer physical condition in late winter and can be heat stressed when temperatures reach -5 to 0 ° c ( r e n e c k e r a n d h u d s o n 1 9 8 6 , timmermann and mcnicol 1988). many authors have concluded that moose seek coniferous shelter in late winter (desmeules 1964, brusnyk 1983, thompson and euler 1987), whereas crête (1988) doubted its necessity when deciduous browse is abundant. our results showed that moose reduced their movements during that period, used cover in a greater proportion, and clear-cuts in a lesser proportion, than in summer or in early winter, but they still preferred mixed stands. consequently, we suggest that dense cover is not a major component of late winter habitat, at least in regions where snow depth is usually < 90 cm as in our study area (paré and courtois 1990). similar findings were made in an agro-forested site and in black spruce clearcuts near our study area (girard and joyal 1984, joyal and bourque 1986) and also on the north shore of the st. lawrence river (courtois et al. 1993). moose decrease movements in deep snow but do not necessarily confine themselves under dense coniferous cover stands. for example, they can take advantage of small conifer patches within mixed stands. forbes and theberge (1993) also noted differences in habitat preference at different scales in algonquin park in ontario. at a regional scale (>1,000 km2), moose preferred sites with at least 33% of their area disturbed by spruce budworm epidemics or cuts (mainly uniform shelterwood and selection cutting). at a finer scale (<100 km2), they preferred conifer stands, mostly hemlock (tsuga canadensis), presumably because their study area was dominated by deciduous stands providing low cover. in our case, mature mixed stands were the most preferred habitat class both at the coarser and the finer scales in all seasons. moreover, moose searched for diversified landscapes with high edge, interspersion and number of cores areas per surface area, and a lower size of core areas relative to habitat patch size. consequently, preference at the coarser scale could result from habitat selection in cut areas courtois et al. alces vol. 38, 2002 188 a trade-off between predator avoidance and access to ample food supply (dussault 2002). at the finer scale, habitat selection seemed to be oriented towards maximizing energy gains and counterbalancing adverse effects of the environment (dussault 2002). presumed relationship between habitat selection and limiting factors we expected that moose would avoid clear-cut areas at the coarser scale because they are considered more vulnerable to hunting and predation in these environments (eason et al. 1981, girard and joyal 1984, dalton 1989, eason 1989). however, clear-cuts were not avoided in either sector, and mortality and productivity were not related to the abundance of clear-cuts. establishment of hr within the landscape may be more tightly related to sociological behaviour, namely philopatry for juvenile females and dispersal for males (cederlund et al. 1987, labonté et al. 1998) than habitat constraints. however, adopting philopatry as opposed to a more flexible strategy overcoming the influence of the factors affecting survival and reproduction could not be evolutionarily stable. habitat selection probably has some portion of genetic programming. our results suggest that avoidance of clear-cuts acts at the finer scale rather than at the coarser scale. it is also possible that moose may have tried to reduce the impact of several limiting factors at both scales. for example, it would not necessarily be a good strategy for a moose to choose a home range with very adverse environmental conditions even if there is no predator in that area. our approach of coarser scale habitat selection presupposes a top-down driven process due to limiting factors such as predation (rettie and messier 2000). an alternative hypothesis may be a bottom-up process where the establishment of the home range is the result of daily decisions in locating food patches and encounters with conspecifics and potential predators, like wolves and hunters, along with other variables such as daily and seasonal weather patterns. our data could support this second interpretation since analyses demonstrate equivocal results for clear-cut avoidance at the coarser scale, and the fact that moose are unlikely to explore large landscapes before selecting a home range. behavioural changes that could diminish the impact of clear-cuts seemed to operate mainly at the finer scale, i.e., within the hr. many studies have shown that moose, and particularly females with calves, avoid open habitats (peek et al. 1987, courtois and crête 1988, dalton 1989, eason 1989, hundertmark et al. 1990, dussault 2002). in late winter, females with calves rarely go 60 m beyond the forest cover, whereas other moose browse up to 80 m from the forest fringe (thompson and euler 1987). remaining close to cover could reduce predation risk. at the finer scale, cuts were one of the less preferred habitat classes, except in early winter. moose preferred the forested part of their hr which probably contributed to minimizing the influence of clear-cuts on mortality. it is not known whether the impact of clear-cuts depends on moose density. at a higher density and particularly near to carrying capacity, the impact of habitat modifications may have been higher. in such situations, more animals would have been forced to frequent marginal habitats and would have been exposed to higher mortality risks (sinclair and arcese 1995). management implications our work adds more evidence that moose and forest management can co-exist. however, recent clear-cuts are avoided except in early winter. the type of stands and the magnitude of forest harvesting determine the impact of logging (joyal 1987). alces vol. 38, 2002 courtois et al. habitat selection in cut areas 189 limited diameter cuttings carried out in shade-tolerant hardwood stands are usually favourable to moose (crête 1977). openings created by such forest harvesting techniques, in which usually 40% of the basal area of the deciduous trees and 75% of the conifers are removed, rapidly stimulate the growth of browse. in shade-intolerant hardwood, > 75% of the basal area of deciduous and conifers is usually removed. in coniferous stands where all commercial trees (> 9 cm) are harvested; the influence of cuts is more pronounced and depends on the time lag between cutting and regeneration of browse species and re-establishment of cover. an acceptable threshold for the proportion of cuts in the landscape can be estimated from the examination of the hr composition. in this study, moose hr comprised on average 30-40% of mixed, deciduous and spruce budworm-affected stands, 30-35% of coniferous stands, 2025% of clear-cut areas, and 10-15% of other habitats. in northeastern minnesota, peek et al. (1976) suggested creating landscapes comprising 40-50% of cutovers and only 5-15% of conifers in order to favour moose. the spatial organization of the landscape has been studied on some occasions. considering the area of track networks seen during aerial surveys, residual blocks < 20-50 ha should be unattractive to moose (courtois et al. 1998b). eason (1989) suggested 70 ha as the minimum size for an attractive block of forest, with observed moose densities increasing with the size of residual blocks. courtois et al. (1998b) and potvin et al. (1999) suggested alternating 50-100 ha cut and leave blocks. based on landscape pattern indices, this study shows that small cuts regularly interspersed with residual stands should be beneficial. acknowledgements we would like to thank michel crête from société de la faune et des parcs du québec (fapaq), jean-pierre ouellet and luc sirois from université du québec à rimouski (uqar), and two anonymous referees for providing helpful comments on earlier versions of this paper. alain caron of uqar assisted in the preparation of habitat maps. we are particularly indebted to aldée beaumont, laurier breton, and andré gaudreault from fapaq for their assistance in field work. references aebischer, n.j., p.a. robertson, and r.e kenward. 1993. compositional analysis of habitat use from animal radiotracking data. ecology 74:1313-1325. alldredge, j.r., and j.t. ratti. 1992. further comparison of some statistical techniques for analysis of resource selection. journal of wildlife management 56:1-9. ballard, w.b. 1992. bear predation on moose: a review of recent north american studies and their management implication. alces supplement 1:162-176. brassard, j.m., e. audy, m. crête, and p. grenier. 1974. distribution and winter habitat of moose in québec. naturaliste canadien 101:67-80. brusnyk, l.m., and f.f gilbert. 1983. use of shoreline timber reserves by moose. journal of wildlife management 47:673-685. cederlund, g., f. sandegren, and k. larsson. 1987. summer movements of female moose and dispersal of their offspring. journal of wildlife management 51:342-352. co u r t o i s , r., and m. cr ê t e. 1988. déplacements quotidiens et domaines vitaux des orignales du sud-ouest du québec. alces 24:78-89. , , and f. barnard. 1993. habitat selection in cut areas courtois et al. alces vol. 38, 2002 190 productivité de l’habitat et dynamique d’une population d’orignaux du sud de la taïga québécoise.publication 2144. ministère du loisir, de la chasse et de la pêche du québec, québec, canada. , j. labonté, and j.-p. ouellet. 1998a. déplacements et superficie du domaine vital de l’orignal, alces alces, dans l’est du québec. canadian fieldnaturalist 112:602-610. , j.-p. ouellet, and b. gagné. 1998b. characteristics of cutovers used by moose (alces alces) in early winter. alces 34:201-211. cowan, i. mct., w.s. hoar, and j. hatter. 1950. the effect of forest succession upon the quantity and nutritive values of woody plants used as food by moose. canadian journal of research 28:249-271. crête, m. 1977. importance de la coupe forestière sur l’habitat hivernal de l’orignal dans le sud-ouest du québec. canadian journal of forest research 7:241-257. . 1988. forestry practices in québec and ontario in relation to moose population dynamics. forestry chronicle 64:246-250. . 1989. approximation of k carrying capacity for moose in eastern québec. canadian journal of zoology 67:373-380. dalton, w.j. 1989. use by moose (alces alces) of clear-cut habitat where 100% or 50% of the production forest was logged. cofdra project 32001. ontario ministry of natural resources, toronto, canada. delvaux, h., r. courtois, l. breton, and r. patenaude. 1999. relative efficiency of succinylcholine, xylazine, and carfentanil/xylazine mixtures to immobilize free-ranging moose. journal of wildlife diseases 35:38-48. desmeules, p. 1964. the influence of snow on the behavior of moose. report 3. ministère du tourisme, de la chasse et de la pêche du québec, québec, canada. du s s a u l t , c. 2002. influence des contraintes environnementales sur la sélection de l’habitat de l’orignal (alces alces). ph.d. thesis. université laval, québec, canada. , r. courtois, j. huot, and j.-p. ouellet. 2001. the use of forest maps for the description of wildlife habitats: limits and recommendations. canadian journal of forest research 31:12271234. eason, g. 1989. moose response to hunting and 1-km2 block cutting. alces 25:6374. , e. thomas, r. jerrard, and k. oswald. 1981. moose hunting closure in a recently logged area. alces 17:111125. eddy, w.f. 1977. a new convex algorithm for planer sets from acm. transactions on mathematical software 3:398403. edenius, l. 1992. moose feeding in relation to position of food plants. alces supplement 1: 132-138. efron, b., and r.j. tibshirani. 1993. an introduction to the bootstrap. chapman and hall, new york, new york, usa. elkie, p.r., r.r. rempel, and a.p. carr. 1999. patch analyst user’s manual. publication tm-002. northwest science and technology, ontario ministry of natural resources, thunder bay, ontario, canada. (esri) environmental systems research institute inc. 1996. arcview gis. the geographic systems for everyone. environmental systems research institute inc., redlands, california, usa. forbes, g.j., and j.b. theberge. 1993. multiple landscape scales and distribution of moose, alces alces, in a forest alces vol. 38, 2002 courtois et al. habitat selection in cut areas 191 ecotone. canadian field-naturalist 107:201-207. franzmann, a.w., and c.c. schwartz. 1985. twinning rates: a possible population condition assessment. journal of wildlife management 49:394-396. girard, f., and r. joyal. 1984. l’impact des coupes à blanc mécanisées sur l’orignal dans le nord-ouest du québec. alces 20:40-53. hall, l.s., p.r. krausman, and m.l. morrison. 1997. the habitat concept and the plea for standard terminology. wildlife society bulletin 25:173-182. hooge, p.n., and b. eichenlaub. 1997. a n i m a l m o v e m e n t e x t e n s i o n t o arcview. ver. 1.1. alaska biological science center, u.s. geological survey. anchorage, alaska. hundertmark, k.j., w.l. eberhardt, and r.e. ball. 1990. winter habitat use by moose in southeastern alaska: implications for forest management. alces 26:108-114. johnson, d.h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61:65-71. joyal, r. 1987. moose habitat investigations in québec and management implications. swedish wildlife research supplement 1: 139-152. , and c. bourque. 1986. variations, selon la progression de l’hiver, dans le choix de l’habitat et du régime a l i m e n t a i r e c h e z t r o i s g r o u p e s d’orignaux (alces alces) en milieu agroforestier. canadian journal of zoology 64:1475-1481. krefting, l.w. 1974. moose distribution and habitat selection in north central north america. naturaliste canadien 101:81-100. labonté, j., j.-p. ouellet, r. courtois, and f. bélisle. 1998. moose dispersal and its influence in the maintenance of harvested populations. journal of wildlife management 62:225-235. lamontagne, g., h. jolicoeur, and r. lafond. 1999. plan de gestion de l’ours noir, 1998-2002 publication 3960. société de la faune et des parcs du québec, québec, canada. larivière, s., h. j olicoeur, and m. crête. 2000. status and conservation of the gray wolf (canis lupus) in wildlife reserves of québec. biological conservation 94:143-151. loranger, a.j., t.n. bailey, and w.w. larned. 1991. effects of forest succession after fire in moose wintering habitats on the kenai peninsula, alaska. alces 27:100-109. mcclean, s.a., m.a. rumble, r.m. king, and w.l. baker. 1998. evaluation of resource selection methods with different definitions of availability. journal of wildlife management 62:793-801. mcgarigal, k., and b.j. marks. 1995. fragstats. spatial pattern analysis program for quantifying landscape structure. oregon state university, corvalis, oregon, usa. (mer) ministère de l’énergie et des r e s s o u r c e s . 1 9 8 4 . n o r m e s d’inventaire forestier. ministère de l’énergie et des ressources du québec, québec, canada. paré, m., and r. courtois. 1990. inventaire aérien de l’orignal dans les zones de chasse 12 en janvier 1988 et dans la zone de chasse 13 en janvier 1989.publ i c a t i o n 1 7 6 4 . m i n i s t è r e d e l’environnement et de la faune du québec, québec, canada. peek, j.m., r.j. mackie, and g.l. dusek. 1992. over-winter survival strategies of north american cervidae. alces supplement 1:156-161. , d.j. pierce, d.c. graham, and d.l. davis. 1987. moose habitat use and implications for forest management habitat selection in cut areas courtois et al. alces vol. 38, 2002 192 in northcentral idaho. swedish wildlife research supplement 1:195-199. , d.l. urich, and r.j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. potvin, f. 1998. la martre d’amérique (martes americana) et la coupe à blanc e n f o r ê t b o r é a l e : u n e a p p r o c h e télémétrique et géomatique. ph.d. thesis. université laval, québec, canada. , r. courtois, and l. bélanger. 1999. short-term response of wildlife to clear-cutting in québec boreal forest: multiscale effects and management implications. canadian journal of forest research 29:1120-1127. , k. lowell, m.-j. fortin, and l. bélanger. 2001. how to test habitat selection at the home range scale: a resampling random windows technique. écoscience 8:399-406. rempel, r.s., p.c. elkie, a.r. rodgers, and m.j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. renecker, l.a., and r.j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64:322-327. rettie, w.j., and p.d. mcloughlin. 1999. overcoming radiotelemetry bias in habitat-selection studies. canadian journal of zoology 77:1175-1184. , and f. messier. 2000. hierarchical habitat selection by woodland caribou: its relationship to limiting factors. ecography 23:466-478. sas institute inc. 1989. sas/stat user’s guide. sas institute incorporated, cary, north carolina, usa. schwartz, c.c. 1992. physiological and nutritional adaptations of moose to northern environments. alces supplement 1:139-155. sinclair, a.r.e., and p. arcese. 1995. population consequence of predationsensitive foraging: the serengeti wildebeest. ecology 76:882-891. thompson, i.d., and d.l. euler. 1987. moose habitat in ontario: a decade of change in perception. swedish wildlife research supplement 1:181-193. , and r.w. stewart. 1998. management of moose habitat. pages 377402 in a. w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. timmermann, h.r., and j.g. mcnicol. 1988. moose habitat needs. forestry chronicle 64:238-245. van ballenberghe, v. 1989. twenty years of moose immobilization with succinylcholine chloride. alces 25:25-30. wiens, j.a. 1989. spatial scaling in ecology. functional ecology 3:385-397. wilson, s.f., d.m. shackleton, and k.l. campbell. 1998. making habitat-availability estimates spatially explicit. wildlife society bulletin 26:626-631. f:\alces\vol_39\p65\3914.pdf alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 177 moose-vegetation-soil interactions: a dynamic system john pastor1 and kjell danell2 1department of biology and natural resources research institute, university of minnesota, duluth, mn 55812, usa; 2department of animal ecology, swedish university of agricultural sciences, se-90183, umeå, sweden abstract: we review the processes by which moose (alces alces) interact with vegetation at the module (leaf and shoot) and genet (individual plant) levels of organization, and show the consequences of these interactions for plant population, community, ecosystem, and landscape dynamics. moose forage selectively on photosynthetic and meristematic tissues of a few preferred species. these species respond with compensatory growth and often tissues of higher forage quality, leading to a positive feedback at the module and genet level. however, height growth of browsed genets is often reduced or even curtailed by browsing, leading to higher levels of mortality and eventual replacement of browsed species by unbrowsed ones. these unbrowsed species (predominantly conifers) grow more slowly and have litter of low nutrient content that decomposes slowly. consequently, even though moose browsing stimulates growth and browse availability at module and genet levels, ecosystem productivity and nitrogen cycling decline. such feedbacks eventually lead to spatial patterns in the landscape. genotypic and phenotypic differences within forage species modify these responses somewhat, and plant responses to moose browsing all differ somewhat along productivity gradients. other herbivores, notably invertebrates, are also affected by these changes in vegetation. we conclude by suggesting some unanswered questions and new directions for future research. alces vol. 39: 177-192 (2003) key words: browsing, ecosystems, moose, soil, vegetation through their foraging behavior, moose (alces alces), the largest extant herbivore in boreal regions, exert many changes to plants, plant communities, and ecosystems. the purpose of this paper is to review the interactions between moose, vegetation, and soil at a number of hierarchical levels and to suggest gaps in our knowledge, and some experimental and modeling approaches to fill them. these complex interactions between moose and plants occur at several nested hierarchical levels (danell et al. in press). t h e l o w e s t i s t h e m o d u l e l e v e l o f meristematic tissues of leaves and shoots, the level of the primary foraging decisions of moose. the genet, or individual plant l e v e l , r e s p o n d s t o t h e r e m o v a l o f meristematic tissue through reallocations of carbon and nutrients to compensate for consumptive removals. in addition, the plant’s responses to browsing affect the intensity and probability of future browsing. the genet must contend not only with removal of portions of its tissues, but also competition with its neighbors that may or may not have been browsed by moose. these interactions between the browsed genet and its neighbors affect both population dynamics (through changes in distributions of genotypes, sex ratios, reproductive potential, mortality, and age structure) and community composition (through shifts in d o m i n a n c e b e t w e e n b r o w s e d a n d unbrowsed species). to the extent that browsed and unbrowsed species differ in moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 178 functional attributes important to ecosystem processes, such as litter quality or n fixation rates, the changes in community structure caused by differential browsing translate into changes in ecosystem productivity and nutrient cycling rates. these higher order changes, especially in the cycling of limiting nutrients such as nitrogen, in turn feed back on all individuals in the plant community through changes in soil fertility. this cascade of interactions has both beneficial and detrimental consequences to the energy and nutrient balances of moose. therefore, the effects of moose on vegetation are best viewed as a continuum of interactions (hjältén et al. 1993) that constitute a dynamic system of multiple feedbacks of different temporal and spatial scales. interactions between moose and plant meristems: the modular level moose behavioral decisions at the modular level browsing of leaves and shoots, both meristematic tissues, are the decisions made by moose that most immediately affect plants. whether to browse shoots or leaves or both depends on plant species and growth form (deciduous vs. coniferous), season, and whether the plant has been previously browsed. deciduous trees, especially salix, populus, betula, and sorbus spp., are generally the preferred food of moose, especially in summer. browsing of these species during summer consists almost entirely of leaf stripping. in winter, the bare current shoots of these species are also browsed, along with the current shoots (twigs and needles) of two conifers, pinus sylvestris in fennoscandia and russia and abies balsamea, particularly in the maritime provinces and new england, in north america. moose maximize energy intake per unit time at the modular level by browsing larger shoots (belovsky 1978, spalinger and hobbs 1992). as bite size increases, so does energy intake rate, but digestibility decreases because larger shoots contain higher proportions of lignified woody material (vivas and saether 1987). moose therefore optimize shoot diameter to maximize energy intake rate within the constraints of decreasing digestibility; the optimal shoot diameter selected across a wide range of species appears to be 3 – 5 mm, depending on plant species (shipley et al. 1999). pastor et al. (1999a) showed that moose maximize energy intake rates per unit time by taking single bites most often and browsing approximately 3.5 g per bite. this bite size is similar to measurements independently made by gross et al. (1993), risenhoover (1987), and others reviewed by renecker and schwartz (1998). average dry matter intake rates by moose range between 30 – 45 g per minute (belovsky and jordan 1978; belovsky 1981; renecker and hudson 1985, 1986; shipley and spalinger 1992; spalinger and hobbs 1992; gross et al. 1993; renecker and schwartz 1998; pastor et al. 1999a), or approximately 10 bites per minute of foraging. given that foraging bouts last approximately 30 – 60 minutes and moose have 4 – 5 foraging bouts to meet daily requirements of 5 – 10 kg per day (belovsky and jordan 1978, renecker and hudson 1986, renecker and schwartz 1998), moose take approximately 1,500 – 3,000 bites or more per day. such a large number of bites each day can have a substantial impact on plants, especially since the bites are concentrated on photosynthetic or apical meristems. however, moose distribute these bites over a large portion of the landscape and rarely consume all available bites within a patch, usually taking only approximately 20% of bites available or 1 – 2 bites per individual plant (shipley et al. 1998), except when a alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 179 food species is rare, whereupon it is heavily and repeatedly browsed (brandner et al. 1990). nonetheless, by making a large number of biting decisions per day and concentrating their impact on the browsing tips of a few plant species, moose have a great ability to affect growth of plants, competitive abilities between browsed and adjacent unbrowsed plants, plant succession, and ecosystem properties. therefore, decisions made by moose at the modular level are also reflected at higher levels of vegetation organization. responses of modules to browsing deciduous trees whose leaves have been stripped have some capacity to regrow leaves the same season, but the ability to regrow leaves depends on shoot morphology, growth strategy, intensity and frequency of stripping, and site fertility. experimental leaf stripping from betula pendula resulted in lower standing leaf biomass by the end of the growing season, but refoliation produced a second crop of leaves such that total leaf production during the growing season (removed leaves plus regrown leaves) did not differ between stripped and control trees (bergström and danell 1995). however, 1 year after defoliation, total leaf biomass was lower than controls because defoliation decreased the number of shoots. responses of birches to current shoot removal during the non-growing season differ substantially from the above responses to defoliation and also depend on the type of shoot browsed. birches and many other deciduous species (especially in the betulaceae, salicaceae, and rosaceae) produce two types of shots, termed “long” and “short” for obvious reasons. long shoots are exclusively vegetative growth and are characterized by extreme apical dominance (hormonal suppression of bud emergence lower on the stem by the apical bud). short shoots are smaller side shoots with reproductive structures such as catkins, fruits, etc., that often emerge at their base. long shoot dry mass, leaf dry mass, leaf number, leaf area, and chlorophyll and n content of leaves on long shoots are higher on moderately browsed b. pendula and b. pubescens than on slightly browsed trees (danell et al. 1985, bergström and danell 1987). some short shoots on moderately browsed birches can sometimes develop into long shoots, thus compensating for leaf biomass declines when short shoots are browsed (danell et al. 1985). although the number of long shoots sometimes declines following moderate browsing, the frequency of branched shoots increases (bergström and danell 1987). senn and haukioja (1994) show that these responses are primarily the result of removal of the apical buds and reduction in hormonal suppression of buds lower on the stem. shoot browsing also increases leaf nitrogen and chlorophyll contents in birches (bergström 1984, danell and huss-danell 1985, danell et al. 1985, danell et al. in press) and feltleaf willow (salix alaxensis; kielland et al. 1997). in contrast, defoliation decreases food quality of b. pubescens ssp. tortuosa leaves because of induced defenses (haukioja et al. 1985, hanhimäki 1989, ruohomäki et al. 1992). therefore, shoot browsing results in higher quality foliage the following year in birch and willow but defoliation results in lower quality leaves in subsequent crops during the same year. because shoot and leaf browsing occur at different times of the year, the effect of removal of either tissue on subsequent leaf quality may also be a seasonal effect. therefore, new experiments are needed in which shoots are removed in summer as well as winter to determine the interactive effects of shoot removal per se and season of removal on leaf chemistry. like deciduous species, pinus sylvestris moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 180 also compensates for winter (shoot plus needles) browsing, but the compensation can be delayed for one or more years (edenius et al. 1995) because growth in these pines is determinate rather than indeterminant as in the birches (millard et al. 2001). edenius et al. (1995) also found that densely grown pines compensate more for lost biomass than open grown pines. balsam fir (abies balsamea) generally shows very little compensatory growth following browsing, and browsed stems eventually become progressively weaker and then die (brandner et al. 1990, thompson and curran 1993). feedbacks between modular responses and moose foraging the different responses of modules of deciduous and coniferous species to moose browsing affect the frequency and intensity of subsequent browsing on individual genets. birches whose shoots were previously browsed by moose have a higher probability of being browsed again than unbrowsed or slightly browsed birches because of the higher leaf and stem chemical quality, large long shoots, and greater proportion of shoots within reach of moose (bergström 1984, danell and huss-danell 1985, danell et al. 1985, danell et al. in press). thus, the compensatory growth response of birches to browsing at the modular level establishes a positive feedback loop at the genet level, resulting in even greater consumption in subsequent years. these positive feedbacks between shoot browsing during winter, the higher quality of regrown leaves, and subsequent browsing are the opposite of that noted for leaves that had been stripped by moose during summer (miquelle 1983) or wounded by insects during summer (haukioja et al. 1985, hanhimäki 1989, ruohomäki et al. 1992). in summer, leaf stripping induces defenses that reduce probabilities of being consumed again. thus, leaf stripping in summer results in a negative feedback between moose and the browsed plant while winter browsing on shoots gives positive feedbacks. as noted above, new experiments are needed to sort out the direct effects of browsing vs. stripping per se from the indirect effects of season of browsing on changes in food quality. in pines, the delay in compensatory response may temporarily release browsed pines from subsequent browsing, allowing some recovery (edenius et al. 1995). as we shall see below, these different compensatory responses between deciduous species and conifers greatly affect competition with neighboring plants. balsam fir is also repeatedly browsed by moose, leading to a highly pruned growth form (brandner et al. 1990). however, because compensatory regrowth is very small, many balsam fir escape from repeat browsing, especially on sites of high fir and low moose density (brandner et al. 1990). interactions between moose and whole plants: the genet level here, we consider how the decisions made by moose at the individual plant level depend in part on responses of the plant modules to browsing. we also consider the effect of genotype and phenotype on browsing intensity and recovery of plants, and the effect of browsing on whole plant growth and height growth. moose foraging decisions at the genet l e v e l besides decisions made at the modular level reviewed above, decisions made at the genet (individual tree) level are also important. danell et al. (1991a) showed that moose foraging decisions are made at this level (consumption and preferences of individuals of a given species within stands are alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 181 the same regardless of stand composition) more than at the stand level (consumption and preference of individuals did not depend on stand composition, and so stand composition was not the primary decision on where to forage). because of the compensatory and generally higher quality regrowth reviewed above, browsed plants, especially deciduous plants, have a higher probability of being rebrowsed, resulting in even greater consumption in subsequent years (bergström 1984, danell and hussdanell 1985, danell et al. 1985, danell et al. in press). genotypic and phenotypic differences in susceptibility of genets to browsing in a provenance study, danell et al. (1991b) found that, when transplanting pines from different sites and exposing them in cafeteria tests to free-ranging moose, p. sylvestris individuals from fertile habitats were browsed more intensively than pines from infertile habitats because of their larger shoot size and higher quality food. jia et al. (1995) showed that moose adjust bite diameters among phenotypes of p. sylvestris shoots according to shoot characteristics such as growth rate and nutrient contents: in general bite diameters are larger from phenotypes of higher productivity than lower productivity. such differences in browsing preference and bite size may be at least partly genetically based because slowly growing northern genotypes of p. sylvestris are less preferred by moose than more rapidly growing southerly genotypes, even when grown in a common garden (niemelä et al. 1989). balsam fir from the open habitats of thinned stands are more intensively browsed than those from unthinned stands because of higher crude fat, protein, and nutrient concentrations (thompson et al. 1989). the fact that these chemical differences were responses to thinning rather than phenotypic sorting along a fertility gradient shows that some of the phenotypic susceptibility of genets to browsing is environmentally rather than genetically based. on the other hand, balsam fir from the maritime provinces are much less defended than balsam fir from continental north america, and these broad geographic differences are believed to be genetically based (hunt 1993). this may account for the greater proportion of balsam fir in the diet of moose in the maritime provinces (bergerud and manuel 1968) than in continental north america (krefting 1974, risenhoover and maass, 1987, mcinnes et al. 1992). within several birch species, genets also differ in susceptibility to browsing depending on biogeographic origin. in a common garden experiment, bryant et al. (1989) found that icelandic b. pubescens was more preferred by snowshoe hares (lepus americanus) and finnish mountain hares (l. timidus) than finnish b. pubescens or b. pendula, which in turn were more preferred than siberian b. middendorffi. this rank order is inversely related to the concentrations of resins and triterpene acids in internodes of these provenances. moreover, iceland did not have a resident mammalian population until settlement by vikings in the 9th century, and finnish hare populations are less dense than siberian. this relationship between susceptibility to browsing and long-term coexistence with hare populations strongly suggests that biogeographic differences between genets in browse preference have resulted from co-evolution between the plants and browsing animals. although no studies have demonstrated this specifically, the discrimination of moose between genotypes within a species may also lead to changes in the distribution of genotypes in the population, thus making moose a selection pressure on its forage species (danell et al. in press). moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 182 growth responses of genets to browsing although compensatory growth of new shoots and side shoots often results in little or no decrease in total dry matter production in birches (danell et al. 1985, bergström and danell 1987, hjältén et al. 1993), height growth of birch can be greatly affected by browsing, depending on the type of tissue consumed. defoliation of b. pubescens decreases height growth by almost 50% (hjältén et al. 1993, bergström and danell 1995), presumably because of decreased s h o o t g r o w t h d u e t o r e d u c t i o n s i n photosynthate. for established trees, shoot browsing decreases height growth of b. pubescens and b. pendula when large proportions of current shoots are removed (bergström and danell 1987, hjältén et al. 1993). height growth of p. sylvestris was also similarly decreased when 100% of current shoots were clipped (edenius et al. 1995). the height growth of seedlings and small suckers and ramets of deciduous species such as betula, populus, and salix is almost always curtailed by browsing because such individuals consist almost entirely of a single shoot with one apical meristem. browsing of this apical meristem releases the shoot from apical dominance (senn and haukioja 1994) and often kills the main stem (krefting 1974, heinen and sharik 1990, mcinnes et al. 1992). although the total dry matter production of an individual genet is usually not curtailed outright by browsing, height growth of browsed seedlings and suckers is then curtailed in favor of increased production of new side shoots or new ramets. effect of moose browsing on plant population dynamics the effects of moose at the module and genet levels are translated to the population level by their consequences for reproduction and establishment, mortality, and seed dispersal. reproductive potential of genets in response to browsing browsing generally decreases reproduction. the number of female catkins (and hence seed production) is reduced in b. pendula and b. pubescens with higher browsing intensity, although mean viable seed mass increased slightly, suggesting partial compensation in potential seed germination success to the reduction of seed number (bergström and danell 1987). in part, this may be due to reduced short shoot production of browsed plants, as noted above. beaked hazelnut (corylus cornuta), an important winter forage species for moose in north america, also usually does not set seed when browsed (trottier 1981). browsing also decreases cone production in p. sylvestris, especially with severe browsing on productive sites (edenius et al. 1995). moose damage seedlings by trampling and uprooting, resulting in severe repression of establishment in some areas (bergerud and manuel 1968). mortality of genets in response to browsing in many studies, mortality of individuals was directly proportional to browsing intensity and the more preferred deciduous plant species had higher rates of mortality (krefting 1974, heinen and sharik 1990, mcinnes et al. 1992, edenius et al. 1995, danell et al. in press). mortality rates are particularly high in the smaller size classes. mortality in response to browsing is higher on infertile soils, presumably because limiting nutrients and water are scarce enough to severely inhibit compensatory growth after browsing (edenius et al. 1995). severe bark-stripping by moose almost always causes the stem to die because the alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 183 cambial tissue is removed when the stem is girdled (miquelle and van ballenberghe 1989). in contrast, shoot or leaf browsing rarely appears to directly cause mortality, at least in deciduous species. instead, mortalities of leaf stripped or shoot-browsed plants are directly proportional to the suppression of their height growth (risenhoover and maass 1987, danell et al. in press). this suggests that browsed individuals die because of increased light limitations as they become overtopped by adjacent unbrowsed individuals. the reduced height growth of seedlings of shade intolerant species puts them at a disadvantage when they are overtopped by unbrowsed neighbors, especially of less preferred conifer species. the recruitment of browsed deciduous seedlings and suckers into larger tree size classes is greatly reduced because of suppression of their growth by light limitations imposed by unbrowsed neighbors. support for this comes from experiments performed by hjältjén et al. (1993), who found that compensatory height growth of b. pubescens decreases with increasing stand density, presumably because of increased severity of light competition to regrowing shoots from neighbors. mclaren (1996) also found that shoot-browsed balsam fir is more likely to die when canopy cover was greater than 60% compared to browsed individuals grown under a more open canopy. changes in sex ratios in response to browsing some of the most preferred species (populus and salix) are dioecious, that is, the male and female reproductive organs occur on different individuals. there is some limited evidence that some herbivores appear to prefer male individuals (ågren et al. 1999), although whether this is enough to alter sex ratios and if so, if the alteration is large enough to be important evolutionarily, is not known. this is an area deserving further investigation. changes in forest age structure because moose populations cycle, they can change age structure of plant species as their foraging pressure waxes and wanes. for example, moose populations on isle royale cycle with a period of 38 ± 13 years (peterson et al. 1984). age structure of balsam fir also cycles coincident with this period (mclaren and peterson 1994). mclaren and peterson (1994) suggest that the moose-wolf predator-prey cycle results in periodic suppression (when moose populations are high) and release (when moose populations are low) of balsam fir, resulting in age gaps in the balsam fir population. snyder and janke (1976) also show dependence of age class distributions on moose population densities, not only of forage species, but of non-preferred species such as picea, indicating that cycles in populations of preferred species affect competitors through imposition and release of competition for light and perhaps soil resources. effects of moose browsing on the plant community and ecosystem processes early successional boreal forests are often dominated by deciduous species such as populus, betula, and salix, that not only are the preferred species of moose but also respond positively to moose browsing with regrowth of leaves that are higher in nitrogen, protein, and other nutrients. these leaves decompose more quickly than leaves from unbrowsed individuals of the same species (irons et al. 1991, kielland et al. 1997), leading to a temporary increase in ecosystem carbon turnover and nutrient cycling rates (molvar et al. 1993, kielland et al. 1997). in the long run, however, numerous moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 184 studies comparing plant community composition inside and outside exclosures have shown that the abundances of preferred species, especially the deciduous species, decline as they are replaced by non-preferred species, especially picea (krefting 1974, risenhoover and maass 1987, mcinnes et al. 1992, thompson et al. 1992). in north america, there is a consistent shift from an aspen-birch-spruce-fir community to a more open and unbrowsed spruce (picea glauca, p. mariana) community with an understory of heavily browsed preferred species with increased moose population density. in many areas, unbrowsed spruce is often the only species able to grow above browse height (janke et al. 1978, bryant and chapin 1986, mcinnes et al. 1992, thompson and curran 1993). the primary reason for this species replacement is the suppression of height growth of plants that are repeatedly browsed by moose, leading to their shade-induced mortality as noted above. these long-term successional shifts in plant community composition towards nonpreferred species greatly depress rates of ecosystem properties such as net primary productivity (mcinnes et al. 1992) and nitrogen cycling (pastor et al. 1993). the decline in productivity occurs for two reasons. first, non-preferred species grow more slowly than preferred species (danell et al. 1985, bryant and chapin 1986, mcinnes et al. 1992, pastor and naiman 1992). secondly, non-preferred species have litter that is difficult to decompose because of low nitrogen and high lignin contents, the same reasons why moose, with microbially mediated ruminant digestion avoid them (bryant and chapin 1986, pastor and naiman 1992). the same chemical properties of tissues that cause moose to forage selectively also result in a depression of soil nitrogen availability of up to half that for forests without moose (pastor et al. 1988, 1993; pastor and cohen 1997). this depression of soil nitrogen availability because of vegetation changes is not offset by local increases in nitrogen from fecal and urine deposition (pastor et al. 1993, 1996; pastor and cohen 1997). thus, decisions made by moose at the modular and genet level are reflected at the ecosystem level because the same plant chemical properties affect both digestive rate and nutrient cycling. such effects on ecosystem properties are widely distributed across the landscape in characteristic patterns: patches of high density browse are heavily browsed, allowing unbrowsed conifers to invade and create coincident patches with lower available nitrogen (pastor et al. 1998). moose must therefore contend not only with changes at modular and individual plant levels, but also with the effect of these plant responses on the distribution of food across the landscape and the cycling of limiting nutrients to support that food. thus, at the module and genet levels, shoot browsing by moose, especially on deciduous species, has a positive effect on growth and tissue quality, which leads to repeated browsing. the repeated browsing keeps these shoots and individuals within browse height, thus temporarily increasing browse supply for moose. but in the long run, reproductive potential is decreased, height growth is suppressed, and these species yield to unbrowsed species that depress availability of soil nitrogen, leading to a decrease in the food supply, nitrogen cycling, and net primary production of the entire ecosystem. therefore, the effects of moose on vegetation appear to be a continuum of responses at several levels, sometimes positive (especially at the finest levels of vegetation organization) and other times negative (especially at plant community and ecosystem levels over the long run). alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 185 secondary effects of moose browsing on other animal species moose browsing on vegetation has secondary effects on other animals, especially insects, at all levels of plant organization. danell and huss-danell (1985) found greater abundance of a wide variety of insect types on shoots of moderately browsed b. pendula and b. pubescens than on shoots of unbrowsed individuals because of the higher nitrogen and chlorophyll contents of regrown leaves on browsed shoots. surprisingly, mountain hare (lepus timidus) did not appear to discriminate between browsed and unbrowsed birch shoots (danell and huss-danell 1985). recently, suominen et al. (1999a,b) showed that moose browsing increases light penetration to the forest floor, that in turn increases temperature and decreases soil moisture, thus favoring some invertebrate species, such as carabid beetles and disfavoring others, such as gastropods. increased leaf litter nitrogen and changes in tannin chemistry of browsed alaskan paper birch (betula papyrifera var. humulis, formerly b. resinifera) result in faster decay when these leaves are deposited in streams by trees growing in riparian zones (irons et al. 1991). the effects of moose on litter chemistry of terrestrial species are thus translated into adjacent aquatic ecosystems as litter is transported across ecological boundaries. such secondary effects of moose on other trophic levels are only beginning to be recognized and deserve much more attention. for example, it would be particularly interesting to determine if bird communities differ between areas heavily impacted by moose compared with areas of low moose population. some unanswered questions and suggestions for future research t h e g e n e t i c b a s i s o f f o r a g i n g d e c i s i o n s b y m o o s e a n d p l a n t responses: co-evolution of moose and plant species? a number of the studies reviewed above (e.g., danell et al. 1985, 1991b; bryant et al. 1989; niemelä et al. 1989) have suggested that there are genotypic differences between individuals within the same species with respect to moose browsing preferences and plant responses. in general, faster growing genotypes are more preferred, have higher quality plant tissues, and show more compensatory growth than slower growing phenotypes. to determine the relative contributions of genetic and environmental controls on these plant properties, common garden experiments are needed across a wide range of site fertilities. it is entirely possible that some plant responses are controlled more by genotypic differences and others are controlled more by environmental differences. the relative expression of genetic and environmental factors may also differ along fertility gradients. the same genetic mechanisms that may determine quality of forage for moose also determine quality and decomposability of litter. we therefore have the fascinating possibility that moose selection among genotypes has consequences at the ecosystem level (cohen et al. 2000). plant community dynamics may also exert a selection pressure on moose populations. geist (1974) proposed that there are two phenotypes of moose: large bodied individuals with a high consumption rate that are at a selective advantage after large disturbances such as fire that generate much high quality food, such as regenerating populus, and small-bodied individuals that persist when disturbances are small. geist suggested that these phenotypes have moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 186 a genetic basis, but thus far this has not been demonstrated. it is also not clear whether these phenotypic differences in moose are a response to differences in quantity of food vs. quality of food in these two disturbance regimes. nevertheless, there are some differences between resident and migratory moose populations in foraging behavior that suggest a genetic basis: histøl and hjeljord (1993) showed that migratory moose in norway used habitats of lower quality, had a higher proportion of poorer quality p. sylvestris and b. pubescens in their diet, and rebrowsed previously browsed individuals more so than did resident moose. histøl and hjeljord (1993) suggest that this implies some genetic differences in foraging behavior between partly sympatric subpopulations with overlapping ranges. although it is possible that genetic differences underlie these different behaviors in sympatric populations, other mechanisms are also possible (and not mutually exclusive), including learned behavior, stochastic differences in previous experience between populations, exclusion of migratory individuals from better habitats by resident individuals, etc. nonetheless, the genetic basis of moose foraging clearly deserves more experimental attention. there are also intriguing patterns in the fossil record that suggest co-evolution between moose and the plant species that constitute the current boreal forest biome. moose arose fairly recently in the fossil record during the northern glacial advances of the late pleiocene and early pleistocene (bubenik 1998), represented by the genera cervalces and libralces. the genus alces arose from libralces represented by at least 4 species, of which only a. alces survives. a. alces is believed to have originated in far eastern siberia from ancestral stocks of alces gallicus that previously had migrated east from europe (bubenik 1998, hundertmark et al. 2002). a. alces migrated east into the rest of north america and west into scandinavia and the rest of europe shortly after the last advance of the wisconsin/würm ice sheet, although the exact time is not certain (bubenik 1998, hundertmark et al. 2002). during these migrations, a. alces radiated into 8 subspecies, of which 7 survive (groves and grubb 1987). at the same time that a. alces was undergoing these migrations and adaptive radiations, boreal species of the genera populus, betula, picea, and abies that constitute its present habitat also migrated northward following the retreat of the ice sheet, at least in north america (larsen 1980). hundertmark et al. (2002) propose that a cool climate during the last glacial period contracted and fragmented moose habitat, which led to sparse and isolated populations from which the current subspecies evolved during the expansion and reorganization of habitat in the subsequent current interglacial period. therefore, it is interesting to speculate that the current complex relationships between moose and vegetation reviewed above arose when the circumpolar boreal forest became assembled, suggesting possible co-evolution between moose and the plant species they require for food and shelter (bryant et al. 1989). moose-vegetation interactions along productivity gradients studies reviewed above (e.g., danell et al. 1991a,b; mcinnes et al. 1992; edenius et al. 1993) show that moose browsing affects plant performance and responses differently along site productivity gradients. however, moose population densities also vary along productivity gradients (e.g., peterson et al. 1984, mclaren and peterson 1994) and so the performance and responses of plants along productivity gradients is confounded with moose population density alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 187 and hence foraging pressure. to sort out these confounding factors, controlled experiments comparing simulated browsing at different intensities to natural browsing are required along productivity gradients. we and our colleagues and students have been engaged in one such experiment for the past 4 years. exclosures, 70 x 70 m in size, have been established on 8 sites representing a productivity gradient in northern sweden. the sites were all clearcut 810 years ago and planted to p. sylvestris. on the least productive sites, the pines coexist mainly with dwarf shrubs and lichens while on the most productive sites the pines coexist with deciduous species such a s b . p e n d u l a , p . t r e m u l a , r u b u s chamaemorus, grasses, and herbs. we have subdivided each exclosure into 4 subplots. within each subplot (25 x 25 m) we are simulating moose browsing by clipping at an intensity corresponding to 0, 1, 3, and 5 moose per 100 ha, distributed amongst species according to dietary preferences of moose on these sites. we also add urine and fecal material at rates corresponding to these population densities. our colleagues and we are measuring changes in plant growth, litter fall, soil temperature and moisture, soil nitrogen availability, and soil invertebrate diversity and population density. preliminary results after 3 years of the experiment indicate reductions in litter fall with increasing browsing pressure and proportionally greater reductions of litter fall on productive than unproductive sites (ingalill persson et al., unpublished data). moose-landscape interactions: how d o m o o s e f o r a g e i n l a n d s c a p e s previously affected by other moose? older studies of moose-habitat relations (phillips et al. 1973, peek et al. 1976) have assumed that habitats change independently of moose and moose adjust their movements accordingly. the research reviewed here suggests otherwise, namely that through their foraging decisions moose greatly affect the abundance and distribution of both browsed and unbrowsed plant species and ecosystem properties, and that such interactions produce patterns in the landscape (e.g., pastor et al. 1998). this raises the interesting question of how moose and their progeny deal with landscape patterns that previous generations of moose have made. to answer this question, moen et al. (1997a, 1998) developed a simulation model of the energetics of a foraging moose in a spatially explicit landscape, parameterized by many of the empirical relations and data in the above papers and in the moose physiology literature (hudson and white 1985). foraging decisions are made on the basis of several rules: how much to eat at a given spot; when to stop eating; and where to move to next. the model has been well validated against independent data (moen et al. 1997a, 1998). moen et al. (1997a, 1998) show that the foraging strategy of moose creates different patterns in the landscape, which in turn affect the energy balance of a moose. in particular, foraging according to the marginal value theorem generates a landscape that cannot support moose populations in the long run: areas of high browse density are browsed heavily to reduce them to the average browse density of a landscape and areas of browse density lower than the marginal value criteria are bypassed and then grow out of reach. the net result is a landscape of low forage availability because of suppression of growth in areas of heavy browsing coupled with areas where forage has grown out of reach when not browsed. a strategy which seems to sustain moose populations at positive energy balances appears to be to browse 20-30% of that available at a feeding station then move to the best nearest patch until the rumen is full, then begin randomly at some moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 188 other point in the landscape for the next foraging bout (moen et al. 1998), similar to that found for free-ranging moose by shipley et al. (1998). pastor et al. (1999b) showed that this strategy produces a simulated landscape whose pattern matches that of real landscapes on isle royale (pastor et al. 1998). therefore, there may be a strong feedback between the generation of landscape pattern and the energetics of foraging moose and how they expend energy to search for food. if so, the effect of moose on landscape patterns could be a strong selection pressure for genotypes of certain foraging behaviors. testing of the predictions of moose movement patterns requires the sampling intensity, continuity, and accuracy of gps collars (moen et al. 1996, 1997b). our colleagues and we are now engaged in analyzing such data gathered over the past 7 years or so, and data are also being gathered by others. further advances in gps collar technology and model development should help clarify the answers to this question, but may also raise others. for example, do some moose “cheat” on other moose – that is, can some moose employ the marginal value strategy if other moose employ other strategies that maintain a landscape that enables a positive energy balance? how do moose simultaneously adjust their movement patterns in response to both their effects on the landscape and also to avoid predators? in conclusion, the interactions of moose with vegetation are a continuum of nested responses and feedbacks, some positive and some negative, and are more extensive than previously thought. studies of these interactions have caused us to revise older, more static ideas of moose-habitat relationships. instead the current picture is a more dynamic one of nested responses at different levels of ecological organization. acknowledgements we thank an anonymous reviewer and dr. ed addison for helpful comments on this manuscript. this paper was delivered at the international moose conference held in hafjell, norway in august 2002. we thank the organizers of this conference for an enjoyable and stimulating meeting in a beautiful country. references ågren j., k. danell, t. elmquist, l. ericson, and j. hjältén. 1999. sexual dimorphism and biotic interactions. pages 217-246 in m. a. gerber, t. e. dawson, and l. f. delph, editors. gender and sexual dimorphism in flowering plants. springer-verlag, berlin, germany. belovsky, g. e. 1978. diet optimization in a generalist herbivore: the moose. theoretical population biology 14:105-134. . 1981. food plant selection by a generalist herbivore: the moose. ecology 62:1020-1030. , and p. a. jordan. 1978. the timeenergy budget of a moose. theoretical population biology 14:76-104. bergerud, a. t., and f. manuel. 1968. moose damage to balsam fir –white birch forests in central newfoundland. journal of wildlife management 32:729746. bergström, r. 1984. rebrowsing on birch betula pendula and b. pubescens stems by moose. alces 12:870-896. , and k. danell. 1987. effects of simulated winter browsing by moose on morphology and biomass of two birch species. journal of ecology 75:533544. , and . 1995. effects of simulated summer browsing by moose on leaf and shoot biomass of birch, betula pendula. oikos 72:132-138. brandner, t. a., r. o. peterson, and k. l. alces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 189 risenhoover. 1990. balsam fir on isle royale: effects of moose herbivory and population density. ecology 71:155164. bryant, j. p., and f. s. chapin iii. 1986. browsing – woody plant interactions during boreal forest plant succession. pages 213-225 in k. van cleve, f. s. chapin iii, p. w. flanagan, l. a. viereck, and c. t. dyrness, editors. forest ecosystems in the alaskan taiga. springer-verlag, new york, new york, usa. , j. tahvanainen, m. sulkinoja, r. julkunen-tiitto, p. reichardt, and t. green. 1989. biogeographic evidence for the evolution of chemical defense by boreal birch and willow against mammalian browsing. american naturalist 134:20-34. bubenik, a. r. 1998. evolution, taxonomy and morphophysiology. pages 77-124 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. cohen, y., j. pastor, and t. vincent. 2000. nutrient cycling in evolutionary stable ecosystems. evolutionary and ecological research 6: 719-743. danell, k., r. bergström, l. edenius, and g. ericsson. in press. ungulates as drivers of tree population dynamics at module and genet levels. forest ecology and management. , l. edenius, and p. lundberg. 1991b. herbivory and tree stand composition: moose patch use in winter. ecology 72:1350-1357. , and k. huss-danell. 1985. feeding by insects and hares on birches earlier affected by moose browsing. oikos 44:75-81. , , and r. bergström. 1985. interactions between browsing moose and two species of birch in sweden. ecology 66:1867-1878. , p. niemelä, t. varvikko, and t. vuorisalo. 1991a. moose browsing on scots pine along a gradient of plant productivity. ecology 72:1624-1633. edenius, l., k. danell, and r. bergstöm. 1993. impact of herbivory and competition on compensatory growth in woody plants: winter browsing by moose on scots pine. oikos 66:286-292. , , and h. nyquist. 1995. effects of simulated moose browsing on growth, mortality, and fecundity in scots pine: relations to plant productivity. canadian journal of forest research 25:529-535. geist, v. 1974. on the evolution of reproductive potential in moose. naturaliste canadien 101:527-537. gross, j. e., l. a. shipley, n. t. hobbs, d. e. spalinger, and b. a. wunder. 1993. functional response of herbivores in food-concentrated patches: tests of a mechanistic model. ecology 74:778791. groves, c. p., and p. grubb. 1987. relationships of living deer. pages 21-59 in c. m. wemmer, editor. biology and m a n a g e m e n t o f t h e c e r v i d a e . smithsonian institution press, washington, d.c., usa. hanhimäki, s. 1989. induced resistance in mountain birch: defense against leafchewing insect guild and herbivore competition. oecologia 81:242-248. haukioja, e., p. niemelä, and s. siré. 1985. foliage phenols and nitrogen in relation to growth, insect damage, and ability to recover after defoliation in the mountain birch, betula pubescens ssp. tortuosa. oecologia 65:214-222. heinen, j. t., and t. l. sharik. 1990. the influence of mammalian browsing on tree growth and mortality in the pigeon river state forest, michigan. amerimoose, vegetation, and soil – pastor and danell alces vol. 39, 2003 190 can midland naturalist 123:202-206. histøl, t., and o. hjeljord. 1993. winter feeding strategies for migrating and nonmigrating moose. canadian journal of zoology 71:1421-1428. hjältén, j., k. danell, and l. ericsson. 1993. effects of simulated herbivory and intraspecific competition on the compensatory ability of birches. ecology 74:1136-1142. hudson, r. j., and r. g. white. 1985. bioenergetics of wild herbivores. crc press, incorporated, boca raton, florida, usa. hundertmark, k. j., g. f. smith, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. schwartz. 2002. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. molecular phylogenetics and evolution 22: 375387. hunt, r. s. 1993. abies. pages 354-362 in flora of north america editorial committee, editors. flora of north america. vol. 2. pteridophytes and gymnosperms. oxford university press, oxford, u.k. irons iii, j. g., j. p. bryant, and m. w. oswood. 1991. effects of moose browsing on decomposition rates of birch leaf litter in a subarctic stream. canadian journal of fisheries and aquatic science 48:442-444. janke, r. a., d. mckaig, and r. raymond. 1978. comparison of presettlement and modern upland boreal forests on isle royale national park. forest science 24:115-121. jia, j., p. niemelä, and k. danell. 1995. moose alces alces bite diameter selection in relation to twig quality on four p h e n o t y p e s o f s c o t s p i n e p i n u s sylvestris. wildlife biology 1:47-55. kielland, k., j. p. bryant, and r. w. ruess. 1997. moose herbivory and carbon turnover of early successional stands in interior alaska. oikos 80:2530. krefting, l. w. 1974. the ecology of the isle royale moose with special reference to the habitat. technical bulletin 297-1974, forestry series 15. agricultural experiment station, university of minnesota, minnesota, usa. larsen, j. a. 1980. the boreal ecosystem. academic press, new york, new york, usa. mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litterfall of the boreal forest, isle royale, michigan, usa. ecology 73:2059-2075. mclaren, b. e. 1996. plant-specific response to herbivory: simulated browsing of suppressed balsam fir on isle royale. ecology 77:228-235. , and r. o. pe t e r s o n. 1994. wolves, moose, and tree rings on isle royale. science 266:1555-1558. millard, p., a. hester, r. wendler, and g. baillie. 2001. interspecific defoliation responses on sites of winter nitrogen storage. functional ecology 15:535543. miquelle, d. g. 1983. browse regrowth and consumption following summer defoliation by moose. journal of wildlife management 47:17-24. , and v. van ballenberghe. 1989. impact of bark stripping by moose on aspen-spruce communities. journal of wildlife management 53:577-586. moen, r., y. cohen, and j. pastor. 1998. evaluating foraging strategies with a moose energetics model. ecosystems 1:52-63. , j. pastor, and y. cohen. 1997a. a spatially-explicit model of moose foraging and energetics. ecology 78:505521. , , and . 1997b. inalces vol. 39, 2003 pastor and danell – moose, vegetation, and soil 191 terpreting behavior from activity counters in gps collars on moose. alces 32:101-108. , , , a n d c . c . schwartz. 1996. effect of moose movement and habitat use on gps collar performance. journal of wildlife management 60:659-668. molvar, e. m., r. t. bowyer, and v. van ballenberghe. 1993. moose herbivory, browse quality, and nutrient cycling in a n a l a s k a n t r e e l i n e c o m m u n i t y . oecologia 94:472-479. niemelä, p., m. hagman, and k. lehtilä. 1989. relationship between pinus sylvestris l. origin and browsing prefe r e n c e b y m o o s e i n f i n l a n d . scandinavian journal of forest research 4:239-246. pastor, j., and y. cohen. 1997. herbivores, the functional diversity of plant species, and the cycling of nutrients in ecosystems. theoretical population biology 51:165 -179. , , and r. moen. 1999b. the generation of spatial patterns in boreal landscapes. ecosystems 2:439450. , b. dewey, and d. christian. 1996. carbon and nutrient mineralization and fungal spore composition of vole fecal pellets in minnesota. ecography 19:5261. , , r. moen, m. white, d. mladenoff, and y. cohen. 1998. spatial patterns in the moose-forest-soil ecosystem on isle royale, michigan, usa. ecological applications 8:411424. , , r. j. naiman, b. dewey, and p. mcinnes. 1988. moose, mic r o b e s , a n d t h e b o r e a l f o r e s t . bioscience 38:770-777. , , , p. f. mcinnes, and y. cohen. 1993. moose browsing and soil fertility in the boreal forests of isle royale national park. ecology 74:467-480. , and r. j. naiman. 1992. selective foraging and ecosystem processes in boreal forests. american naturalist 139:690-705. , k. standke, k. farnsworth, r. moen, and y. cohen. 1999a. further development of the spalinger-hobbs mechanistic foraging model for freeranging moose. canadian journal of zoology 77:1505-1512. peek, j. m., d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. o., r. e. page, and k. m. dodge. 1984. wolves, moose, and the allometry of population cycles. science 244:1350-1352. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37:266-278. renecker, l. a., and r. j. hudson. 1985. estimation of dry matter intake of freeranging moose. journal of wildlife management 49:785-792. , and . 1986. seasonal foraging rates of free-ranging moose. journal of wildlife management 50:143147. , and c. c. schwartz. 1998. food habits and feeding behavior. pages 403-440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. risenhoover, k. l. 1987. winter foraging strategies of moose in subarctic and boreal forest habitats. ph.d. thesis, michigan technological university, houghton, michigan, usa. moose, vegetation, and soil – pastor and danell alces vol. 39, 2003 192 , and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17:357-364. ruohomäki, k., s. hanhimäki, e. haukioja, l. iso-iivari, s. neuvonen, p. niemelä, and j. suomela. 1992. variability in the efficacy of delayed inducible resistance in mountain birch. entomologia experimentalis et applicata 62:107-115. senn, j., and e. haukioja. 1994. reactions of the mountain birch to bud removal: effects of severity and timing, and implications for herbivores. ecology 75:494-501. shipley, l. a., s. blomquist, and k. danell. 1998. diet choices made by free-ranging moose in northern sweden in relation to plant distribution, chemistry, and morphology. canadian journal of zoology 76:1722-1733. , a. w. illius, k. danell, n. t. hobbs, and d. e. spalinger. 1999. predicting bite size selection of mammalian herbivores: a test of a general model of diet optimization. oikos 84:5568. , and d. e. spalinger. 1992. mechanics of browsing in dense food patches: effects of plant and animal morphology on intake rate. canadian journal of zoology 70:1743-1752. snyder, j. d., and r. a. janke. 1976. impact of moose browsing on borealtype forests of isle royale national park. american midland naturalist 95:79-92. spalinger, d. e., and n .t. hobbs. 1992. mechanisms of foraging in mammalian herbivores: new models of functional response. american naturalist 140:325348. suominen, o., k. danell, and r. bergström. 1999a. moose, trees, and ground-living invertebrates: indirect interactions in swedish pine forests. oikos 84:215226. , , and j. p. bryant. 1999b. indirect effects of mammalian browsers on vegetation and ground-dwelling insects in an alaskan floodplain. ecoscience 6:505-510. thompson, i. d., and w. j. curran. 1993. a reexamination of moose damage to balsam fir – white birch forests in central newfoundland: 27 years later. canadian journal of forest research 23:1388-1395. , , j. a. hancock, and c. e. butler. 1992. influence of moose browsing on successional forest growth on black spruce sites in newfoundland. forest ecology and management 47:2937. , r. e. mcqueen, p. b. reichardt, d. g. trenholm, and w. j. curran. 1989. factors influencing choice of balsam fir twigs from thinned and unthinned stands by moose. oecologia 81:506-509. trottier, g. c. 1981. beaked hazelnut – a key browse species for moose in the boreal forest region of western canada. alces 17:257-281. vivas, h. j., and b.-e. saether. 1987. interactions between a generalist herbivore, the moose, alces alces, and its food resources: an experimental study of winter foraging behaviour in relation to browse availability. journal of animal ecology 56:509-520. alces37(2)_457.pdf alces36_105.pdf f:\alces\vol_39\p65\3930.pdf alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 109 status of moose populations and challenges to moose management in fennoscandia sten lavsund1, tuire nygrén2, and erling j. solberg3 1skrindvägen 25 s-756 47 uppsala, sweden; 2finnish game and fisheries research institute, ilomantsi game research station, fi-82900 ilomantsi, finland; 3norwegian institute for nature research, tungasletta 2, n-7485 trondheim, norway abstract: in the fennoscandian countries, norway, sweden, and finland, moose (alces alces) populations began to increase rapidly in the 1960s and have since then been among the most productive and heavily harvested moose populations in the world. at the start of the 20th century, the total annual harvest was < 10,000 moose, whereas in 2000, the annual kill reached about 200,000. the winter population was estimated to be about 500,000. in sweden and finland, the highest harvest numbers (and presumably population density) were recorded in the first half of the 1980s and in finland again in the late 1990s and during the beginning of the 2000s. in norway, the 1990s was the decade of the highest harvest numbers. the current regional moose density during winter varies from < 0.2 to about 2 moose/km2 within fennoscandia. locally, the density may far exceed this level in typical wintering areas (e.g., 5-6 moose/km2). in general, the current densities are lower in the north than in the south and higher in norway and sweden than in finland. the strong increase in harvest and the present high densities are explained by several factors. first, modern forestry clear-cutting practices have provided fennoscandian moose with prime habitats in the form of early succession stages. accordingly, the current carrying capacity is likely to be relatively high compared to the situation 50-100 years ago. the current trend, however, is towards less activity in the forest and a decreasing proportion of forests found at an early successional stage. this may increase the food limitation already seen in several populations; i.e., in all three countries, body mass and recruitment rates have been found to decrease with increasing density. second, the introduction of sex and age-specific harvesting in the early 1970s has increased the general productivity of the populations. by focusing the harvest on calves, yearlings, and adult males, the proportion of productive females, the mean age of females, and the annual recruitment rate have increased. simultaneously, the proportion and mean age of males have decreased, and in some populations, this has been associated with delayed parturition dates and lower fecundity; i.e., due to inadequate number of males for timely reproduction. third, mortality other than hunting is low, and only near the eastern border of finland with russia has predation by wolves and bears had a notable effect on productivity figures. this situation is about to change with increasing populations of large carnivores in all of fennoscandia during the 1990s. the management principles have been quite similar within fennoscandia, although differences in legislation have resulted in national and regional differences in management performance. in general, moose managers take advantage of data collected by hunters during the hunting season (e.g., hunting statistics, number, sex, and age of moose observed) to monitor population development and determine hunting quotas. moreover, in all three countries, the issues of traffic accidents and damage to forestry and agriculture play a central role in moose management and discussions concerning optimum population sizes. alces vol. 39: 109-130 (2003) key words: alces alces, damage to forestry, fennoscandia, finland, harvest, management, moose, moose accidents, norway, population status, selective harvesting, sweden moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 110 during the last 100 years, the status of moose has changed from being a relatively rare species to becoming a widely distributed and dominating species all over the forested part of fennoscandia: norway, sweden, and finland (fig.1). less than 10,000 individuals were harvested annually at the start of the 20th century. in contrast, more than 200,000 (0.3 moose/km2) were harvested annually during the most recent years. this strong increase in harvest, and presumably population size, appears to have several causes, with changing practice in forestry (increasing frequency of clear cuts) and introduction of sex and age-specific harvesting probably being the most important. the absence of bears (ursus arctos) and wolves (canis lupus) was another factor that facilitated the increase in the populations (e.g., cederlund and markgren 1987, østgård 1987). the increasing density of moose has been appreciated by hunters, but has also caused frustration among landowners and the authorities because of the negative impact on commercial forests and highway traffic safety (e.g., solbraa 1998, hakkila and kärkkäinen 1999, edenius et al. 2002). in addition, there has been growing concern about the impact of high densities and the intensive harvesting on the moose population itself. high densities are commonly associated with decreasing body condition, fecundity, and survival (e.g., sæther 1997, gaillard et al. 1998), and intensive harvesting may have consequences for the population dynamics beyond the direct effect on the population growth rate. the proportion of adult males has, for instance, seriously decreased in many populations following the introduction of a sexand age-specific harvesting regime in the early 1970s, with possible demographic and genetic effects (e.g., nygrén 1986, 1990; ericsson 1999; sæther et al. 2003). there were several attempts to halt the population growth and stabilise the population size in fennoscandia during the 1980s and 1990s, but these attempts often resulted in large population fluctuations. legislative, administrative, and social factors were involved in the failures (nygrén 1998a). in addition, there was a lack of experience with the high density and productivity of the strongly female-biased populations, as well as limited data on population performance. thus, to be able to perform sustainable moose management, there is an evident need for consecutive information on moose population dynamics southern götaland eastern götaland western götaland western svealand eastern svealand southern norrland northern norrland lapland oulu district inland finland coastal finland 1 2 3 4 5 6 7 89 10 11 12 13 14 15 16 17 18 fig. 1. the fennoscandian area: norway, sweden, and finland. counties in norway: 1 = r o g a l a n d , 2 = v a g d e r , 3 = a a g d e r , 4 = t e l e m a r k , 5 = v e s t f o l d , 6 = ø s t f o l d , 7=akershus, 8=buskerud, 9=hordaland, 1 0 = s o g n a n d f j o r d a n e , 1 1 = o p p l a n d , 12=hedmark, 13=møre and romsdal, 14=strøndelag, 15=n-trøndelag, 16=nordland, 17=troms, 18=finnmark. regions in sweden: southern götaland, eastern götaland, western götaland, eastern svealand, western svealand, southern norrland, northern norrland. regions in finland: coastal finland, inland finland, oulu district, lapland. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 111 and demography at each of the local, regional, and national scales. part of the process towards a better understanding and management of the moose population is to recall and analyse past population development and experience with various methods of moose management. in the present paper, therefore, we summarise the development and current status of the fennoscandian moose populations with respect to harvest, population density, and population structure. previously, papers on the status of moose in fennoscandia were presented as a part of the proceedings from the 2nd international moose symposium in uppsala, sweden in 1984 (cederlund and markgren 1987, nygrén 1987, østgård 1987), and accordingly, we will mainly focus on developments during the last 20 years in the present paper. we also provide a brief overview of present day moose management systems in fennoscandia, and finally, present some future challenges for moose management in the three countries. methods the status of fennoscandian moose populations is described mainly by the use of hunting statistics and systematic moose observations performed by moose hunters (“moose observation monitoring”). traditionally, harvest statistics have been assumed to provide a reasonably proxy to the variation in moose density in norway and sweden (cederlund and markgren 1987, østgård 1987), recognizing time lags due to delays in the decision-making process (e.g., to settle the right quota size; cederlund and markgren 1987, solberg et al. 1999). the relationship between the variation in harvest and moose density has since been confirmed in several independent studies (e.g., solberg et al. 1997, 1999). since the introduction of moose observation monitoring in the early 1970s in finland and in the mid-1980s in norway and sweden, variation in moose density and population structure have been estimated from the numbers of moose observed during the hunting season. the observation monitoring is a systematic recording and collecting of sex and age (calf or adult) of moose observed by moose hunters during the hunting season. several indices of population structure are calculated from the observation monitoring data (e.g., nygrén and nygrén 1976, solberg and heim 2002). the most important indices are “calves/ adult”, “calves/cow”, and “cows/bull” as indices of recruitment rate and adult sex ratio, respectively. in addition, indices of population density are calculated in norway and sweden as “moose seen per hunter day” and in finland as “moose seen per team-hunting day”. in finland, hunters also provide estimates of the number of moose left on their hunting grounds after the hunting season. despite the crude sampling procedure and a high number of likely confounding variables (variation in weather, hunting skills, number of hunters, hunting methods, etc.), the observation indices are found to provide precise information on the temporal development in population size and structure within a given area (ericsson and wallin 1994, 1999; solberg and sæther 1999; solberg et al. in press), provided that the number of observations is relatively high (ericsson and wallin 1994, sylvén 2000). in finland, more than 5,000 hunting clubs annually record 200,000 – 300,000 moose observations and, in norway, approximately 200,000 observations are added annually to the database (rolandsen et al. 2004). in sweden, these data were not yet available in 2002 for analysis on a national level (j. kindberg, svenska jägareförbundets viltövervakning, personal communication) despite having been collected on a local level since the early 1980s. the only reliable and systematic statistics available in sweden are therefore from the harvest moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 112 0 20 40 60 80 100 1964 68 72 76 80 84 88 92 96 00 h ar ve st x 1 00 0 0 50000 100000 150000 200000 19 71 19 75 19 79 19 83 19 87 19 91 19 95 19 99 20 03 sweden finland norway s t a t i s t i c s ( s v e n s k a j ä g a r e f ö r b u n d e t s viltövervakning 2002). to get a rough estimate of the moose density within and between countries, we used data from moose observation monitoring and harvest data, and to some extent data from aerial and other types of surveys during winter. in finland, retrospective analysis of previous years´ population estimates was also used for the period 1983 – 1996 (nygrén 1984, nygrén and pesonen 1993). results and discussion harvest, traffic accidents, and population development since the early 1970s, fennoscandian moose populations, as indexed by the annual harvest, have varied widely both temporally and regionally. in sweden and finland, the highest harvests were recorded in the first half of the 1980s, and in finland again in the late 1990s and early 2000s (figs. 2, 3, and 4). in norway, the harvest increased during the 1970s and 1980s until it more or less stabilized during the 1990s (figs. 2 and 5). the all-time annual record of moose kills was 174,709 in sweden (in 1982), 84,524 in finland (in 2002), and 39,309 in norway (in 1999) (fig. 2). in 2003, the total harvest was about 225,000 moose in fennoscandia: norway 38,600, sweden 103,185, and finland 84,466. the number of moose traffic accidents covaries to a large extent with the annual harvest during the period with available f i g . 2 . n u m b e r o f m o o s e h a r v e s t e d i n fennoscandia, 1971 – 2003. 0 50000 100000 150000 200000 19 72 19 75 19 78 19 81 19 84 19 87 19 90 19 93 19 96 19 99 fig. 3. annual variation in the accumulated harvest of moose in different regions of sweden during the period 1972-2001. the regions are from bottom-up (see fig. 1): southern g ö t a l a n d , e a s t e r n g ö t a l a n d , w e s t e r n g ö t a l a n d , e a s t e r n s v e a l a n d , w e s t e r n svealand, southern norrland, and northern norrland. fig. 4. annual variation in the accumulated harvest of moose in 4 regions of finland, 19642003. the regions are from bottom-up (see fig. 1): coastal finland, inland finland, oulu district, and lapland. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 113 data in fennoscandia (figs. 6-8). although this index varies with several factors (e.g., the severity of the winter; andersen et al. 1991), it is commonly assumed to be closely associated with moose density (e.g., lavsund and sandegren 1991; solberg et al. 1997, in press; haikonen and summala 2000). in norway, the number of moose killed in traffic accidents peaked in 1993 (fig. 6). a similar relationship is present in sweden from the period 1972-1999 (seiler 2003; fig. 7). the number of accidents peaked in 1980, 2 years prior to the peak in the moose harvest, whereas another peak appeared in the late 1980s, 2 years prior to the second peak in the number of harvested moose. this time difference is consistent with the assumption that harvesting is the main driver of population fluctuations and that the number of road-kills is a fair index of population density (solberg et al. in press). in finland, 1,100 – 3,000 moose accidents were reported annually during the period 1976-2003 (fig. 8). the number of fig. 5. annual variation in the accumulated harvest of moose in different counties of norway during the period 1971-2003. the counties are from bottom-up: østfold, akershus, hedmark, vestfold, buskerud, oppland, telemark, a-agder, v-agder, s-trøndelag, n-trøndelag, nordland, troms, finnmark, rogland, hordaland, sogn og fjordane, møre og romsdal (in counties 14-18 the annual harvest is still very modest). concerning counties cf. fig. 1. year 1988 1990 1992 1994 1996 1998 2000 m o o s e k il le d i n t ra ff ic a c c id e n ts 0 200 400 600 800 1000 1200 1400 1600 car train h a rv e s t 24000 26000 28000 30000 32000 34000 36000 38000 40000 42000 harvest fig. 6. annual variation in the number of moose harvested and reported dead in accidents on roads and railways in norway, 1 april 1987 to 31 march 2002. data are reported for the hunting year from 1 april to 31 march in year t+1. accidents does not correlate as well as in sweden and norway (figs. 6-7) with the large annual fluctuations in harvest, but seems to be a rather good index of populamoose status in fennoscandia – lavsund et al. alces vol. 39, 2003 114 tion density in finland (fig. 8). the winter moose population in finland was estimated to be between 66,700 (1996) and 113,000 –125,000 animals (2002) in the period 1981-2002 (nygrén 1996a, unpublished data; ruusila et al. 2002), whereas in norway, solberg et al. (in press) estimated the total norwegian moose population during winter to be between 90,000 (1995) and 117,000 (2000) in the period 1991-2000. the swedish moose population in the winter of 2000/2001 was estimated to be around 250,000. thus, almost 500,000 moose may have roamed the forests of fennoscandia during winter at the start of the new millennium. regional population density the winter density of moose in 2000/ 2001 for all countries combined (forested areas in fennoscandia cover approximately 650,000 km2; global resource assessment 2000 fao, www.fao.org) was 0.7-0.8 moose/km2 or slightly less than 1.0 moose/ km2 in norway and sweden and approximately 0.5 moose/km2 in finland. in norway, the highest densities of moose are found in the southeastern and central parts (fig. 9), with an average winter density between 1 and 2 moose/km2, and 0 20000 40000 60000 80000 100000 120000 140000 160000 180000 200000 19 72 197 5 197 8 198 1 198 4 19 87 19 90 19 93 19 96 19 99 0 1000 2000 3000 4000 5000 6000 7000 moose harvest left moose vehicle collisions right fig. 7. annual variation in the number of moosevehicle collisions and harvested moose in sweden, 1972 -1999. trends in swedish moose harvest correlate significantly with the number of police-reported vehicle collisions with moose (r2 = 0.77, n = 30, p <0.0001) (after seiler 2003). 0 1000 2000 3000 4000 1976 1981 1986 1991 1996 2001 n u m b e r o f m o o s e a c c id e n ts 0 20 40 60 80 100 x1000 n u m b e r o f h ar ve st ed m o o s e accidents moose bag fig. 8. annual variation in the number of moosetraffic accidents and harvested moose in finland, 1976-2003. data on moose accidents are from the finnish road administration/road traffic accident data bank. fig. 9. mean annual moose harvest in norway per km2 of forest and bogs below the timberline in different municipalities within counties (demarcated by black lines; see fig. 1), 19992001. white: no hunting; light gray: 0.01 – 0.10; gray: 0.11 – 0.40; dark gray: 0.41 – 0.70; and black: 0.71 – 1.20 moose/km2, respectively. note that the coastline is demarcated in black. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 115 for a few municipalities slightly above 2 moose/km2 (solberg et al. 2003). in the more continental parts of southern norway and inner parts of northern norway, the average moose density is lower, but may in effect be much higher during winter due to concentrations in restricted wintering areas. in particular, in areas with deep snow, moose tend to congregate on the valley floor or in low elevation areas with less snow during winter (sæther et al. 1992, hjeljord 2001). in these areas, densities may far exceed 2 moose/km2 over large areas (e.g., sæther et al. 1998). for instance, in the central parts of troms (area 17, fig. 1), moose from the mainland part of the county tend to concentrate in 2 valleys during winter where the average density may reach as high as 5-6 moose/km2 in the core distribution areas (20-50 km2) (b.–e. sæther, j. solberg, and m. heim, unpublished data). the lowest moose density is found along the west coast, particularly in the more central coastal regions, where moose are still a rare sight and hunting is prohibited (fig. 9, solberg et al. 2003). to some extent these low densities may be a temporary phenomenon as moose quite recently have colonized these areas. in sweden, the highest winter densities of moose (calculated using harvest data) during recent years (2001) are found in central sweden (svealand, figs. 1 and 10) with densities of 1.1 -1.2 moose/km2. slightly lower densities are found in southern sweden (götaland, figs. 1 and 10, 0.6 0.9/km2) and northern sweden (norrland, figs. 1 and 10, 0.4 0.9 moose/km2). the lowest m o o s e d e n s i t i e s a r e f o u n d i n t h e southernmost part of sweden, an area dominated by farmland, and in the northernmost part of the country, which has very low forest productivity. in these northern regions, winters are usually 1-2 months longer than in the south and snow is deeper, (between 0.5 and 1 m; sveriges nationalatlas skogen 1996, sveriges nationalatlas klimat, sjöar och vattendrag 1997). as in northern norway, seasonal migrations to wintering areas are common in these regions, and may locally exceed the average density by as much as 5 times (sweanor 1987, ball et al. 2001). in finland, the most preferred habitats for moose are in coastal finland (fig. 1), especially in the southwestern archipelago and the west coast where the snow depth does not restrict moose mobility and the growing season is longer than in other parts of finland (fig. 11). historically, these areas have had the greatest moose densities in spite of efforts to reduce population density in areas where the human population and traffic densities are the highest. the highest densities reached to date were in southern finland during the winter 1977-78 (fig. 12), with average densities of some game management districts exceeding 1.1 moose/km2. subsequently, densities in both 0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8 2 19 72 19 75 19 78 19 81 19 84 19 87 19 90 19 93 19 96 19 99 southern götaland eastern götaland western götaland eastern svealand western svealand southern norrland northern norrland fig. 10. moose post-hunting population density (moose/km2) in relation to forest land in sweden, 1972 – 2001. the figures must be regarded only as a rough index of the population density. regions as given in fig. 1. moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 116 coastal and inland finland were reduced to 0.4 – 0.5 moose/km2 (fig. 12). in the oulu district, where the population increase during the 1970s was much slower (fig. 12), the densities increased to the same level as in coastal and inland finland. in lapland, the average densities of moose were lower, but even there, the highest densities before 1996-97 were achieved in the mid-1980s. the stable period ended after a change in legislation during the second half of the 1990s. at first the densities decreased, especially in inland and eastern finland where the density decreased below 0.25 moose/km2 in some game management associations. some associations in the easternmost areas of finland, where large carnivore populations exist and have a significant effect on moose productivity (nygrén 1980), protected the moose for 13 years (nygrén 1998b). data comparable to density figures for 1974-96 in finland are not available for the late 1990s. however, according to moose density indices (fig. 13), there were about 0.6 moose/km2 in coastal finland, 0.5 moose/ km2 in inland finland and 0.3 moose/km2 in the oulu district (nygrén et al. 2000). since then, densities first increased and then decreased to an average level of 0.35 moose/ km2 in the winter of 2003/2004 (ruusila et al. 2002, 2004). changing population structure and population condition there has been a significant change in moose population structure in fennoscandia during the last 30 years. prior to 1970, moose were mainly harvested as yearlings or older, with both sexes almost equally represented. however, as part of the strategy to increase population density in the early 1970s, ageand sex-specific harvesting was introduced in norway (østgård fig. 11. density gradients of the finnish moose population, 1977-1996 (moose/km2 of dry land area). modified from nygrén 1996b. moose densities as calculated in fig. 12 alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 117 coastal finland 0 2 4 6 8 10 12 1975 1981 1987 1993 1999 ob s. p er h u n tin g d ay 0 0,2 0,4 0,6 0,8 m o o se /k m ² observations per hunting day moose/km² inland-finland 0 2 4 6 8 1975 1981 1987 1993 1999 o b s. p er h un tin g d ay 0 0,2 0,4 0,6 m o o se /k m ² oulu district 0 2 4 6 8 1975 1981 1987 1993 1999 o b s. p er h un tin g d ay 0 0,1 0,2 0,3 0,4 m oo se /k m ² 0 0,2 0,4 0,6 0,8 1974 1978 1982 1986 1990 1994 m oo se /k m 2 coastal-finland inland-finland oulu district lappi 1 9 8 7 ) a n d s w e d e n ( c e d e r l u n d a n d markgren 1987), and similarly, new hunting principles were adopted in finland after the protection of moose from 1969-1971 (nygrén 1987). the focus of the harvest was put on bulls and juveniles (calves and yearlings), leaving an increasing proportion of productive adult cows in the population. the result was an immediate strong increase in the harvesting of adult males (fig. 14), whereas the calf harvest was initially low due to a reluctance to shoot calves (fig. 15). since then, however, there has been a gradual change of attitude and at present calves are harvested at a rate approximately proportional to their presence in the population in norway, and at even higher proportions in sweden and finland (fig. 15). in contrast to the proportion of calves, the proportion of males in the adult harvest has decreased compared to the 1970s (fig. 14). this is likely a reaction to the previously intensive harvesting of males, leaving a decreasing proportion of adult males in the population. for instance in norway, adult males comprised 40-50% of the yearling fig. 12. variation of moose densities (moose/ km2 of dry land area) in 4 regions of finland. from nygrén 1996b. regions as given in fig. 1. the corresponding density of moose/km2 forested area was calculated by multiplying the number of moose/km2 dry land area by 1.40 in coastal finland, 1.20 in inland finland, 1.10 in oulu district, and 1.03 in lapland. fig. 13. moose density indices (estimation of hunters of moose/km2 of dry land area and moose seen/team hunting day) in 3 regions of finland, 1975-1999 (data from finnish game and fisheries research institute). moose densities as calculated in fig. 12. regions as given in fig. 1. moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 118 and adult population in the early 1970s (sæther et al. 2001), whereas the proportion of males is currently less than 30% in many populations and in some populations closer to 20% (fig. 16). similarly, the proportion of males decreased quickly in southern finland during the 1970s and early 1980s (e.g., nygrén 1986), but later that trend was slowed as a result of changes in harvest recommendations (fig. 17). no comparable figures are available for swedish populations. however, since on average there are a similar proportion of adult males in the harvest in sweden as in finland (fig. 14), we believe the proportion of adult males in the swedish population to be higher than in norway and comparable to the situation in finland. following the change of harvest systems there was an increase in the production of calves all over fennoscandia (e.g., koivisto 1963; nygrén 1984, 1987; solberg et al. 1999; nygrén et al. 2000). a part of this was probably due to an increase in the p r o p o r t i o n o f a d u l t f e m a l e s i n t h e populations. however, by relaxing the harvest of adult moose females from an intensive harvest pressure and by intensifying the calf harvest, the average age of females in the populations also increased, which positively influences fecundity (nygrén 1990, solberg et al. 1999). selective harvesting of adult females may have further enhanced this development (cederlund and markgren 1987, wallin 1992, solberg et al. 2000). many hunters select females based on the numbers of calves accompanying the females during the hunt and try to select those without calves in preference to those with 1 or 2 calves. highly reproductive females may consequently experience higher survival. accordingly, ericsson (1999) showed that the cost of reproduction in sweden was reversed for the high-reproductive female segment aged 5-10 years. entering the hunt with 2 calves was more beneficial for survival than entering with 1 calf or no calves. females 5-10 years old not giving birth had a 3.2 times higher risk of being killed during the hunt. this selective harvest resulted in a 2.5 times higher potential growth rate for the population versus a random harvest of adult females. probably as a result of selective harvesting, the present productivity of the finnish moose population is the highest ever recorded. in autumn 1999, the average number of calves/female was 1.01 in coastal 0 0,5 1 1,5 2 2,5 19 72 19 76 19 80 19 84 19 88 19 92 19 96 20 00 sweden finland norway fig. 14. bulls per cow in the moose harvest in sweden, finland, and norway, 1972 – 2001. 0 0,1 0,2 0,3 0,4 0,5 0,6 19 72 19 75 19 78 19 81 19 84 19 87 19 90 19 93 19 96 19 99 sweden finland norway fig. 15. the proportions of calves in the moose harvest in sweden, finland, and norway, 1972 – 2001. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 119 finland, 1.02 in inland finland, 0.91 in oulu district, and 0.78 in lapland (figs. 1 and 18; nygrén et al. 2000). until 2002, there was no significant change (ruusila et al. 2001, 2002). in norway, the numbers of observed calves/female also increased during the initial phase through to a peak in the early 1990s and have since been decreasing over large areas. the decrease has been associated with increasing population density and decreasing body masses of calves and yearlings and is generally assumed to be caused by density-dependent food limitation (solberg et al. 1997, 2002). in spite of high moose densities in fennoscandia, no clear density effect has been found that could explain the large 1986 1988 1990 1992 1994 1996 1998 m a le s p e r fe m a le 0,3 0,4 0,5 0,6 0,7 1986 1988 1990 1992 1994 1996 1998 m o o s e s e e n / h u n te rd a y 0,2 0,4 0,6 0,8 1,0 1,2 1,4 year 1986 1988 1990 1992 1994 1996 1998 c a lv e s p e r fe m a le 0,5 0,6 0,7 0,8 0,9 year 1986 1988 1990 1992 1994 1996 1998 t w in r a te 1,2 1,3 1,4 1,5 a b c d østfold hedmarkvestfold buskerud oppland telemark a-agder v-agder s-trøndelag n-trøndelagnordlandtroms akershus 0,30 0,50 0,70 0,90 1,10 1975 1979 1983 1987 1997 1995 1999 m a le s p e r fe m al e coastalfinland inlandfinland oulu district lapland fig. 17. annual variation in the number of males per female in 4 regions of finland, 1975-1999. data from moose observation monitoring; finnish game and fisheries research institute. regions as given in fig. 1. fig. 16. norway. annual variation in (a) moose seen per hunter-day, (b) adult (> 1-year-old) males per female, (c) calves per female, and (d) calves per calf-rearing female (twinning rate) in different counties of norway during the period 1986-1998. data from “moose observation monitoring”. counties as given in fig. 1. moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 120 (through maternal effects or vegetation), or alternatively, that the range quality (carrying capacity) may be significantly reduced (e.g., solberg et al. 2002, broman 2003). factors behind population development recognizing that harvesting is among the most important factors behind the variation in moose density in fennoscandia, it is of interest to describe some of the underlying processes leading to the observed population development during the last 30 years. in general, the increasing moose densities are assumed to be a reaction to the introduction of age and sex-specific harvesting in the early 1970s and relatively light harvesting due to a general desire to increase moose hunting opportunities (nygrén 1984, cederlund and markgren 1987, østgård 1987). as the population density increased, however, the number of moose killed on the roads and railroads by cars and trains (figs. 6, 7, 8) and damage to forestry also increased (lavsund 1987, lavsund and sandegren 1991, seiler 2003). in sweden and finland, these problems resulted in public opinions that the moose population had to be reduced, which led to the peak harvest in the early 1980s. similarly, in sweden, the second peak in the late 1980s is regarded to be a reaction to a still ongoing discussion concerning moose-forest interactions. in 1988, a special report, “älgen och skogen” (moose and forestry) was published (rülcker 1988), and in 1990 proposals concerning population goals for moose were presented by the government (fransson 1990). forestry concerns in the late 1990s were again reflected in increased harvest (svenska jägareförbundets viltövervakning 2002) driven at least to some extent by forestry stake-holders (carlestål 2000). special survey methods were introduced to measure the level of damage to commerscale fluctuations in the population density. in all three countries, however, some density-related factors have been studied and do seem to locally have an effect on population dynamics. in norway, moose body mass has been decreasing since the 1970s, following increasing densities (solberg et al. 1997, hjeljord and histøl 1999) and, as mentioned above, calves/female ratios have decreased in many populations during the last 30 years (e.g., solberg et al. 1997, 2000). in sweden, body mass and fecundity decreased in several populations following the high moose densities in the early 1980s (sand et al. 1996). similarly, in finland, the average productivity and body mass decreased at the beginning of 1980s after the peak density years, but increased again during lower density periods of the 1990s (nygrén 1997). in other high density populations in southern sweden and southern norway, body mass and recruitment rates decreased following increasing density, but currently show no sign of increase again despite significant reduction (> 50%) in population density (e.g., solberg et al. 2002, broman 2003). this indicates that long time lags in the effect of high density 0,40 0,60 0,80 1,00 1,20 1975 1979 1983 1987 1991 1995 1999 c a lv e s p e r fe m a le coastalfinland inlandfinland oulu district lapland fig. 18. annual variation in the number of calves per female in 4 regions of finland, 1975-1999. data from moose observation monitoring; finnish game and fisheries research institute. regions as given in fig. 1. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 121 cially valuable trees, as well as to species i m p o r t a n t t o b i o l o g i c a l d i v e r s i t y (skogsstyrelsen 2002) and by the late 1990s forest agencies had presented goals concerning acceptable levels of different kinds of forestry damage. one goal was that yearly levels of certain types of damage to pine saplings must not exceed 2 %, and that willows (salix spp.), aspen (populus tremula), and rowan (sorbus aucuparia) must be able to regenerate. following similar problems in norway, there were several attempts to stabilize the population size during the 1980s, especially in the southeastern part of the country where the growth rate was exceptionally high during the 1970s and early 1980s. in the more western and northern counties, where the density increased more slowly due to a slower introduction of the new hunting regime (østgård 1987), a similar response was observed in the early and mid-1990s. during the initial stabilizing phase the effort to halt the population increase often resulted in overand under-harvesting due to local inexperience with the dense and highly productive female-biased populations. this was exacerbated by a lack of appropriate census data on moose density, as data from moose observation monitoring was of little use in moose management until the late 1980s. since then, population indices derived from moose observation monitoring are increasingly used as a tool in local moose management in norway, and the frequency of apparently uncontrolled fluctuations in the annual harvest have decreased. management of moose in finland has been very target-oriented. the public discussion about management goals began in the mid-1970s and the first density goals were set in 1976 (nygrén 1987). in coastal finland, the maximum tolerable density was thought to be 0.7 moose/km2 of dry land (lakes excluded) compared to 1.0 moose/ km2 of forestry land. in short order this was considered to be too many moose and the density goals were adjusted down in 1980 and revised again in 1984, 1988, 1994, and 1995. among the principal early reasons for lower density goals was the negative effect of high densities on moose calf production. later, the goals for maximum densities were based on the tolerable number of traffic accidents (fig. 8) and damage to forestry and agriculture. unlike in sweden and norway, moose damage on private land in finland is compensated for with money from license fees. up to 1993, management decisions were made at a centralized level. the finnish game and fisheries research institute (fgfri) played an important role and provided annual recommendations for license numbers and selective hunting after the density goals were set (nygrén and pesonen 1993). moose observations and retrospective population analyses had an important position in population monitoring. later, the hunting legislation was reformed and a system of locally operated moose management areas was adopted in 1993. the game management districts and associations became more independent, and the management of the moose population less coordinated. a couple of years later, the first problems with decreasing densities were experienced (nygrén 1996a). the number of licenses was cut drastically (nygrén et al. 1999), and, after the second rapid population increase, the amount of forest damage and the number of traffic accidents (fig. 8) increased to intolerable levels. in 2002 and 2003, a larger number of moose were harvested than ever before in finland (fig. 4). moose hunters, hunting rights, and hunting methods in fennoscandia hunting is a very popular activity in fennoscandia, and depending on the number moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 122 of moose licenses issued, a large number of hunters are hunting moose. for instance in finland, about 100,000 hunters hunted moose in 1999, whereas 2 years earlier, when the moose quota was significantly lower, there were only 69,000 hunters (koskela and nygrén 2002). in norway, about 200,000 persons hunt annually, of which 56,000 are currently hunting moose (statistics norway 2002; http://www.ssb.no). the similar figure in sweden is about 300,000 hunters, of which 80% take part in moose hunting (ekman 1992). the moose itself has no owner in fennoscandia (e.g., nygrén 2000), but the landowner holds the hunting rights. the landowner can in turn lease the hunting rights to hunters. the start and the duration of the hunting season vary within and between countries. in finland, the season begins on the last weekend of september and continues until december 15th. in norway, the season is currently from 25 september to 30 october, but with local variation at both ends of the season. the longest seasons are found in sweden, where moose can be hunted for a minimum of 70 days and, in some areas, for as long as 3.5 months. moose hunting begins in the first week of september in northern sweden and in the second week of october in southern sweden. in addition to this system, there is a system of short (5 days or less) open seasons for small areas in which usually only one moose may be harvested. moose hunting is a social activity that often involves one or several dogs and a large group of people, of which several may not carry a gun (beaters). in finland, for instance, an average hunting team has 18 members and a hunting area of 5,600 hectares (koskela and nygrén 2002). hunting with dogs is most popular (73 % of hunting days), but also flushing with beaters is common (19 % of moose hunters do not carry a weapon) (koskela and nygrén 2002). although less detailed information is available from norway and sweden, similar hunting methods are common in both countries. current moose management, harvest regulations, and management goals the general mechanism for regulating the number of moose to be harvested is a licence system, with licences issued by local or regional authorities. in norway, the number of moose hunting licences is set by the municipality wildlife management authority in accordance with an established “minimum area” for each licence (danielsen 2001). until 2001, the county governor settled the “minimum area” (jaren 1992), but this responsibility is now delegated to the municipality. the size of the “minimum area” may vary among municipalities, and even within municipalities, depending on the local moose density and the planned development of the moose population in the municipality. moose hunting can only occur within the legal hunting season set by the national hunting authorities (directorate for nature management) and on land defined as a moose hunting area by the municipal wildlife management authorities. the number of licences is issued in accordance with the size of the hunting area and the local “minimum area”. the minimum hunting area for one licence (one moose) is the same size as the “minimum area” in that municipality or part of municipality. licenses may be specified as to sex and age categories (calf, adult female, adult male) or, alternatively, as a number of un-specified animals in cases where the hunting area has an approved population management plan of 3-5 years duration. the population management plans have to describe in detail the desired number and proportion of each sex and age-category of moose to be harvested during the planning period. to be approved, the manalces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 123 agement plan must also be compatible with municipal moose management goals; i.e., to what extent the local authorities want the moose density to increase or decrease. changes in population density and structure in norway are, in most municipalities, determined by the use of data from moose observation monitoring and to some extent by irregular winter aerial surveys (solberg and saether 1999). since the introduction of new hunting regulations in 2002, hunting areas with an approved population management plan are given the complete hunting quota for the planning period at the outset. the approval may be withdrawn or amended by the municipality if the harvest deviates significantly from the approved plan and/or if the status of the moose population radically changes. the new practice of locally based moose population management plans is part of a gradual decentralization of wildlife management in norway. the intention is to provide more precise moose management in accordance with local management goals (danielsen 2001). jaren (1992) and danielsen (2001) provide more detailed information about moose management in norway. in sweden, moose hunting licences are issued by the county authorities and specify the number of adults and calves that may be harvested. however, in specific “moose management areas”, which due to their large size and shape are assumed to hold their ‘own’ moose population, the number of moose to be harvested is decided by the landowners and hunters themselves. these “moose management areas” have been introduced to decrease the administrative work of the county authorities, and to inspire hunters to be more responsible in managing their local moose population. however, a shortcoming of this management system is that the information concerning the density and other traits of the populations often are inadequate to set regulations to achieve specific density goals (e.g., broman 2003). the main methods used to follow changes in population density and structure are by the use of data from moose observation monitoring and in some areas by winter aerial surveys. in finland, the 15 game management districts each determine the number of hunting permits issued. since 1993, a single licence has granted the right to shoot either 1 adult moose or 2 moose calves. the minimum area needed for a permit is 10 km2. in most cases, only one hunting club can hunt in a hunting area, except in large state-owned areas in northern finland where local inhabitants have hunting-rights in their own municipality and several hunting clubs can hunt moose simultaneously in the same area. earlier, the number of licences issued for hunting clubs was based on information from local hunters, hunting authorities (game management associations and districts), and fgfri moose researchers, as well as annual negotiations between hunting authorities and stakeholders. at present, the game management districts and associations produce the information needed for management, as well as deciding management goals and numbers of permits more independently, on the understanding that the density margins from 1995 are to be maintained. the main management goal in norway, sweden, and finland has been to maintain a highly productive moose population that tolerates large annual harvest quotas. also, stability of the population has been a common goal for all countries; in norway as stability of harvest and in finland as stability of the post-harvest population. finland is the only country that has official density goals set by the ministry of agriculture and forestry. since 1995, the goal has been to have 0.2 – 0.5 moose/km2 of dry land (compared to 0.26 – 0.65 moose/km2 of forestry moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 124 land) in most of the country (nygrén 1997). in the northernmost areas the goal has been lower, 0.05 – 0.3 moose/km2 of dry land (i.e., about the same as per forestry land in northern finland where more than 90 % of areas consist of forestry land). in norway and sweden, the goals are more local and less uniform, and higher moose densities are more readily tolerated than in finland. future challenges future challenges for moose management in fennoscandia are numerous. an important challenge concerns the recent change from a centralized to a decentralized system of moose management. in all fennoscandian countries, local decisionmaking has been accepted as the future system of moose management. to what extent this will improve or worsen the management is not yet known, but the critical comments are many (e.g., nygrén and nygrén 1994, nygrén 1998a, broman 2003). a commonly asked question is what are the capabilities of local hunters to apply their knowledge of moose population dynamics? in finland, for instance, local hunters (=game management associations) now are responsible for collecting information, for analyzing and determining the population status, for deciding population goals and license numbers, and for determining the composition of the harvest. in small hunting areas, this is an impossible task for the local hunters because moose summer and winter areas are usually larger than the area of the association. in addition, the goals can differ extensively between neighboring associations. professional managers and researchers in sweden and norway have expressed similar concerns. the general view is that moose hunters may have direct interest in the resource itself, but not necessarily have deep insight into moose population dynamics or interest in moose management. this may change in the future if local communities, or larger aggregations of local communities, are willing to invest the necessary resources in developing local expertise. alternatively, moose management may be delegated to traditional management agencies, while local involvement is restricted to formulation of goals (broman 2003). another challenge to future moose management concerns the negative effects of selective harvesting and high population density. in all three countries, but especially in norway, the proportion of adult males in the population has decreased significantly during the last 30 years. in populations with extreme sex-bias, sæther et al. (2003) and solberg et al. (2002) reported delayed parturition and lower fecundity, most likely due to an inadequate number of experienced males for timely reproduction. the same type of decrease in adult male proportions and calf production was experienced in finland after the peak harvest years in the early 1980s before the number of adult males/females was adjusted over a couple of years by femaledominated harvesting (nygrén 1986, nygrén and pesonen 1993). the mechanism needed to get more males back in the populations is simple in theory – a better balance in the adult kill. however, in practice this appears to be a difficult task as the hunters´ will to protect productive cows is strong, and may even be opposed by judicial impediments. for instance in finland, it is forbidden by law to kill a cow with a calf. another factor that is going to complicate moose management in the future is the increasing densities of both wolves and brown bears in all three countries. in scandinavia, wolves were “functionally extinct” from the early 1960s until late 1970s – the first breeding (since 1964) was recorded in northern sweden in 1978, and again in 1983 at the border between norway and sweden further south (wabakken et al. 2001). in alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 125 finland, stray wolves live all around the country, but the strongest populations have in the course of the 20th century existed along the south-eastern border zone (nyholm 1996). the present population in scandinavia probably descend from a few dispersing individuals from this border population between finland and russia (flagstad et al. 2003). although current numbers in fennoscandia are still low (around 100-120 wolves in norway and sweden (wabakken et al. 2004) and at least 150-165 in finland (kojola 2004)), the overall trend is for an increase. similarly, brown bears are slowly increasing in numbers in finland (at least 800-830 in 2003; kojola 2004) and sweden (around 1,000 in the mid-1990s; sandegren and swenson 1997) and significantly more in 2004 (kindberg et al. 2004), and are slowly recolonizing norway along the swedish border (< 50 individuals in 2003; solberg et al. 2003). in core bear areas, predation by bears on moose calves can be significant (swenson et al. 2001), and although it is unlikely that there will be social acceptance for high wolf densities in fennoscandia in the near future (e.g., palviainen 2000), their presence will have to be taken into account locally when setting moose harvest quotas (kojola and nygrén 1998, solberg et al. 2003). probably the largest challenge in the future, at least in norway and sweden, is how to deal with the impact of high moose densities on the forest ecosystem. the effects of dense moose populations on forest biodiversity are a focus for research and discussion in both sweden (persson et al. 2000, edenius et al. 2002) and norway (e.g., solbraa 1998). similarly, reports of decreasing body condition and reproduction in high-density areas receive increasing attention. particularly in parts of southern scandinavia, where calf production and body masses have not increased despite significant reductions in moose densities, there is growing concern that chronic high moose densities have created permanent or longterm changes in the forest that may take a long time to recover (e.g., punsvik 2004). alternatively, the frequency of forestry activity may be a primary stimulus of mooseforest management imbalances. during the last 30-40 years, modern forestry practices have provided fennoscandian moose with prime habitats in the form of early succession stages created by clear-cutting. accordingly, the current carrying capacity is assumed to be high compared to the situation 50-100 years ago (sæther et al. 1992). the current trend, however, is towards a change in the activity in forestry – more cleaning and less clear-cutting which means that a decreasing proportion of forests are found at an early successional stage (rolstad et al. 2002, skogsstyrelsen 2004). in finland the situation is different, as the densities of moose are much lower and forest statistics do not indicate any decrease in the area of forest clear-cutting. still, the increasing numbers of traffic accidents and forest damage have generated a wish to also decrease the moose density in finland. references andersen, r., b. wiseth, p. h. pedersen, and v. jaren. 1991. moose-train collisions: effects of environmental conditions. alces 27: 79-84. ball, j. p., c. nordengren, and k. wallin. 2001. partial migration by large ungulates: characteristics of seasonal moose ranges in northern sweden. wildlife biology 7:39-47. broman, e. 2003. environment and moose population dynamics. doctoral thesis, department of environmental sciences and conservation, göteborg university, göteborg, sweden. carlestål, b., editor. 2000. är älgen ett hinder för att nå de skogspolitiska målen? (is moose an obstacle to reach moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 126 the goals of the forestry policy?). kungliga skogs och lantbruksakademiens tidskrift 139:2:1-97. cederlund, g., and g. markgren. 1987. the development of the swedish moose population, 1970-1983. swedish wildlife research supplement 1:55-62. danielsen, j. 2001. local community based moose management plans in norway. alces 37:55-60. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forest. silva fennica 36:57-67. ekman, h. 1992. social and economic roles of game and hunting. pages 64-71 in r. bergström, h. huldt, and u. nilsson, editors. swedish game – biology and management. svenska jägareförbundet, stockholm, sweden. ericsson, g. 1999. demographic and life history consequences of harvest in a swedish moose population. ph.d. thesis, swedish university of agricultural sciences, umeå, sweden. , and k. wallin. 1994. antal älgar som ses bara en fråga om hur många som finns? att observera älg – en fråga om täthet, rörelser och synbarhet. mimeo, 31 pp. swedish university of agricultural sciences, department of animal ecology, umeå, sweden. (in swedish). , and . 1999. hunter observations as an index of moose alces alces population parameters. wildlife biology 5:177-185. flagstad, ø., c. w. walker, c. vilà, a. k. sundqvist, b. fernholm, a. k. hufthammer, ø. wiig, i. koyola, and h. ellegren. 2003. two centuries of the scandinavian wolf population: patterns of genetic variability and migration during an era of dramatic decline. molecular ecology 12:869-880. fransson, j. 1990. skada av vilt. sou 1990:60. (in swedish). gaillard, j-m, m. festa-bianchet, and n. g. yoccoz. 1998. population dynamics of large herbivores: variable recruitment with constant adult survival. trends in ecology and evolution 13:5863. haikonen, h., and h. summala. 2000. hirvikanta, liikenne ja hirvikolarit. liikenneministeriön julkaisuja/publications of the ministry of transport and communications 20: 1-105. edita ltd., edita, finland. (in finnish with english summary). hakkila, p., and m. kärkkäinen. 1999. hirvestäjä metsänomistajan kukkarolla. metsätieteen aikakauskirja 1:139-146. (in finnish). hjeljord, o. 2001. dispersal and migration of northern forest deer – are there unifying concepts? alces 37:353-370. , and t. histøl. 1999. range-body mass interactions of a northern ungulate – a test of hypothesis. oecologia 119:326-339. jaren, v. 1992. monitoring norwegian moose populations for management purposes. alces supplement 1:105111. kindberg, j., j. swenson, s. brunberg, and g. ericsson. 2004. prelimiär rapport om populationsutveckling och –storlek av brunbjörn i sverige, 2004. en rapp o r t t i l l n a t u r v å r d s v e r k e t f r å n skandinaviska björnprojektet 31 maj 2004.12.14. (in swedish). k o i v i s t o , i. 1963. hirvikantamme r a k e n t e e s t a , l i s ä ä n t y m i s e s t ä j a verotuksesta/ composition, productivity and kill of the finnish moose (alces alces) population. suomen riista 16:722. (in finnish with english summary). kojola, i. 2004. suurpetojen lukumäärä ja l i s ä ä n t y m i n e n v u o n n a 2 0 0 3 . riistantutkimuksen tiedote 194:1-7. (in finnish). alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 127 , and k. nygrén. 1998. karhun ja suden vaikutus hirvikantaan. jahti 1:89. (in finnish). ko s k e l a , t., and t. ny g r é n . 2002. hirvenmetsästysseurueet suomessa vuonna 1999./moose hunting clubs in finland 1999. suomen riista 48:65-79. (in finnish with english summary). lavsund, s. 1987. moose relationships to forestry in finland, norway and sweden. swedish wildlife research supplement 1:229-244. , and f. sandegren. 1991. moosevehicle relations in sweden: a review. alces 27:118-126. nygrén, k. 1980. susien vaikutuksesta hirvikantaan / effect of the wolf on the moose population. suomen riista 28:7178. (in finnish with english summary). . 2000. kenelle riista kuuluu. jahti/ jakt 4: 6-9. (in finnish). . and t. nygrén. 1976. hirvi ja hirvenmetsästys suomessa. riistantutkimusosaston tiedonantoja 2: 1-33. (in finnish). nygrén, t. 1984. hirvikannan inventointi ja verotuksen suunnittelu suomessa / moose population census and planning of cropping in finland. suomen riista 31:74-82. (in finnish with english summary). . 1986. hirvitiheydet pienentyneet, kannan rakenne edelleen vääristynyt. riistantutkimusosaston monistettu tiedote 47:1-4. (in finnish). . 1987. the history of moose in finland. swedish wildlife research supplement 1:49-54. . 1990. the relationship between reproduction rate and age structure, sex ratio and density in the finnish moose population. proceedings of the xvi congress of the international union of g a m e b i o l o g i s t s , v y s o k é t a t r y , štrebské pleso, èssr. . 1996a. hirvikanta pienimmillään 1 9 v u o t e e n , r a k e n n e e n t i s t ä k i n naarasvaltaisempi. riistantutkimuksen tiedote 145:1-28. (in finnish). . 1996b. hirvi. pages 103-108 in h. lindén, m. hario, and. m. wikman, editors. riistan jäljille. riistaja kalatalouden tutkimuslaitos, edita, helsinki, finland. (in finnish with english summary). . 1997. hirvikanta ja sen säätely. pages 39-52 in j. kairikko, j. aatolainen, p. louhisola, t. nygrén, and s. t a k a m a a , e d i t o r s . h i r v i e l ä i n t e n metsästyksen käsikirja. gummerys kirjapaino oy, jyväskylä, finland. (in finnish). . 1998a. voimistunut hirvikanta tuottavampi kuin koskaan – taustalla muutokset lainsäädännössä, menettelytavoissa ja tavoitteissa. riistantutkimuksen tiedote 154:1-17. (in finnish). . 1998b. metsä kasvattaa hirvet, ihmiset ja pedot korjaavat sadon. pages 64-65 in sellua sivistystä sahanpurua. metsäklusteri pohjois-karjalassa. p o h j o i s k a r j a l a n m e t s ä k e s k u s & pohjois-karjalan liitto. (in finnish). , and k. nygrén. 1994. 20 vuotta hirvihavaintoja. riistantutkimuksen tiedote 129:3-15. (in finnish). , and m. pesonen. 1993. the moose population (alces alces l.) and methods of moose management in finland, 1975-89. finnish game research 48:46-53. , , r. tykkyläinen, and m. wallén. 1999. hirvijahdin kohteena rakenteeltaan kuntoutunut ja erittäin hyvätuottoinen kanta. riistantutkimuksen tiedote 160:1-13. (in finnish). , r. tykkyläinen, and m. wallén. 2000. syksyn suurjahdin kohteena erittäin tuottava, nopeasti kuntoutunut hirvikanta. riistantutkimuksen tiedote moose status in fennoscandia – lavsund et al. alces vol. 39, 2003 128 168: 1-16. (in finnish). nyholm, e. 1996. susi (canis lupus). pages 38-41 in h. lindén, m. hario, and m.wikman, editors. riistan jäljille. riistaja kalatalouden tutkimuslaitos, edita. helsinki, finland. (in finnish with english summary). østgård, j. 1987. status of moose in norway in the 1970’s and early 1980’s. swedish wildlife research supplement 1:63-68. palviainen, s. 2000. suurpedot pohjoisk a r j a l a s s a – p o h j o i s k a r j a l a i s t e n l u o n n o n k ä y t t ä j i e n k o k e m u k s i a suurpedoista/large terrestrial carnivores in north karelia – experiences of north-karelian nature-users concerning large terrestrial carnivores. pohjoiskarjalan liiton julkaisu 51:38-154. (in finnish with english summary). persson, i.-l., k. danell, and r. bergström. 2000. disturbance by large herbivores in boreal forests with special reference to moose. annales zoologici fennici 37:251-263. punsvik, t. 2004. fylkesmannens plass i f r a m t i d a s h j o r t e v i l t f o r v a l t n i n g ? hjorteviltet 14:41-43. (in norwegian). rolandsen, c. r., e. j. solberg, and v. grøtan. 2004. ’sett elg’ –materialet i norge 1984-2002. hjorteviltet:6-13. rolstad, j., e. framstad, v. gundersen, and k. storaunet. 2002. naturskog i norge. definisjoner, økologi og bruk i norsk skogog miljøforvaltning. aktuelt fra skogforskningen 1-2002:1-53. ruusila, v., m. pesonen, s. heikkinen, a. karhapää, r. tykkyläinen, and m. wallén. 2004. hirvikannan koko ja vasatuotto pienenivät vuonna 2003. riistantutkimuksen tiedote 196:1-9. (in finnish). , , r. tykkyläinen, and m. wallén. 2001. hirvikannan kasvu pysähtyi, mutta naaraita säästävä verotus pitänyt vasatuoton korkeana. riistantutkimuksen tiedote 173:1-11. (in finnish). , , , and . 2002. hirvikanta lähes ennallaan suurista kaatomääristä huolimatta. riistantutkimuksen tiedote 180:1-12. (in finnish). rülcker, j. 1988. älgen och skogen. problemställningar och förslag till l ö s n i n g a r . s l u t r a p p o r t f r å n ä l g / skoggruppen. (in swedish). sæther, b.-e. 1997. environmental stochasticity and population dynamics of large herbivores: a search for mechanisms. trends in ecology and evolution 12:143-149. , r. andersen, o. hjeljord, and m. heim. 1998. ecological correlates of regional variation in life history of the moose alces alces: reply. ecology 79:1938-1939. , m. heim, e. j. solberg, k. s. jacobsen, r. olstad, j. stacy, and m. sviland. 2001. effekter av rettet avskytning på elgbestanden på vega. norwegian institute for nature research, fagrapport 049. (in norwegian). , e. j. solberg, and m. heim. 2003. effects of altering adult sex ratio and male age structure on the demography of an isolated moose population. journal of wildlife management 67:455466. , k. solbraa, d. p. sødal, and o. hjeljord. 1992. sluttrapport elg-skogsamfunn. norwegian institute for nature research, forskningsrapport 28. (in norwegian). sand, h., r. bergström, g. cederlund, m. östergren, and f. stålfelt. 1996. density-dependent variation in reproduction and body mass in female moose alces alces. wildlife biology 2:233245. sandegren, f., and j. swenson. 1997. alces vol. 39, 2003 lavsund et al. – moose status in fennoscandia 129 b j ö r n e n – v i l t e t , e k o l o g i n o c h människan. svenska jägareförbundet, stockholm, sweden. seiler, a. 2003. the toll of the automobile: wildlife and roads in sweden. doctoral thesis, swedish university of agricult u r a l s c i e n c e s , u m e å , s w e d e n . silvestria 295. skogsstyrelsen. 2002. enkel älgbetningsinventering äbin. skogsstyrelsen http://www.svo.se. (in swedish). . 2004. skoglig statistikinformation. h t t p : / / w w w . s v o . s e / f a k t a / s t a t / default.htm. solberg, e., v. grøtan, c. m. rolandsen, h. brøseth, and s. brainerd. in press. change in sex ratio as an estimator of population size for norwegian moose. wildlife biology. , and m. heim. 2002. monitoring moose in norway: see them, shoot them, measure them and eat them. pages 1619 in moose and deer, a special issue of “hjorteviltet” periodical for moose and deer in norway. , , b-e. sæther, and f. holmström. 1997. oppsummeringsrapport, overvåkningsprogram for hjortevilt. norwegian institute for nature research fagrapport 30. (in norwegian). , a. loison, b-e. sæther, and o. strand. 2000. age-specific harvest mortality in a norwegian moose alces alces population. wildlife biology 6:4152. , t. h. ringsby, b-e. sæther, and m. heim. 2002. biased adult sex ratio can affect fecundity in primipareous moose. wildlife biology 8:109-120. , and b-e. sæther. 1999. hunter observations of moose alces alces as a management tool. wildlife biology 5:4353. , , o. strand, and a. loison. 1999. dynamics of a harvested moose population in a variable environment. journal of animal ecology 68:186-204. , h. sand, j. linnell, s. brainerd, r. andersen, j. odden, h. brøseth, j. swenson, o. strand, and p. wabakken. 2003. store rovdyrs innvirkning på h j o r t e v i l t e t i n o r g e : ø k o l o g i s k e prosesser og konsekvenser for jaktuttak og jaktutøvelse. norwegian institute for nature research fagrapport 63. (in norwegian). solbraa, k. 1998. elg og skogsbruk, biologi, økonomi, beite, taksering, forvaltning. skogsbrukets kursinstitutt, biri, norway. s t a t i s t i c s n o r w a y . 2 0 0 2 . h t t p : / / www.ssb.no. sv e n s k a j ä g a r e f ö r b u n d e t s vi l t ö v e r vakning. 2002. avskjutningsstatistik 1960-2001. (in swedish). sveriges national atlas. 1996. skogen. (in swedish). . 1997. klimat, sjöar, och vattendrag. (in swedish). sweanor, p. y. 1987. winter ecology of a swedish moose population: social behavior, migration and dispersal. swedish university of agricultural sciences, department of wildlife ecology, report 13. swenson, j. e., b. dahle, and f. sandegren. 2001. bjørnens predasjon på elg. nina fagrapport 048: 1-22. sylvén, s. 2000. effects of scale on hunter moose alces alces observation rate. wildlife biology 6:157-165. wabakken, p., å. aronson, h. sand, t. h. strømseth, and i. kojola. 2004. ulv i skandinavia. statusrapport for vinteren 2003-2004. høgskolen i hedmark. oppdragsrapport nr. 5. , h. sa n d, o. li b e r g, and a. bjärvall. 2001. the recovery, distribution, and population dynamics of wolves on the scandinavian peninsula, 1978-1998. canadian journal of zoolmoose status in fennoscandia – lavsund et al. alces vol. 39, 2003 130 ogy 79:710-725. wallin, k. 1992. how to model moose population ecology? alces supplement 1:121-126. 139 vulnerablity of yearling and 2-year-old bull moose to two antler based harvest regulations in british columbia daniel a. aitken1, ian w. hatter2, roy v. rea3, and kenneth n. child4 1college of new caledonia, 3330 22nd avenue, prince george, british columbia v2n 1p8; 249-640 upper lakeview road, invermere, bc v0a 1k3, canada; 3natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia v2n 4z9, canada; 46372 cornell place, prince george, british columbia v2n 2n7 canada. abstract: a spike-fork (s/f) general open season (gos) for bull moose (alces alces) was introduced with a lottery draw, limited entry hunting (leh) in the omineca (1981), thompson (1993), and okanagan (1993) regions of british columbia. the s/f regulation permitted harvest of a bull having no more than two tines on one antler, including the tines on the main antler and brow palms; the leh controlled the harvest of bulls with antlers >s/f. in the peace region, the s/f regulation was implemented (1996) as part of soft regulations which permitted harvest of bulls with spike, fork, or antlers with 3 or more points on either brow palm; in 2003, soft10 regulations permitted the harvest of bull moose with ≥10 points on one or both antlers. these combinations with the s/f regulation were meant to control annual harvest of bulls, maintain herd social structure, and maximize recreational opportunity. we used age and antler point data collected through a voluntary tooth return program (vtrp) from 1988 to 2003 (n = 39,325) to assess vulnerability of yearlings (n = 12,743) and 2-year-olds (n = 8,712) to the s/f regulation as well as a hypothetical spike-only regulation. for each age class, we defined potential vulnerability to the s/f regulation as the proportion of bulls in the harvest with s/f antlers when no antler-based restrictions were in place. we similarly defined potential vulnerability to the spike-only regulation as the proportion of bulls in the harvest with at least one spike antler. potential vulnerability across british columbia to the spike-fork regulation was 43% for yearlings and 10% for 2-year-old bulls, whereas potential vulnerability to the spike-only regulation was 8% for yearlings and 1% for 2-year-old bulls. realized vulnerability to harvest of each age class was defined as the proportion of that age class with spike-fork antlers when there were spike-fork regulations combined with either leh or other antler-based restrictions. similarly, realized vulnerability to harvest for spike-only bulls in each age class was the proportion of harvested bulls with at least one spike antler when spike-fork regulations were combined with either leh or as part of the soft or soft10 regulations. realized vulnerability across british columbia to the s/f regulation was 49% for yearlings and 7% for 2-year-old bulls; realized vulnerability to the spike-only regulation was 9% for yearlings and 1% for 2-year-old bulls. potential vulnerabilities and realized vulnerabilities varied regionally and annually, which may reflect different subspecies of moose (a. a. shirasi, a. a. andersoni, a. a. gigas) with different antler architectures, but more likely, differences related to habitat quality across the latitudinal breadth of british columbia. the s/f regulation provides hunting opportunity, but combined with other hunting seasons/regulations, may not provide adequate protection of yearling and 2-year-old bulls in some regions. the spike-only regulation exposes fewer yearling and 2-year-old bulls to harvest and offers an alternative to regulate bull harvests while maintaining hunter opportunity. alces vol. 57: 139–166 (2021) key words: alces alces, antler regulations, bull moose, hunting, spike-fork, spike-only, vulnerability, yearling bull spike-fork antler regulations – aitken et al. alces vol. 57, 2021 140 licensed hunting of moose (alces alces) in north america has traditionally focused on harvesting bull moose (bulls) with temporal restrictions on harvest of antlerless animals in a population (timmermann 1987). a departure from this tradition was implemented in the omineca region of central british columbia (macgregor and child 1981) when regulations were modified to institute selective harvest of a. a. andersoni. these regulations were designed to address differential harvest of sexes and age classes focusing on calves, cows, and immature and mature bulls (bubenik 1971, child 1983, child and aitken 1989). beginning in 1981, the general open season (gos) afforded hunters the opportunity to harvest either a calf or spike-fork (s/f) bull, with other bulls and cows harvested through a lottery draw or limited entry hunt (leh). the s/f regulation permitted harvest of bulls having no more than 2 tines on one antler, including tines on the main antlers and brow palms (bc ministry of environment 1981–1982); essentially, this approach directed hunters to focus harvesting on smaller antlered yearlings and 2-year-old bulls (child et al. 2010a, 2010b). in combination, these regulations were intended to control harvest of bulls, maximize recreational opportunity, and maintain a balanced social structure (i.e., maintain prime bulls) to improve herd productivity (child 1983, bubenik 1985, 1987, aitken and child 1992, boer 1992, timmermann 1992, child 1996). while focussed harvest on young/ immature bulls combined with controlled harvest of older/mature bulls is a desired outcome of selective harvesting, it is also important to ensure that yearlings and 2-year-old bulls are in sufficient supply to be recruited into older age classes to maintain desired breeding sex ratios and age structure. previous assessments of the s/f regulation in the omineca region indicated that this antler-based regulation, in combination with a controlled harvest of older/mature bulls, held promise as an effective harvest strategy to control harvest of bulls while providing moderate levels of hunting recreation (hatter and child 1992, hatter 1998, 1999, demarchi and hartwig 2008). the s/f regulation had been used in alaska since 1987 as part of a selective harvest system (shs) (schwartz et al. 1992) to regulate harvests of a. a. gigas, but was suspended in 2011 and 2012 over concern of skewed bull:cow ratios. it was replaced with a spike-only regulation in several areas in 2013 (robertia 2013) remaining in use to the present day (alaska department of fish and game 2020–2021). the predominant sub-species of moose throughout british columbia is alces alces andersoni with a. a. shirasi found in the extreme southeastern corner of the province (eastman and ritcey 1987). although a. a. gigas was thought to occur at the extreme northwestern corner of british columbia (eastman and ritcey 1987), recent genetic analyses (hundtermark et al. 2006, colson 2013, decesare et al. 2020) suggests that a. a. andersoni extends further northwest than previously believed. gasaway et al. (1987) reported that mean antler size of prime bulls was smallest in a. a. shirasii, intermediate in a. a. andersoni, and largest in a. a. gigas. however, decesare et al. (2020) suggested that environmental factors rather than genetic differences were the primary influence of observed differences in antler size between these subspecies. since antlers vary with age and body size (gasaway et al. 1987, stewart et al. 2000, bowyer et al. 2002, child et al. 2010a, jensen et al. 2013, andreozzi et al. 2015), and presumably between subspecies (gasaway et al. 1987), we expect that the prevalence of s/f antlers would be highest in a. a. shirasi, intermediate in a. a. andersoni, and lowest in a. a. gigas. furthermore, we expect more s/f antlers in yearlings alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 141 compared to 2-year-olds, as well as in moose from southern versus northern areas across the distribution of a. a. andersoni. the purpose of this study was to examine vulnerability to the s/f regulation from 1988 to 2003 of both yearling and 2-year-old bull moose within 7 specific wildlife management regions and across the province of british columbia. we examined whether the s/f regulation was equally effective in all years and regions given the presence of different subspecies of moose, as well as the extensive geographical ranges and the 15-year period of study. the vulnerabilities of the subspecies found within british columbia were examined by comparing the vulnerabilities of a. a. shirasi in the eastern portions of the kootenay region to those of a. a. andersoni in the western portions of the kootenay region, and by comparing the vulnerabilities of a. a. gigas in the northwestern portion of the skeena region with those of a. a. andersoni in adjacent areas of that region. differences in vulnerabilities were compared in the range of a. a. andersoni from south (49th parallel) to north (60th parallel) across british columbia. we also evaluated temporal trends in vulnerabilities of yearlings and 2-year-old bulls to the s/f regulation within regions. we report on the historical use of the s/f regulation within british columbia, and make recommendations about its continued use. we also examined the vulnerability of yearlings and 2-year-old bulls to a hypothetical spike-only regulation (a bull having no more than 1 tine on one antler) to evaluate that as an alternative to the s/f regulation if overharvest of yearling or 2-year-old bull moose occurred under the s/f regulation. study area we assessed vulnerability of yearling and 2-year-old bull moose to antler-based regulations from 1988 to 2003 in 7 wildlife administrative regions of british columbia: kootenay, okanagan, thompson, cariboo, omineca1, skeena, and peace1 (fig. 1). these regions exhibit a wide variety of landforms from high mountain peaks interspersed with rolling plateaus to alluvial valleys (church and ryder 2010). the majority of our study area lies within the montane cordillera and boreal cordillera ecozones, with the taiga plains and boreal plains in the far northeast and the semi-arid plateau to the southwest (canadian council on ecological areas 2014). the administrative regions vary physio-graphically, ecologically, and climatically with forest cover types and moose habitat use patterns varying by region (eastman and ritcey 1987). 1note: although the omineca and peace are zones a and b within region 7, we refer to them as the omineca region and peace region throughout the article. fig. 1. wildlife administrative regions of british columbia, ministry of forests, lands, natural resource operations and rural development (from the bc hunting and trapping regulations synopsis 2018–2020). the study area included 7 regions: 3 thompson, 4 kootenay, 5 cariboo, 6 skeena, 7 zone a omineca, 7 zone b peace, and 8 okanagan. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 142 harvests of bulls throughout the study regions were regulated with a mix of gos, leh, and/or s/f seasons (hatter 1998, 1999, demarchi and hartwig 2008; table 1). the s/f antler point restriction was employed in the omineca region throughout the study and was later adopted in the thompson (1993), okanagan (1993), and peace (1996) regions. the omineca, thompson, and okanagan regions combined the s/f gos with a leh that did not employ antler point restrictions. in contrast, starting in 1996, the peace region combined a s/f gos with a gos that included antler point restrictions. prior to 1996, an any-bull gos was advertised in the peace region from 15 august to 31 october each year; from 1996 to 2003, the any-bull gos was retained for the period of 15 to 31 august. the harvest in september and october from 1996 to 2002 was subject to the soft regulation which permitted the harvest of bulls with spike or fork antlers (sof) or having at least one antler with a brow palm bearing ≥3 points (t) (bc ministry of environment 1996–1997; fig. 2). in 2003, soft was modified to the soft10 regulation (poole and demars 2015) which permitted harvest of bulls with at least one antler with a minimum of 10 points, in addition to those with spike or fork antlers (sof), or with ≥3 points on either brow palm (t) (bc ministry of water, land and air protection 2003–2004). the cariboo and skeena regions did not employ antler point restrictions with either gos or leh. methods age and antler point data for assessing the effects of antler regulations on yearling and 2-year-old bulls were obtained from the voluntary tooth return program (vtrp) table 1. temporal and regional summary of any bull general open season (gos), any bull limited entry hunt (leh), and spike-fork (s/f) general open season hunting regulations used in regions of british columbia (1988 to 2003). region any bull gos or leh s/f gos combined with leh kootenay 1988–2003 not used okanagan 1988–1992 1993–2003 thompson 1988–1992 1993–2003 cariboo 1988–2003 not used omineca not used 1988–2003 skeena 1988–2003 not used peace1 1988–1995, 1996–2003 1996–2003 1any bull gos august to october 1988–1995, any bull gos august 1996–2003, soft bull september to october 1996–2002, soft10 bull september to october 2003. fig. 2. antler architectures with configurations labelled for yearling or 2-year-old bull moose in british columbia (adapted from the hunting and trapping regulations synopsis in bc ministry of environment 2018–2020). alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 143 that operated from 1988 to 2003, except for 1999 when it was temporarily suspended. a harvest data card envelope was provided to hunters upon purchase of their moose hunting licence or mailed with leh authorization. hunters could record specifics of their hunt including the management unit (mu), date and sex of kill, and the number of antler points on the left and right antlers. instructions were provided for removing a lower incisor tooth that was sealed within the envelope and mailed (postage paid) to the provincial wildlife management agency. hunters submitting a tooth received a “management participant” jacket crest. age was determined from cementum annuli inspections (sergeant and pimlott 1959) performed by regional technicians and contractors. we used the vtrp data to separately calculate the vulnerability of 2 age classes of bull moose (yearling or 2-year-old) to two regulations (s/f or spike-only) in two different ways (potential or realized vulnerability). generally, vulnerability was calculated as the proportion of the harvest of each age class that consisted of bull moose with antlers of the specified configuration. potential vulnerability to each regulation was calculated when all bulls were legal to hunt without antler restrictions, either under a gos or a leh (table 1). potential vulnerability of each age class to the s/f regulation was defined as the proportion of bulls in the harvest with s/f antlers when no antler based restrictions to harvest were in place during a gos or leh. similarly, potential vulnerability of each age class to the spike-only regulation was defined as the proportion of bulls in the harvest with spike-only antlers when no antler based restrictions to harvest were in place during a gos or leh (table 2). realized vulnerability to each regulation was calculated for the okanagan, thompson, omineca, and peace regions when the s/f regulation was in effect with other regulations allowing harvest of bull moose (table 1). for the okanagan, thompson, and omineca regions, realized vulnerability of each age class to the s/f regulation was defined as the proportion of the harvest with s/f antlers when s/f regulations were combined with leh regulations. similarly, realized vulnerability of each age class to the spike-only regulation was defined as the proportion of the harvest with spike-only antlers when s/f regulations were combined with leh regulations (table 2). realized vulnerability to each regulation for the peace region was similarly calculated when the s/f regulation was in effect as part of the soft or soft10 regulations (table 1). when the soft regulations were in effect the total harvest of bulls of each age class consisted of bulls with s/f antlers and bulls with table 2. equations used to calculate potential vulnerability and realized vulnerability to the s/f regulations and to the spike-only regulation for both yearling and 2-year-old bull moose in british columbia. regulation potential vulnerability any bull gos or leh realized vulnerability s/f gos and any bull leh soft s/f = x x sf in age class all age class = + = x x x x x sf in age class sf in age class any bull in age class sf in age class soft bulls in age class spike only = x x spike only in age class all age class = + = x x x x x spike only in age class sf in age class any bull in age class spike only in age class soft bulls in age class 1the soft regulation was used in the peace region from 1996 to 2002 while the soft10 regulation was introduced in 2003. for 2003, the realized vulnerability for the peace region was calculated by replacing the number of soft bulls in the denominator with the number of soft10 bulls. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 144 tri-palm antlers. total harvest of each age class under the soft10 regulations consisted of bulls with s/f or tri-palm or 10 point antlers. since the vtrp listed only total point counts, and not counts of points on the brow palms, we assumed all non-s/f bulls harvested in peace region from 1996 to 2002 during the soft period were tri-palms, since only tri-palms were legal. in 2003 during the soft10 period, no yearling (n = 11) and only one 2-year-old (n = 4) bull had ≥10 points on one antler; consequently, we assumed the yearlings and other 2-year-old bulls were tri-palms. analyses were restricted to potential vulnerability in the kootenay, cariboo, and skeena regions, and to realized vulnerability in the omineca region due to the consistency of regulations in place throughout the vtrp period (table 1). as the s/f regulation was implemented part way through the vtrp period in the okanagan (1993) and thompson (1993) regions, we were able to determine both potential vulnerability (from any-bull seasons from 1988 to 1992) and realized vulnerability (from s/f seasons combined with leh bull seasons from 1993 to 2003). potential vulnerability in the peace region was determined using vtrp records for august to october gos harvests from 1988 to 1995 along with gos harvest records for august only from 1996 to 2003. realized vulnerability for the peace region was determined using vtrp records from september through october in each year from 1996 to 2003 when soft and soft10 regulations were in effect. we combined samples across years to calculate overall or pooled estimates of potential vulnerability and realized vulnerability of each age class to each of the s/f regulation and the spike-only regulation for each region and the entire province. we also combined samples across years to calculate potential vulnerabilities to both regulations for each age class of a. a. andersoni and a. a. shirasi within the kootenay region using the geographic ranges of each as described by stent (2010). similarly, we combined samples across years to calculate potential vulnerabilities for moose in the north-western portion of the skeena region, assuming wildlife management zone (wmz) 6f atlin (bc ministry of forests, lands and natural resource operations 2015) matched the geographic range for a. a. gigas as described by eastman and ritcey (1987). we calculated 95% confidence intervals for each estimate of vulnerability based on the normal approximation to the binomial distribution (zar 1984). we used a 2 × 2 chi-square contingency table with yates continuity correction to test for significant differences in potential vulnerabilities of both age classes to both the s/f and spike-only regulations between moose of different subspecies in two areas of british columbia. first, using pooled samples we compared moose in the kootenay region within the ranges of a. a. andersoni and a.a. shirasi. second, using pooled samples we compared moose in the skeena region within the ranges of a. a. gigas (wmz 6f atlin) with moose (a. a. andersoni) from adjacent ranges (wmz 6e stikine). we combined samples across years to compare potential vulnerabilities of a. a. andersoni among 3 broad geographical zones across british columbia: the southern zone included the kootenay, okanagan, and thompson regions, the central zone included the cariboo region, and the northern zone included the skeena and peace regions. a 3 × 2 chi-squared contingency table was used to test for differences between areas for each combination of age class and regulation. similarly, for each combination of age class and regulation, we used a 6 × 2 alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 145 chi-square contingency table to test for differences between regional estimates of potential vulnerability from the pooled sample for the thompson, kootenay, cariboo, skeena, peace, and okanagan regions. similarly, from the pooled sample for each of the okanagan, thompson, omineca, and peace regions, we used a 4 × 2 chi-square contingency table to test for differences between regional estimates of realized vulnerability for each combination of age class and regulation. we used a 2 × 2 chi-square contingency table with yates continuity correction to test for differences between potential vulnerability and realized vulnerability of each age class from the pooled sample within the okanagan, thompson, and peace regions. we determined annual vulnerability, either potential or realized, by region in a given year for each age class of bull for those years where there were ≥25 bulls of that age class with antler point information in the vtrp sample (table 3). consequently, we only calculated potential vulnerability of yearlings for 39 of 70 combinations of year and region, and realized vulnerability of yearlings for 33 of 42 combinations of year and region. similarly, for 2-year-old bulls we calculated potential vulnerability for 47 of 70 combinations of year and region, and realized vulnerability for 20 of 42 combinations of year and region. central tendency and dispersion of annual vulnerabilities for each combination of regulation/ age/region were described by median vulnerability and range of vulnerabilities, respectively. for each estimate of annual vulnerability, we calculated 95% confidence intervals based on the normal table 3. sample sizes for (a) yearling and (b) 2-year-old bull moose from the voluntary tooth return program (vtrp) for each region/year used in the analyses, british columbia. year (a) yearling bull moose (b) 2-year-old bull moose ko1 ok th ca om sk pe total ko ok th ca om sk pe total 1988 3 2 27 30 202 56 63 383 8 3 43 61 162 99 69 445 1989 3 3 26 18 199 96 94 439 16 4 24 51 138 113 74 420 1990 11 0 34 20 181 87 62 395 14 3 31 51 119 106 44 368 1991 9 0 51 10 158 103 14 345 14 5 58 27 137 95 36 372 1992 14 12 81 267 214 80 16 684 34 9 53 136 151 71 27 481 1993 14 14 39 225 216 82 14 604 19 5 19 151 189 61 29 473 1994 18 9 46 278 250 98 28 727 19 8 16 182 215 77 57 574 1995 15 7 31 261 310 217 62 903 25 3 15 164 219 132 68 627 1996 24 16 50 251 370 209 18 938 17 5 24 181 247 163 27 664 1997 17 16 49 377 357 231 62 1,109 31 5 22 231 276 157 22 744 1999 n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a n/a 2000 11 39 65 281 308 253 11 968 18 14 36 139 190 159 3 559 2001 9 72 115 352 396 330 38 1,312 12 16 34 174 216 185 16 649 2002 11 64 98 316 435 356 24 1,304 8 16 24 230 223 180 8 689 2003 16 87 99 369 418 343 30 1,362 26 26 34 251 252 209 12 810 total 195 376 887 3,434 4,460 2,781 610 12,743 293 126 471 2,266 3,028 1,999 529 8,712 1region names are abbreviated: ko = kootenay, ok = okanagan, th = thompson, ca = cariboo, om = omineca, sk = skeena, and pe = peace. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 146 approximation to the binomial distribution (zar 1984). we used the kruskal-wallis single factor analysis of variance to test for differences in annual potential vulnerability of each age class among the kootenay (only for 2-yearolds), thompson, cariboo, skeena, and peace regions. similarly, we used the kruskal-wallis single factor analysis of variance to test for differences in annual realized vulnerability of yearlings among the okanagan, thompson, omineca, and peace regions. we used the mann-whitney test to identify differences in annual realized vulnerabilities for 2-year-olds between the thompson and omineca regions; the okanagan and peace regions were omitted due to small sample size, hence, vulnerability could not be determined for the peace region and only in a single year for the okanagan region. stata 12 was used for all analyses with significance set at α = 0.05; statistical testing procedures followed zar (1984). regional trends in annual vulnerability of yearlings to each regulation were illustrated in the cariboo (potential vulnerability, 1992–2003), skeena (potential vulnerability, 1988–2003), thompson (realized vulnerability, 1993–2003) and omineca (realized vulnerability, 1988–2003). these regions were chosen as there were ≥10 years of consistent regulations with n ≥ 25 yearlings in the vtrp each year. similarly, regional trends in annual vulnerability of 2-year-old bulls were assessed in the cariboo, skeena, and omineca regions; the thompson region was omitted as only 4 of 10 years had an adequate sample size. long-term trends in annual vulnerabilities were illustrated by fitting a third degree polynomial to the data points. the polynomial was employed as it was more sensitive to change than either a linear or log-linear line (kuzyk 2016, arsenault et al. 2019). results teeth and antler point counts were received through the vtrp program for 39,325 bulls from 1 to 23 years of age. of those, 12,743 (32.4%) were yearling bulls (table 3a) of which 45% had s/f antlers (n = 5,779) with 1,004 spike-only. antler point counts and teeth were received from a total of 8,712 bulls 2-years old (table 3b) of which ~ 9% had s/f antlers (n = 776) with 111 spikeonly. only 2.9% of the remaining 17,825 samples (3 to 23 years old) were s/f. therefore, we used only yearlings and 2-year-old bulls to determine vulnerability. the majority of the vtrp samples came from 3 regions (cariboo, skeena, and omineca) for both yearling (80%) and 2-year-old bulls (84%). provincial vulnerability to the s/f regulation and spike-only regulation across british columbia, potential vulnerability of yearlings to the s/f regulation was 42.7% (n = 7,123; 95% ci = 41.6–43.9%) and realized vulnerability was 48.6% (n = 5,620; 95% ci = 47.3–50.0%). potential vulnerability of 2-year-old bulls was 10% (n = 5,274; 95% ci = 9.2–10.9%) and realized vulnerability was 7.2% (n = 3.439; 95% ci = 6.3–8.0%). vulnerability to the spike-only regulation was lower for both age classes across british columbia. potential vulnerability of yearlings was 7.8% (95% ci = 7.2–8.4%) and realized vulnerability was 8.7% (95% ci = 7.9–9.4%). potential vulnerability of 2-year-olds was 1.3% (95% ci = 0.9–1.6%) and realized vulnerability was 1.3% (95% ci = 0.9–1.7%). regional vulnerability to the s/f regulation based on the pooled sample, the potential vulnerability of yearlings to the s/f regulation ranged from 39.2% (n = 2,781; 95% ci = alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 147 37.4–41.1%) in the skeena region to 59.5% (n = 195; 95/% ci = 52.3–66.7%) in the kootenay region (fig. 3a). potential vulnerability of yearlings was different among the thompson, kootenay, cariboo, skeena, peace, and okanagan regions (χ2 = 678.4, df = 5, p < 0.001). the realized vulnerability of yearlings ranged from 39.5% (n = 4,460; 95% ci = 38.0–40.9%) in the omineca region to 92.5% (n = 133; 95% ci = 88.4–97.4%) in the peace region (fig. 3a). realized vulnerability was different (χ2 = 747.2, df = 3, p < 0.001) among the thompson, omineca, peace, and okanagan regions. realized vulnerability was greater than potential vulnerability in the thompson (χ2 = 58.5, df = 1, p < 0.001), peace (χ2 = 99.3, df = 1, p < 0.001), and okanagan regions (χ2 = 38.3, df = 1, p < 0.001). based on the pooled sample, the potential vulnerability of 2-year-olds to the s/f regulation ranged from 6.8% (n = 1,999; 95% ci = 5.7–7.9%) in the skeena region to 37.5% (n = 24; 95% ci = 15.6–59.4%) in the okanagan region (fig. 3b). potential vulnerability was different among the thompson, kootenay, cariboo, skeena, peace, and okanagan regions (χ2 = 185.2, df = 5, p < 0.001). the realized vulnerability ranged from 5.2% (95% ci = 4.4–6.0%) in the 195 17 219 3434 2781 477 359 668 4460 133 0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% kootenay okanagan thompson cariboo omineca skeena peace v ul ne ra bi lit y region potential realized potential bc realized bc (a) 49% 43% 294 24 209 2266 1999 482 102 262 3028 47 0% 10% 20% 30% 40% 50% 60% 70% kootenay okanagan thompson cariboo omineca skeena peace v ul ne ra bi lit y region potential realized potential bc realized bc (b) 10% 7.2% fig. 3. vulnerability to the s/f regulation from pooled estimates of (a) yearling bull moose and (b) 2-year-old bull moose in british columbia. vertical bars show 95% ci. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 148 omineca region (n = 3,028) to 42.6% (n = 47; 95% ci = 27.2–57.9%) in the peace region (fig. 3b). realized vulnerability was also different among the thompson, omineca, peace, and okanagan regions (χ2 = 198.9, df = 3, p < 0.001) . realized vulnerability was greater than potential vulnerability in the peace region (χ2 = 8.09, df = 1, p = 0.008), but not in the thompson (χ2 = 2.22, df = 1, p = 0.142) or okanagan regions (χ2 = 0.76, df = 1, p = 0.459). regional vulnerability to the spike-only regulation based on the pooled sample, the potential vulnerability of yearlings to the spike-only regulation ranged from 5.9% (n = 17; 95% ci = −0.9–20.4%) in the okanagan region to 28.2% (n = 195; 95% ci = 21.6–34.8%) in the kootenay region (fig. 4a). potential vulnerability was different among the thompson, kootenay, cariboo, skeena, peace, and okanagan regions (χ2 = 152.8, df = 5, p < 0.001). the realized vulnerability ranged from 6% (n = 4,460; 95% ci = 5.3– 6.7%) in the omineca region to 23.3% (n = 133; 95% ci = 15.7–30.9%) in the peace region (fig. 4a). realized vulnerability was different among the thompson, omineca, peace, and okanagan regions (χ2 = 202.3, df = 3, p < 0.001). realized vulnerability was greater than potential vulnerability in the peace region (χ2 = 29.8, df = 1, p < 0.001) but not in the thompson (χ2 = 0.0027, df = 1, p = 0.958) and okanagan (χ2 = 1.96, df = 1, p = 0.215) regions. based on the pooled samples, the potential vulnerability of 2-year-olds to the spikeonly regulation ranged from 0% (n = 24; 95% c = −2.1–2.1%) in the okanagan region to 3.8% (n = 209; 95% ci = 1.0–6.7%) in the thompson region (fig. 4b). potential vulnerability was different among the thompson, kootenay, cariboo, skeena, peace, and okanagan regions (χ2 = 44.8, df = 5, p < 0.001). the realized vulnerability ranged from 1.1% in the omineca region (n = 3.028; 95% ci = 0.7–1.5%) to 6.4% in the peace region (n = 47; 95% ci = −1.7–14.5%) (fig. 4b). realized vulnerability was different among the thompson, omineca, peace, and okanagan regions (χ2 = 13.0, df = 3, p = 0.005). realized vulnerability was not greater than potential vulnerability in the thompson (χ2 = 1.60, df = 1, p = 0.261), peace (χ2 = 0.79, df = 1, p = 0.420), and okanagan regions (χ2 = 0.72, df = 1, p = 1.000). subspecies differences in vulnerability to the s/f regulation and spike-only regulation within the kootenay region, the potential vulnerability of yearling a. a. shirasii bulls was 62.2% (n = 148; 95% ci = 54.0–70.3%) and that of a. a. andersoni was 50% (n = 48; 95% ci = 34.7–65.3%); no difference was found (χ2 = 2.22, df = 1, p = 0.14). in contrast, potential vulnerability to the s/f regulation was higher (χ2 = 4.354, df = 1, p = 0.037) in 2-year-old a. a. shirasii (18.0%; n = 228, 95% ci = 12.8–23.2%) than a. a. andersoni bulls (7.5%; n = 67, 95% 95% ci = 3.8– 14.5%). similarly, potential vulnerability to the spike-only regulation was higher (χ2 = 5.72, df = 1, p = 0.017) in yearling a. a. shirasii (32.4%; n = 228, 95% ci = 24.5–40.3%) than a. a. andersoni bulls (14.6%; n = 67, 95% ci = 3.5–25.7%). potential vulnerability to the spike-only regulation was similar (χ2 = 0.13, df = 1, p = 0.72) in 2-year-old a. a. shirasii (2.2%; n = 48, 95% ci = 0.01–4.3%) and a. a. andersoni bulls (1.5%; n = 67, 95% ci = −0.02–5.2%). the potential vulnerability of yearlings to the s/f regulation within the purported ranges of a. a. gigas (36.2%; n = 260, 95% ci = 30.1–42.2%) was not different alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 149 (χ2 = 0.039, df = 1, p = 0.84) than that of yearlings in adjacent areas with a. a. andersoni (35.3%; n = 260, 95% ci = 29.2– 41.4%). likewise, there were no differences (χ2 = 0.05, df = 1, p = 0.82) between 2-yearold moose within the purported ranges of a. a. gigas (4.1%; n = 269, 95% ci = 1.5–6.6%) and in adjacent areas with a. a. andersoni (4.5%; n = 245, 95% ci = 1.7–7.3%). the potential vulnerability of yearling moose to the spike-only regulation within the purported ranges of a. a. gigas (10.8%; n = 260, 95% ci = 6.8–14.7%) was similar (χ2 = 0.028, df = 1, p = 0.87) to that in adjacent areas with a. a. andersoni (10.3%; n = 252, 95% ci = 6.4–14.3%). the potential vulnerability of 2-year-old bulls was identical (zero) within the purported ranges of a. 195 17 219 3434 2781 477 359 668 4460 133 0% 5% 10% 15% 20% 25% 30% 35% 40% kootenay okanagan thompson cariboo omineca skeena peace v ul ne ra bi lit y region potential realized potential bc realized bc (a) 8.7% 7.8% 294 24 209 2266 1999 482 102 262 3028 47 0% 2% 4% 6% 8% 10% 12% 14% 16% kootenay okanagan thompson cariboo omineca skeena peace v ul ne ra bi lit y region potential realized potential bc realized bc (b) 1.3% 1.3% fig. 4. vulnerability to the spike-only regulation from pooled estimates for (a) yearling bull moose and (b) 2-year-old bull moose in british columbia. vertical bars show 95% ci. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 150 a. gigas (0%; n = 269, 95% ci = −0.2–0.2%) and in adjacent areas with a. a. andersoni (0%; n = 252, 95% ci = −0.2% – 0.2%). geographical differences in vulnerability to the s/f and spike-only regulations the potential vulnerability of a. a. andersoni yearlings to the s/f regulation was different (χ2 = 17.64, df = 2, p = 0.0001) among the southern (kootenay, okanagan, thompson regions) (51.1%; n = 284, 95% ci = 45.1–57.1%), central (cariboo regions) (43.9%; n = 3,434, 95% ci = 42.2–45.6%), and northern (skeena and peace regions) geographical zones (40.2%; n = 2.998, 95% ci = 38.4–42.0%). potential vulnerability of 2-year-old a. a. andersoni bulls was also different (χ2 = 35.94, df = 2, p < 0.0001) among the southern (18.7%; n = 300, 95% ci = 14.1–23.2%), central (8.1%; n = 2.266, 95% ci = 6.9–9.2%), and northern zones (10.8%; n = 2.212, 95% ci = 9.5–12.1%). similar vulnerability patterns were found for yearling and 2-year-old a. a. andersoni bulls to the spike-only regulation. potential vulnerability of yearling a. a. andersoni bulls differed significantly (χ2 = 42.84, df = 2, p < 0.0001) among the southern (16.6%; n = 284, 95% ci = 12.0–21.1%), central (7.3%; n = 3.434, 95% ci = 6.4–8.2%), and northern zones (6.1%; n = 2,998, 95% ci = 5.2–7.0%). potential vulnerability of 2-year-old a. a. andersoni bulls was also different (χ2 = 9.54, df = 2, p = 0.008) among the southern (3.0%; n = 300, 95% ci = 0.9–5.1%), central (0.9%; n = 2,266, 95% ci = 0.5–1.3%), and northern zones (1.4%; n = 2,212, 95% ci = 0.9–1.9%). temporal differences in vulnerability to the s/f regulation the annual potential vulnerability of yearlings across british columbia to the s/f regulation varied from 28 to 90% (median = 43%) (table 4a). the annual median was ~ 10% higher in the thompson region (53%) than in the peace (40%), skeena (41%), and cariboo (44%) regions, but was not different (h = 4.85, df = 3, p = 0.18). the annual vulnerability was highly variable (2–3 x): peace = 33–71%, thompson = 42–81%, skeena = 28–71%, and cariboo = 35–90%. the annual realized vulnerability of yearlings across british columbia to the s/f regulation varied from 26 to 96% (median = 69%) (table 4a). the annual median was lowest in the omineca (40%), and ~2× higher in the thompson (81%), okanagan (84%), and peace (93%) regions, and differed among the regions (h = 25.41, df = 3, p = 0.001). the annual realized vulnerability was less variable than the potential vulnerability: peace = 85–96%, okanagan = 69–89%, thompson = 65–86%, and omineca = 26–51%. vulnerability of yearlings to the s/f regulation varied over time in the cariboo, skeena, thompson, and omineca regions (fig. 5a and b). in the cariboo and skeena regions, potential vulnerability was generally higher (>50%) in the late 1980s and early 1990s than in later years. in contrast, realized vulnerability was generally more consistent in the omineca region over the entire study period (30–50%), and highest in the thompson region (70–90%). the annual potential vulnerability of 2-year-olds across british columbia to the s/f regulation ranged from 0 to 61% (median = 9%) (table 4b). the median vulnerability of 2-year-olds was lowest in the skeena region (5.5%) and highest in the peace region (26%). no differences were found (h = 12.743, df = 4, p = 0.26) among the thompson, kootenay, cariboo, skeena, and peace regions. the largest range of vulnerability was in the peace region (6.8 to 61.4%) and the smallest in the skeena region (1.4–16.2%). alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 151 the annual realized vulnerability of 2-year-olds across british columbia to the s/f regulation ranged from 2.1 to 27% (median = 5.5%) (table 4b). the median vulnerability was ~ 2.5 × lower in the omineca region (median 4.9%, range = 2.1– 9.8%) than the thompson region (median 12.2%, range 2.9% to 14.7%), but not different (h = 2.89, df = 1, p = 0.09). vulnerability of 2-year-olds to the s/f regulation varied over time in the cariboo, skeena, thompson, and omineca regions (fig. 6a and b). annual potential vulnerability in the cariboo and skeena regions was generally higher in the late 1980s and early 1990s (>10%) than later in study period. annual realized vulnerability in the omineca region varied between 2 and 10%. temporal differences in vulnerability to the spike-only regulation annual potential vulnerability for yearlings varied from 1.6 to 37% (median = 7.0%) (table 5a). annual potential vulnerability was different among regions (h = 11.071, df = 3, p = 0.011); the thompson (median = 19.2%) was higher than the cariboo (median = 6.8%), skeena (median = 6.3%), and peace (median = 8.1%) regions which were similar. the annual range of vulnerability was lowest in the cariboo region (5.7–10%) and highest in the thompson region (9.9–37%). annual realized vulnerability for yearlings varied from 3.3 to 37% (median = 12%) (table 5a). the annual range was lowest in the omineca region (3.3–9.7%) and highest in the peace region (15–37%). it differed among regions (h = 23.46, df = 3, p = 0.0001) and was lower in the omineca (median = 6.6%) than in the thompson (median = 16.6%), okanagan (median = 18.7%), and peace regions (median = 27.6%). annual potential vulnerability of yearlings varied over time in the cariboo, skeena, thompson, and omineca regions (fig. 7a), generally from 5 to 10% in the cariboo and skeena regions. annual realized vulnerability (fig. 7b) varied from 3 to 10% in the omineca region and 10 to 25% in the thompson region. annual potential vulnerability of 2-yearolds ranged from 0 to 16.1%, (median = 0.6%) table 4. medians and ranges of annual % vulnerabilities to the s/f regulation by region for (a) yearling bull moose and (b) 2-year-old bull moose in british columbia. vulnerabilities were calculated for each combination of region/year from 1988 to 2003 when there were ≥25 moose/year in the vtrp for that combination of region/year. region (a) yearling bull moose (b) 2-year-old bull moose potential realized potential realized n1 median (range) n median (range) n1 median (range) n median (range) kootenay 0 n/a 5 15 (0–19) n/a okanagan 0 5 84 (69–89) 0 1 27 thompson 5 53 (42–81) 10 81 (65–86) 4 17 (13–45) 4 12 (2.9–15) cariboo 12 44 (35–90) n/a 15 6.6 (2.6–41) n/a omineca n/a 15 40 (26–51) n/a 15 4.9 (2.1–9.8) skeena 15 41 (28–71) n/a 15 5.5 (1.4–16) n/a peace 7 40 (33–71) 3 93 (85–96) 8 26 (6.8–61) 0 total 39 43 (28–90) 33 69 (26–96) 47 9.1 (0–61) 20 5.5 (2.1–27) 1n is the number of years with ≥25 moose/year in the vtrp. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 152 (table 5b). the smallest range of vulnerability was in the kootenay region region (0%, 5 years) and the largest was in the thompson region (0–16.1%, 4 years). although annual potential vulnerability was low overall, it was different among the regions (h = 16.062, df = 4, p = 0.0029), highest in the thompson (median = 2.7%) and peace regions (median = 2.8%), and similar in the kootenay (median = 0%), cariboo (median = 0.7%), and skeena regions (median = 0.6%). annual realized vulnerability of 2-yearolds ranged from 0 to 5.6% (median = 1.3%) (table 5b). annual realized vulnerability was not different among regions (h = 2.403, df = 1, p = 0.12). annual vulnerability of 0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y cariboo skeena poly. (cariboo) poly. (skeena) (a) 0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y omineca thompson poly. (omineca) poly. (thompson) (b) fig. 5. regional trends in (a) potential vulnerability of yearlings to the s/f regulation in the cariboo and skeena regions, and (b) realized vulnerability of yearlings to the s/f regulation in the omineca and thompson regions, british columbia. vertical bars show 95% ci. the trends are illustrated by 3rd degree polynomials fit to the data. alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 153 2-year-olds varied over time in the cariboo, skeena, and omineca regions (fig. 8a and b). the potential (cariboo and skeena) and realized vulnerabilities (omineca) were generally <2% each year. discussion we found that ~43% of yearling bull moose in british columbia were potentially vulnerable to the s/f regulation, whereas only ~8% were vulnerable to the spike-only regulation. as expected, the potential vulnerability of 2-year-old bulls across the province was lower for both regulations, 10% and 1%, respectively. our yearling vulnerabilities were similar to those calculated for yearlings in 1980–1991 (some overlap with this study) from vtrp antler point 0% 10% 20% 30% 40% 50% 60% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y cariboo skeena poly. (cariboo) poly. (skeena) (a) 0% 5% 10% 15% 20% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y omineca poly. (omineca) (b) fig. 6. regional trends in (a) potential vulnerability of 2-year-olds to the s/f regulation for the cariboo and skeena regions, and (b) realized vulnerability of 2-year-olds to the s/f regulation for the omineca region, british columbia. vertical bars show 95% ci. the trends are illustrated by 3rd degree polynomials fit to the data. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 154 count data for the s/f (52%) and spike-only (10%) regulations (hatter 1993). higher vulnerabilities for yearling bulls likely reflect their smaller antler size (gasaway et al. 1987, stewart et al. 2000, bowyer et al. 2002, jensen et al. 2013, andreozzi et al. 2015) and lower point counts (child et al. 2010a) compared to 2-year-olds. potential vulnerabilities for both yearling and 2-year-old bulls differed within the ranges of a. a. shirasii and a. a. andersoni in the kootenay region, and among southern, central, and northern geographical zones within the range of a. a. andersoni. these differences in both age classes across british columbia presumably reflects the differential antler size for the subspecies (gasaway et al. 1987), and/or differences in habitat quality (geist 1987) along the latitudinal breadth of british columbia. based on pooled samples, yearling moose in the kootenay region had the highest potential vulnerability to both regulations (59% s/f and 28% spike-only). higher vulnerability to both regulations was observed in areas with a. a. shirasi (62% s/f and 32% spike-only) than in areas with a. a. andersoni (50% s/f and 15% spike-only). similarly, hatter (1993) found higher vulnerability to the s/f (82%) and spike-only (54%) regulations in yearlings in the eastern portion of the kootenay region where a. a. shirasi were during the 1980–1991 period. in adjacent areas to the west with a. a. andersoni, hatter (1993) found lower vulnerability of yearlings to the s/f (50.8%) and spike-only (9.4%) regulations. we found vulnerability to both regulations was higher for 2-year-old a. a. shirasi (18% s/f and 2.2% spike-only) than 2-year-old a. a. andersoni (7.5% s/f and 1.5% spike-only). stent (2010) found higher incidence of s/f antlers in yearling a. a. shirasii (69%) than yearling a. a. andersoni (48%) in the kootenay region. similarly, in bulls 2 years and older, the incidence of s/f antlers, although low overall, was higher in a. a. shirasii (6.5%) than a. a. andersoni (2.4%). combined, these studies indicate that vulnerability to the s/f regulation was higher for both age classes of a. a. shirasii than a. a. andersoni, and consequently, vulnerability in the kootenay region was higher than in regions without a. a. shirasi. table 5. medians and ranges of annual % vulnerabilities to the spike-only regulation by region for (a) yearling bull moose and (b) 2-year-old bull moose in british columbia. vulnerabilities were calculated for each region/year from 1988 to 2003 when there were ≥25 moose/year in the vtrp for that combination of region/year. region (a) yearling bull moose (b) 2-year-old bull moose potential n1 median (range) realized n median (range) potential n1 median (range) realized n median (range) kootenay 0 n/a 5 0 (0–0) n/a okanagan 0 5 19 (7.7–24) 0 1 3.8 thompson 5 19 (9.9–37) 10 17 (12–24) 4 2.7 (0–16) 4 2.8 (0–5.6) cariboo 12 6.8 (5.7–10) n/a 15 0.7 (0–7.4) n/a omineca n/a 15 6.6 (3.3–9.7) n/a 15 1.1 (0–2.8) skeena 15 6.3 (3.8–18) n/a 15 0.6 (0–2.1) n/a peace 7 8.1 (1.6–18) 3 28 (15–37) 8 2.8 (0–11) 0 total 39 7.0 (1.6–37) 33 12 (3.3–37) 47 0.6 (0–16) 20 1.3 (0–5.6) 1n is the number of years with ≥25 moose/year in the vtrp. alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 155 in the purported range of a. a. gigas in the extreme north-western portion of british columbia (skeena), potential vulnerability to the s/f regulation was 36% for yearlings and 4.1% for 2-year-olds (vtrp samples). in adjacent areas with a. a. andersoni, we found similar potential vulnerability to the s/f regulation for yearlings (35%, n = 252) and 2-year-olds (4.5%, n = 245). schwartz et al. (1992) reported that ~ 50% of a. a. gigas yearling bulls on the kenai peninsula (alaska), but nearly no 2-year-old bulls, were vulnerable to the s/f regulation. although the vulnerabilities of yearlings in north-western british columbia were lower than those in alaska, the vulnerabilities of 2-year-old bulls were intermediate of values in alaska and british columbia. unfortunately, we cannot determine whether moose from the extreme north-western 0% 5% 10% 15% 20% 25% 30% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y cariboo skeena poly. (cariboo) poly. (skeena) (a) 0% 5% 10% 15% 20% 25% 30% 35% 40% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y omineca thompson poly. (omineca) poly. (thompson) (b) fig. 7. regional trends in (a) potential vulnerability of yearlings to the spike-only regulation for the cariboo and skeena regions, and (b) realized vulnerability of yearlings to the spike-only regulation for the omineca and thompson regions, british columbia. vertical bars show 95% ci. the trends are illustrated by 3rd degree polynomials fit to the data. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 156 portion of british columbia were a. a. gigas (eastman and ritcey 1987) or a. a. andersoni (hundtermark et al. 2006, colson 2013, decesare et al. 2020). we found that potential vulnerabilities of yearling and 2-year-olds to the s/f and spike-only regulations were significantly different among regions within the range of a. a. andersoni. in general, potential vulnerabilities were generally higher in the southern (thompson, okanagan, and kootenay) than the central (cariboo) and northern regions (peace and skeena). these differences may reflect the general size increase of moose from south to north (geist 1987). the vulnerabilities reported by hatter (1993) for yearlings from the vtrp in 1980–1991 also varied among regions across the range of a. a. andersoni, although a distinct, south-to-north pattern was not evident. our study differed from the work of hatter (1993) in timing and lengths of 0% 5% 10% 15% 20% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y cariboo skeena poly. (cariboo) poly. (skeena) (a) 0% 1% 2% 3% 4% 5% 6% 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 v ul ne ra bi lit y omineca poly. (omineca) (b) fig. 8. regional trends in (a) potential vulnerability of 2-year-olds to the spike-only regulation for the cariboo and skeena regions, and (b) realized vulnerability of 2-year-olds to the spike-only regulation for the omineca region, british columbia. vertical bars show 95% ci. the trends are illustrated by 3rd degree polynomials fit to the data. alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 157 hunting season, sample sizes, and geographic boundaries. the realized vulnerabilities of both yearling and 2-year-old bulls to the s/f and spike-only regulations were also different among regions within the range of a. a. andersoni. beyond differences in antler structure, the realized vulnerabilities could reflect the variation in availability and harvest rate of >s/f bulls among regions. the realized vulnerability in the omineca region was much lower than in the thompson and okanagan regions, and harvest strategies differed. these 3 regions combined the s/f season with leh gos for bulls >s/f, but the s/f season was longer in the omineca region (bc ministry of water, land and air protection 2003– 2004), and a larger number of leh permits were offered there than in the thompson and okanagan regions (bc ministry of water, land and air protection 2004– 2005). larger numbers of leh permits in the omineca likely lead to larger harvests of bulls with >s/f antlers, and lower realized vulnerabilities. although the timing and length of the hunting seasons were nearly identical in the omineca and peace regions, realized vulnerabilities were higher in the peace region than the omineca. because harvest opportunities for non s/f bulls in the peace region under soft or soft10 regulations were very restrictive (only bull moose with tri-palm or 10-point antlers were available), lower harvest of bulls with >s/f antlers increased realized vulnerabilities in the peace region. the annual estimates of vulnerability within regions in the range of a. a. andersoni indicated that the potential and realized vulnerabilities of yearlings to both the spikefork and spike-only regulations were higher in the late 1980s and late 1990s than in the early 1990s and early 2000s; patterns for 2-year-old bulls were less distinct. this temporal pattern in antler morphology may reflect the influence of browse quantity and quality related to moose density (boertje et al. 2007), climate/weather conditions (murray et al. 2006), and forest practices, including the conversion of moose seasonal ranges to pine and spruce plantations (rea et al. 2017). for example, herbicide applications (connor 1992) and mechanical brush cutting (rea and gillingham 2001) impose nutritional effects (negative and positive) by affecting browse quality and quantity that influence antler development of yearling and 2-year-old bulls (young and boertje 2018). we are currently exploring relationships among moose density, weather/climate, and antler architectures of yearling and 2-yearold moose throughout british columbia to better understand regional differences in vulnerabilities. implementation of the s/f regulation did not result in overharvest of young bulls in the omineca region. hatter and child (1992) found that teen bulls (mostly yearlings and some 2-year-olds) comprised ≥30% of the observed bull population in the omineca region after the hunting season (gos for s/f combined with leh for other bulls) in 6 of 9 years from 1982 to 1990. likewise, bull:cow ratios increased significantly following implementation of the selective harvesting system (shs) on the kenai peninsula, alaska (schwartz et al. 1992). adoption of the s/f regulation in combination with leh (thompson and okanagan regions) or as part of soft10 regulations (peace region) caused regional changes in harvest. by combining moose harvest data (bc data catalogue 2021) with age and antler point data from the vtrp, we determined that the average annual bull harvest declined in the thompson (−50%) and peace (−72%) regions, but remained relatively unchanged (+4%) in the okanagan region following adoption of the s/f regulation. the yearling spike-fork antler regulations – aitken et al. alces vol. 57, 2021 158 proportion of the harvest increased in each region (okanagan 15–53%, thompson 30–47%, peace 21–55%) as the proportion of 2-year-old bulls decreased (okanagan 21–10%, thompson 28–18%, peace 17– 6.6%). interestingly, in each region the average annual harvest of yearlings with s/f antlers increased, while conversely, harvest of yearlings with >s/f antlers and 2-year olds with ≥s/f antlers declined. further, these shifts in harvest resulted in significantly higher realized than potential vulnerability for yearlings in the 3 regions. higher realized than potential vulnerabilities in 2-year-olds also occurred in the peace region, but not in the okanagan or thompson regions. similar changes occurred following adoption of the s/f regulation on the kenai peninsula, alaska where the yearling proportion in the harvest increased from 40% pre-shs to 64% during the shs; at the same time, 2–3 year-old bulls declined in the harvest from 38 to 17% (schwartz et al. 1992). in general, we expected realized vulnerabilities to increase following adoption of the s/f regulation since effort was focused on harvesting bulls with s/f antlers, while the opportunity to harvest larger antlered bulls was reduced by leh or soft regulations. small sample sizes for 2-year-old bulls during the soft period in the peace region and for yearling and 2-year-old bulls in the okanagan region prior to adoption of the s/f regulation precluded analysis and interpretation of regional data. implementation of the s/f regulation in the kootenay region in 2009 did not result in overharvest of young bulls. despite the high potential vulnerability of yearlings (61%) to the s/f regulation, post-hunt surveys during the first 2 years of the gos for s/f bulls showed that s/f bulls accounted for 3–5% of all moose observed, which was similar to s/f proportions pre-gos (szkorupa 2013). furthermore, there was a similar proportion of mature bulls in the harvest preand postgos (szkorupa 2013). in contrast, recent surveys in the thompson (bc ministry of forests, lands and natural resource operations and rural development 2018, 2019) and okanagan (gyug 2013) regions suggest that the harvest of s/f bulls during the gos, combined with the leh bull harvest and an unknown male-biased harvest from unlicensed hunting, reduced bull:cow ratios below population objectives in certain areas. changes to s/f season dates and duration combined with managing hunter access are currently used to better control harvest of yearlings in these regions (c. procter, pers. comm.). hunting regulations and access restrictions are the two management levers most readily available to moose managers throughout british columbia (bc mflnro 2015). further adjustments to hunting season dates, access restrictions, and antler point restrictions may be required to minimize the risk of overharvesting yearling and 2-year-old bulls. however, further access restrictions and reductions in season length could concentrate hunters in time and space, and possibly increase harvest of s/f bulls on open hunting areas. if further restrictions are required, consideration should be given to shifting from s/f to a spike-only regulation since realized vulnerabilities of yearlings to the spike-only regulation are much lower than to the s/f, and 2-year-olds are essentially not vulnerable to a spike-only regulation. the spike-only regulation offers an alternative for controlling bull harvests while continuing to provide moderate levels of hunting opportunity, and has been used since 2013 to reduce harvest of bulls in several areas of alaska (robertia 2013, alaska department of fish and game 2020–2021). our analyses were limited to age and antler point data collected through the vtrp (voluntary) of which the quantity and quality alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 159 of submissions may have influenced certain results. in addition, lack of accuracy in tooth aging (rolandsen et al. 2007) can influence the calculation of vulnerabilities. however, we believe ages determined by our technicians were accurate since this study dealt with yearlings and 2-year-olds, age classes for which determination of age is highly accurate (rolandsen et al. 2007, boertje et al. 2015). the number of submissions to the vtrp varied regionally, but adequate samples were collected from the kootenay, thompson, cariboo, omineca, and skeena regions throughout the study where submissions for all bulls ranged from 43.5% (thompson region) to 57.0% (kootenay region) of the estimated total harvest of bulls in each region (bc ministry of forests, lands, and natural resources operations, unpublished hunter survey). in contrast, sample size was much lower in the okanagan region prior to adoption of the s/f regulation – the low submission rate (23.3%) from the small harvest produced inadequate sample sizes (n = 17 yearlings and 24 2-year-olds). we also considered vtrp records from the peace region to be inadequate given the reporting rate was only 11.0% of the estimated bull harvest during the study period. interestingly, the rates of submission markedly increased in both the thompson (29.7 to 57.4%) and okanagan (23.3 to 66.2%) regions following adoption of the s/f regulation, whereas the rate of submissions for the peace region changed little (12.4 to 9.0%) following adoption of the soft regulation. we assessed the quality of submissions by comparing the similarity of vulnerabilities determined in the omineca region from the vtrp (hatter 1993, this study) with studies based on antler inspections by provincial biologists (hatter and child 1992, child et al. 2010b). we found realized vulnerabilities of yearlings in the omineca region to the s/f and spike-only regulations were 39.5% and 6.0%, respectively, while vulnerabilities of 2-year-olds were 5.2% and 1.1%, respectively. hatter (1993) reported 43.5% vulnerability to the s/f regulation and 7% vulnerability to the spike-only regulation for yearlings (n = 1,490) from 1980 to 1991. in the omineca region (1982–1988), hatter and child (1992) examined antlers submitted by hunters and found realized vulnerability of yearlings (n = 166) was 22.9% to the s/f regulation and 5.4% to the spikeonly regulation; realized vulnerability of 2-year-old bulls (n = 150) was 2.0% to the s/f regulation and 0.3% to the spike-only regulation. child et al. (2010b) examined 1,686 sets of antlers from moose harvested in 1982– 1989 and found that the vulnerability of yearling bulls to the s/f regulation was 46%, and the vulnerability of 2-year-old bulls was 8%. the lower vulnerabilities reported by hatter and child (1992) may reflect their smaller sample sizes. it was speculated that s/f yearlings were underreported in the earlier studies but no corrections were made, whereas child et al. (2010b) corrected their sample for underreporting of bulls with s/f antlers. despite the differences in timing, methods, and sample size, the vulnerabilities calculated from the vtrp by hatter (1993) and examinations of antlers by child et al. (2010b) are reasonably similar to those we report here. we believe that the vtrp records indicate that hunters were accurately reporting counts of antler points in the omineca region, and have no reason to believe otherwise for the other regions. non-representative submissions from hunters could be a potential source of bias in the vtrp data and possibly occurred with submissions from the peace region for yearling and 2-year-old bulls. potential vulnerability of yearlings region (s/f 43.8%, spike-only 6.9%, n = 477) to both spike-fork antler regulations – aitken et al. alces vol. 57, 2021 160 regulations was similar to that in the adjacent in the skeena region (s/f 39.2%, spikeonly 6.4%, n = 2,781), but potential vulnerabilities for 2-year-old bulls were higher in the peace region (s/f 23.7%, spike-only 3.7%, n = 482) than in the skeena region (s/f 6.4%, spike-only 0.7%, n = 1,999). the higher potential vulnerability of 2-year-old bulls in the peace region may reflect submission of a higher than expected number with s/f antlers, or lower than expected number with >s/f antlers compared to the skeena region. realized vulnerability to the s/f regulation in the peace region for yearlings (92.5%, n = 133) and 2-year-olds (42.6%, n = 47) was lower than expected, particularly for 2-year-olds. the vtrp records show 7.5% of yearlings (n = 133) and 57.4% of 2-year-olds (n = 47) had >s/f antlers, which we assume were tri-palm antlers as only bulls with s/f or tri-palm antlers were legal to harvest during the soft period. in contrast, inspections of antlers from the omineca (child et al. 2010a) indicated <1% of yearlings and 5% of 2-yearolds had tri-palm antlers. schwartz et al. (1992) reported almost no yearlings or 2-year-olds had tri-palm antlers in alaska. thus, lower than expected realized vulnerabilities in the peace region could reflect submission of lower than expected numbers with s/f antlers, or higher than expected numbers with >s/f antlers. although the vtrp had no information on the frequency of tri-palm antlers, it did provide information on number of antler points which allowed us to examine data from the peace region from another perspective. the vtrp records showed 0.7% of yearlings (n = 477) and 1.2% of 2-year-olds (n = 482) in the peace region had antlers with ≥10 points during periods when any bull gos regulations were employed. by comparison, in the adjacent skeena region (any bull gos or leh), the vtrp records showed 0.04% of yearlings (n = 2,781) and 1.25% of 2-year-olds (n = 1,999) had antlers with ≥10 points. similarly, in the adjacent omineca region (leh and s/f gos), 0.07% of yearlings (n = 4,460) and 0.40% of 2-yearolds (n = 3,028) had antlers with ≥10 points. this examination suggests similarities in the frequency of antlers with ≥10 points among these adjacent regions during gos or leh seasons, and supports the observations of child et al. (2010a) that antlers with ≥10 points are rare in yearling and 2-year-old bulls. in the omineca region, child et al. (2010a) reported that 46% of yearling bulls had s/f antlers with <1% having tri-palm or 10-point antlers; 8% of 2-year-old bulls had s/f antlers, and 5% had tri-palm antlers with <1% with 10-point antlers. applying these antler architectures (child et al. 2010a) to the peace region during the periods when soft regulations were in use would produce realized vulnerability near 100% for yearlings and ~ 60% for 2-year-olds. likewise, few yearling or 2-year-old bulls in alaska have tri-palm antlers (schwartz et al. 1992), and applying alaskan data to the peace region produced realized vulnerability of nearly 100% for both yearling and 2-yearold bulls. both examples produced higher estimates of realized vulnerability for both yearling and 2-year-old bulls than those calculated in the peace region. we suggest that the unexpected vulnerability of bulls in the peace region is most likely due to the submission of a number of non-representative samples rather than different antler architectures in the region. interestingly, the higher than expected potential vulnerability for 2-year-old bulls could arise from an excessive number of 2-year-old bulls with s/f antlers and/or an insufficient number with >s/f antlers in the vtrp. in contrast, the lower alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 161 than expected value for realized vulnerability for both yearling and 2-year-old bulls could arise from the opposite scenario. unfortunately, the low rates of submission limited our analyses of data from the peace region, particularly during the periods of soft (1996–2002) and soft10 regulations (2003). we believe that the vulnerability of bull moose to the s/f and spike-only regulations likely lies somewhere between the potential and realized vulnerabilities we documented. the potential vulnerabilities we report may be low if hunters could not identify spike or fork antlers, if hunters selected against s/f antlered bulls, or if hunters did not report s/f bulls. bulls with small spike or fork antlers may be under-sampled if they are more difficult to recognize as a bull compared to bulls with larger antler architectures. in addition, hunters may choose to harvest larger antlered bulls in the belief that such a choice would provide more meat or a trophy; lower reporting rates for spike or s/f moose were suspected in earlier studies (hatter and child 1992, hatter 1993, child et al. 2010b). hunter selection and/or under reporting could result in under-sampling of s/f bulls in both age classes, thereby lowering their potential vulnerability. the realized vulnerability estimates we report are high, in some cases approaching 100%. these high values reflect focussed efforts to harvest s/f bulls during a s/f gos, combined with restricted harvest of larger antlered bulls. if s/f bulls are under reported as described above, even the values we report here may be low. the variation of realized vulnerability among the okanagan, thompson, omineca, and peace regions may be partly due to different timing and length of s/f gos between and within the regions. also, part of the variation may be due to different numbers of permits and different lengths and timing of the leh seasons between and within these regions (bc hunting and trapping regulations synopses and leh regulations synopses accessible at 100.gov.bc.ca). realized vulnerability will be higher when more s/f bulls or fewer bulls with >s/f antlers are harvested. larger numbers of s/f bulls may be harvested if longer s/f seasons are offered, or if seasons coincide with periods of increased vulnerability of s/f bulls. conversely, offering smaller numbers of leh tags, or timing leh seasons to match periods of reduced vulnerability of >s/f bulls may result in fewer large antlered bulls in the harvest. obviously, if only s/f bulls were available for harvest, the realized vulnerability would be 100%. in the thompson, omineca, and okanagan regions, the harvest of bulls with ≥3 points was restricted to varying degrees by offering different numbers of leh tags. in the peace region, the soft and soft10 regulations restricted the harvest of non s/f bulls to those with ≥3 points on the brow palm of either antler or those with ≥10 points on either antler. as described above, applying the antler architectures for the omineca region (child et al. 2010a, 2010b) to the peace region would produce realized vulnerabilities to the s/f regulation approaching 100% for both yearlings and 2-year-old bulls. more accurate estimates of potential and realized vulnerabilities of yearling and 2-year-old bulls to the s/f regulation can be obtained by requiring mandatory reporting and/or compulsory inspection of all harvested animals. management implications and recommendations based on our analysis of harvest data and others (hatter and child 1992, hatter 1993, child et al. 2010b), we contend that the s/f regulation applied across british columbia provided adequate protection of yearlings http://gov.bc.ca spike-fork antler regulations – aitken et al. alces vol. 57, 2021 162 and nearly complete protection of 2-year-old bulls under most conditions by focusing harvest on small antlered yearlings and 2-yearolds. we have also shown that the spike-only regulation would have been more conservative than the s/f regulation by focusing harvest on an even smaller proportion of the yearling age class, while providing nearly complete protection for 2-year-old bulls. in the years since our study period (1988–2003), extensive landscape changes have occurred across british columbia (kuzyk and heard 2014). a mountain pine beetle (dendroctonus ponderosae) outbreak followed by salvage logging has resulted in extensive road construction and large harvested areas throughout the interior of british columbia (ritchie 2008), especially in the omineca, cariboo, and skeena regions. forested areas outside the salvage logging zone have also realized increased access resulting from continued resource development, predominantly industrial forest harvesting. an outbreak of spruce bark beetle (dendroctonus rufipennis) since 2014 (province of british columbia 2020) has resulted in further salvage logging operations, mostly in the omineca region. these ongoing landscape changes have likely raised the vulnerability of moose to harvest by increasing hunter access and reducing security cover (kuzyk and heard 2014, rea et al. 2017). the combination of these impacts that occurred largely after our study period, and the geographical and temporal differences illustrated in this study, point to the continued need for monitoring bull harvests where the s/f regulation is utilized. young and boertje (2018) stressed the importance of understanding the relationship between antler architecture and age, particularly in young bulls in which nutritional stress may retard antler development and subsequently influence sex and age classifications. understanding these relationships is particularly important when harvests are regulated by antler architectures. we encourage further investigation of the relationships between antler growth and nutritional status of yearling bulls (e.g., adams and pekins 1995) given the changes to, and relationships among, landscape, climate, and range/forage nutrition in british columbia that influence growth and antler architecture. managers considering adopting the s/f regulation or any other antler point-based regulation, should collect and analyse antler architecture and age data to identify those animals vulnerable to antler point-based regulations. following adoption, data collection should continue to monitor trends in vulnerability and that the regulations target the desired age classes. furthermore, we recommend that managers quantify the proportion of yearling bulls when conducting population inventory and monitoring. this proportion is always important when monitoring population dynamics (adams and pekins 1995, hatter 2011, boertje et al. 2019), but especially so under the s/f regulation that effectively focuses harvest on yearlings, yet may impact recruitment into older age classes. population-level inventories and antler architecture measurements are critically important data that help guide effective conservation and sustainable management of moose populations. acknowledgements we would like to thank the thousands of hunters for their voluntary submissions of teeth and antler point-count data which made this study possible. we also thank the following personnel from the bc ministry of forests, lands, natural resources operations and rural development: k. schurmann and g. kuzyk provided the alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 163 vtrp data set for moose that formed the basis of our analysis. g. kuzyk, c. procter, s. marshall, and a. walker provided constructive comments on earlier drafts which greatly improved this manuscript. we also thank p. pekins (editor), e. bergman (associate editor), and 2 anonymous reviewers for their reviews and comments which resulted in further improvements to this manuscript. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53–59. aitken, d. a., and k. n. child. 1991. relationship between in-utero productivity of moose and population sex ratios: an exploratory analysis. alces 28: 175–187. alaska department of fish and game. 2020–2021. identifying a legal moose. pages 30–31 in alaska hunting regulations. www.adfg.alaska.gov/ static/regulations/wildlife regulations/ pdfs/mooseid.pdf (accessed april 2021). andreozzi, h. a., p. j. pekins, and l. e. kantar. 2015. analysis of age, body weight and antler spread of bull moose harvested in maine, 1980–2009. alces 51: 45–55. arsenault, a. a., a. r. rodgers, and k. whaley. 2019. demographic status of moose populations in the boreal plain ecozone of canada. alces 55: 43–60. bc (british columbia) data catalogue. 2021. hunter sample moose survey estimates 1976 to current. https://catalogue.data.gov.bc.ca/dataset/c49778723 4 7 b 4 0 a 1 8 0 b 3 a 5 b 6 0 c d 4 d d a 8 / resource/ccaeab19-fdd6-4693-ad2513574e14d0dd/download/hs-moosesurvey-estimates-1976-to-current.xlsx (accessed march 2021). bc (british columbia) ministry of environment. 1981–1982. hunting regulations synopsis. www.a100.gov. bc.ca/appsdata/acat/documents/r22560/ bchunt-reg-1982.pdf (accessed june 2020). _____. 1996–1997. hunting and trapping regulations synopsis. www.a100.gov. bc.ca/appsdata/acat/documents/r22560/ bchunt-reg-1997.pdf (accessed june 2020). _____. 2016–2018. hunting and trapping regulations synopsis. tourism publications, ministry of forests, lands and natural resource operations, victoria, british columbia, canada. _____. 2018–2020. hunting and trapping regulations synopsis. www.a100.gov. bc.ca/appsdata/acat/documents/r22560/ bchunt-2018-20.pdf bc (british columbia) ministry of forests, lands and natural resource operations. 2015. provincial framework for moose management in british columbia. british columbia ministry of forests, lands, and natural resource operations, fish and wildlife branch, victoria, british columbia, canada. bc (british columbia) ministry of forests, lands and natural resource operations and rural development. 2018. moose factsheet. www.env.gov. bc.ca/fw/wildlife/management-issues/ docs/2018 moose factsheet.pdf (accessed june 2020). _____. 2019. moose factsheet. www.env. gov.bc.ca/fw/wildlife/management-issues/docs/2019 moose factsheet.pdf (accessed june 2020). bc (british columbia) ministry of water, land and air protection. 2003–2004. hunting and trapping regulations synopsis. www.a100. gov.bc.ca/appsdata/acat/documents/ r 2 2 5 6 0 / b c h u n t r e g 2 0 0 4 . p d f (accessed june 2020). _____. 2004–2005. limited entry hunting regulations synopsis. www.a100.gov. bc.ca/appsdata/acat/documents/r22560/ leh_04_05.pdf (accessed march 2021). http://www.adfg.alaska.gov/static/regulations/wildlife%20regulations/pdfs/mooseid.pdf http://www.adfg.alaska.gov/static/regulations/wildlife%20regulations/pdfs/mooseid.pdf http://www.adfg.alaska.gov/static/regulations/wildlife%20regulations/pdfs/mooseid.pdf https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx https://catalogue.data.gov.bc.ca/dataset/c4977872-347b-40a1-80b3-a5b60cd4dda8/resource/ccaeab19-fdd6-4693-ad25-13574e14d0dd/download/hs-moose-survey-estimates-1976-to-current.xlsx http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1982.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1982.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1982.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1997.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1997.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-1997.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-2018-20.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-2018-20.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-2018-20.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.env.gov.bc.ca/fw/wildlife/management-issues/docs/2018%20moose%20factsheet.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-2004.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-2004.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/bchunt-reg-2004.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/leh_04_05.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/leh_04_05.pdf http://www.a100.gov.bc.ca/appsdata/acat/documents/r22560/leh_04_05.pdf spike-fork antler regulations – aitken et al. alces vol. 57, 2021 164 boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1: 1–10. boertje, r. d., m. m. ellis, and k. a. kellie. 2015. accuracy of moose age determinations from canine and incisor cementum annuli. wildlife society bulletin 39: 383–389. doi: 10.1002/ wsb.537 _____, g. g. frye, and d. d. young, jr. 2019. lifetime sex-specific moose mortality during an intentional population reduction. journal of wildlife management 84: 6–19. doi: 10.1002/ jwmg.21782 _____, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. doi: 10.2193/2006-159 bowyer, r. t., k. m. stewart, b. m. pierce, k. j. hundertmark, and w. c. gasaway. 2002. geographical variation in antler morphology of alaskan moose: putative effects of habitat and genetics. alces 38: 155–162. bubenik, a. b. 1971. social well-being as a special agent of animal sociology. international conference on the behavior of ungulates and its relation to management. 2–5 november 1971, university of calgary, calgary, alberta, canada. _____. 1985. reproductive strategies in cervidae. pages 367–373 in p. f. fennessy and k. r. drew, editors. biology of deer production. royal society of new zealand bulletin 22. _____. 1987. behaviour of moose (alces alces) of north america. swedish wildlife research supplement 1: 333–366. canadian council on ecological areas. 2014. ecozones of canada. http://www. ccea.org/downloads/shapefiles/ca_ e c o z o n e s _ 1 m _ v 5 _ f i n a l _ m a p % 2 0 v20140213.pdf (accessed january 2019). child, k. n. 1983. selective harvest of moose in the omineca: some preliminary results. alces 19: 162–177. _____. 1996. moose harvest management in british columbia: regulation simplification and strategy harmonization. report to the wildlife branch, ministry of environment, lands and parks, victoria, british columbia, canada. _____, and d. aitken. 1989. selective harvest, hunters, and moose in central british columbia. alces 25: 81–97. _____, _____, and r. v. rea. 2010a. morphometry of moose antlers in central british columbia. alces 46: 123–134. _____, _____, _____, and r. a. demarchi. 2010b. potential vulnerability of bull moose in central british columbia to three antler-based hunting regulations. alces 46: 113–121. church, m., and j. m. ryder. 2010. physiography of british columbia. pages 17–45 in. r. g. pike, t. e. redding, r. d. moore, r. d. winkler, and k. d. bladon, editors. compendium of forest hydrology and geomorphology in british columbia. land management handbook 66, volume 1 of 2. ministry of forests and range, forest science program, victoria, british columbia, canada. colson, k. 2013. genetic population structure of moose (alces alces) at multiple spatial scales. m. s. thesis. university of fairbanks, fairbanks, alaska, usa. connor, j. f. 1992. impacts of the herbicide glyphosate on moose browse and moose use of four paired treatedcontrol cutovers near thunder bay, ontario. m.s. thesis. school of forestry, lakehead university, lakehead, ontario, canada. http://www.ccea.org/downloads/shapefiles/ca_ecozones_1m_v5_final_map%20v20140213.pdf http://www.ccea.org/downloads/shapefiles/ca_ecozones_1m_v5_final_map%20v20140213.pdf http://www.ccea.org/downloads/shapefiles/ca_ecozones_1m_v5_final_map%20v20140213.pdf http://www.ccea.org/downloads/shapefiles/ca_ecozones_1m_v5_final_map%20v20140213.pdf alces vol. 57, 2021 spike-fork antler regulations – aitken et al. 165 decesare, n. j., b. v. weckworth, k. l. pilgrim, a. b. d. walker, e. j. bergman, k. e. colson, r. corrigan, r. b. harris, m. hebblewhite, b. r. jesmer, j. r. smith, r. b. tether, t. p. thomas, and m. k. schwartz. 2020. phylogeography of moose in western north america. journal of mammalogy 101: 10–23. doi: 10.1093/jmammal/ gyz163 demarchi, r. a., and c. l. hartwig. 2008. towards an improved moose management strategy for british columbia. habitat conservation trust fund report cat07-0-0325. victoria, british columbia, canada. eastman, d., and r. ritcey. 1987. moose habitat relationships and management in british columbia. swedish wildlife research supplement 1: 101–117. gasaway, w. c., d. j. preston, d. j. reed, and d. j. roby. 1987. comparative antler morphology and size of north american moose. swedish wildlife research supplement 1: 311–325. geist, v. 1987. on the evolution and adaptations of alces. swedish wildlife research supplement 1: 11–23. gyug, l. w. 2013. okanagan moose inventory 2012–2103. report for fish and wildlife branch, british columbia ministry of forests, lands and natural resource operations, penticton, british columbia, canada. hatter, i. w. 1993. yearling moose vulnerability to spike-fork regulation. memo report. wildlife branch, british columbia environment, victoria, british columbia, canada. _____. 1998. moose conservation and harvest management in central and northern british columbia. draft for stakeholder discussion. wildlife branch, british columbia environment, victoria, british columbia, canada. _____. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91–103. _____, and k. n. child. 1992. an evaluation of a spike-fork bull moose antler regulation in central british columbia. proceedings of the 1991 moose harvest workshop, kamloops, british columbia. wildlife branch, british columbia environment, victoria, british columbia, canada. hatter, j. 2011. early ecology and management of the moose in central british columbia. reprint of “the moose of central british columbia”, 1950 ph. d. thesis, state college of washington, pullman, washington, usa. island blue print/printorium bookworks, victoria, british columbia, canada. hundtermark, k. j., r. t. bowyer, g. f. shields, c. c. schwartz, and m. h. smith. 2006. colonization history and taxonomy of moose alces alces in southeastern alaska inferred from mtdna variation. wildlife biology 12:331–338. doi: 10.2981/0909-6396( 2006)12[331: chatom]2.0.co;2 jensen, w. f., j. r. smith, j. j. maskey jr., j. v. mckenzie, and r. e. johnson. 2013. mass, morphology, and growth rates of moose in north dakota. alces 49: 1–15. kuzyk, g. 2016. provincial population and harvest estimates of moose in british columbia. alces 52: 1–11. _____, and d. heard. 2014. research design to determine factors affecting moose population change in british columbia: testing the landscape change hypothesis. wildlife bulletin no. b-126. british columbia ministry of forests, lands, and natural resource operations, victoria, british columbia, canada. macgregor, w. g., and k. n. child. 1981. changes in moose management in british columbia. alces 17: 64–76. spike-fork antler regulations – aitken et al. alces vol. 57, 2021 166 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. doi: 10.2193/0084-0173(2006)166[1:pnd aci]2.0.co;2 poole, k., and c. demars. 2015. review of moose management in the peace region of british columbia. report prepared for british columbia ministry of forests, lands and natural resource operations, fish and wildlife section, fort st. john, british columbia, canada. province of british columbia. 2020. omineca spruce beetle outbreak. www2.gov.bc.ca/gov/content/industry/ forestry/managing-our-forest-resources/ forest-health/forest-pests/bark-beetles/ spruce-beetle/omineac-spruce-beetle (accessed june 2020). rea, r. v., k. n. child, and d. a. aitken. 2017. seeing the forests for their hoofage and stumpage values. british columbia forest professional 24: 18–19. _____, and m. p. gillingham. 2001. the impact of the timing of brush management on the nutritional value of woody browse for moose alces alces. journal of applied ecology 38: 710–719. doi: 10.1046/j.1365-2664.2001.00641.x ritchie, c. 2008. management and challenges of the mountain pine beetle infestation in british columbia. alces 44: 127–135. robertia, j. 2013. 10 illegal moose taken on alaska’s kenai peninsula as hunters adapt to new rules. anchorage daily news, 16 september 2013. www.adn. com (accessed april 2021). rolandsen, c. m., e. j. solberg, m. heim, f. holmstron, m. i. solem, and b. -e. saether. 2007. accuracy and repeatability of moose (alces alces) age as estimated from dental cement layers. european journal of wildlife research. doi: 10.1007/s10344-007-0100-8 schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula, alaska. alces 28: 1–13. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23: 315–321. doi: 10.2307/3796891 stent, p. 2010. kootenay moose composition surveys: winter 2009/10. ministry of environment, environmental stewardship division, cranbrook, british columbia, canada. stewart, k. m., r. t. bowyer, j. g. kie, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36: 77–83. szkorupa, t. 2013. kootenay moose general open season monitoring. final project report. habitat conservation trust foundation, victoria, british columbia, canada. timmermann, h. r. 1987. moose harvest strategies in north america. swedish wildlife research supplement 1: 565–579. _____. 1992. moose sociobiology and implications for harvest. alces 28: 59–77. young, d. d., and r. d. boertje. 2018. agerelated antler characteristics in an intensively managed and nutritionally stressed moose population. alces 54: 37–44. zar, j. h. 1984. biostatistical analysis. prentice-hall, englewood cliffs, new jersey, usa. http://www2.gov.bc.ca/gov/content/industry/forestry/managing-our-forest-resources/forest-health/forest-pests/bark-beetles/spruce-beetle/omineac-spruce-beetle http://www2.gov.bc.ca/gov/content/industry/forestry/managing-our-forest-resources/forest-health/forest-pests/bark-beetles/spruce-beetle/omineac-spruce-beetle http://www2.gov.bc.ca/gov/content/industry/forestry/managing-our-forest-resources/forest-health/forest-pests/bark-beetles/spruce-beetle/omineac-spruce-beetle http://www2.gov.bc.ca/gov/content/industry/forestry/managing-our-forest-resources/forest-health/forest-pests/bark-beetles/spruce-beetle/omineac-spruce-beetle http://www.adn.com http://www.adn.com alces37(1)_1.pdf 23 grooming and rubbing behavior by moose experimentally infested with winter ticks (dermacentor albipictus) edward m. addison1,2, douglas j. h. fraser3, and robert f. mclaughlin4 1wildlife research and development section, ontario ministry of natural resources and forests, 2140 east bank drive, peterborough, ontario, canada k9j 7b8; 2present address: 26 moorecraig road, peterborough, ontario canada k9j 6v7; 3344 wessex lane, nanaimo, bc v9r 6h5; 4r. r. #3, penetanguishene, ontario, canada l0k 1p0 abstract: rates of grooming, rubbing, and shaking were observed of 12 moose (alces alces) infested with 4 levels of winter ticks (dermacentor albipictus) and 5 uninfested control animals. modes of grooming varied among moose and occurred with the tongue, hind feet, head, ears, antlers, teeth, and neck. only moose with ticks used teeth and ears to groom. uninfested moose and moose prior to being infested groomed and rubbed little. grooming was greater immediately following than before infestation, and initial grooming and rubbing were predominant at the sites of infestation. grooming declined in mid-winter months when nymphs develop slowly and increased in late winter and early spring when nymphs and adults actively feed; rubbing only increased in late winter and early spring. cumulative grooming-rubbing was positively correlated with level of tick infestation and hair loss, and negatively correlated with end body weight of female calves only. intense individual bouts of grooming and rubbing during april lasted 13–141 min. over the entire study, cumulative grooming-rubbing in daylight hours for moose with 21,000–42,000 larvae equaled 6–28 d (μ = 12.7), and from february to april moose with 42,000 ticks groomed and rubbed on average ≥5.0–7.5 min/h. the removal of ticks was high (77–96%) indicating that grooming and rubbing are positive behavioral responses with respect to reducing the impact of winter ticks. alces vol. 55: 23–35 (2019) key words: dermacentor albipictus, winter tick, moose, alces alces, grooming, hair loss, fitness grooming and rubbing behavior by captive moose (alces alces) infested with winter ticks (dermacentor albipictus) were previously described by individual actions and temporal changes in behavior during midlate winter (samuel 1991). welch et al. (1991) further documented grooming behavior throughout the infestation period and found that winter ticks stimulated grooming and rubbing in moose and that these behaviors varied with tick phenology. observations of free-ranging moose in elk island national park, alberta documented that calves groom more than adult moose, limited grooming occurs from october to february, extensive grooming occurs in march and april, and a positive correlation exists between grooming activity and hair loss (mooring and samuel 1998a, 1999). by controlling the level of infestation of winter ticks on captive moose, our research augments previous research by testing 6 null hypotheses: 1) grooming and rubbing is similar for moose infested and not infested with winter ticks, 2) grooming and rubbing is similar regardless of the level of infestation, 3) grooming and rubbing is similar before and after infestation, 4) grooming grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 24 and rubbing is similar regardless of the site of infestation, 5) grooming and rubbing is consistent throughout the phenology of winter ticks, and 6) hair loss is independent of the amount of grooming and rubbing. our study further evaluated the relationship between cumulative grooming and rubbing activity and body mass of moose late in the infestation period. methods calf moose used in this study were raised in captivity in 1980 (n = 2), 1981 (n = 4), and 1982 (n = 18) in algonquin provincial park, ontario (45° 33’n, 78° 35’w) as described by addison et al. (1983). each pen (29.6 × 16.5 m) had an observation booth positioned 3 m high at the back of and straddling the midline of each pair of pens. winter tick (tick hereafter) larvae were collected annually during september and october 1980–1982 by dragging flannel sheets over vegetation, then removing and counting the attached larvae. in year 1 (1980 – 2 calves), one calf was infested with 1000 larvae and the other with 8000 larvae on one side only on 11 november. in year 2 (1981 – 4 calves), 2 calves were infested on 25 and 26 september, one with 20,000 larvae applied to the right side of the body, the other with 22,000–23,000 larvae applied to the dorsal and upper lateral surfaces; 2 calves were maintained as controls. in year 3 (1982 – 18 calves), ticks were applied on the dorsal and upper lateral body surfaces from 17 september – 12 october. the initial half was applied by 24 september – 2 october; 30 september was designated as the date of infestation for all animals. three treatment groups were established: 1) 4 calves received 21,000 larvae, 2) 4 calves received 42,000 larvae, and 3) 4 calves served as controls. the remaining 6 calves (reference animals) served to document growth and developmental stages of ticks in the captive herd (addison and mclaughlin 1988). during application of larvae in 1981 and 1982, calves were tethered on a short lead for 30 min to prevent them from grooming and allow larvae to reach the skin. in 1982–1983, a 1.6 % liquid solution of sendran [(2-propan-2-yloxyphenol) n-methylcarbamate] was applied twice to the 4 control animals in november; later, rotenone was applied liberally and rubbed thoroughly into the hair coat twice in december (4 calves), once in january (2 calves), and once in early march (4 calves). in 1982–83, one female and one male of the same treatment group were assigned to each observation pen; the 6 reference animals were maintained together in a larger pen. each year throughout daylight hours a team of observers recorded behavioral data in 2.5 h intervals after which they were replaced by a new team. each day, 6 moose were observed simultaneously: 4 treated animals, 2 each from the 21,000 and 42,000 tick treatments, and 2 controls. the remaining 6 moose, 4 treatments and 2 controls, were observed the following day; observations occurred 6 days/month. behaviors were recorded for 60, 1-min intervals/h. if moose could not be observed in their entirety (e.g., behind trees or the shed), the interval was discarded from analysis. in october 1982, behaviors of infested moose were first recorded 16–20 days post-infestation. observers were as consistent as possible with all trained similarly. within 1 week after each monthly observation period in 1982–83, ticks were removed from the 6 reference calves to measure and classify the developmental stage (see addison and mclaughlin 1988). the 5 stages of development were: 1) october – larvae detaching, molting, and reattaching as unfed nymphs, 2) november to january – 100% nymphs with slow growth (diapause), 3) february – 90% alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 25 feeding nymphs, 10% feeding adults, 4) march – 50% feeding nymphs and 50% feeding adults, and 5) april – 90–100% feeding adults. all studies were approved under an animal care protocol with close scrutiny by a provincial veterinarian who determined the april termination date each year. in most cases experimental moose were euthanized (see addison et al. 1987) for related study of pathology and to collect and count the remaining ticks. three distinct behavioral activities were defined and measured during observations: grooming, rubbing, and shaking. grooming was defined as the use of one body part applied to the same or other body part; antlers, ears, head, teeth, tongue, neck, and hind feet were used to groom. rubbing was defined as the application of pressure against an extraneous object. only vertical structures (fences, trees, shed) were available for rubbing in year 1; slanted poles secured in each pen allowed moose to stand under or over the pole to rub their upper and lower body in years 2 and 3. there were 5 distinct types of shaking: 1) the head, 2) the body, 3) head then body, 4) body then head, and 5) head and body simultaneously. most quantitative comparisons among grooming, rubbing, and shaking were restricted to year 3 when the data spanned the months from september (pre-infestation) to april, and 4 moose were in each treatment. parts of the body to which grooms and rubs were applied are illustrated in figure 1. fig. 1. areas on moose for which grooming and rubbing were recorded (1-perianal; 2-croup; 3-back and loin; 4-withers; 5-neck; 6-sides of head; 7-forehead; 8-dewlap/chin; 9-thigh/upper hind leg; 10-ribs; 11-shoulder/upper foreleg; 12-posterior belly; 13-chest/anterior belly; 14-penis/scrotum; 15-hind feet and lateral hind leg; 16-forefeet and lateral foreleg; 17-muzzle/nose; 18-medial hind leg; 19-medial foreleg; 20-antlers). grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 26 the elapsed time of grooms and rubs was measured with a stopwatch (nearest second). if the beginning or end of an activity was not observed, length was assigned as equal to the monthly average of that activity. monthly rates of grooming for each moose were expressed as the number of distinctly different grooms/h and total minutes groomed/h. the grooming rate in a monthly observation period was calculated from the number of grooms and the average groom time, applying the average groom time to each groom for which length was unknown, summing a revised total groom time, and dividing it by the hours of available observation. rates of rubbing were calculated similarly. the cumulative time of grooming plus rubbing was also calculated on a monthly basis. hair loss data from 1982–1983 were as described for these same moose by mclaughlin and addison (1986). the cumulative volume of hair loss represents loss only on the dorsal and lateral aspects of the body behind the head; that is, the neck, withers, shoulders, back/loin, ribs, croup, thigh, and perianal areas (fig. 1). the number of ticks that survived through to detachment of engorged females was calculated by retrieving detached ticks from the pens plus counting the ticks remaining on each moose at the end of the experiment (see addison et al. 1979). the pens were checked for ticks morning and evening throughout the adult female detachment period; one-half of the collected ticks were attributed to each of the 2 moose in each pen. data analysis the shapiro-wilk normality test was used to assess the normality of data. student’s t-test was used to compare both the rate and duration of uninfested moose pre-infestation (september) and post-infestation (october). anova was used to test for relationships between rate and duration of grooms among treatment groups during pre-infestation in september and post-infestation in october, as well as of rubs post-infestation. spearman rank correlation was used to test the monthly relationship of mean grooms and mean rubs among all 5 treatment groups. spearman rank correlation was used to test relationships between cumulative grooming-rubbing and hair loss, body weight at death, and total number of detached and remaining ticks at experiment end. anova was used to test for differences in the mean rate of shakes by treatment in april. sample size precluded testing for differences among monthly mean grooms plus rubs by treatment group. results after annual infestations, a total of 5006 h of observation occurred during the 3 years: ~50–110 h/moose/month from november to april in year 1, ~35–50 h from october to april in year 2, and ~27–30 h from october to april in year 3. an additional ~17–23 h/moose of observation occurred prior to infestation in year 3. despite our efforts, some ticks transferred to certain control moose and remained on them throughout the experiments. an untreated control in year 2 had plentiful ticks within its hair and was excluded from the analysis; no ticks were observed on the second control and it was not euthanized to digest hair and count ticks. in year 3 when acaricides were applied, 1 control was tick-free and 4, 21, and 84 ticks were collected from the other 3 control animals. modes of grooming a total of 25,429 grooms were observed of which 18,903 (74%) were identified to specific mode and area. licking with the tongue and scratching with the hind feet were the predominant modes (82%). the tongue was used nearly exclusively (81–99%) to groom the perianal, croup, back/loin, ribs alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 27 and thigh/upper leg areas; likewise, the hind feet were used to groom (~93%) the neck, cheek, forehead, and dewlap. ears, antlers, teeth, and the neck curled back on itself with a pinching-like action were lesser modes of grooming (table 1). modes of grooming varied among moose as certain animals groomed more with their teeth, head, ears, neck, and antlers (table 1). control moose groomed proportionally more with their hind feet than infested moose, and 2 controls groomed with their teeth, just 1–2 × each. only the 10 moose infested with ≥20,000 larvae groomed with their teeth and ears (table 1). areas groomed and rubbed moose with >20,000 larvae groomed most to the thigh and upper hind legs, ribs, shoulder, and upper forelegs, back and loin, neck, withers, and forefeet and lateral forearms (table 2). rubs were directed mainly to the head (34.8%), neck (33.9%), shoulder and upper foreleg (8.9%), and withers (7.8%) (n = 2479). location of rubbing was distributed similarly among treatment groups except that the back/loin and withers were not rubbed by control moose. control moose groomed proportionately more to the sides of the head, forehead, muzzle, feet, and lateral aspects of the limbs. grooming and rubbing were highest on the single infested side during the first post-infestation observation period for 2 of 3 moose (table 3). subsequently, <50% of grooms and rubs were on the infestation side at 1–2 months post-infestation and varied table 1. grooming activity by infested (dermacentor albipictus) and uninfested control moose, algonquin park, ontario, 1980–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. activity is described by groom total and proportionally by mode (body part used to groom). na = not applicable, female. moose id infestation (# ticks) # grooms hind feet antlers teeth head ears tongue neck m4 0 351 32.5 na 0.3 6.2 0 61.0 0 m11 0 58 24.6 0 1.3 3.3 0 71.1 0 m12 0 252 28.2 na 0 9.1 0 62.7 0 m17 0 108 28.1 0.8 0 9.9 0 61.2 0 m18 0 137 36.8 na 0 6.6 0 56.6 0 m5 1000 450 23.8 0.8 3.1 1.1 0.2 69.3 1.6 m6 8000 1582 4.4 4.4 1.1 8.9 1.2 79.5 0.4 m1 20,000 1697 17.9 1.7 1.8 15.3 10.4 47.4 6.1 m9 21,000 844 19.5 0.1 1.1 2.1 0.4 76.8 0.1 m10 21,000 2095 14.6 na 1.0 18.5 2.6 63.2 0 m13 21,000 912 13.3 0.3 0.3 2.4 1.6 82.0 0 m14 21,000 1157 15.8 na 1.0 7.4 5.1 70.5 0.2 m3 22,000–23,000 3025 15.7 9.6 0.5 8.5 2.4 63.2 0 m7 42,000 2020 9.4 0.1 0.6 5.6 6.4 77.6 0.1 m8 42,000 995 15.0 na 0.6 9.8 27.5 46.8 0.3 m15 42,000 1967 10.9 0.2 2.8 6.1 1.3 78.4 0.2 m16 42,000 1253 14.5 na 1.4 2.6 2.0 79.6 0 grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 28 between sides thereafter. a higher proportion of grooms was on the infestation side during the first 4 months post-infestation for the third animal (table 3). rates and duration moose groomed approximately 1–3 times/h in september prior to infestation and no differences were found among treatment groups (p = 0.46; table 4). there was no difference in the rate of grooming (p = 0.22) or in the duration of individual grooms (p = 0.14) before and immediately after infestation in control moose, and their grooming activity (min/h) declined gradually from october to april (table 5). the number of grooms/h increased 2.4–5.3 fold immediately after infestation in moose with 21,000 larvae and 6.1–28.2 × in moose with 42,000 larvae (table 4). immediately following infestation, both grooms/h (p = 0.002) and duration of individual grooms (p = 0.002) were different among treatment groups. grooming time (min/h) by moose with ≥21,000 larvae was stable or declined somewhat from november to january, generally increased in february and march, and was highest in april when table 2. distribution of grooming by infested (dermacentor albipictus) and uninfested control moose, algonquin park, ontario, 1981–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. grooming activity per body location is described by average percent. infestation treatment 0 20,000–23,000 42,000 # moose 5 6 4 # grooms 906 9730 6235 grooming location perianal (1) 1.1 3 3.2 croup (2) 2.2 3.2 3.9 back and loin (3) 9.1 11.1 9.7 withers (4) 1.8 9 5 neck (5) 4.3 10.2 5.9 sides of head (6) 4.3 2 1.1 forehead (7) 2.7 2 1.2 dewlap/chin (8) 0.2 0.6 0.5 thigh/upper hind leg (9) 14.8 12.6 15.8 ribs (10) 14.3 11 11.5 shoulder/upper foreleg (11) 3.9 10.5 12 posterior belly (12) 2.8 2.7 5.3 chest/anterior belly (13) 4 2.6 2.7 penis/scrotum (14) 1.4 1.1 2 hind feet/lateral hind leg (15) 9 5.1 6 forefeet/lateral foreleg (16) 11.7 8.1 6.3 muzzle/nose (17) 9.9 2 1.2 hind leg-medial (18) 2.3 2.4 6.1 foreleg-medial (19) 0.3 0.6 0.7 antlers (20) 0 0.3 0.1 alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 29 adult ticks were feeding. moose infested with 1000 and 8000 larvae in november had more variable rates of grooming through the infestation period. mean time spent grooming was perfectly correlated with level of infestation from december through april (rs = 1, p = 0; table 5). rubbing did not follow a similar pattern as grooming. there was no difference in rubs/h among treatment groups (p = 0.4) shortly after infestation. time spent rubbing was generally more variable than grooming, increased throughout the infestation period (table 5), and was perfectly correlated with level of infestation in march and april (rs = 1, p = 0) for moose with ≥8000 larvae. the combined mean rate of grooming and rubbing (min/h) by treatment group was positively related to level of infestation every month during the period of infestation (table 6). the mean maximum time spent grooming and rubbing was 7.65 min/h (13% of time) during april for the most heavily infested animals. intense individual bouts of grooming and rubbing during april lasted 13–141 min with 0.72–1.62 different grooms rubs/min. differences in grooming and rubbing between table 3. monthly percentage of grooms-rubs to the infested side of 3 moose treated with dermacentor albipictus, algonquin park, ontario, 1980–1983. individual infestations were: moose 1 – 20,000 ticks on 25–26 september; moose 2 – 1000 ticks on 11 november; moose 3 – 8000 ticks on 11 november. only moose 3 continued heavier grooming on the infested side for >1–2 months post-infestation (see bolded data). month moose 1 – r moose 2 – r moose 3 – l october 65.5 november 57.8 63.4 72.7 december 42.7 42.6 73.5 january 55.4 51.5 65.3 february 44.3 41.8 61.6 march 42.7 62.5 37.3 april 39.5 55.2 42 table 4. rate and duration of grooms by moose preand post-infestation with dermacentor albipictus, algonquin park, ontario, 1980–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. infestation moose id # grooms/h duration of grooms (sec) pre-infestation (sept) post-infestation (oct) pre-infestation (sept) post-infestation (oct) 0 11 0.9 0.6 11 10 12 1.4 5.7 8 17 17 1.4 0.9 10 7 18 1.3 1.7 8 12 21,000 9 1.7 4.0 8 23 10 3.0 13.2 7 17 13 1.2 4.1 8 19 14 1.2 6.6 6 19 42,000 7 1.8 16.3 6 31 8 1.2 12.0 7 23 15 0.7 18.3 5 22 16 1.8 11.1 8 22 grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 30 control moose and moose with 1000 larvae were minimal (table 5). the animal with 8000 larvae groomed and rubbed more from january to april than that with 1000 larvae (table 5). the amount of hair loss was positively correlated to the estimated cumulative grooming rubbing calculated at 197 d post-infestation in april (rs = 0.86, p = 0.0003). however, the amount of cumulative grooming-rubbing to realize 2% hair loss was highly variable among moose; ~ 34–81 h for 6 moose and 102 and 204 h for 2 others. tick recovery the number of ticks recovered after daily searches of the pens and boiling hides was 0–85 for control moose, 1522–4565 for 4 moose infested with 21,000 larvae, and 2102–8535 for moose infested with 42,000 larvae (table 7). the number recovered was not related to cumulative grooming-rubbing in moose infested with 21,000 (rs = −0.8, p = 0.2) or 42,000 larvae (rs = −0.4, p = 0.6); however, the recovery estimates should be considered minimal because our protocol could not account for detached ticks consumed by ravens (corvus corax) and canada jays (perisoreus canadensis). the minimum proportion of the infestation dose surviving to detachment varied from 5–22% (µ = 12.4%). body weight relationships the 6 males weighed 200–273 kg at the end of the 1982–1983 experiment, but weight of individuals was not related to the table 5. average of grooming and rubbing time (min/h) by infested (dermacentor albipictus) and uninfested control moose, algonquin park, ontario, 1980–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. *indicates the holding pens had no angular pole for rubbing back or belly; the other trials provided a pole. infestation # moose sep oct nov dec jan feb mar apr grooms 0 4 0.23 0.54 0.20 0.34 0.12 0.11 0.11 0.09 1000* 1 0.11 0.34 0.52 0.42 0.16 0.21 8000* 1 0.06 0.47 1.36 1.67 0.66 2.53 21,000 4 0.21 2.21 2.42 2.03 1.76 1.94 1.74 2.91 42,000 4 0.17 5.92 3.78 3.05 2.88 4.02 3.77 5.71 rubs 0 4 0.108 0.022 0.044 0.017 0.004 0 0.004 0.016 1000* 1 0.006 0.021 0.021 0.004 0 0.017 8000* 1 0.06 0.1 0.07 0.06 0.12 0.5 21,000 4 0.032 0.124 0.018 0.091 0.174 0.252 0.185 0.714 42,000 4 0.069 0.024 0.831 0.689 1.489 1.419 1.623 1.939 table 6. average time (min/h) of combined grooming-rubbing by infested (dermacentor albipictus) and uninfested control moose, algonquin park, ontario, 1980–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. infestation # moose sep oct nov dec jan feb mar apr 0 4 0.34 0.56 0.24 0.36 0.12 0.12 0.12 0.11 21,000 4 0.24 2.33 2.44 2.12 1.93 2.19 1.93 3.63 42,000 4 0.24 5.94 4.61 3.74 4.36 5.44 5.39 7.65 alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 31 cumulative time spent grooming and rubbing (rs= −0.6, p = 0.21). at the absolute scale, the 3 heaviest males (224–273 kg) groomed and rubbed the equivalent of 2.5 days (average over daylight hours) throughout the study, whereas the 3 lightest (200–217 kg) groomed and rubbed an average of 17 days (>7 × more). one male infested with 21,000 larvae was an outlier as it groomed minimally compared to other infested animals. conversely, body weight of the 6 females (174–237 kg) was related to the cumulative time spent grooming and rubbing (rs= −0.81, p = 0.05). the 3 heaviest females (216–237 kg) groomed and rubbed an average of only 3.6 days (daylight hours) in contrast with the 3 lightest (174–198 kg) that groomed and rubbed an average of 12 days. shaking of the 3328 shakes recorded, ~80% were of the body or head alone. the rate of shaking (shakes/h) did not consistently increase immediately after infestation, but occurred much more frequently during april for moose infested with ≥21,000 larvae. significant differences (p = 0.04) in the mean rate of shaking (shakes/h) occurred in april: control (μ = 1.23, 0.67–1.68); 21,000 larvae (μ = 2.96, 2.39–3.93); 42,000 larvae (μ = 3.54, 2.39–6.21). discussion winter ticks are associated with increased grooming and rubbing by moose (glines and samuel 1989, samuel 1991, welch et al. 1991, mooring and samuel 1999) and free-ranging wapiti (cervus canadensis) (mooring and samuel 1998b). our controlled study with captive moose is the first to document grooming and rubbing by moose across a specific range of winter tick infestation. specifically, we found a positive relationship between infestation level and the extent of grooming and rubbing by captive calves. the fact that control calves groomed and rubbed less frequently through the table 7. relationship between cumulative grooming-rubbing time and hair loss on infested (dermacentor albipictus) and uninfested control moose, algonquin park, ontario, 1980–1983. control moose were accidentally exposed to a minimal number of ticks of unknown origin. hair loss was calculated as in mclaughlin and addison (1986) at 187 days post-infestation. recovered ticks were those removed from animals euthanized on 18–29 april plus detached ticks collected daily in holding pens. infestation moose id grooming-rubbing (min) hair loss (%) # ticks recovered 42,000 15 19,723 26.1 2102 7 9061 30.4 5016 8 6862 25.7 8535 16 6417 27.9 4721 21,000 10 6310 23.6 1522 14 5659 27.2 1933 9 3255 24.3 2731 13 2800 0.8 4565 0 12 834 4.7 21 17 489 0 0 11 487 0 85 18 426 0 4 grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 32 winter (1982–1983) reflected, in part, the repeated application of acaricides to prevent accidental infestations of larvae. however, we were unsuccessful in completely eliminating all winter ticks and advise that effective treatments be mandatory during translocations of moose to prevent geographic spread of ticks. wild moose presumably harbor light infestations of winter ticks without overt behavioral response and/or hair loss, since grooming and rubbing was generally similar between control moose and those infested with 1000 larvae. the threshold for discernibly increased grooming and rubbing by moose was no more than 8000 larvae. as expected, we found that grooming and rubbing were consistently highest in march-april during the feeding stages of nymphs and adult ticks (samuel 1991, welch et al. 1991, mooring and samuel 1999). grooming began within minutes of infestation and remained elevated for 2–3 weeks during the period of feeding, detachment, and ecdysis by larvae and their subsequent reattachment as nymphs. early grooming by moose was also observed by welch et al. (1991) but not by others working with both captive and free-ranging wild moose (samuel 1991, mooring and samuel 1999). this discrepancy might simply reflect the timing of observations as the latter studies commenced post-larval feeding. further, our consolidated infestation scheme did not reflect the normal, weeks-long infestation period (mooring and samuel 1999) and may have contributed to an enhanced level of irritation and grooming activity in the study moose. unlike grooming, rubbing increased only in march-april when adult ticks were feeding, as reported by others (samuel 1991, mooring and samuel 1999). although patterns were evident, the individual variation in the amount and modes of grooming and rubbing varied extensively within infestation treatments. possible explanations include differences in immunological sensitivity and/ or ticks redistributing themselves on the body. despite variation in behavior and response among moose in this and other studies, it is clear that the level of grooming and rubbing by moose is influenced by the phenology and feeding by winter ticks. for 2 of 3 moose infested on only one side of their body, grooming and rubbing was greater on the infested side for only the first month, post-infestation. this short period is consistent with newly ecdysed nymphs which are mobile in the hair coat and redistributed themselves prior to the november observation period (see addison and mclaughlin 1988). similarly, nymphs ecdysing to adults would be mobile and able to redistribute themselves. a positive relationship between groomingrubbing and hair loss in moose has been reported previously (glines and samuel 1989, mooring and samuel 1999), and our study animals were used by mclaughlin and addison (1986) in their study of hair loss on tick-infested moose. we calculated cumulative grooming-rubbing during daylight hours only because we observed no grooming rubbing activity during limited nocturnal observations in january. however, we believe that this behavior likely occurs continuously during the intense period of irritation in march-april, and consider the cumulative grooming-rubbing estimates as minimums. the combined results of this study and that of mclaughlin and addison (1986) clearly document that infestation level has a strong and direct relationship with both cumulative grooming-rubbing and hair loss during the latter part of the infestation period. importantly, extensive grooming and rubbing occurred earlier in the infestation period prior to measurable hair loss. thus, alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 33 the timing of hair loss surveys (as an indirect measure of infestation level and mortality) is an important consideration and should be conducted late in the infestation period, and always at the same time in the phenology of ticks. our results support the approach of bergeron and pekins (2014) and others who have used annual hair loss surveys in late winter and early spring to estimate tick infestation among areas and years. shaking was considered separate from grooming and rubbing because hair is not removed from the follicles or sheared, hence it had little effect on differential hair loss between infested and uninfested moose. nevertheless, the higher rate of shaking by infested moose was consistent with shaking as a response to ticks. although geist (1963) noted that moose shook frequently following exposure to water, neither rain nor snow explained the difference in shaking between infested and control moose as all treatment groups were observed simultaneously under the same weather conditions. shaking was also included as a form of grooming in previous studies (samuel 1991, mooring and samuel 1999). proposed previously (glines and samuel 1989), here we document for the first time the negative relationship between cumulative grooming-rubbing and tick infestation on moose. many factors contribute to the proportion of ticks removed during infestation including the marked individual variation in modes of grooming that we observed, and that these modes varied in relative effectiveness. in addition, grooming continues where ticks and most hair have been removed because the sheared off, embedded mouthparts of ticks continue to irritate moose. despite some detached ticks burrowing into the duff layer or being removed by ravens and canada jays (addison et al. 1989) before we could retrieve them, the majority (77– 96%) of ticks originally infested on moose and not present in late april likely were removed by grooming and rubbing. blood extraction by ticks is considered a serious threat to the physiological condition of young moose in late winter and early spring (samuel 2004, musante et al. 2007), as are potential thermoregulatory impacts from hair loss (mclaughlin and addison 1986, addison and mclaughlin 2014). most physiological influence from blood loss and thermoregulatory challenges would occur in late winter-early spring after nymphs engorge in february (see addison and mclaughlin 1988). thus, the lower body weight of infested calves in late autumn and early winter (addison et al. 1994) presumably reflects their higher rates of grooming, rubbing, and immunological response. indirectly, our data also support the acute physiological response and high mortality measured in calves harboring on average >45,000 adult ticks in spring (jones et al. 2019), an adult infestation level exceeding our original larval treatments. in summary, moose with higher infestation of winter ticks groomed and rubbed more than other moose over a range of infestation from 8000–42,000 larvae. increased grooming and rubbing lead to increased hair loss in our infested moose, although considerable variation in the amount and modes of grooming and rubbing existed among moose. in general, these results validate use of hair loss surveys among years or areas as potential measures of variation in the infestation level of winter ticks on moose. acknowledgements we thank ontario ministry of natural resources field staff for assisting with capture of moose calves. we greatly appreciate the efforts of s. fraser, s. gadawski, a. jones, s. mcdowell, l. berejikian, k. long, k. paterson, l. smith, d. bouchard, v. ewing, m. jefferson, grooming and rubbing of ticks – addison et al. alces vol. 55, 2019 34 a. macmillan, a. rynard, n. wilson, c. pirie, m. mclaughlin, and p. methner for their strong commitment to some or all of capturing, raising, and husbandry of moose calves. l. smith and s. oram assisted with behavioral observations in 1980–1981; l. smith, n. wilson, and d. kristensen in 1981–1982; and a. rynard, v. ewing, m. jefferson, and a. macmillan in 1982– 1983. m. mclaughlin transformed most data into digital files and r. bramwell coordinated verification of digital data. p. addison assisted in transformation of data sets for summary of results. fieldwork was conducted at the wildlife research station in algonquin park. we thank d. joachim for technical assistance throughout the study and g. smith and c. macinnes for strong administrative support without which the study could not be done. r. addison and p. pekins provided valuable editorial advice. references addison, e. m., n. a. fish, d. j. h. fraser, and t. j. o’shaughnessy. 1987. postmortem temperatures of moose carcasses. alces 23: 285–299. _____, f. j. johnson, and a. fyvie. 1979. dermacentor albipictus on moose (alces alces) in ontario. journal of wildlife diseases 15: 281–284. doi:10.7589/ 00903558-15.2.281 _____, and r. f. mclaughlin. 1988. growth and development of winter tick, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670–678. doi:10.2307/3282188 _____, and _____. 2014. shivering by captive moose infested with winter ticks. alces 50: 87–92. _____, _____, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and not infested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469–1476. doi:10.1139/z94-194 _____, _____, and d. j. h. fraser. 1983. raising moose calves in ontario. alces 19: 246–270. _____, r. d. strickland, and d. j. h. fraser. 1989. gray jays, perisoreus canadensis, and common ravens, corvus corax, as predators of winter ticks, dermacentor albipictus. canadian field-naturalist 103: 406–408. bergeron, d. h., and p. j. pekins. 2014. evaluating the usefulness of three indices for assessing winter tick abundance in northern new hampshire. alces 50: 1–15. geist, v. 1963. on the behaviour of the north american moose (alces alces andersoni peterson 1950) in british columbia. behaviour 20: 377–416. doi:10.1163/156853963x00095 glines, m. v., and w. m. samuel. 1989. effect of dermacentor albipictus (acari:ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. experimental and applied acarology 6: 197–213. doi:10.1007/ bf01193980 jones, h. j., p. j. pekins, l. kantar, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi:10.1139/cjz-2018-0140 mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus) induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. doi:10.7589/ 00903558-22.4.502 mooring, m. s., and w. m. samuel. 1998a. the biological basis of grooming in moose: programmed versus stimulus driven grooming. animal behaviour 56: 1561–1570. doi:10.1006/anbe.1998. 0915 alces vol. 55, 2019 grooming and rubbing of ticks – addison et al. 35 _____, and _____. 1998b. tick-removal grooming by elk (cervus elaphus): testing the principles of the programmed grooming hypothesis. canadian journal of zoology 76: 740–750. doi:10.1139/ z97-247 _____, and _____. 1999. premature loss of winter hair in free-ranging moose (alces alces) infested with winter ticks (dermacentor albipictus) is correlated with grooming rate. canadian journal of zoology 77: 148–156. doi:10.1139/ cjz-77-1-148 musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m. 1991. grooming by moose (alces alces) infested with the winter tick, dermacentor albipictus (acari): a mechanism for premature loss of winter hair. canadian journal of zoology 69: 1255–1260. doi:10.1139/z91-176 welch, d. a., w. m. samuel, and c. j. wilke. 1991. suitability of moose, elk, mule deer, and white-tailed deer as hosts for winter ticks (dermacentor albipictus). canadian journal of zoology 69: 2300–2305. doi:10.1139/z91-323 alces vol. 46, 2010 lankester – impact of meningeal worm on moose 53 understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations murray w. lankester 101-2001 blue jay place, courtenay, british columbia v9n 4a8, canada. abstract: declines in moose (alces alces) populations have occurred repeatedly during the past century on the southern fringe of moose range in central and eastern north america, generally in the same geo-climatic regions. these prolonged declines, occurring over a number of years, are associated with higher than usual numbers of co-habiting white-tailed deer (odocoileus virginianus). numerous proximate causes have been hypothesized but none has gained widespread acceptance among cervid managers. however, current knowledge of the nature of moose declines and the biology of meningeal worm (parelaphostrongylus tenuis) makes this parasite the most credible explanation. other suggested disease-related causes are rejected, including infection with liver flukes (fascioloides magna). there is no clinical evidence that flukes kill moose. as well, this parasite occurs at only moderate prevalence and intensity in some jurisdictions and is completely absent in others where moose declines are known. winter ticks (dermacentor albipictus), on the other hand, do kill moose but usually have a distinctly different and more immediate impact on populations. it is recognized that moose, albeit at lowered density, can persist for extended periods in the presence of p. tenuis-infected deer at moderate densities. however, it is argued here that parelaphostrongylosis can, when conditions favour sustained high deer densities and enhanced gastropod transmission, cause moose numbers to decline to low numbers or to become locally extinct. short, mild winters favour deer population growth in areas previously best suited for moose. wetter and longer snow-free periods increase the numbers and availability of terrestrial gastropod intermediate hosts and the period for parasite transmission. it is hypothesized that these climatic conditions increase rates of meningeal worm transmission to moose and of disease, primarily among younger cohorts. reports of overtly sick moose are common during declines but may not account for the total mortality and morbidity caused by meningeal worm. other means by which the parasite may further lower recruitment and productivity causing slow declines still needs clarification. managers in areas prone to declines should monitor weather trends, deer numbers, and the prevalence of meningeal worm in deer. moose recovery will occur only after deer numbers are decidedly reduced, either by appropriate management or a series of severe winters. alces vol. 46: 53-70 (2010) key words: alces alces, climate, dermacentor albipictus, fascioloides magna, moose die-offs, moose sickness, odocoileus, parelaphostrongylosis, white-tailed deer. the neurological disease of moose (alces alces) known as “moose sickness” has been reported in eastern and central north america for almost 100 years. its cause, the meningeal or brain worm (parelaphostrongylus tenuis) from white-tailed deer (odocoileus virginianus), has been known for almost 50 years (anderson 1964). but the suggestion made long ago by anderson (1965, 1972) that the resulting disease (parelaphostrongylosis) can cause moose populations to decline is still not generally accepted (lankester and samuel 2007). unequivocal, direct evidence identifying p. tenuis as a main cause of moose declines, admittedly, is limited. this probably has encouraged the continued search for alternative explanations including ticks with an attendant bacterium (klebsiella paralyticum), competition with deer, trace element deficiencies, a proposed virus, declining habitat quality, direct and indirect effects of warming climate and heat stress, liver flukes (fascioloides magna), and a variety of other proposed, humanimpact of meningeal worm on moose – lankester alces vol. 46, 2010 54 induced stressors (see review by lankester and samuel 2007). none, however, can be as strongly argued as the meningeal worm hypothesis, albeit relying heavily on indirect evidence including knowledge of the parasite’s biology in gastropods, deer, and moose, its known pathogenicity, and the reoccurring association of sustained high numbers of infected deer with moose declines. the purpose of this paper is to characterize past and recent moose declines and to highlight the biology of the meningeal worm, p. tenuis, considered here to be the most likely cause of periodic, prolonged moose declines. moose declines declines in moose numbers have occurred repeatedly in several eastern north american jurisdictions during the past century (anderson 1972, whitlaw and lankester 1994a, lankester and samuel 2007). declines typically occur, almost imperceptibly, over a number of years and only in the relatively narrow band of mixed coniferous-deciduous forest ecotone extending from the atlantic, around the great lakes basin, and westward toward the edge of the central great plains. pre-1900, much of this area was covered with mature forests and was the southern extent of recent moose and caribou (rangifer tarandus) range. since then, extensive habitat rejuvenation and/or extended periods of shorter, less severe winters have periodically created exceptional conditions allowing sustained high deer densities. otherwise, deer numbers in moose range are kept relatively low (approximately ≤ 5/km2) by periodic harsh winters, regulated hunting, and predators (whitlaw and lankester 1994b). historical moose declines had certain characteristics; they occurred when moose sickness was being reported and deer numbers were unusually high. examining long-term (80 years) historical data beginning in 1912, whitlaw and lankester (1994a) found an inverse relationship between moose and deer numbers with moose declining when deer densities exceeded 5/km2. noted declines were gradual, with moose population estimates going from high to low values over periods of 7-10 years. analyses showed that high deer numbers, high reporting rates of sick moose (# of cases/# of years with reported cases), and declining moose numbers were coincident in at least 5 of 13 identified declines, despite concerns about the precision of historical densities estimates and doubts that reporting rates were representative. as well, a possible time shift may confound such analyses. increased rates of infection in gastropods may lag behind a buildup in deer numbers and because some snails live 2-3 years (lankester and anderson 1968), the impact on moose could continue after deer decline. despite periodic moose declines, it has long been evident that moose can persist with infected deer for extended periods. karns (1967) suspected that moose could be managed in minnesota at acceptable numbers provided deer did not exceed 12/mi2 (4.6/km2). in ontario, whitlaw and lankester (1994b) documented a 10-year period (up to 1992) of relative deer-moose population stability. managers surveyed in 45 game management units indicated that deer and moose were co-habiting, despite occasional reports of sick moose. moose numbers were thought to be increasing in 36 and declining slightly in only 5 management units. nonetheless, moose densities were inversely related to deer densities and were greatest when deer were <4/km2. moose density was also inversely related to the intensity of p. tenuis larvae in deer feces. other jurisdictions similarly reported moose and deer apparently co-existing in close proximity during the 1970s and 1980s, including maine (dunn and morris 1981), northern new brunswick (boer 1992), southern quebec (dumont and crete 1996), northern new york (garner and porter1991), and voyageurs national park in minnesota (gogan et al. 1997). during this period, however, deer in alces vol. 46, 2010 lankester – impact of meningeal worm on moose 55 these locations were not noted to be unusually abundant and were likely constrained over the longer term by weather, habitat, predators, and/or hunting. it is clear, however, that moose co-habiting with infected white-tailed deer are less productive than elsewhere, although comparisons among identical circumstances are seldom possible. for example, on isle royale where moose are constrained primarily by habitat and wolves, density ranges from approximately 1-2/km2 (vucetich and peterson 2008). mean density approaching 3/km2, and often exceeding 4/km2, is realized in newfoundland where moose exist with bears (ursus americanus) and hunting (mclaren and mercer 2005). in comparison, over much of mainland eastern north america, where moose exist with infected white-tailed deer as well as with hunters and predators, moose densities typically are <0.4/ km2 (timmermann et al. 2002). apparently, a strong interplay of limiting factors, effects of meningeal worm included, is already reflected in lower moose densities where they persist in habitats with infected deer. moose sickness has been reported historically in the canadian provinces of new brunswick, nova scotia, quebec, ontario, manitoba, and the northern states of maine, vermont, new hampshire, michigan, and minnesota (lankester 2001). moose essentially disappeared in wisconsin by the early 1900s (parker 1990) and have re-colonized only the highland areas of northern new york (hickey 2008); both states have had relatively high deer densities with regularity since the early 1900s (culhane 2006). moose continue to struggle in the upper peninsula of michigan despite 2 reintroductions (hickie 1944, dodge et al. 2004). of nearly 500 reports of sick moose (literature to 2001), most occurred in nova scotia (28%) during the 1940-1950s, and northern minnesota (20%) during the 1930-1940s (lankester 2001). more recently, during a prolonged period of climate warming beginning in the late 1980s, these same 2 jurisdictions have again experienced moose declines. in nova scotia, moose were declared an endangered species in 2003 (beazley et al. 2006), and moose in northwestern minnesota have all but disappeared (murray et al. 2006). these recent studies in nova scotia and minnesota, as well as in north dakota (maskey 2008), serve to further characterize the nature of moose declines and confirm their occurrence during prolonged periods of high deer densities. they also provide new information lacking in historical accounts, including estimates of moose mortality rates, necropsy findings, and better estimates of concurrent deer densities. further insight into the plight of moose facing rising deer numbers is provided by dodge et al. (2004) in upper michigan and by anecdotal observations on the health of moose populations in northwestern ontario and neighbouring southeastern manitoba. nova scotia the history of deer is well known (patton 1991); after an introduction in the 1890s and likely immigration from neighboring new brunswick, deer numbers increased across the province and a “boom” occurred in the 1940s. by the early 1950s deer were more plentiful than ever, but 3 hard winters in the late 1950s depressed deer numbers and moose responded positively (benson and dodds 1977). for the next 15 years deer numbers were fairly stable, producing an annual harvest of about 20,000 deer. three recognizable moose declines were identified in the period 1930-1975 (whitlaw and lankester 1994a). during these declines, a total of 137 cases of moose neurological disease (6.5-10/yr) were noted; a record number of cases for any single jurisdiction. by 1985 a series of relatively mild winters had produced another increase in deer numbers and a harvest of 63,000 deer. deer declined to one-quarter that number in 1990 and moose numbers may again have responded positively (pulsifer and nette 1995). impact of meningeal worm on moose – lankester alces vol. 46, 2010 56 parker (2003), in an excellent review of the status of moose in mainland nova scotia, suggested that numbers generally showed a somewhat continuous decline beginning as early as the late 1920s, despite a partial hunting closure in 1937 and a total ban in 1981 (excluding cape breton island). agreeing with earlier authors (dodds 1963, telfer 1967, benson and dodds 1977), he noted that a decline in the numbers of moose harvested clearly showed an inverse trend with increasing deer harvests. presently, moose remaining in mainland nova scotia are largely confined to 5 remnant, localized groups, each with 50-600 animals (beazley et al. 2008). they are mostly in elevated areas with early, deep-snow winters and separated from deer, at least during winter. these areas are thought to serve as partial refugia from parelaphostrongylosis (telfer 1967, pulsifer and nette 1995, lankester 2001, beazley et al. 2006). although data are scant, calf survival in some of these groups was low, and parelaphostrongylosis was observed in the population (beazley et al. 2006). these authors, citing benson and dodds (1977), concluded that sufficient circumstantial evidence exists to suggest that a decline of the mainland moose population began following marked increases in deer numbers in the late 1920s-early 1950s, and continued in association with periodic high deer densities. moose were almost extirpated on cape breton island by the late 1800s, but an introduction of animals from alberta in the late 1940s led to a hunted population of about 5000 (beazley et al. 2006, bridgland et al. 2007). moose continue to prosper on the highlands of cape breton island where deer do not winter, but are absent in southeastern lowland areas. beazley et al. (2008) concluded that moose were not excluded from the lowlands by lack of suitable habitat, but possibly by climatic and geological factors, including a role played by meningeal worm in white-tailed deer. much of nova scotia (excluding the highlands of cape breton) has a climate moderated by proximity to the sea and is warmer and wetter than much of southern, continental moose habitat. although periodic hard winters along with coyote predation are known to impact deer numbers (patterson et al. 2002), extended periods of deer population growth and a wetter climate conducive to gastropod transmission may explain why moose numbers outside of refugia have not recovered. it may be significant that much higher prevalence of p. tenuis infection has been found in gastropods in nova scotia and southern new brunswick (2.6% and 2.3%, respectively) than is found in northern new brunswick and more continental parts of eastern canada (<0.5%; see review by lankester 2001). northwestern minnesota a recent moose decline commencing about 1985 in an area including the agassiz national forest wildlife refuge, the red lake wildlife management area, and adjacent agricultural land was studied by eric cox (deceased). data from related aerial moose surveys, and necropsy of radio-collared (females and neonates) and accidentally killed moose from 1995-2000 were later analyzed by murray et al. (2006). the geography of the general area is rather atypical of moose habitat with much within the northern minnesota wetlands ecoregion that is characterized by standing water and permanent wetlands (murray et al. 2006). it comprises a mosaic of private farmlands and protected areas, with marshes dominating natural areas along with lowland areas of willow (salix spp.), aspen (populus spp.), and black spruce (picea mariana), and uplands with aspen and oaks (quercus spp.). following an earlier decline in the 1940s, moose numbers in northwestern minnesota grew from an estimated 1300 animals in 196061 to about 4000 by 1984-85 (murray et al. 2006). in 1971 the population was deemed high enough to support the first hunt in 49 years alces vol. 46, 2010 lankester – impact of meningeal worm on moose 57 (karns 1972). numbers declined slightly after 1985 only to rise again by the early 1990s. thereafter, moose began a step-wise decline that continued to 2000-01 when numbers may have been as low as 500 (murray et al. 2006, peterson and moen 2009); the limited hunting was stopped after 1996. after declining for almost 2 decades, an aerial survey in 2007 estimated the moose population at <100 (lenarz 2007b). deer numbers began to increase in the 1970s, declined somewhat in the mid-1980s, but peaked again in the mid-1990s. peak aerial estimates of about 9/km2 in 1980 and 8/km2 in 1994 were recorded in the agassiz national forest wildlife refuge (peterson and moen 2009). two severe winters in 1995-97 reduced deer numbers dramatically, but ensuing mild winters and restricted hunting allowed recovery by the early 2000s. simulation models in 2007 estimated pre-fawning deer density at 5-14/km2 in the northwest forest hunting zones (lenarz 2007a). initially during the moose decline (19841997), mid-winter calf survival was relatively high (54-94 calves:100 cows; murray et al. 2006). only in subsequent years (1997-2001) did mid-winter calf:cow ratios fall below 50:100; the annual survival rate of male calves was half that of females. pregnancy and twinning rates in the declining population were low (<50%) for most female age classes and correlated with nutritional state (bone marrow fat), and female reproductive senescence was apparent as early as 8 years old. excessive mortality generally was thought to begin among the yearling age class and worsen progressively with age. population age structure was skewed toward younger age classes and relatively few prime breeding-aged animals (4-7 years old) were present (murray et al. 2006). the population experienced a pooled mean annual mortality rate of 24% of which 87% was attributed to pathology associated with parasitic disease and related malnutrition (murray et al. 2006). the prevalence of liver fluke (fascioloides magna) in the population was 89%, and the authors concluded that up to 32% of parasite-related death was due to this parasite. meningeal worm was believed responsible for 5% of the deaths in a radiocollared sub-sample, and 20% in a second submitted by the public. the cause of up to 25% of deaths was classified as “unknown” but thought to be parasite-related. murray et al. (2006) classified moose as having died from liver flukes if severe organ and tissue damage was evident with no other overt cause of death; those without damage but with numerous flukes were considered “probable liver fluke deaths”; 89% percent were infected. given the absence of published evidence linking the death of moose to fluke infection, their analyses may overstate the importance of this parasite in the decline. it should also be noted that karns (1972) sampled much the same area and found a similar prevalence of fluke infection (87%) 28 years earlier when the population had grown to huntable numbers and continued to do so, thereafter. murray et al. (2006) considered death was due to meningeal worm if at least one p. tenuis was found in the cranium and no other suspected cause of death was evident. known difficulties in locating p. tenuis in moose heads (lankester et al. 2007) suggest that deaths during the decline due to this parasite were likely underestimated. overall, murray et al. (2006) concluded that the recent moose decline was primarily the result of climate change (increasing summer and winter temperatures), possibly directly through summer heat stress or indirectly through ecosystem changes including higher deer numbers and parasite-related disease and malnutrition. inappetence and/or inanition may have caused reduced condition since food was not thought lacking and the growing season over the study period had increased by up to 39 days (murray et al. 2006). impact of meningeal worm on moose – lankester alces vol. 46, 2010 58 northeastern minnesota this area differs from northwestern minnesota in it is largely mixed boreal forest with longer colder winters, and has yet to show clear signs of a prolonged decline in moose numbers, although it may be imminent (lenarz et al. 2009, peterson and moen 2009). aerial survey data (1983-2008; m. lenarz, unpublished in peterson and moen 2009) indicate that the moose population has fluctuated between 4,000-7,500 with no evident long-term trends. over the last 12 years of this interval, calf:cow ratios and hunter success rates have both declined (lenarz 2009). limited hunting has occurred annually since 1971 except in 1991. deer densities are lower than in northwestern minnesota; increasing numbers in the late 1980s and 1990s were severely reduced during 2 harsh winters (1995-97). estimates of pre-fawning densities averaged 2.2/km2 in 1996, rose to about 4/ km2 by 2003, and have remained fairly stable since (lenarz 2007b). an ongoing 6-year study of 116 radiocollared adult bulls and cows indicates that pregnancy and calf survival rates are nearly normal although twinning rates and mid-winter calf:cow ratios trend downward; the calf:cow ratio was 0.32 in january 2009 (lenarz et al. 2009). the estimated annual, non-hunter mortality rate was 21% over the 6-year period, or about twice that expected, most in the southern portion of the study area. point estimates for the finite rate of increase averaged 0.86 indicative of a long-term declining population. about 60% of observed adult mortality was classified as “unknown causes”, a good portion believed to be related to the effects of parasites and disease worsened by warming climate and heat stress (lenarz et al. 2009); most deaths occurred in spring and fall. liver fluke infection is much less common in moose in northeastern minnesota than the northwest and was not thought important, whereas meningeal worm infections were detected in some dead moose. annual mortality estimates of radiocollared moose led lenarz (2009) to suggest that a decline in the northeastern moose population was occurring despite a lack of clear evidence from annual aerial surveys. using data from the radio-collared animals, lenarz et al. (2009) found that warming january temperatures above estimated physiological thresholds for moose were inversely correlated with subsequent annual survival. temperatures exceeding thresholds were considered to constitute heat stress by increasing energy expended to stay cool. they hypothesized that parasitic disease was likely a proximate cause of mortality while lowered productivity and increased mortality due to heat stress best explained what was occurring in this population (lenarz et al. 2009). north dakota maskey (2008) completed a hallmark dissertation study (university of north dakota) of factors likely to have been important in a recent moose population decline in the pembina hills area of northeastern north dakota. his findings are of particular interest because of this population’s close proximity to one in neighboring northwestern minnesota believed by murray et al. (2006) to have declined due to heat stress and liver flukes. moose populations here experienced a period of growth beginning in the 1960s. by the mid-1990s, however, moose in the northeastern pembina hills area began a steady, decade-long decline to very low numbers (johnson 2007). in a sample of 32 moose dying, 19% had f. magna infection while 75% of the moose exhibited signs consistent with meningeal worm infection (maskey 2008). as in minnesota, the decline in the pembina hills coincided with an unprecedented increase in the range and abundance of white-tailed deer in northern north dakota (smith et al. 2007), and with a long-term, wet climate cycle beginning around 1993 (todhunter and rundquist 2004 cited in maskey 2008). interestingly, moose numbers in 2 other areas west of the pembina hills alces vol. 46, 2010 lankester – impact of meningeal worm on moose 59 were either steady or increasing, despite local increases in deer numbers. significantly, deer in these areas had lower rates of meningeal worm infection, presumably because of drier conditions less suitable for parasite transmission (maskey 2008). michigan moose were extirpated from lower michigan by the late 1800s and most were gone from the upper peninsula by 1900 (dodge et al. 2004). an attempt to reintroduce them in the 1930s failed. a second establishment of 61 animals brought from ontario in 1985 and 1987 has persisted but grown more slowly than expected. despite meningeal worm initially causing 38% of observed mortality, this protected population, free of most predators, showed some growth when deer were estimated at 4.3/km2 (aho and hendrickson 1989). however, rather than reaching a predicted population of 1000 by the year 2000, aerial surveys conducted in 1996-1997 estimated <150 moose; radio-collaring subsequently occurred in 1995-2001 to investigate the decline (dodge et al. 2004). of 17 deaths of marked moose, more than half were attributed to meningeal worm and liver flukes. survival rates of adults, yearlings, and calves were all similar to those found in other non-hunted, lightly predated populations. they concluded that poor population growth was due to low pregnancy and twinning rates, and no yearling reproduction caused by less than optimal food quality and supply (dodge et al. 2004). state-wide, deer increased dramatically after about 1985 (frawley 2008) and by 2007 the harvest approached 484,000 animals. in the upper peninsula (up), peak populations of the early 1990s were reduced briefly by 2 severe winters (1995-97) but resumed their increase until another severe winter in 200708. nankervis et al. (2000) found 44% of deer heads sampled in the up had meningeal worm and 0.7% of gastropods contained infective larvae. the moose population on deer-free isle royale has behaved quite differently from that in the up in the past 30 years. reduced wolf numbers, following a canine parvovirus outbreak about 1980, allowed moose numbers to increase to about 2500 (3/km2) by the winter of 1995-96 (vucetich and peterson 2008); however, that severe winter and a delayed spring reduced the population of poorly nourished animals to about 500. the population subsequently doubled by 2002 but declined again despite adequate food. since 2002 moose have been subjected to high numbers of ticks and hair loss in spring, possibly due to a decade-long trend of warmer than average summers; however, late-winter mortality of infested moose was not prominent. instead, much mortality since 2002 is attributed to high predation on calves by a disproportionately high wolf population (vucetich and peterson 2008). changes in the numbers of moose on isle royale over the past 50 years fluctuated, sometime abruptly, largely in relation to winter severity, food constraints, and changes in wolf predation. by comparison, numbers of moose confined in 2 other parks (sleeping giant and algonquin provincial parks, ontario) changed slowly but dramatically when co-habiting with increasing numbers of unhunted deer with meningeal worm (whitlaw and lankester 1994b). northwestern ontario and southeastern manitoba the ranges of moose and deer first overlapped in northwestern ontario in 19001920. thereafter, deer increased provincially until the late 1950s (whitlaw and lankester 1994b) when they were as far north as sioux lookout. moose declined during the 1940s and the hunting season was closed briefly in 1949. a series of deep snow winters began in the early1960s, but deer numbers remained robust until 2-3 extremely severe winters in the mid-1970s reduced their populations by at least 50% (probably closer to 80%; b. ranta, impact of meningeal worm on moose – lankester alces vol. 46, 2010 60 personal communication). moose increased during the early part of this period but apparently declined somewhat by the mid-1970s. from 1980-2007, winters were increasingly warmer and shorter, interrupted by only a few hard winters (1995-97 and 2007-09). surveys in 1980-1992 of management units with both moose and deer indicated that moose numbers were stable to slightly increasing over much of the region, and were highest where deer density was estimated at <4/km2 (whitlaw and lankester 1994b). by the mid-1990s, the kenora district had some of the highest moose populations in the province at about 2/km2 on the aulneau peninsula in lake of the woods where only a black powder and archery hunt was allowed (b. ranta, personal communication.). but by the early 2000s, deer were becoming noticeably more abundant throughout much of northwestern ontario. this period of deer increase was characterized by several years of shorter, milder winters and large tracts of forests in the region subject to blow-down and insect damage (spruce budworm [choristoneura fumiferana] produced a bonanza of readily available forage when the dead and dying balsam was colonized with arboreal lichens, principally usnea sp.). deer peaked at high numbers in the kenora area in the winter of 2006-07 and were again as far north as sioux lookout. by 2007, moose were virtually absent on the aulneau peninsula. by the end of the 1990s, moose numbers in parts of northwestern ontario were in decline, especially south of highway 17 between kenora and thunder bay. little information is available on the demography of the current moose population but the trend in numbers was downward and poor calf productivity and recruitment was notable (a. rodgers, ontario ministry of natural resources, personal communication). moose in southeastern manitoba, east of winnipeg and south of lac du bonnet, have historically shared habitat with an infected and widely fluctuating deer population primarily regulated by winter severity. in the early 1970s, deer were numerous in much of the area and sightings of sick moose with meningeal worm (and f. magna) were common (lankester 1974). more recently (1995-2008), deer increased in number and northern distribution after several easy winters. meanwhile, moose declined and virtually disappeared from the extreme southeast corner of the province, south of highway 1, and licensed hunting was discontinued in 2000 (v. crichton, manitoba wildlife & ecosystem protection branch, personal communication). meningeal worm transmission the importance of climate in understanding rates of transmission of meningeal worm to deer and the risk of infection to moose is under appreciated. firstly, winter length and severity are important determinants of deer numbers at the northern limits of their range. secondly, climate in summer (amount of precipitation and length of summer) determines 1) the survival of the parasite outside its host (as first-stage larvae), 2) the survival, abundance, and mobility of gastropods (the intermediate host), and 3) the suitability and length of the snow-free period when transmission is possible (lankester 2001). thusly, climate determines the density of deer and gastropods, and in turn, the rates at which each becomes infected. as emphasized by wasel et al. (2003), the odds of encounter between an infected gastropod and a white-tailed deer (or moose) depend on the density and degree of spatial overlap of both. meningeal worm in gastropods terrestrial gastropods (both snails and slugs) are required intermediate hosts of meningeal worm. they become infected with p. tenuis by encountering first-stage larvae on deer feces or in soil where they are readily washed by rain and melting snow (lankester alces vol. 46, 2010 lankester – impact of meningeal worm on moose 61 2001). although the prevalence of infection in gastropods can be quite low (e.g., 1/1000), transmission is, nonetheless, very efficient. because of the large amounts of vegetation consumed daily by deer, a low prevalence of infection in gastropods can still result in almost all deer becoming infected and at a young age (slomke et al. 1995, lankester and peterson 1996). a variety of gastropod species can serve as sources of infection but one in particular is most important, the small, ubiquitous dark slug deroceras laeve (lankester and anderson 1968, lankester 2001). it is often the most frequently infected species and the one with the most meningeal worm larvae, probably because it is very mobile. it is one of the first land gastropods to become active in spring and one of the last to cease movement in autumn (lankester and peterson 1996). terrestrial gastropod populations respond to wet climate; moisture (precipitation and dew) determines their reproductive success and survival as well as mobility in ground litter and on low vegetation. hawkins et al. (1997, 1998) demonstrated their potential to become more numerous on surface vegetation during wet periods. for every snail sampled on the surface during moderately dry periods, almost 50 more were present in the first 5 cm of underlying duff and soil (hawkins et al. 1998). further, models describing the potential impact of meningeal worm on moose were most sensitive to changes in the intrinsic rate of increase in gastropods (schmitz and nudds 1994). infection of gastropods is affected by climate; for example, snails and slugs in a wet forested area on navy island, ontario, were >6 x (5.1 versus 0.8%) more likely to be infected with p. tenuis than those in a dry upland forest habitat (lankester and anderson 1968). a number of studies report overall prevalence in gastropods in many deer/moose areas to be much less than 1%, yet higher mean values of 2.6-9.0% occur in the canadian maritime provinces and in deer-only areas of the southeastern united states (lankester 2001). these higher values probably reflect a warmer, moister climate with longer periods of gastropod activity, and probably higher deer densities as well (lankester 2001). presumably, climatic conditions that favor gastropod numbers and mobility not only increase their rates of encountering larvae on feces or in soil, but also the likelihood that they will be ingested by cervids. infection of gastropods also reflects the density of infected deer. there was a 4-fold difference (0.04% versus 0.16%) in the prevalence of meningeal worm larvae between summer range in northern minnesota (4 deer/ km2) and winter range where they aggregate (50 deer/km2) for a few months (lankester and peterson 1996). likewise, gastropods on navy island where deer exist year-round at density of about 90/km2, were 100 times more frequently infected (4.2%) than gastropods on summer range in minnesota. unusually high prevalence in gastropods is realized only where deer density is exceptionally high for long periods (lankester 2001). meningeal worm in deer the importance of winter to deer survival is well known (karns 1980) but there is a lack of reliable data needed to test suspected relationships between climate, deer density, and p. tenuis-infection rate. measuring deer density with acceptable precision is notoriously difficult and continues to be attempted by only a few jurisdictions. as well, accurately measuring the prevalence and intensity of meningeal worm in deer requires proper techniques and an understanding of the parasite’s developmental biology (slomke et al. 1995, forrester and lankester 1997). of exclusive interest here is the biology of meningeal worm in deer near the northern limits of their distribution where they periodically share habitat in large numbers with moose. valuable information comes from a study of heads and feces from road-kill deer collected in impact of meningeal worm on moose – lankester alces vol. 46, 2010 62 midto late winter within a deer wintering area near grand marais, minnesota (slomke et al. 1995). deer density (pre-fawning) at that time was estimated conservatively at about 2/km2 (lenarz 1993) and the moose population was deemed fairly stable. maturing worms were detected in the heads of a large percentage of fawns in early winter (december-february) indicating that deer are exposed to the parasite early in life (slomke et al. 1995); >90% had encountered the parasite by their second autumn. infected deer retain the same live worms in their cranium for life, young animals pass twice as many larvae as older animals, and larval output is highest in spring. in deer, the parasite requires 3-4 months before first-stage larvae occur in feces. traditionally, deer heads used for assessing infection rates are collected during the autumn deer hunting season, but this is not the ideal time to obtain an accurate estimate of infection rates because a large portion of the harvest could be fawns and yearlings with recent infections. because some worms will still be developing inside the spinal cord and be difficult to detect, heads of fawns in hunter-killed samples should either be excluded or analyzed separately. deer feces to be examined for larvae are best collected in late winter when most viable infections will be patent, and feces can be collected off snow and not be contaminated with soil nematodes. only low numbers of adult worms are found in the heads of deer living in moose range (e.g., x = 3.2, range = 1-13; slomke et al. 1995) because infection rates in gastropods are low and many have only a single larva. and, shortly after becoming infected, deer develop an immune protection against further infection even if exposure is to only a single infective larva (slomke et al. 1995). only those worms acquired within a few months of the first exposure are able to reach the cranium and mature before protection develops against further infection. hence, fawns infected in late summer or autumn are likely to be immune to further infection by the time snow melts the following spring when gastropods resume activity. this so-called concomitant immunity no doubt protects deer from acquiring too many worms that will eventually reside close to the brain. many deer initially pick up only a single infective larva (or 2 of the same sex) and do not encounter another before the immune response becomes protective. as a result, up to one-third of all infected deer may never pass larvae because a mature worm of the opposite gender cannot gain foothold (slomke et al. 1995). hence, the rates of infection in deer, mean numbers of mature worms, and the proportion of sterile infections in a population are determined by the rate of initial acquisition of infective larvae by fawns and yearlings. therefore, year-to-year changes in infection rates and the factors responsible can only be detected by examining successive fawn cohorts (peterson et al. 1996). or, if the protection against re-infection is as strong as believed, past differences in annual transmission rates might be revealed using cohort studies of worm numbers in adult deer. considerable data support the prediction that the prevalence of meningeal worm in deer increases with increased deer density (karns 1967, behrend and witter 1968, slomke et al. 1995, peterson et al. 1996, wasel et al. 2003, maskey 2008). the mean number of larvae in feces of fawns after their first winter (intensity of infection) is also positively correlated with deer density (whitlaw and lankester 1994b, slomke et al. 1995, peterson et al. 1996). the importance of climate in meningeal worm transmission deserves more emphasis. the prevalence in deer is positively correlated with summer and autumn precipitation that presumably favours gastropod infection rates and availability (gilbert 1973, brown 1983, bogaczyk 1990, bogaczyk et al. 1993, peterson et al. 1996, wasel et al. 2003, maskey 2008). using 9 years of fecal data, peterson et al. (1996) found that the prevalence in alces vol. 46, 2010 lankester – impact of meningeal worm on moose 63 10-month-old fawns correlated positively with the number of days in autumn when deer could still access ground vegetation. low rainfall, and possibly lower deer density, probably determine the westernmost limit of meningeal worm and have prevented its spread to vulnerable cervid communities in western canada (wasel et al. 2003). unfavourable conditions for transmission in western manitoba and central north dakota resulted in low rates of infection (<20% with worms in the head). of equal importance was the remarkably high proportion (44%) of deer that had only a single worm in the cranium; this is an expected consequence of low transmission rates. a large number of deer with only a single worm are effectively immunized and thought never to develop patent infections (slomke et al. 1995). meningeal worm in moose the meningeal worm rarely matures to produce larvae in moose, hence, infection depends on the presence of infected deer. infected moose show behavioral and neuromotor disease of varying severity (lankester et al. 2007). some animals show almost imperceptible or intermittent motor deficiencies with slight toe-dragging or stumbling. severely affected animals can exhibit profound weakness of the hind quarters and may be unable to rise. others that become laterally recumbent and flail their legs probably die. some animals exhibit chronic debility including loss of fear of humans and weight loss, and may remain within a restricted area for an extended period; those escaping predation may recover eventually. overtly sick moose show a typical suite of neurological signs, but the severity of signs manifest is not necessarily reflective of the number of worms found on examination. moose usually acquire <10 worms (x = 2.5 ± 0.6; lankester et al. 2007). those with >3-4 in the cranium invariably show severe neurological impairment, but severe signs are observed in moose with only a single worm, and sometimes none. failure to locate worms may relate to the body site and/or killed worms, or host inflammatory response (i.e., meningitis and perineuritis). in addition, it is suspected that some infected animals may experience unobservable, physiological or behavioural abnormalities. the difficulty of finding adult worms at necropsy often makes a definitive diagnosis elusive. for example, in a sample of 34 moose showing typical clinical signs of parelaphostrongylosis, adult meningeal worms were found in only 44% (lankester et al. 2007). moose of all ages are affected but younger animals certainly predominate. the mean age of animals showing signs in the above study was 3.6 years (range = 0.6-14). those with worms detectable in the cranium were younger (1.8 ± 0.5 years) than those with signs but without worms in the head (5.2 ± 1.2 years). females made up 76% of sick animals >3 years old. the sexes were more balanced (10 male:7 female) among younger animals (<3 years). there was no indication that females acquire fewer worms than males, but results suggest that over time, worms are overcome by moose and that females may survive infection longer. results were similar in a smaller sample of 10 sick moose from southeastern manitoba (lankester 1974). only one experimental study has investigated the response of moose to low numbers of infective larvae, similar to that encountered in nature (lankester 2002). all 5 moose (5-9.5 months) given 3-10 larvae showed abnormal locomotory signs after 20-28 days. symptons became progressively more pronounced with front limb lameness and hindquarter weakness. however, after 77-130 days, marked signs persisted in only one animal, were diminished markedly in 2, and disappeared in 2. two animals were challenged with 15 larvae (199 days after initial infection) with no noticeable effects; at necropsy, one had a single worm believed to be from the challenge. results impact of meningeal worm on moose – lankester alces vol. 46, 2010 64 indicate that all young moose ingesting a few larvae show some impairment, even if intermittent and temporary. worms in the cranium were overcome and some animals recovered, at least for the short-term, and such animals appeared to have a degree of protection that may reduce the impact of subsequent infections. how long-lasting such protection might be in naturally infected moose is unknown. an enzyme-linked immunosorbent assay (elisa) developed using excretory products of p. tenuis larvae was positive for all of the animals infected above (ogunremi et al. 2002). the level of antibody response was strongly correlated with the inoculating dose, and levels remained high in all animals that still had worms in the cranium when euthanized. but, titers diminished in 2 and were undetectable in a third animal that had overcome worms by the time of necropsy (186 days after infection). antibodies became elevated after challenge infection. more serological study and accompanying, competent necropsy of sick moose is needed to fully appreciate the utility of the elisa in measuring and monitoring the impact of meningeal worm on moose populations. other potential disease-causing parasites of several parasites known in moose (lankester and samuel 2007), only 2 (liver fluke and winter tick [dermacentor albipictus]) have come to be associated with dead or sick animals during moose population declines. for several reasons, neither is likely to be the cause of gradual moose declines discussed here. liver fluke the giant liver fluke, or deer fluke, has a spotty and limited distribution across moose range. it occurs in the great lakes basin including central and northwestern ontario, southeastern manitoba, northern minnesota, the upper peninsula of michigan, quebec, and a few locations in western canada (pybus 2001). the parasite has a water-based life cycle and is found only where aquatic snails of a particular genus (lymnaeus spp.) occur. white-tailed deer and elk (cervus elaphus) are its principal hosts and the source of infection to moose. it is acquired by eating an intermediate larval stage that emerges from snails and encysts on aquatic vegetation. moose are a dead-end host and do not propagate the parasite. their risk of infection presumably is directly related to the density of co-habiting infected deer and to the densities of suitable aquatic snails. there is no clinical evidence that flukes kill or debilitate moose. however, the considerable tissue pathology seen in some heavily infected livers has led to the suggestion that flukes may cause mortality when moose are stressed (see reviews by pybus 2001, lankester and samuel 2007). in moose, like in domestic cattle, migrating flukes cause bloody tracts, extensive fibrosis and compensatory liver tissue hypertrophy; infected livers may be more than double normal size. moose experimentally infected with f. magna and observed for >12 months showed no outward signs of disease (m. lankester and w. foreyt, unpublished). two calves and a yearling were given 50-225 metacercariae and observed for up to 16 months. the liver of animals infected as calves were enlarged and contained bloody tracks, extensive fibrosis, and walled capsules; 1 and 11 flukes were recovered. that of the yearling had 3 large, thick-walled cysts but no flukes were found at necropsy. growth, weight gain, and behaviour of all 3 were similar to uninfected, farm-reared moose. in most hosts studied (deer, elk, caribou and cattle [bovus spp.]), the prevalence of fluke infection increases with host age and plateaus in older age classes (pybus 2001); young-of-the-year are rarely infected. mean intensity generally is similar within each infected age class suggesting that an acquired, alces vol. 46, 2010 lankester – impact of meningeal worm on moose 65 immunological resistance to further infection develops. as well, flukes have a highly aggregated distribution in normal cervid hosts. most have only a few flukes, while a small number of animals may carry large numbers. in dead end hosts like moose and cattle, long-standing chronic infections are characterized by large paste-filled, thick-walled, closed cysts with few recoverable live flukes (lankester 1974). flukes are not considered important to the health of cattle despite infections that resemble those in moose (wobeser et al. 1985). liver fluke infections were noted during earlier moose declines in minnesota (fenstermacher and olsen 1942), and flukes were found in livers of moose taken in the first hunt in many years; 17 and 87% were infected from the northeastern and northwestern regions, respectively (karns 1972). moose populations in minnesota continued to grow even with hunting, and more recently, as moose in northwestern minnesota experienced a marked decline (from about 1995-2005), the prevalence of fluke infection was essentially unchanged at 89% (murray et al. 2006). flukes were common in the wetter habitat of northwestern minnesota during population growth and decline. flukes are less common in northeastern minnesota and not thought to play a primary role in any decline there (lenarz et al., unpublished data). in a study in adjacent northeastern north dakota, flukes were found in only18% of sick moose and were not considered responsible for declining moose numbers (maskey 2008). in nova scotia, with a history of moose declines and where moose have been declared “endangered”, liver fluke has never been present. in total, these examples and data indicate that the deer liver fluke is not a significant factor in moose declines. winter tick the winter tick can be found on virtually all moose wherever they occur in north america, with the exception of the island of newfoundland and north of approximately 60° n latitude (samuel 2004). disease results when moose acquire unusually large numbers and are subjected to a long, severe winter and possibly diminished food quality or availability. infested animals exhibit increased grooming and restlessness due to the irritation of blood-sucking ticks, lose weight, and experience extensive hair loss in late winter, all of which can contribute to death. die-offs attributed to moose ticks have particular characteristics. relatively large numbers of infested moose, many with premature hair loss, are found dead in late winter/ spring. such conspicuous mortality usually occurs after a prolonged cold winter, and is often reported in unusually high density populations such as those protected in parks. calves and yearlings are thought to be the most severely affected but older animals may not be spared. die-offs often are widespread and rapid but short-lived (samuel et al. 2000), continuing for only 1-2 consecutive springs. such epizootics are independent of deer density. moderate tick infestations presumably have an on-going, sub-clinical impact on moose, but conspicuous die-offs most often occur when moose are dense, tick numbers are high, and nutritionally stressed animals have experienced a severe winter. warmer, shorter winters result in increased survival of adult females dropping off moose onto litter in march and april and increased tick populations on moose the following winter (drew and samuel 1986, samuel 2007). warm, snow-free octobers increase the survival of seed ticks and their likelihood of infesting moose (samuel and welch 1991). tick numbers generally correlate positively with moose density (samuel 2007). over a 12-year study in elk island national park, alberta, there was a 1-year lag in tick numbers relative to moose numbers. die-offs occurred when moose approached 3/km2 and mean numbers of ticks on moose reached 50-60,000. since ticks generally have their impact of meningeal worm on moose – lankester alces vol. 46, 2010 66 greatest impact on individual moose when populations are high, this parasite is unlikely to be the cause of prolonged, relentless declines in moose populations. conclusions pronounced declines in moose numbers on the southern limits of their distribution in eastern north america have occurred repeatedly over the past century. the most conspicuous reductions have occurred generally in the same geo-climatic regions, the milder and moister parts of eastern canada (ns) and a central mixedwood, wetlands area, west of lake superior (comprising parts of northwestern minnesota, northeastern north dakota, southeastern manitoba, and northwestern ontario). the most recent declines were accompanied by a warmer and wetter period. earlier declines (1930s-1950s) followed large-scale successional renewal of harvested mature forests, but also coincided with a lesser-known warm period (le mouel et al. 2008). it is concluded here that extended periods of warmer, and possibly wetter climate provide conditions conducive to moose declines resulting from increased winter survival of white-tailed deer and increased transmissibility of diseasecausing meningeal worm. reports of overtly sick moose are common during declines, but the number of recognizably sick animals may not represent the total mortality and morbidity caused by meningeal worm. additional means by which the parasite may lower recruitment and productivity causing slow, insidious declines still require clarification. managers in areas prone to declines can lessen impending harm to moose by monitoring weather trends, deer numbers, and the prevalence of meningeal worm in deer. if retention of high quality moose hunting is desired, deer numbers can be reduced where indicated by adjusting deer harvest and banning supplemental winter feeding that artificially elevates deer populations relative to habitat. acknowledgements i gratefully acknowledge the hard work and inspiration of several graduate students and close colleagues in pursuing a better understanding of an important and re-occurring wildlife problem. references aho, r. w., and j. hendrickson. 1989. reproduction and mortality of moose translocated from ontario to michigan. alces 25: 75-80. anderson, r. c. 1964. neurologic disease in moose infected experimentally with pneumostrongylus tenuis from white-tailed deer. veterinary pathology 1: 289-322. _____. 1965. an examination of wild moose exhibiting neurologic signs, in ontario. canadian journal of zoology 43: 635639. _____. 1972. the ecological relationships of meningeal worm and native cervids in north america. journal of wildlife diseases 8: 304-310. beazley, k., m. ball, l. isaacman. s. mcburney, p. wilson, and t. nette. 2006. complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada. alces 42: 89-109. _____, h. kwan, and t. nette. 2008. an examination of the absence of established moose (alces alces) populations in southeastern cape breton island, nova scotia. alces 44: 81-100. behrend, d. f., and j. f.witter. 1968. pneumostrongylus tenuis in white-tailed deer in maine. journal of wildlife management 32: 963-966. benson, d. a., and d. g. dodds. 1977. deer of nova scotia. nova scotia department of lands and forests, halifax, nova scotia, canada. boer, a. h. 1992. history of moose in new brunswick. alces, supplement 1: 16-21. alces vol. 46, 2010 lankester – impact of meningeal worm on moose 67 bogaczyk, b. 1990. a survey of metastrongyloid parasites in maine cervids. m.sc. thesis, university of maine, orono, maine, usa. _____, w. b. krohn, and h. c. gibbs. 1993. factors affecting parelaphostrongylus tenuis in white-tailed deer (odocoileus virginianus) from maine. journal of wildlife disease 29: 266-272. bridgland, j., t. nette, c. dennis, and d. quann. 2007. moose on cape breton island, nova scotia: 20th century demographics and emerging issues in the 21st century. alces 43: 111-121. brown, j. e. 1983. parelaphostrongylus tenuis (pryadko and boev) in the moose and white-tailed deer of nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. culhane, e. 2006. wisconsin herd control takes back seat to trophies. the moose call 20: 30-31. dodds, d. g. 1963. the present status of moose (alces alces americana) in nova scotia. transactions of the northeast wildlife conference, portland, maine, usa. dodge, w. b., s. r. winterstein, d. e. beyer, and h. campa iii. 2004. survival, reproduction, and movements of moose in the western upper peninsula of michigan. alces 40: 71-85. drew, m. l. and w. m. samuel. 1986. reproduction in winter tick, dermacentor albipictus, under field conditions in alberta, canada. canadian journal of zoology 64: 714-721. dumont, a., and m. crete. 1996. the meningeal worm, parelaphostrongylus tenuis, a marginal limiting factor for moose, alces alces in southern quebec. canadian field naturalist 11: 413-418. dunn, f. d., and k. i. morris. 1981. preliminary results of the maine moose season (1980). alces 17: 95-110. fenstermacher, r., and o. w. olsen. 1942. further studies of diseases affecting moose. iii. cornell veterinarian 32: 241-254. forrester, s. g., and m. w. lankester. 1997. extracting protostrongylid nematode larvae from ungulate feces. journal of wildlife diseases 33: 511-516. frawley, b. j. 2008. michigan deer harvest survey report, 2007 seasons. wildlife report no. 3485. michigan department of natural resources, lansing, michigan, usa. garner, d. l., and w. f. porter. 1991. prevalence of parelaphostrongylus tenuis in white-tailed deer in northern new york. journal of wildlife diseases 27: 594-598. gilbert, f. f. 1973. parelaphostrongylus tenuis (dougherty) in maine: i – the parasite in white-tailed deer (odocoileus virginianus, zimmermann). journal of wildlife diseases 9: 136-143. gogan, p. j. p., k. d. kozie, e. m. olexa, and n. s. duncan. 1997. ecological status of moose and white-tailed deer at voyageurs national park, minnesota. alces 33: 187-201. hawkins, j. w., m. w. lankester, r. a. lautenschlager, and f. w. bell. 1997. effects of alternative conifer release treatments on terrestrialgastropods in northwestern ontario. forestry chronicle 73: 91-98. _____, _____, and r. a. nelson. 1998. sampling terrestrial gastropods using corrugated cardboard sheets. malacologia 39: 1-9. hickey, l. 2008. assessing re-colonization of moose in new york with his models. alces 44: 117-126. hickie, p. f. 1944. michigan moose. michigan department of conservation, lansing, michigan, usa. johnson, r. e. 2007. moose and elk population study. north dakota game and fish department report number a-168, north dakota game and fish department, impact of meningeal worm on moose – lankester alces vol. 46, 2010 68 bismarck, north dakota, usa. karns, p. d. 1967. pneumostrongylus tenuis in minnesota and implications for moose. journal of wildlife management 31: 299-303. _____. 1972. minnesota’s 1971 moose hunt: a preliminary report on the biological collections. proceeding of the north american moose conference and workshop. 8: 115-123. _____. 1980. winter – the grim reaper. pages 47-53 in r. l. hine and s. nehls, editors. white-tailed deer population management in the north-central states. proceedings of the 1979 symposium of the north central wildlife society, the wildlife society, urbana, illinois, usa. lankester, m. w. 1974. parelaphostrongylus tenuis (nematoda) and fascioloides magna (trematoda) in moose of southeastern manitoba. canadian journal of zoology 52: 235-239. _____. 2001. extrapulmonary lungworms of cervids. pages 228-278 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals, 2nd edition. iowa state university press, ames, iowa, usa. _____. 2002. low-dose meningeal worm (parelaphostrongylus tenuis) infections in moose (alces alces). journal of wildlife diseases 38: 789-795. _____, and r. c. anderson. 1968. gastropods as intermediate hosts of meningeal worm, pneumostrongylus tenuis, dougherty. canadian journal of zoology 46: 373-383. _____, and w. j. peterson. 1996. the possible importance of deer wintering yards in the transmission of parelaphostrongylus tenuis to white-tailed deer and moose. journal of wildlife diseases 32: 31-38. _____, _____, and d. ogunremi. 2007. diagnosing parelaphostrongylosis in moose (alces alces). alces 43: 49-59. _____, and w. m. samuel. 2007. pests, parasites and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. le mouel, j.-l., v. courtillot, e. blanter, and m. shnirman. 2008. evidence for a solar signature in 20thcentury temperature data from the usa and europe. comptes rendus geoscience 340: 421-430. lenarz, m. 1993. the white-tailed deer of minnesota’s forested zone: population trends and modelling. annual report. minnesota department of natural resources, grand rapids, minnesota, usa. _____. 2007a. population trends of whitetailed deer in the forest zone – 2007. pages 103-112 in m. h. dexter, editor. status of wildlife populations, 2007. unpublished report, division of fish and wildlife, minnesota department of natural resources, st. paul, minnesota, usa. _____. 2007b. 2007 aerial moose survey. pages 113-119 in m. h. dexter, editor. status of wildlife populations, 2007. unpublished report, division of fish and wildlife, minnesota department of natural resources, st. paul, minnesota, usa. _____. 2009. 2009 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. (accessed december 2009). _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. maskey, j. j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph.d. thesis, university of north dakota, grand forks, alces vol. 46, 2010 lankester – impact of meningeal worm on moose 69 north dakota, usa. mclaren, b. e., and w. e. mercer. 2005. how management unit license quotas relate to population size, density, and hunter access in newfoundland. alces 41: 75-84. murray, d. l., w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1-30. nankervis, p. j., w. m. samuel, s. m. schmitt, and j. g. sikarskie. 2000. ecology of meningeal worm, parelaphostrongylus tenuis (nematoda), in white-tailed deer and terrestrial gastropods of michigan’s upper peninsula with implications for moose. alces 36: 163-181. ogunremi, o. a., m. w. lankester, s. j. dergousoff, and a. a. gajadhar. 2002. detection of anti-parelaphostrongylus tenuis antibodies in experimentally infected moose (alces alces). journal of wildlife diseases 38: 796-803. parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. report prepared for the nova scotia department of natural resources, halifax, nova scotia, canada. parker, l. r. 1990. feasibility assessment for the reintroduction of north american elk, moose, and caribou into wisconsin. wisconsin department of natural resources, bureau of research, madison, wisconsin, usa. patterson, b. r., b. a. macdonald, b. a. lock, d. g. anderson, and l. k. benjamin. 2002. proximate factors limiting population growth of white-tailed deer in nova scotia. journal of wildlife management 66: 511-521. patton. 1991. deer history. province of nova scotia @ 2009. (accessed december 2009). peterson, r., and r. moen. 2009. report to the minnesota department of natural resources (dnr) by the moose advisory committee. (accessed december 2009). peterson, w. j., m. w. lankester, and m. r. riggs. 1996. seasonal and annual changes in shedding of parelaphostrongylus tenuis larvae by white-tailed deer in northeastern minnesota. alces 32: 61-73. pulsifer, m. d., and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. pybus, m. j. 2001. liver flukes. pages 121149 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals, 2nd edition. iowa state university press, ames, iowa, usa. samuel, w. m. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1, federation of alberta naturalists, edmonton, alberta, canada. _____. 2007. factors affecting epizootics of winter ticks and mortality of moose. alces 43: 39-48. _____, m. s. mooring, and o. i. aalangdong. 2000. adaptations of winter ticks (dermacentor albipictus) to invade moose and moose to evade ticks. alces 36: 183-195. _____, and d. a. welch. 1991. winter ticks on moose and other ungulates: factors influencing their population size. alces 27: 169-182. schmitz, o. j., and t. d. nudds. 1994. parasite-mediated competition in deer and moose: how strong is the effect of meningeal worm on moose? ecological applications 4: 91-103. slomke, a. m., m. w. lankester, and w. j. peterson. 1995. infrapopulation dynamics of parelaphostrongylus tenuis in white-tailed deer. journal of wildlife diseases 31: 125-135. impact of meningeal worm on moose – lankester alces vol. 46, 2010 70 smith, j.r., r. a. sweitzer, and w. f. jensen. 2007. diets, movements, and consequences of providing food plots for white-tailed deer in central north dakota. journal of wildlife management 71: 2719-2726. telfer, e. s. 1967. comparison of moose and deer winter range in nova scotia. journal of wildlife management 31: 418-425. timmermann, h. r., r. gollat, and h. a. whitlaw. 2002. reviewing ontario’s moose management policy – 1980-2000 – targets achieved, lessons learned. alces 38: 11-45. todhunter, p. e. and b. c. rundquist. 2004. terminal lake flooding and wetland expansion in nelson county, north dakota. physical geography 25: 68-85. vucetich, j. a., and r. o. peterson. 2008. ecological studies of moose on isle royale: annual report 2008-2009. michigan technological university, houghton, michigan, usa. (accessed december 2009). wasel, s. m., w. m. samuel, and v. crichton. 2003. distribution and ecology of meningeal worm, parelaphostrongylus tenuis (nematoda), in northcentral north america. journal of wildlife diseases 39: 338-346. whitlaw, h. a., and m. w. lankester. 1994a. a retrospective evaluation of the effects of parelaphostrongylosis on moose populations. canadian journal of zoology 72: 1-7. _____, and _____. 1994b. the co-occurrence of moose, white-tailed deer and parelaphostrongylus tenuis in ontario. canadian journal of zoology 72: 819-825. wobeser, g., a. a. gajadhar, and h. m. hunt. 1985. fascioloides magna: occurrence in saskatchewan and distribution in canada. canadian veterinary journal 26: 241-244. p139-146_4205.pdf alces vol. 41, 2005 dungan and wright moose summer diet composition 139 summer diet composition of moose in rocky mountain national park, colorado jason d. dungan1 and r. gerald wright2 1department of fish and wildlife resources, university of idaho, p.o. box 441136, moscow, id 83844-1136, usa; 2usgs idaho cooperative fish and wildlife research unit, department of fish and wildlife resources, university of idaho, p.o. box 441136, moscow, id 83844-1136, usa abstract: summer diet composition of habituated adult moose (alces alces) in rocky mountain national park, colorado, was determined using direct observations and fecal analysis. direct observations determined moose ate 20 different plant species, including 6 willow (salix spp.) species, which comprised 91.3% of the overall diet from june through mid-september. geyer willow (salix geyeriana) accounted for 45.1% of summer diets. other species included mountain alder (alnus incana, 2.5%), quaking aspen (populus tremuloides, 1.1%), and bog birch (betula glandulosa, 1.0%). aquatic plants accounted for 1.9%, forbs 1.1%, and grasses 0.9%. moose ate 11 different species of woody browse, which comprised 96.9% of the diet. species diversity in the diet peaked in july with 18 different species, including 7 species of non-woody browse. fecal analysis showed moose consumed carex spp. major genera (> 1%) contributing to moose summer diets that were indicated by direct observations, except quaking aspen (1.1%). alces vol. 41: 139-146 (2005) key words: alces alces, diet, food habits, moose, willow, woody browse historically, moose (alces alces shirasi) were rare in colorado and early sightings recorded in colorado were believed to be moose that wandered into the state from northwestern wyoming (bailey 1944). the colorado division of wildlife (cdow) introduced two groups of 12 moose near the town of rand, colorado, in 1978 and 1979; 13 km west of rocky mountain national park (rmnp). the objective was to establish a viable resident moose population in the area. ter in the park was recorded in 1985-1986 (stevens 1988). presently between 61-66 moose are estimated to summer in the park (j. dungan, university of idaho, moscow, idaho, unpublished data). large mammalian herbivores can cause major changes in plant community composition and structure (augustine and mcnaughton 1998). herbivory is a major concern of rmnp where large numbers of rocky mountain elk (cervus elaphus) occur. managers are particularly concerned about effects on riparian willow (salix spp.) and upland shrub communities, based on visual appearance of short-hedged willow on elk winter range. in a recent study, elk herbivory was found to suppress heights, leader lengths, and annual production of willow, and herbaceous productivity of willow sites within the park (zeigenfuss et al. 1999). high densities of moose have also been of ecosystems. bark stripping by moose in denali national park, alaska, may increase the rate of succession in aspen-spruce commoose summer diet composition dungan and wright alces vol. 41, 2005 140 munities by killing trees (miquelle and van ballenberghe 1989). similar results were found in isle royale national park, michigan, where high rates of moose browsing depressed nitrogen mineralization and net primary production of boreal forest ecosystems (pastor et al. 1993). furthermore, mcinnes et al. (1992) showed that moose herbivory reduced tree biomass and production, and increased shrub and herb biomass at isle royale. winter food supply is generally considered the limiting factor in some moose populations (crete 1989, kufeld and steinert 1990, maccraken et al. 1997, zheleznov-chukotsky and votiashova 1998). woody browse is usually the only food supply available for moose during the winter, and therefore moose diet and habitat studies have focused on use, availability, and quality of winter browse (leresche and davis 1973). few studies have examined use of summer habitat, although summer diets are generally 1.5 – 3 times more nutritious than winter diets (schwartz 1992), and summer is a key period in which moose build up fat reserves that take them through the winter. knowledge of moose summer decisions on managing moose habitat. peek (1974) reviewed 41 food habit studies in north america, of which 18 were from the intermountain west, but none as far south as colorado. five of the 18 studies included summer food habits, and revealed even greater variation between areas than those on the winter range (peek 1974). summer diet studies have increased (joyal and sherrer 1978, butler 1986, van ballenberghe et al. 1989), but none have investigated the southern extent of the rocky mountains. in north park, willow was the most commonly selected habitat by moose in all seasons between 1991 and 1995 (kufeld and bowden 1996). similarly, willow comprised about 80-85% of the diet of wild adult moose in denali national park during the summer (van ballenberghe et al. 1989). the longterm effects of moose browsing on riparian communities from alaska to rmnp, is not yet understood. the purpose of this study was to document summer diet composition of moose in rmnp. study area this study was conducted in rmnp during the summers of 2003 and 2004. rmnp is located in north-central colorado just west of estes park. rmnp covers an area of 1,075 km2 and ranges from 2,389 m to 4,345 m in elevation. the park lies astride the continental divide and the west and east side differ in climate. annual precipitation ranges from 37.6 to 51.7 cm. temperatures range from highs in july and august of 24°c to lows in december-february of –17°c (monello and johnson 2003). we conducted this study predominantly on the west side of the park within the colorado river drainage and within higher elevation meadows east of the continental divide. the study was conducted at 2 distinct elevational strata, below 3,000 m (low elevation sites) and above 3,000 m (high elevation sites), because of differences in plant species composition, distribution, and plant morphology. lower elevation sites tend to be comprised of a greater variety of plant species. plants tend to be more sparsely distributed, and to be taller with longer leader lengths and greater biomass, than plants at high elevation sites. low elevation riparian meadows are characterized by large stands of geyer willow (salix geyeriana), mountain willow (salix monticola), drummond willow (salix drummondiana), plane-leaf willow (salix planifolia), and smaller stands of whiplash willow (salix lasiandra), and wolf willow ( ). other common species are beaked sedge (carex utriculata), bog birch (betula glandulosa), mountain alder (alnus incana), marsh reed grass (calamagrostis canadensis), white clover (trifolium repens), alces vol. 41, 2005 dungan and wright moose summer diet composition 141 and strawberry (fragaria ovalis). these areas are surrounded by stands of ponderosa pine (pinus ponderosa pseudotsuga menziesii), quaking aspen (populus tremuloides), and narrowleaf cottonwood (populus angustifolia). high elevation meadows are characterized by large stands of plane-leaf willow, wolf willow, and bog birch. surrounding trees include quaking aspen, lodgepole pine (pinus contorta abies lasiocarpa) (beidleman et al. 2000). methods over 3 million people visit rmnp annually. viewing wildlife is a major visitor animals such as moose are visible from park roads. as a result, moose within rmnp are accustomed to people and their habituation enabled us to directly observe feeding behavior and estimate diet composition. from 1 june through 15 september we observed the feeding behavior of moose at distances between 5-20 m. we recorded feeding data using hand-held voice recorders. feeding data were grouped into feeding periods (bouts) which the moose fed continuously. miquelle and jordan (1979) reported more than 95% of all bites recorded occurred during such bouts. we counted individual bites taken by plant species when possible. individual moose were followed as long as possible, and each continuous span was considered a single observation set. bite counts were not conducted at night for observer’s safety and compliance with park regulations. during feeding periods the sizes of all bites were recorded to estimate intake leaves), medium (5-10 leaves), or large (> 10 leaves). after each observation set, simulated moose bites were collected by clipping 10-20 samples/species in each of the three sizes as closely to the observed bite size as possible. these samples were bagged, oven-dried at 60ºc for 48 hours, and weighed in accordance with methods described by renecker and hudson (1985). average dry weight per bite (g/bite) for each size and species was calculated. diet composition was based on percentage dry weight of species consumed by moose. this was derived by multiplying the number of bites taken of each species in each size class by the average dry weight per bite (van ballenberghe et al. 1989). fresh fecal samples were collected from observed moose during foraging bouts. sample size ranged from 6 to 17 per month, and included pellets collected in october and november, while walking low elevation transects, to look at early winter diets. these samples were frozen and sent to the wildlife habitat lab at washington state university for microhistological analysis. sub-samples of each fecal pellet group were combined with others from the same month and year. fecal samples were blended with water, stained with a lactophenol blue stain. relative cover (korfhage 1974, davitt 1979) of plant cuticle and epidermal fragments were scope views on each of four slides (total 100 views) per month. a 10 square x 10 square grid was used to measure area covered by area covered were recorded by plant genus and species. percent diet composition was calculated by dividing cover of each plant by total cover observed for all species, then multiplying by 100. we compared differences in percentages of major plant species between years, elevations, months, and sexes for bite count data using a 1-way anova. salix species were pooled, and all percentages were arcsine transformed in accordance with krebs (1999). experimental units were observations (n = 54), and pair-wise differences were located moose summer diet composition dungan and wright alces vol. 41, 2005 142 using the tukey hsd procedure. differences and all statistics were performed using sas 8.3 (sas institute inc., cary, north carolina) statistical software. results we recorded feeding data on 11 female and 43 male moose for a total of 54 observation sets (table 1). over 75,000 bites were counted during 177 feeding bouts (table 1). data from direct observations showed no difference between sexes and elevations (p (f = 3.51; df = 3,48; p = 0.02) among months for ly more mountain alder in late summer (september) than early summer (june). likewise, years for western dock (rumex aquaticus) (f = 4.15; df = 1,48; p = 0.04) and triangular leaf senecio (senecio triangularis) (f = 5.71; df = 1,48; p = 0.02). moose ate 2004. other than the forementioned species, moose ate similar diets among months (p > 0.05) and years (p > 0.05). all bite count data were therefore pooled, for all animals, sexes, months, and years to estimate summer diet composition (table 2). moose consumed 11 different species of woody browse, which comprised 96.9% of the diet. six willow species comprised 91.3 % of moose summer diets. geyer willow accounted for 45.1% of summer diets followed by plane-leaf willow (22.7%), mountain willow (11.7%), and drummonth site (elevation) observations female male feeding bouts bites june low 15 4 11 27 9,807 july low 17 4 13 60 20,324 high 3 0 3 15 11,013 august low 7 3 4 32 15,277 high 4 0 4 17 10,544 september low 6 0 6 21 7,692 high 2 0 2 5 2,322 totals 54 11 43 177 76,979 table 1. summary of moose feeding data collected in rocky mountain national park, colorado from june 1st through september 15th, 2003 and 2004. species diet % salix geyeriana 45.10 s. planifolia 22.70 s. monticola 11.70 s. drummondiana 9.90 alnus incana 2.50 1.50 aquatic spp. 1.30 populus tremuloides 1.10 betula glandulosa 0.97 grasses 0.91 cirsium spp. 0.71 rumex aquaticus 0.47 senecio triangularis 0.32 s. lasiandra 0.29 nuphar lutea ssp. polysepala 0.21 0.16 psychrophila leptosepala 0.03 mentha spicata 0.02 pincus contorota trace1 shepherdia canadensis trace1 table 2. percentage of plant species consumed by moose from june 1st through september 15th 2003 and 2004 in rocky mountain national park, colorado. percentages were based on sociated weights, measured in grams/bite. 1trace species represented less than 0.01 % of the diet. alces vol. 41, 2005 dungan and wright moose summer diet composition 143 mond willow (9.9%). geyer willow ranked it ranked third behind mountain willow and drummond willow. other woody species included mountain alder (2.5%), quaking aspen (1.1%), and bog birch (1.0%). aquatic species accounted for 1.9%, forbs 1.1%, and grasses at 0.9%. species diversity peaked in july with moose eating 18 different species, including 7 species of non-woody browse. monthly fecal analysis for 2003 showed moose relied more heavily on species other than willow in june, most notably carex spp. (46.4%, table 3), then increased willow consumption in july with a peak in august at 90%. in september, willow consumption began to decrease and early winter diets included species other than willow, most notably conifer needles (34.1%), shrubs (18.1%), and carex spp. (11.6%, table 3). fecal analysis data were pooled for all months and years, with the exception october/november 2003, to show summer moose diet composition (table 4). fecal analysis showed moose consume 79.3% willow, which is 11.9% less than direct observations carex spp. as a major contributor to moose by direct observation. fecal analysis was not able to identify forbs, willow, or shrubs tions with the exception of quaking aspen (1.1%, table 4). with the exception of carex spp., both techniques showed similar results (table 4). discussion moose summer diets in rmnp consisted of 11 different species of woody browse, which accounted for roughly 97 % of the overall diet. these results are similar to those found by van ballenberghe et al. (1989) in denali national park, alaska, where woody species made up 96% of moose summer diets, and joyal and scherrer (1978) in mont-tremblant park, quebec, where moose summer diets consisted of 100% woody browse. moose in rmnp use riparian willow communities during the summer, which contain little aquatic vegetation or forbs. moose were observed eating 9 different nonwoody species. studies in less mountainous habitats have found moose consume larger proportions of forbs (25%, leresche and davis 1973; 70.6%, knowlton 1960) and aquatics (9.3%, mcmillan 1953) than those in rmnp. forbs and aquatics may contain higher concentrations of important minerals (belovsky 1978), and have higher digestibility levels than woody browse (leresche and davis 1973). forbs and aquatics only accounted for 2.4% of the overall summer diet of moose in rmnp. moose ate 6 willow species comprising 91.3% of their summer diets, with geyer willow accounting for 45.1%. results were similar to those found by mcmillan (1953) for plants jun-03 jul-03 aug-03 sep-03 oct/nov 03 2004 grasses 2.10% 0.00% 0.60% 0.80% 8.90% 7.70% carex spp. 46.40% 5.00% 0.10% 3.80% 11.60% 3.60% salix spp. 48.00% 88.20% 90.00% 82.90% 25.30% 81.40% shrubs 0.40% 6.00% 9.30% 11.50% 18.10% 6.90% conifer needle 0.70% 0.00% 0.00% 0.00% 34.10% 0.20% forbs 2.40% 0.60% 0.00% 1.00% 2.00% 0.00% sphagnum moss 0.00% 0.00% 0.00% 0.00% 0.00% 0.20% insect 0.00% 0.20% 0.00% 0.00% 0.00% 0.00% table 3. percent of major plant species consumed by moose in rocky mountain national park, colorado per month for 2003, and by year for 2004, as indicated by fecal analysis. moose summer diet composition dungan and wright alces vol. 41, 2005 144 moose in yellowstone national park (willow 88.5%) and van ballenberghe (1989) in denali national park (81.5%). similarities of diet composition of moose in rmnp and moose in denali national park suggest use of very similar habitats during the summer. willow habitats not only provide a high quality woody browse for consumption (leresche and davis 1973) but are also used extensively for cover (kufeld and bowden 1996). fecal analysis showed moose rely more heavily on grasses, sedges, and forbs (50.9%) in early summer than woody browse (49.1%). peek (1974) found that grasses and sedges were seldom consumed other than in spring, when digestibility and nutrient content are high. consumption of willow species increased in july, peaked in august, and started to decline in september. both techniques showed the same trend, with exception of june bite count data, suggesting moose eat willow when available biomass and nutrients are high (stumph 2005) and rely less heavily on willow in spring and fall when available biomass and nutrients are lower. early winter fecal analysis showed willow comprised only 25.3 % of the diet, whereas conifer needles made up 34.1%, and grasses, forbs, and sedges made up 22.5 %. forbs and grasses comprised relatively larger percentages of fall diets than winter diets (peek 1974). during both years we observed moose moving out of the lower riparian willow communities, with the onset of the rut, and moving into higher more heavily forested areas of the park, possibly giving a reason for the large percent of conifer needles in early winter diets. in a study conducted from december 1991 to november 1995, moose from north park, colorado, tended to winter at lower elevations and move to higher elevations during spring, summer, and fall (kufeld and bowden 1996). winter transect data showed little use of low elevation riparian willow communities from october through december for both years. willows were the primary forage plants in 5 out of 6 winter diet studies reviewed by peek (1974), and risenhoover (1987) reported that willow comprised 94.3% of winter diets in denali national park. stevens (1970) reported timber types received 82% of total use during the winter in the gallatin region of montana. october/november fecal analyses coupled with winter transect data suggest that willow may not constitute a large proportion of moose winter diets in rmnp, although a comprehensive study of winter diets has not been performed. similar results were achieved using bite count data and fecal analysis. direct observations failed to detect the large amount of sedges consumed by moose in early summer, which was largely affected by adult movements during june. older, more habituated moose had not yet moved to summer range from table 4. percent diet composition of major plant species consumed by moose in rocky mountain national park, colorado from june 1st through september 15th, 2003 and 2004, using direct observations and fecal analysis. all months and years were pooled for both techniques. 1shrubs consisted of alnus incana, betula glandulosa, , and shepherdia canadensis. 2aquatics consisted of aquatic spp., nuphar lutea spp. polysepala, and rumex aquaticus. 3forbs consisted of cirsium spp., mentha spicata, psychrophila leptosepala, and senecio triangularis. 4trace species represented less than 0.01 % of the diet. species fecal analysis bite count salix spp. 79.3 91.3 carex spp. 8.7 shrubs 6.8 3.61 grasses 4.3 0.91 aquatics 0.11 1.982 populus tremuloides 1.12 forbs 0.51 1.083 conifer needles 0.19 trace4 insects 0.02 alces vol. 41, 2005 dungan and wright moose summer diet composition 145 winter range, necessitating that our sample be focused on younger, less habituated moose. this caused us to miss a large proportion of feeding data per day, which could account for some of the variation. grasses were also slightly higher (3.4%) in fecal analysis than direct observations. consumption of grasses and sedges is hard to observe because of their low growth form, and bites were not recorded which could also lead to differences between the two techniques. wallmo et al. (1973) reported bite count methods estimated more use of shrubs as a class and less use of grass and forbs than actually occurred for mule deer (odocoileus hemionus). fecal analysis could not differentiate willow, forb, or shrub species, and failed to detect populus tremuloides. other species not detected by fecal analysis included cirsium spp., rumex aquaticus, senecio triangularis, and nuphar lutea ssp. polysepala, but these species were less than 1 % of the observed diet. monthly fecal sample size was smaller than similar studies, which could have contributed to not detecting these species. acknowledgements john skinner, basil iannone, and mandy cluck who spent countless hours observing moose night and day. park service staff were more than helpful all 3 years of the study, with special thanks going to ryan monello. funding for this project was provided by rocky mountain national park through the usgs idaho cooperative fish and wildlife research unit. this manuscript is dedicated to the memory of dr. francis singer. references augustine, d. j., and s. j. mcnaughton. 1998. ungulate effects on the functional species composition of plant communities: herbivore selectivity and plant tolerance. journal of wildlife management 62:1165-1183. bailey, a. 1944. records of moose in colorado. journal of mammalogy 25:192193. beidleman, l. h., r. g. beidleman, and b. e. willard. 2000. plants of rocky mountain national park. falcon, helena, montana, usa. belovsky, g. 1978. diet optimization in a generalist herbivore: the moose. theoretical population biology 14:105-134. butler, c. e. 1986. summer food utilization and observations of a tame moose, alces alces. canadian field naturalist 100:85-88. crete, m. 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67:373-380. davitt, b. b. 1979. elk summer diet composition and quality on the colockum multiple use research unit, central washington. m.s. thesis. washington state university, pullman, washington, usa. joyal, r., and b. scherrer. 1978. summer movement and feeding by moose in western quebec. canadian field naturalist 92:252-258. knowlton, f. f. 1960. food habits, movement, and populations of moose in gravely mountains, montana. journal of wildlife management 24:162-170. korfhage, r. c. 1974. summer food habits of elk in the blue mountains of northeastern oregon based on fecal analysis. m.s. thesis. washington state university, pullman, washington, usa. krebs, c. j. 1999. ecological methodology. second edition. benjamin/cummings, new york, new york, usa. kufeld, r. c., and d. c. bowden. 1996. movement and habitat selection of shiras moose (alces alces shirasi) in colorado. alces 32:85-99. _____, and s. f. steinert. 1990. an estimate of moose carrying capacity in willow moose summer diet composition dungan and wright alces vol. 41, 2005 146 habitat in north park, colorado. colorado division of wildlife unpublished report. fort collins, colorado, usa. leresche, r. e., and j. l. davis. 1973. importance of nonbrowse foods to moose on the kenai peninsula, alaska. journal of wildlife management 37:279-287. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal south-central alaska. wildlife monographs 136. mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73:2059-2075. mcmillan, j. f. 1953. some feeding habits of moose in yellowstone national park. ecology 34:102-110. miquelle, d. g., and p. a. jordan. 1979. the importance of diversity in the diet of moose. proceedings of the north american moose conference and workshop 15:54-79. _____, and v. van ballenberghe. 1989. impact of bark stripping by moose on aspen-spruce communities. journal of wildlife management 53:577-586. monello, r., and t. johnson. 2003. the elk herd in rocky mountain national park. rocky mountain national park unpublished report. estes park, colorado, usa. pastor, j., b. dewey, r. j. naiman, p. f. mcinnes, and y. cohen. 1993. moose browsing and soil fertility in the boreal forest of isle royale national park. ecology 74:467-480. peek, j. m. 1974. a review of moose food habit studies in north america. naturaliste canadien 101:195-215. renecker, l. a., and r. j. hudson. 1985. estimation of dry matter intake of freeranging moose. journal of wildlife management 49:785-792. risenhoover, k. l. 1987. winter foraging strategies of moose in subarctic and boreal forest habitats. ph.d. dissertation. michigan technical university, houghton, michigan, usa. schwartz, c. c. 1992. physiological and nutritional adaptations of moose to northern environments. alces supplement 1:139-155. stevens, d. r. 1970. winter ecology of moose in the gallitin mountains, montana. journal of wildlife management 34:37-46. _____. 1988. moose in rocky mountain national park. rocky mountain national park unpublished report. estes park, colorado, usa. stumph, b. p. 2005. the summer forage qualence on moose forage ecology in rocky mountain national park, colorado. m.s. thesis. university of idaho, moscow, idaho, usa. van ballenberghe, v., d. g. miguelle, and j. g. maccracken. 1989. heavy utilization of woody plants by moose during summer at denali national park, alaska. alces 25:31-35. wallmo, o. c., r. b. gill, l. h. carpenter, and d. w. reichert. 1973. accuracy of of wildlife management 37:556-562. zeigenfuss, l. c., f. j. singer, and d. bowden. 1999. vegetation responses to natural regulation of elk in rocky mountain national park. biological science report usgs/brd 1999 003. u.s. governusa. zheleznov-chukotsky, n. k., and e. s. votiashova. 1998. comparative analysis of moose nutrition of the anadyrsky and omolonsky populations (far north east) in different seasons. alces 34:445-451. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 4308.pdf alces vol. 43, 2007 kochan seasonal adaptations of moose metabolism 123 seasonal adaptations of moose (alces alces) metabolism tatyana i. kochan institute of physiology, komi science center, ural division of the russian academy of sciences, syktyvkar 167982, pervomayskaya, 50, russia abstract: the experiments were conducted on two yearling moose (alces alces) with rumen alces vol. 43: 123-128 (2007) key words: adaptation, alces alces alces) (schwartz et al. 1984, 1985, 1987, 1988; rewinter decreases naturally in response to a times higher than the mean winter hp. similar patterns in consumption and dinutrients and energy, water exchange, rumen ological adaptation to changing conditions methods research was conducted with 2 yearling seasonal adaptations of moose metabolism kochan alces vol. 43, 2007 124 feces and urine were collected in plastic to chemical analysis. the water content, dry (betula pubestris), aspen (picea obovata), mountain ash (sorbus aucuparia), willow (salix sp.) chamaerion angustifolium pinus sylvestris vaccinium vitis-idaea 23:30:17:30. vfa. ( student t-test was used to determine statistical results p < cellulose, and energy were higher in summer 0.75) 0.75 p < 0.001) nutritional parameter dry matter ± 9.5 84.2 ± cellulose ± 1.7 18.2 ± sugar 19.5 ± 1.2 9.1 ± 108.8 ± 7.4 47.9 ± 15.9 ± 0.2 7.3 ± 18.8 ± 1.1 8.7 ± gross energy 2.98 ± 0.27 1.45 ± 2.2 ± 0.2 0.83 ± alces vol. 43, 2007 kochan seasonal adaptations of moose metabolism 125 p and urinary water output changed similarly. (p p lower (p p p < 0.001) times (p increased 3 times (p 5 times (p < 0.001), and acetate 2 times (p < discussion adaptations to cold seasons with limited that aid thermal homeostasis; e.g., when energy . 0.75) in p p < 0.01), p < 0.001). measurement 333.3 ± 83.0 ± ± 108.3 ± 10.8 440.9 ± 191.3 ± 128.8 ± 1.9 ± urine secretion ± 24.5 ± approximate water retention 148.7 ± 20.3 73.0 ± (p p < 0.01). item cellulose 22.7 ± 3.5 ± 0.1 sugar 4.0 ± 0.1 3.0 ± acetic acid ± 1.3 47.2 ± 2.5 propionic acid 12.8 ± 0.9 8.7 ± butyric acid 15.8 ± 3.2 ± 78.2 ± 2.0 ± p < 0.05), p p < 0.001). ± 0.42 3.30 ± acetone + acetoacetate, 2.35 ± ± 4.00 ± 0.80 2.00 ± ± 0.28 ± acetic acid, ± 0.14 ± propionic acid, 0.03 ± 0.00 0.08 ± butyric acid, 0.01 ± 0.00 0.02 ± ± 0.14 ± seasonal adaptations of moose metabolism kochan alces vol. 43, 2007 thermal homeostasis, it is also important to reduce heat loss. 2 and respired co2 oxidation-reduction process. the decrease calorimetry (regelin et al. 1985), suggesting data in this study, which indicated an increase declined in the rumen indicates their lower and proteins. reduction processes leads to less production acceptor) and lower atp production. lower atp synthesis in winter is apparently explained endogenic water in reindeer during cold was maintained or reduced during winter and to reduce heat loss. energy and water sources can occur only with i assumed oxidation was compensated the increased the the reduced (i.e, greater utilization than ) was increased glycolytic intensity prior to winter. in summer, when consumption energy, and water are highest, vfa produced production. th low alces vol. 43, 2007 kochan seasonal adaptations of moose metabolism 127 than in ) and pointing , when there was reduced consumption , and lower water retention, the total vfa concentration in the rumen decreased as increased, while glucose and concentration . changes suggest c adaptation moose to limit heat loss during winter include , enhancement the supply, and reduced water exchange. acknowledgements references cameron, r. d., and j. r. luick. 1972. chermnykh, n. a., m. p. roshchevsky, and e. a. novozhilova. 1980. kopytnye khmylov, a. v. knorre, e. p., and e. k. knorre. 1959. kochan, t. i nogo opredeleniya pitatel’noy cennosti _____ kochanov, n. e., g. m. ivanova,a. e. weber, and a. f. symakov. 1981. wild ruminants – reindeer and moose). _____, and g. v. tulupov. 1979. gazozhidpochinok, k. h. n regelin, w. l., c. c. schwartz, and a. w. franzmann. 1985. seasonal energy _____, _____, and _____. renecker, l. a., and r. j. hudson. 1985. management 49:785-792. schwartz, c. c., m. e. hubbert, and a. w. franzmann. 33. _____, w. l. regelin, and a. w. franzmann. 1984. in moose. alces 20:223-244. seasonal adaptations of moose metabolism kochan alces vol. 43, 2007 128 _____, _____, _____, and m. hubbert. sokolov, v. e., and o. f. chernova. 1987. alces alces weber, a. e., and a. v. chalyshev. 1990. _____, a. f. simakov, n. i. chuv’yurova, a. v. chalyshev, l. p. badlo, t. i. kochan, and n. n. mongalev. 1992. fisiologia pitania yazan, j. p., and m. v. kozhukhov. moregulation in moose). trudy pechoro4305.pdf alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 61 trace elements status of moose and white-tailed deer in nova scotia beth pollock1 and erin roger2,3 1p.o. box 226, norris point, nl, canada a0k 3v0; 2department of natural resources, 136 exhibition st., kentville, ns, canada b4n 4e5 abstract: the province of nova scotia is considered to have two distinct moose populations: mainland and cape breton island. in 2003, moose of the mainland area of the province were formally of the decline of this population have not been determined. factors impacting health, including trace element imbalances, have been considered as potential limiting factors for the mainland population. liver and kidney samples were collected from moose and white-tailed deer throughout nova scotia during the fall and winter 2000 – 2002 to compare trace element concentrations between the two species, in relation to age, gender, and geographical location, and to other areas outside the province. all samples were analysed for arsenic, cadmium, cobalt, copper, lead, manganese, nickel, selenium, and zinc. tissue concentrations of trace elements in deer and moose in nova scotia are generally similar to levels reported in cervid populations elsewhere in north america and europe with the exception of zinc and cobalt, which appear to be lower in nova scotia. kidney cadmium concentrations are high in some nova scotia moose (geometric mean = 60.4 g/g dry weight, 95% ci = 40.3 – 90.6, n = 21), however, similar or higher concentrations have been reported in other regions. relative to reference may impact the health of individual animals either directly or through interactions with other factors (e.g., infectious and non-infectious diseases, harsh environmental conditions, habitat limitations) cannot be dismissed. some considerations for continued monitoring of trace element concentrations in these populations are discussed. alces vol. 43: 61-77 (2007) key words: alces alces, environment, health, kidney, liver, monitoring, nova scotia, trace elements the province of nova scotia is considered to have two moose (alces alces) populations. experiences normal population growth and is harvested by recreational hunters. it was supplemented by 18 moose translocated from alberta in 1947 and 1948 (pulsifer and nette 1995) and, not surprisingly, subsequent genetic analysis (broders et al. 1999) has demonstrated the population’s genetic structure to be most closely related to that of moose from alberta. the second mainland population of moose is indigenous to the region and is representative of the eastern moose subspecies (alces alces americana). it is made up of three subpopulations corresponding to the cumberland, tobeatic, and guysborough regions of the province. this population has been in decline since the mid-1920s, despite being protected from legal hunting since 1981. in 2000, it was be a species at risk of extirpation or extinc3present address: school of biological earth and environment sciences, university of new south wales, sydney, nsw 2052, australia trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 62 tion under the general status of nova scotia wildlife assessment process. following the completion of an independent commissioned status report (parker 2003) in october of 2003, moose of the mainland area of the province were formally listed as “endangered” under the nova scotia endangered species decline of the mainland moose population have size and distribution of moose populations in nova scotia may include disease (e.g., parelaphostrongylosis), habitat suitability, illegal hunting, and possibly predation of calves by black bears (ursus americanus) or coyotes (canis latrans) (beazley et al. 2006). trace element imbalances may also be a potential etiology for this population decline in part because such imbalances have been cited as possible contributing factors in population declines/mortality events of moose in other areas. examples include possible copper northwestern minnesota (flynn et al. 1977, o’hara et al. 2001, custer et al. 2004) and sweden (frank et al. 1994, 2000). results from preliminary trace element analyses of moose kidney and liver tissues have raised the possibility that high cadmium (cd; roger 2002) and/or low cobalt (co; frank et al. 2004) levels are impacting the health of moose on mainland nova scotia. the white-tailed deer (odocoileus virginianus) arrived in and dispersed throughout nova scotia in the period 1890 – 1920, and dance (pulsifer and nette 1995). in 2001, the population estimate for the province was approximately 51,500 (1.22 deer/km2), which is below the optimal size of 80,000 animals, but still considered healthy (t. nette, nova scotia department of natural resources, personal communication). as with moose, white-tailed deer are present on the mainland and cape breton island regions of nova scotia. as the only other cervid species in the province, white-tailed deer may serve as a comparison population for trace elements concentrations in moose with the caution that inter-species differences in trace element concentrations can exist. the interpretation of trace element levels in free-ranging cervid populations can be established and sources of variability within and between species and populations are known to exist due to differences in diet, age, gender, geographical location, and other factors (frøslie et al. 1984, ohlson and staaland 2001, o’hara et al. 2003, custer et al. 2004, gamberg et al. 2005). however, with this caution in mind, comparisons of trace element levels within and between populations and species can be helpful in identifying particular areas for further monitoring or study. the objectives of this study were to: (1) determine trace element concentrations in moose (liver and kidney) and deer (liver) tissue from nova scotia; (2) for each trace element, examine tissue concentration and age, gender, and geographic region; (3) compare trace element concentrations found in nova scotia moose and deer to each other, to other free-ranging cervid populations, and to reference values for domestic cattle; and (4) discuss whether trace element imbalances may be contributing to the decline of the mainland population of moose in nova scotia. methods sample collection during fall and winter 2000 – 2002, samples of liver (n = 48) and kidney (n = 21) from moose and liver from deer (n = 54) were collected from animals killed in vehicle accidents or by hunters, or animals killed and submitted to the department of natural resources because of property intrusion or illness. all tissue samples were placed in plastic bags, labeled, and frozen at -15 to -20°c. animal age was alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 63 determined by size (calves [moose] and fawns [deer]), and/or by analysis of cementum annuli of the lower canine and lower incisor teeth (matson’s laboratory, milltown, montana, available only for calves or fawns and those animals aged by tooth cementum analysis. as calves or fawns (< 12 months), yearlings (12 – 24 months), and adults (> 24 months) for descriptive and statistical analyses. location data were obtained for 51/54 (94%) deer and 20/48 (42%) moose. exact locations of hunted moose in cape breton were not recorded; however hunting is limited to two counties in cape breton: inverness and victoria. the directional location relative to cape breton highlands national park was also recorded. moose on the mainland are found in three subpopulations corresponding to the cumberland, guysborough, and tobeatic regions of the province. for comparisons among geographical locations, moose were categorized according to three regions: mainland east (me) which includes moose from cumberland and guysborough sub-populations, mainland west (mw) which includes the tobeatic sub-population, and cape breton (cb) (fig. 1). similar regions were used to categorize deer (fig. 2). analytical methods all samples were analyzed for trace element concentrations (arsenic [as], cadmium, cobalt, copper, lead [pb], manganese [mn], nickel [ni], selenium [se], and zinc [zn]) at the environmental quality laboratory, environment canada, moncton, new brunswick in 2002. subsequently, samples were placed in and pre-frozen for a minimum of 12 hours. after freezing, samples were placed on the fig. 1. map of nova scotia illustrating locations of moose and number of moose sampled in each region. fig. 2. map of nova scotia illustrating locations of deer and number of deer sampled in each region. trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 64 freeze-dry system (labconco freezone 4.5 limited, ontario, canada) for approximately 48 hours after which they were homogenized using a mortar and pestle. once ground, samples were weighed out (approximately 0.15 g of tissue) into the xp-1500 plus vessels. five ml of nitric acid and 2 ml of deonized water were added to each vessel. each batch of samples contained a blank spike, duplicate, or spiked duplicate, and two quality control samples and was digested in the microwave accelerated reaction system for extraction (mars-x, cem corporation, québec, canada). once samples were at room temperature they were diluted to the 50 ml mark, homogenized, and transferred into 15 ml vials for icp-ms analysis (inductively coupled mass spectrophotometer) (perkin-elmer elan 6000, perkin-elmer life and analytical sciences, ontario, canada). the minimum detection limit (mdl) for analysis was 0.25 g/g for as and se and 0.05 g/g for cd, co, cu, mn, ni, pb, and zn. concentrations are reported in g/g dry weight (dw). statistical analysis all transformations and analyses were carried out using statistical software (stata v.8.0). for descriptive statistical analyses, (95% ci), and ranges were calculated for trace elements. values less than the minimum detection limits (mdl) were replaced with ½ of the mdl. trace elements detected in at least 20% of the samples by species and tissue (all trace elements measured except for as and pb in deer liver and as in moose liver) were included in univariate statistical analyses described below. for each species, simple associations among trace elements and demographic variables (gender, age, geographical location) were examined. trace element concentrations were successfully transformed (using the natural log [ln]) to render them normally distributed in order to meet the assumptions of the following parametric statistical tests: 2 statistics for categorical variables (gender, age group, geographical location) and pair-wise t-tests and one-way anova (followed by bonferroni’s adjustments for pair-wise comparisons) for normally-distributed continuous variables. for those variables for which ln transformation did not result in a normal distribution (co and zn in deer liver, and cd, mn, and pb in moose kruskal-wallis tests were used. differences between trace element concentrations in deer and moose liver samples were examined using t-tests. results of p < 0.05 were reported as for the purposes of comparison, all concentrations reported in wet weight (ww) in other studies were converted to dry weight by estimating the moisture content of liver and kidney tissues as 71.4% (puls 1994), resulting in the conversion equation g/g ww x 3.5 = for tissue trace element levels in free-ranging populations of moose and white-tailed deer are unavailable for comparison. in lieu of ence values were used for comparisons with the caution that species differences in trace element concentrations have been reported. results demographic data moose – forty-eight liver samples and 21 kidney samples were collected from 48 moose. the geographic distribution of sampled moose is shown in fig. 1. the majority of moose sampled were males (28 males, 8 females, 2 unknown gender). ages ranged from 4 to 123 months. based on age category, there were 5 calves, 11 yearlings, 25 adults, and 7 unknown age. males were older than females (male geometric mean age: 38 months, 95% ci = alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 65 26 – 54, n = 23; females: 18 months, 95% ci = 10 – 33, n = 12; t = 2.07, p = 0.023). no age group differences ( 2 = 8.92, df = 4, p = 0.063) or gender differences ( 2 = 2.47, df = 2, p = 0.29) were found among geographical locations. deer – liver samples were collected from 54 white-tailed deer. the geographic distribution of all deer sampled is shown in fig. 2. gender and age data were collected for 51/54 animals. the majority of deer sampled were females (33 females, 21 males, 3 unknown gender). ages ranged from 8 to 126 months (based on tooth cementum analysis). based on age category, there were 2 fawns, 19 yearlings, 30 adults, and 3 unknown age. the gender distribution differed among geographical locations (males: cb = 10/17, me = 8/24, mw = 0/10; 2 = 9.6, df = 2, p = 0.008). no age group differences were found among geographical locations ( 2 = 3.09, df = 4, p = 0.54) or between genders ( 2 = 4.48, df = 2, p = 0.11). tissue concentrations of trace elements table 1 provides a summary of the occurrence and concentrations of trace elements for deer liver, moose liver, and moose kidney. for calculations involving ni concentrations, one sample was dropped because of an erroneously high value (44.5 g/g dw). domestic cattle reference values for marginal and toxic levels of each trace element are also listed where applicable. in general, based on reference values for domestic cattle, concentrations of as, cu, pb, and ni in deer and moose tissues were within normal ranges (puls 1994). results for other trace elements are listed below. analytical results moose liver samples had higher cd and cu and lower co, mn, and se levels than deer liver samples (table 1). some regional differences were seen and are listed below along with detailed age and gender results for each trace element. cadmium – relative to reference values for domestic cattle, cd concentrations in moose and deer were below the chronic toxicity level (liver: 175 g/g dw; kidney: 350 g/g dw; converted from ww) (puls 1994). regional differences were seen with cd in moose kidney (f = 4.29, df = 20, p = 0.030). bonferroni comparison indicated that mw concentrations (geometric mean = 166.6 g/g dw, 95% ci = 48.5 – 572.1, n = 4) were greater than me (45.0 g/g dw, 95% ci = 17.0 – 119.2, n = 7) (p = 0.043) or cb (49.5 g/g dw, 95% ci = 34.6 – 70.9, n = 10) (p = 0.048). a similar relationship was seen with liver concentrations (mw: 27.5 g/g dw, 95% ci = 12.4 – 60.7, n = 4; me: 4.6 g/g dw, 95% ci = 1.1 – 19.1, n = 10; cb: 5.2 g/g dw, 95% ci = 3.8 – 7.0, n = 34; p = 0.005). female moose had higher kidney cd concentrations than males (female geometric mean = 92.4 g/g dw, 95% ci = 43.5 – 196.0, n = 10; male geometric mean = 41.1 g/g dw, 95% ci = 29.7 – 57.3, n = 11; t = -2.30, p = 0.017). between age (in months) or age groups with respect to cd concentrations in deer or moose. cobalt – in relation to reference values for domestic cattle, 29.2% of moose liver samples and 7.4% of deer liver samples were g/g dw) (fig. 3a) (puls 1994). regional differences were found in moose liver co concentrations (f = 3.90, df = 47, p = 0.028). bonferroni comparison indicated that me moose had higher liver co concentrations (geometric mean = 0.14 g/g dw, 95% ci = 0.08 – 0.26, n = 10) than cb moose (0.07 g/g dw, 95% ci = 0.06 – 0.09, n = 34) (p negative association between age (in months) and co concentrations in kidney (pearson’s r = -0.53, p <0.05): as age increased, kidney co concentrations decreased. copper – in relation to reference values trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 66 trace element concentration1 reference values2 deer liver moose liver moose kidney liver kidney n 54 48 21 marginal toxic marginal toxic arsenic na na 0.19 (0.15 – 0.24) na > 24.5 na > 17.5 [< 0.25 – 0.60] [< 0.25 – 1.42] [< 0.25 – 0.59] 13.00% 6.30% 33.30% cadmium 1.1 (0.8 – 1.5)3 5.8 (4.0 – 8.3)4 60.4 (40.3 – 90.6) na > 175 na > 350 [0.05 – 28.1] [< 0.05 – 51.9] [14.3 – 346.1] 100% 97.90% 100% cobalt 0.15 (0.13 – 0.17)3 0.08 (0.07 – 0.10)4 0.10 (0.08 – 0.13) 0.06 > 17.5 < 0.05 > 105 [< 0.05 – 0.41] [< 0.05 – 0.39] [< 0.05 – 0.21] 96.30% 85.40% 95.20% copper 122.0 (88.9 – 167.5)3 251.6 (188.2 – 336.5)4 16.1 (14.8 – 17.5) < 35 def > 875 [2.0 – 506.3] [2.9 – 625.5] [12.5 – 23.6] 35 – 87.5 100% 100% 100% manganese 11.6 (10.1 – 13.4)3 9.2 (7.2 – 11.7)4 13.0 (11.3 – 15.1) < 3.5 def 3.3 – 4.2 [1.8 – 27.9] [0.3 – 57.8] [8.5 – 29.7] 3.5 – 10.5 100% 100% 100% nickel 0.57 (0.41 – 0.80)3 0.37 (0.26 – 0.52)3 0.56 (0.33 – 0.95) < 0.035 < 0.035 [< 0.05 – 4.46] [< 0.05 – 5.24] [0.13 – 5.76] 94.40% 93.80% 100% lead na 0.07 (0.05 – 0.09) 0.05 (0.04 – 0.08) na > 17.5 na > 17.5 [< 0.05 – 0.51] [< 0.05 – 2.69] [< 0.05 – 0.19] 14.80% 56.30% 47.60% table 1. mean trace element concentrations ( g/g dry weight) in white-tailed deer liver and moose liver and kidney in nova scotia in 2000–2001. 1 geometric mean limit. 2 marginal and toxic reference values for cattle in g/g dry weight (puls, 1994). concentrations converted from wet weight based on estimate of 71.4% moisture (conversion factor of 3.5); 3,4 different, p < 0.05. na = not applicable. alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 67 table 1 (continued). mean trace element concentrations ( g/g dry weight) in white-tailed deer liver and moose liver and kidney in nova scotia in 2000–2001. trace element concentration1 reference values2 deer liver moose liver moose kidney liver kidney n 54 48 21 marginal toxic marginal toxic selenium 1.4 (1.2 – 1.6)3 0.7 (0.6 – 0.8)4 2.9 (2.5 – 3.3) < 0.6 def > 4.4 < 3.5 > 8.8 [0.6 – 4.0] [0.3 – 3.6] [1.8 – 5.3] 0.6 – 0.88 100% 100% 100% zinc 79.7 (70.9 – 89.6)3 75.0 (60.7 – 92.6)3 99.7 (87.4 – 113.7) < 70 def 56 – 70 > 455 [33.3 – 461.9] [9.5 – 476.9] [68.6 – 159.5] 70 – 87.5 def/marg 100% 100% 100% 1 geometric mean limit. 2 marginal and toxic reference values for cattle in g/g dry weight (puls, 1994). concentrations converted from wet weight based on estimate of 71.4% moisture (conversion factor of 3.5); 3,4 different, p < 0.05. na = not applicable. for domestic cattle, 4.2% and 2.1% of moose g/g dw) and g/dw), respectively, and 7.4% and 3.7% of deer liver cient, respectively (fig. 3b) (puls 1994). regional differences were found in moose kidney cu concentrations (f = 5.07, df = 20, p = 0.018). bonferroni comparison indicated that mw moose had higher cu (geometric mean = 19.0 g/g dw, 95% ci =15.3 – 23.6, n = 4) compared to cb moose (14.5 g/g dw, 95% ci =13.4 – 15.7, n = 10) (p = 0.025). between age categories in moose (f = 7.33, df = 40, p = 0.002). bonferroni comparison indicated that yearling moose had lower cu in liver (geometric mean = 95.2 g/g, 95% ci = 31.4 – 288.5, n = 11) compared to adult moose (geometric mean = 338.4 g/g dw, 95% ci = 280.6 – 408.0, n = 25) (p = 0.002). manganese – in relation to reference values for domestic cattle, 4.2% and 41.7% g/g dw) , respectively, and 3.7% and 27.8% (puls 1994). in deer, females had higher liver mn concentrations (geometric mean = 12.9 g/g dw, 95% ci = 11.1 – 14.9, n = 33) compared to males (geometric mean = 9.1 g/g dw, 95% ci = 6.7 – 12.5, n = 18; t = -2.36, p = 0.011). nickel – the marginal reference value for liver for domestic cattle (< 0.035 g/g dw) (puls, 1994) is lower than the mdl for moose and deer liver samples (0.05 g/g dw), thus the number of samples below the marginal reference value could not be calculated. in moose, regional differences were found in liver ni concentrations (f = 29.24, df = 46, p < 0.0001). bonferroni comparison indicated that ni in cb was higher (geometric mean = 0.65 g/g dw, 95% ci = 0.52 – 0.80, n = 33) than me (0.08 g/g dw, 95% ci = 0.04 – 0.13, n = 10) (p < 0.001) and mw trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 68 fig. 3. box plots representing trace element concentrations of cobalt (a), copper (b), manganese (c), selenium (d), and zinc (e) in white-tailed deer and moose liver from nova scotia. shaded areas represent the distribution of the middle 50% of the data, short horizontal lines show the median domestic cattle. (0.18 g/g dw, 95% ci = 0.01 – 3.60, n = 4) (p = 0.012). similar differences were seen with kidney samples (f = 12.54, df = 20, p = 0.0004): cb: 1.37 g/g dw, 95% ci = 0.72 – 2.60, n = 10; me: 0.21 g/g dw, 95% ci = 0.16 – 0.28, n = 7; p < 0.001; and mw: 0.33 g/g dw, 95% ci = 0.06 – 1.88, n = 4; p = 0.023). selenium – in relation to reference values for domestic cattle, 39.6% and 33.3% of alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 69 ties (cd) of these trace elements in the form of clinical signs or pathological changes in individual animals has not been consistently found (beazley et al. 2006). the following discussions for each trace element of concern include tissue concentrations found in relation to levels in other studies on nova scotia moose and cervid populations elsewhere. however, caution must again be used when making these comparisons as many trace elements have been shown to vary with age, gender, season, locality, and health status. some considerations for continued monitoring of trace element concentrations in these populations are also discussed. cobalt overall, liver co concentrations in the present study were higher in deer compared to moose. between regions, it was found that me moose had higher liver co levels than cb moose. few other studies were found for comparison of tissue co concentrations. one recent study (frank et al. 2004), cites co/ in observed “moose sickness” in nova scotia. however, recent investigations into causes of morbidity and mortality in moose from the mainland population have not substantiated these conclusions as gross or microscopic (beazley et al. 2006). median concentrations of co reported by frank et al. (2004) were 0.09 g/g dw (range = 0.01 – 0.29) in liver and 0.06 g/g dw (range = 0.01 – 0.19) in kidneys (converted from ww). similar concentrations were found in both moose and deer in the present study in mainland and cape breton populations (table 1). based on criteria for domestic cattle (puls 1994, radostits et al. 2000), of the moose livers sampled by frank et al. (2004), 7/17 (< 0.02 g/g wet weight) compared to 29% of moose livers and 7% of deer livers in the g/g g/g dw), respectively, 3.7% and 11.1% of deer g/g dw) (puls 1994). association between age (in months) and se concentrations in liver (r = 0.49, p<0.05). male deer had higher concentrations of se in liver (geometric mean = 1.70 g/g dw, 95% ci = 1.33 – 2.17, n = 18) compared to females (1.28 g/g dw, 95% ci = 1.13 – 1.45, n = 33; t = 2.36, p = 0.011). zinc – in relation to reference values for domestic cattle, 58.3% and 12.5% of moose g/g dw) g/g dw), respectively, and 37.0% and 35.2% of deer (56 – 70 g/g dw) (puls 1994). discussion based on our analyses of tissue concenor toxicities of trace elements in nova scotia moose or deer populations is lacking. moose had higher liver concentrations of cd and cu and lower co, mn, and se levels than deer. in general, trace element concentrations among mainland and cape breton populations were similar; however, some regional differences were seen in moose for cd, co, cu, and ni. caution must be used when making these comparisons as many trace elements have been shown to vary with age, gender, season, locality, and health status. compared with cattle, it was found that co, cu, mn, se, and zn levels in some nova trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 70 present study. liver cobalt levels in both moose and deer in nova scotia appear to be low compared to moose in the yukon (mean = 0.46 g/g dw, sd = 0.28; converted from ww) (gamberg et al. 2005), and sweden (median = 0.42 g/g dw, range = 0.26 – 0.60; converted from ww) (frank et al. 2000). ciency in nova scotia moose made by frank et al. (2004) appears to have been based on historical reports of diseased animals with signs of weakness, emaciation, and neurological lesions, and they cite mcburney et al. (2001) to support the diagnosis. however, the moose mcburney et al. (2001) described with neurological disease of uncertain etiology were in excellent body condition in contrast to the emaciation described by frank et al. (2004). also, any emaciated moose diagnosed with neurological disease in nova scotia have had parelaphostrongylosis and meningoencephalitis to account for their debilitated physical condition (beazley et al. 2006). therefore, nova scotia moose as presented by frank et al. (2004) appears to be unsubstantiated. as the levels in both moose and deer in nova scotia appear to be lower than those in other areas, continued monitoring of co levels in these species is warranted. this could be accomplished through analysis of serum and liver samples from both healthy and diseased animals throughout the province. as supis lacking at this point in time, it is strongly recommended that data also be collected on the health status, particularly body condition, of sampled animals to help determine if an association is present between low liver co levels and poor health in moose populations in nova scotia. copper been reported in free-ranging moose from alaska (flynn et al. 1977, o’hara et al. 2001), northwest minnesota (custer et al. 2004), and sweden (frank et al. 1994, frank 1998). in general, liver cu concentrations in the present study appeared to be adequate to high, although a small number of moose and deer relative to reference values for domestic cattle. cape breton moose had lower liver cu levels than mw moose and among age groups, yearlings had lower liver cu concentrations than adults. comparable levels of cu in liver were found in other studies on apparently healthy moose from norway (means = 80.5 354 g/g dw; converted from ww) (frøslie et al. 1984) and the yukon (mean = 141.1 g/g dw, sd = 167.0; converted from ww) (gamberg et (frank et al. 1994, o’hara et al. 2001, custer et al. 2004) had lower liver cu levels than moose and deer in the present study. free-ranging cervids include antler deformities (gogan et al. 1988), faulty hoof keratinization and reduced reproductive rates (flynn et al. 1977), and diarrhea, anorexia, emaciation, osteoporosis, and loss of hair colour (frank et al. 1994, frank 1998). at the population level, custer et al. (2004) found an association between lower liver cu concentrations in moose and reduced calf-to-cow ratios in minnesota. in general, young animals and fetuses are (smith 1990, radostits et al. 2000) notably, yearling moose in the present study had lower liver cu concentrations than were found in only a small number of animals. since antler and hoof deformities have been reported in nova scotian moose (beazley et al. 2006), continued monitoring of the cu status of healthy and diseased moose, particularly yearlings, may be of value. where samples for sulfur and molybdenum may aid alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 71 in diagnosis and interpretation of cu results. zinc zinc concentrations in nova scotia moose and deer appear on the lower end of, or lower than, the range of values reported in free-ranging cervids elsewhere. in relation to reference values for domestic cattle, approximately 70% of moose and deer livers concentrations of zn. liver zn concentrations from moose and deer in the present study were lower than concentrations in moose from a previous study on moose in nova scotia (median = 161 g/g dw; converted from ww) (frank et al. 2004) from the yukon (mean = 122.0 g/g dw; converted from ww) (gamberg et al. 2005), northern alaska (mean = 232.8 g/g dw) (o’hara et al. 2001), northwestern minnesota (geometric means = 167 g/g dw and 219 g/g dw) (custer et al. 2004) and were similar to the lower range of concentrations found in moose from norway (means = 73.5 112.0 g/g dw; converted from ww) (frøslie et al. 1984) and caribou (rangifer tarandus) from northern alaska (geometric means = 77.0 246.4 g/g dw; converted from ww]) (o’hara et al. 2003) and the northwest territories (means = 75.8 114.1 g/g dw) (elkin and bethke 1995). it is unclear whether the lower levels seen in individual animals, which generally manifest as hair and skin lesions (radostits et al. 2000), or at the population level. however, in moose or deer in nova scotia to date. zinc may also interact with other trace elements including cd and cu (puls 1994). continued monitoring of zn levels in moose may be of value, particularly in debilitated or diseased animals. manganese based on reference values for domestic cattle, approximately 30% of deer liver samples and 45% of moose liver samples had tions. however, it appears unlikely that mn moose in nova scotia as mn levels in the present study were similar to those of apparently healthy caribou from the northwest territories (liver means = 8.6 15.9 g/g dw) (elkin and bethke 1995) and moose from northwestern minnesota (liver geometric means = 7.9 and 8.0 g/g dw) (custer et al. 2004). selenium in the present study, moose had lower liver se concentrations than deer. based on reference values commonly reported for cattle, 15% of deer liver samples and 73% of moose 1994, radostits et al. 2000). overall, tissue se concentrations in moose and deer in the present study were similar to those reported in cervid populations elsewhere. liver se concentrations in deer and moose were similar to those of caribou and reindeer from greenland (geometric means = 0.30 3.4 g/g dw; converted from ww) (aastrup et al. 2000), and moose from norway (means = 0.28 3.2 g/g dw; converted from ww) (frøslie et al. 1984), northwestern minnesota (geometric means = 1.2 and 2.7 g/g dw) (custer et al. 2004), and sweden (medians = 0.33 1.02 g/g dw; converted from ww) (galgan and frank 1995). soils throughout eastern north america, including nova scotia, are known to be generhealth is complex. in domestic animals, seen in juveniles in the form of weakness, muscle stiffness, inability to stand, and possible sudden death (nutritional muscular dystrophy) (radostits et al. 2000). in a study of free-ranging black-tailed deer (odocoileus hemionus columbianus) in northern california, flueck (1994) found that pre-weaning fawn trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 72 survival increased with se supplementation of adult females compared to unsupplemented controls. ciencies of se may have on the health of nova scotia moose and/or deer either directly or through interaction with other trace elements, however continued monitoring of se levels in these species appears warranted. where observed) concentration in liver is a better indicator of se status than kidney. cadmium cadmium concentrations in moose kidneys analyzed in the present study are among the highest reported for the species. although the highest concentrations found were in adults from the mw region, calves and yearlings from all regions sampled had high kidney concentrations when compared to those found in young animals in other studies. as discussed in other studies, age (glooschenko et al. 1988, paré et al. 1999, o’hara et al. 2001), gender (scanlon et al. 1986, gustafson et al. 2000), geographical location (paré et al.1999), species (gamberg and scheuhammer 1994), and season (crête cd concentrations. thus, direct comparison of tissue concentrations among studies can caution. relative to other regions, overall geometric mean kidney concentrations in moose from the present study are higher than those reported in newfoundland (mean = 19.3 g/g dw; converted from ww) (brazil and ferguson 1989), new england (geometric means by age and gender = 31.3 41.8 g/g dw; converted from ww) (gustafson et al. 2000), maine (mean = 23.8 g/g dw), and norway (means by geographical location = 8.4 20.5 g/g dw) (scanlon et al. 1986), and similar to or lower than mean kidney concentrations reported in québec (means by gender and geographical location = 31.8 100.0 g/g dw, crête et al. 1987; mean = 72.4 g/g dw, paré et al. 1999), new brunswick (means by county = 31.8 225.8 g/g dw; converted from ww) (ecobichon et al. 1988), ontario (means by age group = 2.1 179.9 g/g dw; converted from ww) (glooschenko et al. 1988), alaska (geometric means by geographical location = 4.9 70.7 g/g dw; converted from ww) (arnold et al. 2006), and the yukon (mean = 98.4 g/g dw; converted from ww) (gamberg et al. 2005). in the present study, kidney cd concentrations in moose calves (16.6 and 83.3 g/g dw, n = 2) and yearlings (geometric mean 54.4 g/g dw, 95% ci = 25.0 – 83.7, n = 6) are among the highest reported for these age groups (gamberg et al. 2005), however, the number of young moose sampled was low. comparable levels of kidney cd were reported in moose calves and yearlings sampled from the area surrounding the community of rouynnoranda, québec (paré et al. 1999), and in the yukon (gamberg et al. 2005). paré et al. (1999) attributed the high level of cd contamination found in moose of all ages surrounding rouyn-noranda to natural mineralization of the soil and historical and recent human contributions of cd to the environment (including a cu smelter). gamberg et al (2005) concluded that high levels of cd in yukon moose are a result of naturally occurring geological sources, likely via the ingestion of cd-accumulating plants such as willow, a preferred browse species for moose (renecker and schwartz 1998). other explanations put forth for regional differences seen in cd accumulation in free-ranging cervids also include: differences in buffering capacity of the soil, the et al. 1986, crête et al. 1987, glooschenko et al. 1988), the composition and diversity of forage species in an area (ohlson and staaland 2001), and other habitat differences (custer et al. 2004). in relation to the present study, a discussion of bedrock and soil composition alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 73 in relation to cd levels in moose, porcupines, and willow spp. in nova scotia can be found in roger (2002). liver cd concentrations in deer in the present study were similar to those found in white-tailed deer in other regions of north america (ecobichon et al. 1988, glooschenko et al. 1988, crichton and paquet 2000). the primary target of chronic cd toxicity in mammals and birds is the kidney (scheuhammer 1987, alden and frith 1991) and the earliest light microscopic change in this organ is proximal tubular necrosis (alden and frith 1991). in addition to the renal toxicity, there is also evidence that exposure to cd can result in disturbances in calcium balance and decreases in bone density (taylor et al. 1999, larison et tubular damage in mammals and birds from cd accumulation is generally reported as 100 – 200 g/g ww (approximately 350 – 700 g/g dw) (cooke and johnson 1996), although a renal threshold of 30 g/g ww (approximately 105 g/g dw) for mammals has also been published (outridge et al. 1994). among published studies on cd in ungulates, evidence for biological effects associated with high concentrations is lacking. paré et al. (1999) found no lesions characteristic of renal disease in 33 kidney samples with a mean cd concentration of 123.1 (± 17.98) g/g dw submitted for histopathological examination. o’hara et al. (2003) found no histopathologic evidence of renal lesions in caribou kidneys with cd concentrations of 1.9 – 115.5 g/g dw (converted from wet weight) from northern alaska. kidneys from two moose sampled in the present study (cd concentrations 96.1 and 346.2 g/g dw) were submitted to the atlantic veterinary college diagnostic services for histopathological examination and no evidence of renal proximal tubular lesions was found, although autolysis and freezing artifacts may have masked subtle changes in the tissues (s. mcburney, canadian cooperative wildlife health centre, personal communication). it may be that large ungulates such as moose that tend to accumulate high levels of cd are less susceptible to the toxic effects of cd than experimental animals; however, the possibility that elevated cd concentrations in individual animals may lead to subclinical or clinical disease cannot be dismissed. in the future, monitoring of moose tissue cd concentrations in nova scotia should be carried out in conjunction with detailed health assessment of individuals including histopathological examination of kidneys. conclusions and recommendations a good foundation of trace element data has been collected for moose and deer in nova scotia, however, some age groups and regions are under-represented. some differences were found in trace element status between the mainland and cape breton moose populations which warrant continued monitoring. tissue concentrations of trace elements in moose and deer in nova scotia appear to be generally similar to levels reported elsewhere in north america and europe with the exception of zn and co. although kidney cd concentrations in some nova scotia moose are high, particularly in the mw region, similar or higher concentrations have also been reported elsewhere. to date, individual or population-level health effects in relation to elevated tissue cd levels have not been reported in free-ranging ungulates in north america, including nova scotia. in relation to reference values for domestic co, mn, and zn were found in some moose and deer in nova scotia. at the present time, there appears to be little supporting evidence elements are occurring in nova scotia moose or deer populations. however, the possibilor other trace elements and high levels of cd trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 74 may impact the health of individual animals either directly or through interactions with other factors that may be contributing to the decline of the mainland moose population (e.g., infectious and non-infectious diseases, harsh environmental conditions, habitat limitations) cannot be dismissed. continued monitoring of trace element concentrations in both populations of nova scotia moose is of value because of the endangered status of the mainland population. future monitoring efforts should include relevant demographic data and health assessment for all sampled animals in order to help establish “normal” values for these populations and to identify possible health effects in (co, cu, se, and zn in particular) or toxicities (cd). for each animal sampled, demographic data including geographical location, age, gender, and species are essential. as well, body condition should be assessed as many disease conditions related to trace element tion or emaciation. the timely necropsy of debilitated or diseased animals is also strongly recommended. collection of this supporting data is highly recommended and is essential for meaningful interpretation of trace element data collected in the future. acknowledgements we gratefully acknowledge the cooperation of hunters and employees of the nova scotia department of natural resources (nsdnr) from throughout the province for providing the samples. ken eagle (new brunswick department of natural resources) helped with age estimates for the sampled animals in addition to the tooth cementum analysis. the trace element analyses of tissue samples were completed at and funded by the environmental quality laboratory (eql), environment canada, moncton, new brunswick. the technical assistance provided by marc bernier and kim davis during the sample analyses was very much appreciated as was the help with follow up questions and the eql. tony nette from the nsdnr and pierre-yves daoust and scott mcburney from the canadian cooperative wildlife health centre (ccwhc), atlantic region, were instrumental in realizing the completion of the paper and provided helpful input and editorial comments throughout. thanks also to s. taylor for his assistance with the maps. funding for the data analysis and completion of the manuscript was generously provided by the ccwhc and nsdnr. references aastrup, p., f. riget, r. dietz, and g. asmund. 2000. lead, zinc, cadmium, mercury, selenium and copper in greenland caribou and reindeer (rangifer tarandus). science of the total environment 245:149-159. alden, c. a., and c. h. frith. 1991. urinary system. pages 333-361 in w. m. haschek and c. g. rousseaux, editors. handbook of toxicologic pathology. harcourt brace jovanovich, san diego, california, usa. arnold, s., r. zarnke, t. lynn, m.-a. chimonas, and a. frank. 2006. public health evaluation of cadmium concentrations in liver and kidney of moose (alces alces) from four areas of alaska. science of the total environment 357:103-111. beazley, k., m. ball, l. isaacman, s. mc-l. isaacman, s. mc-s. mcburney, p. wilson, and t. nette. 2006. complexity and information gaps in recovery planning for moose (alces alces americana) in nova scotia, canada. alces 42:89-109. brazil, j., and s. ferguson. 1989. cadmium concentrations in newfoundland moose. alces 25:52-57. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic varialces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 75 ability of moose, alces alces, in canada. molecular ecology 8:1309-1315. cooke, j. a., and m. s. johnson. 1996. cadmium in small mammals. pages 389-404 in w. n. beyer, g. h. heinz, and a. w. redmon-norwood, editors. environmental contaminants in wildlife: interpreting tissue concentrations. lewis publishers, crc press, incorporated, boca raton, florida, usa. crête, m., r. nault, p. walsh, j.-l. benedetti, m. lefebvre, j.-p. weber, and j. gagnon. 1989. variation in cadmium content of caribou tissues from northern québec. science of the total environment 80:103-112. _____, p. potvin, p. walsh, j.-l. benedetti, m. lefebvre, j.-p. weber, g. paillard, and j. gagnon. 1987. pattern of cadmium concentration in the liver and kidneys of moose and white-tailed deer in québec. science of the total environment 66:4553. crichton, v., and p. paquet. 2000. cadmium in manitoba’s wildlife. alces 36:205216. custer, t. w., e. cox, and b. gray. 2004. trace elements in moose (alces alces) found dead in northwestern minnesota, usa. science of the total environment 330:81-87. ecobichon, d. j., r. hicks, and g. redmond. 1988. a survey of cadmium concentrations in liver and kidney of deer and moose in new brunswick. department of natural resources, fish and wildlife branch, fredericton, new brunswick, canada. elkin, b., and r. w. bethke. 1995. environmental contaminants in caribou in the northwest territories, canada. science of the total environment 160-1:307-321. flueck, w. t. 1994. selenium deficiency in free-ranging black -tailed deer. ecology 75:807-812. flynn, a., a. w. franzmann, p. d. arneson, and j. l. oldemeyer. 1977. indications of copper deficiency in a subpopulation of alaskan moose. journal of nutrition 107:1182-1189. frank in sweden; similarities to copper deficiency and/or molybdenosis in cattle and sheep; biochemical background of clinical signs and organ lesions. science of the total environment 209:17-26. _____, r. danielsson, and b. jones. 2000. the concentrations of trace elements in liver and kidneys and clinical chemistry; comparison with experimental molybdenosis and copper deficiency in the goat. science of the total environment 249:107-122. _____, v. galgan, and l. r. petersson. 1994. secondary copper deficiency, chromium deficiency and trace element imbalance in the moose (alces alces l.): effect of anthropogenic activity. ambio 23:315-317. _____, j. mcpartlin, and r. danielsson. 2004. nova scotia moose mystery--a moose sickness related to cobalt and vitamin b12 deficiency. science of the total environment 318:89-100. frøslie, a., a. haugen, g. holt, and g. norheim. 1986. levels of cadmium in liver and kidneys from norwegian cervides. bulletin of environmental contamination and toxicology 37:453-460. _____, g. norheim, j. p. rambaek, and e. steinnes. 1984. levels of trace elements in liver from norwegian moose, reindeer and red deer in relation to atmospheric deposition. acta veterinaria scandinavia 25:333-345. galgan, v., and a. frank. 1995. survey of bioavailable selenium in sweden with the moose (alces alces l.) as monitoring animal. science of the total environment 172:37-45. gamberg, m., m. palmer, and p. roach. 2005. temporal and geographic trends trace elements in moose in nova scotia – pollock and roger alces vol. 43, 2007 76 in trace element concentrations in moose from yukon, canada. science of the total environment 351-352:530-538. _____, and a. m. scheuhammer. 1994. cadmium in caribou and muskoxen from the canadian yukon and northwest territories. science of the total environment 143:221-234. glooschenko, v., r. downes, r. frank, h. e. braun, e. m. addison, and j. hickie. 1988. cadmium levels in ontario moose and deer in relation to soil sensitivity to acid precipitation. science of the total environment 71:173-186. gogan, p. j., d. a. jessup, and r. h. barrett. 1988. antler anomalies in tule elk. journal of wildlife diseases 24:656-662. gustafson, k., k. m. bontaites, and a. major. 2000. analysis of tissue cadmium concentrations in new england moose. alces 36:35-40. larison, j. r., g. e. likens, j. w. fitzpatrick, and j. g. crock. 2000. cadmium toxicity among wildlife in the colorado rocky mountains. nature 406:181-183. mcburney, s., d. brannen, and t. nette. 2001. neurological disease in moose in nova scotia. canadian cooperative wildlife health centre newsletter 8(1). saskatoon, saskatchewan, canada. o’hara, t. m., g. carroll, p. barboza, k. mueller, j. blake, v. woshner, and c. willetto. 2001. mineral and heavy metal status as related to a mortality event and poor recruitment in a moose population in alaska. journal of wildlife diseases 37:509-522. _____, j. c. george, j. blake, k. burek, g. carroll, j. dau, l. bennett, c. p. mccoy, p. gerard, and v. woshner. 2003. investigation of heavy metals in a large mortality event in caribou of northern alaska. arctic 56:125-135. ohlson, m., and h. staaland. 2001. mineral diversity in wild plants: benefits and bane for moose. oikos 94:442-454. outridge, p. m., d. d. macdonald, e. porter, and i. d. cuthbert. 1994. an evaluation of the ecological hazards associated with cadmium in the canadian environment. environmental review 2:91-107. paré, m., r. prairie, and m. speyer. 1999. variations of cadmium levels in moose tissues from the abitibi-temiscamingue region. alces 35:177-190. parker, g. 2003. status report on the eastern moose (alces alces americana clinton) in mainland nova scotia. an independent commissioned report for nova scotia department of natural resources, kentville, nova scotia, canada. puls, r. 1994. mineral levels in animal health: diagnostic data. second edition. sherpa international, clearbrook, british columbia, canada. pulsifer, m. d., and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31:209219. radostits, o. m., c. c. gay, d. c. blood, and k. w. hinchcliff. 2000. veterinary medicine: a textbook of the diseases of cattle, sheep, pigs, goats and horses. ninth edition. harcourt publishers limited, london, u.k. renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behaviour. pages 403-440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. roger, k. e. 2002. plant and mammal tissue cadmium concentrations in nova scotia, and possible effects on moose (alces alces) populations. b.sc. honors thesis. acadia university, wolfville, nova scotia, canada. scanlon, p. f., k. i. morris, a. g. clark, n. fimreite, and s. lierhagen. 1986. cadmium in moose tissues: comparison of data from maine, usa and from telemark, alces vol. 43, 2007 pollock and roger trace elements in moose in nova scotia 77 norway. alces 22:303-312. scheuhammer, a. m. 1987. the chronic toxicity of aluminum, cadmium, mercury and lead in birds: a review. environmental pollution 46:263-295. smith, b. p. 1990. large animal internal medicine: diseases of horses, cattle, sheep, and goats. the c.v. mosby company, st. louis, missouri, usa. taylor, j., r. dewoskin, and f. k. ennever. 1999. toxicological profile for cadmium (update). agency for toxic substances and disease registry, u.s. department of health and human services, atlanta, georgia, usa. 4214(111-114).pdf alces vol. 42, 2006 base et al. status of moose in washington 111 history, status, and hunter harvest of moose in washington state dana l. base1, steve zender1, and donny martorello2 1washington department of fish and wildlife, 2315 north discovery place, spokane, wa 99216, usa; 2washington department of fish and wildlife, 600 capitol way north, olympia, wa 98501, usa abstract: since the middle 20th century, moose have expanded their range and population in washington, especially within the northeastern part of the state. the washington department of fish and wildlife opened a limited-entry hunting season on moose in 1977. permit numbers gradually increased from 3 in 1977 to 98 permits offered in the 2005 hunting season. hunter harvest is believed to be well within the reproductive capacity of washington’s moose population. moose abundance and range are expected to at least remain at current levels into the future. alces vol. 42: 111-114 (2006) key words: alces alces, antler widths, bull / cow / calf ratio, limited-entry hunting, management goals and guidelines, population status, range, tooth cementum aging, washington until the early 1970s there were few records for moose (alces alces) within the state included a photograph of an adult bull taken by hunter pete lemery on november 16, 1929 near twin lakes in ferry county, washington on the colville indian reservation (scheffer and dalquest 1944). in 1954, washington department of game (later renamed washington department of fish and wildlife, wdfw) personnel found a shed moose antler in the selkirk mountains of pend oreille county in the northeastern corner of the state. the following year, 1955, two agency biologists found the carcass of a calf moose in the same general vicinity (s. guenther, wdfw, unpublished data). by 1972 a well-established resident population of moose was documented in pend oreille county that consisted of an estimated 60 animals (poelker 1972). this population grew to 850-1,000 animals over the next 30 years and greatly expanded in range (wdfw 2003). the assumed subspecies of moose in washington is shira’s, alces alces shirasi, as this subspecies comprises the closest moose population to washington in both idaho and british columbia (poelker 1972, compton and oldenburg 1994). figure 1 illustrates the estimated range of moose as of 1997 based upon modeling accomplished by johnson and cassidy (1997). moose are still expanding in distribution within washington as numerous range in northeastern washington have been made since 1997. in 1977, the washington state wildlife hunt of moose within the state. three tags fig. 1. range of moose in washington state, usa, as of 1997 (indicated by shaded area: from johnson and cassidy 1997). status of moose in washington – base et al. alces vol. 42, 2006 112 issued by lottery-type drawing were awarded that year. as both the population and range of moose have expanded since 1977, the number of special hunt permits has gradually increased to a high of 98 permits in 2005 (fig. 2). on a statewide basis the wdfw has the following goals for managing moose: 1. preserve, protect, perpetuate, and manage moose and their habitats to ensure healthy, productive populations. 2. manage moose for a variety of recreational, educational, and aesthetic purposes and ceremonial uses by native americans, wildlife viewing, and photography. 3. manage statewide moose populations for a sustained yield (wdfw 2003). in 2003, the wdfw developed guidelines for managing the hunter harvest of moose in washington (table 1). these guidelines are generally averaged over a 3-year period (1997). management philosophy is directed at providing a high-quality hunting experience with good opportunity for harvesting a mature bull. field observations, aerial surveys, hunter success rates, antler widths, and moose ages (compton and oldenburg 1994). as the range of moose has expanded, the number of game management units (gmus) with allocated moose permits has increased from 1 in 1977 to 10 in 2004. likewise, of the 39 counties within washington state, the number in which moose can be hunted has increased from 1 in 1977 to 6 in 2005. the annual hunter harvest success rate on both bull and cow moose has been consistently high, ranging from 67% to 100% with an average of 92% and a mode of 100%. a total of 748 moose were legally harvested between 1977 and 2005, including 556 bulls and 192 antlerless moose (cows and calves). the annual average age of harvested bull moose as determined by tooth cementum analysis was 5.2 years (range 3.9 6.9) from 1990 through 2004 (n = 373 ) (fig. 3). the oldest bull moose taken by hunters in washington was aged at 15.4 years. this bull was harvested in 2003. the average antler spread of harvested bulls from 1990 through 2005 was 94 cm (37 inches) 0 10 20 30 40 50 60 70 80 90 100 110 19 77 19 79 19 81 19 83 19 85 19 87 19 89 19 91 19 93 19 95 19 97 19 99 20 01 20 03 20 05 year n um be r permits harvest fig. 2. allocation of permits and hunter harvest of moose in washington state, usa, 1977 – 2005. 0 20 40 60 80 100 120 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 20 00 20 01 20 02 20 03 20 04 20 05 year m ea n an tl er sp re ad (c m ) 0 1 2 3 4 5 6 7 8 m ea n ag e mean spread mean age fig. 3. annual average antler spread (cm) and age of hunter-harvested bull moose in washington state, usa, from 1990 to 2005. guideline liberalize harvest level acceptable harvest level restrict harvest access average bull : 100 cow ratio > 75 bulls 60 – 75 bulls < 60 bulls average calf : 100 cow ratio > 50 calves 30 – 50 calves < 30 calves median age of harvested bulls > 6.5 years 4.5 – 5.5 years < 4.5 years table 1. guidelines for managing the hunter harvest of moose in washington state, usa. alces vol. 42, 2006 base et al. status of moose in washington 113 with an annual mean ranging between 84 and 104 cm (33 – 41 inches; n = 440 ) (fig. 3). the widest antler spread of any hunter-harvested moose in washington was 147 cm (58 inches) from a bull taken in 2000. bull and calf moose ratios as determined from early winter helicopter surveys ranged from 63 to 128 bulls and 26 to 74 calves per 100 cows from 1994 through 2005 (table 2). the calf ratio appears to be indicative of a stable to increasing population. percentages of bull moose tallied by age class using criteria outlined by timmermann (1993) and bubenik et al. (1977) has shown fairly equal proportions of adult and sub-adult bulls since 2000 (fig. 4). in addition there has been an increase in the proportion of yearling bulls since 2000, probably indicative of a moose population continuing to grow. references bubenik, a. b., o. williams, and h. r. timmermann. 1977. visual estimation of sex and social class in moose (alces alces) from the ground and the plane. a preliminary study. proceedings of the north american moose conference and workshop 13: 157-176. compton, b. b., and l. e. oldenburg. 1994. the status and management of moose in idaho. alces 30: 57-62. courtois, r., and g. lamontagne. 1997. management systems and current status of moose in quebec. alces 33: 97-114. johnson, r. e., and k. m. cassidy. 1997. terrestrial mammals of washington state. location data and predicted distributions. volume 3 in k. m. cassidy, c. e. grue, m. r. smith, and k. m. dvornich, editors. washington state gap analysis – final report. washington cooperative fish year number seen during survey bulls:100 cows calves:100 cowsbulls cows calves 1994 14 17 5 82 29 1995 17 20 6 85 30 1996 17 24 8 71 33 1997 58 65 21 89 32 1998 33 47 12 70 26 1999 27 36 22 75 61 2000 55 59 29 93 49 2001 31 49 17 63 35 2002 59 46 34 128 74 2003 62 63 35 98 56 2004 39 47 21 83 45 2005 34 48 20 71 42 table 2. bull and calf moose ratios per 100 cows as determined from early winter helicopter surveys in washington state, usa, 1994 – 2005. 0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% 2000 2001 2002 2003 2004 2005 year pe rc en t % yearlings % sub-adults % adults fig. 4. percentages of bull moose tallied by age winter helicopter surveys in washington state, usa, 2000 2005. the sample size (n) of total of each bar. 55 31 59 62 39 34 status of moose in washington – base et al. alces vol. 42, 2006 114 and wildlife research unit, university of washington, seattle, washington, usa. poelker, r. j. 1972. the shiras moose in washington. unpublished administrative report. washington department of game, olympia, washington, usa. scheffer, v. b., and w. w. dalquest. 1944. records of mountain goat and moose from washington state. journal of mammalogy 25: 412-413. timmermann, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29: 35-46. (wdfw) washington department of fish and wildlife. 2003. moose, game management plan, wildlife program. washington department of fish and wildlife, olympia, washington, usa. alces37(1)_55.pdf alces37(1)_97.pdf alces36_183.pdf 4104.p65 alces vol. 40, 2004 cobb et al. sympatric moose and deer 169 relative spatial distributions and habitat use patterns of sympatric moose and white-tailed deer in voyageurs national park, minnesota mccrea a. cobb1,2, peter j.p. gogan3, karin d. kozie4, edward m. olexa3, rick l. lawrence5, and william t. route6 1department of ecology, montana state university, bozeman, mt 59717, usa; 3usgs – northern rocky mountain science center, forestry sciences laboratory, montana state university, bozeman, mt 59717, usa; 4voyageurs national park, 3131 highway 53 south, international falls, mn 56649, usa; 5 department of land resources and environmental sciences, montana state university, bozeman, mt 59717, usa; 6national park service, great lakes network office, 2800 lake shore drive, ashland, wi 54806, usa abstract: we examined the distribution and home range characteristics of moose (alces alces) and white-tailed deer (odocoileus virginianus) at voyageurs national park, minnesota. pellet count transects revealed low densities of moose and higher densities of white-tailed deer, and provided evidence of partial spatial segregation between moose and white-tailed deer possibly due to habitat heterogeneity. there was limited interspecific overlap in the relatively large annual home ranges of radio-collared moose and white-tailed deer. both moose and white-tailed deer exhibited significant selection for spruce (picea spp.) and balsam fir (abies balsamea) vegetation types at the home range scale. white-tailed deer significantly selected a 12-20 m canopy height over all others while moose significantly selected 5-11 m and 21-30 m canopy heights over the 12-20 m canopy height. moose significantly selected open/discontinuous canopy cover and white-tailed deer selected both closed/continuous and open/discontinuous canopy covers over dispersed/ sparse canopy cover. differential habitat selection between moose and white-tailed deer at voyageurs national park might be related to the differences between these species' abilities to cope with a northern mid-continental climate. spatial segregation between moose and white-tailed deer at voyageurs national park may allow moose to persist despite the presence of meningeal worm (parelaphostrongylus tenuis) in white-tailed deer. alces vol. 40: 169-191 (2004) key words: alces alces, compositional analysis, ecology, home range, meningeal worm, moose, odocoileus virginianus, parelaphostrongylus tenuis, pellet groups, sympatric, whitetailed deer 2present address: department of environmental science, policy, and management, university of california, berkeley, ca 94720-3114, usa moose inhabit a circumpolar region of northern boreal forests dominated by spruce (picea spp.), pine (pinus spp.), and fir (abies spp.). the range of moose in north america has expanded since 1955 (peterson 1955) while numbers throughout the range increased from approximately 940,000 to 975,000 between 1960 and 1990 (karns 1998). moose numbers in minnesota increased eight fold from approximately 1,500 animals in 1960 to 12,000 in 1990 (karns 1998). numbers of moose in northern minnesota may have peaked prior to the 1990 estimates as moose abundance in adjacent ontario started to decline in the mid-1980s (thompson and euler 1987). moose at sympatric moose and deer – cobb et al. alces vol. 40, 2004 170 voyageurs national park (vnp), minnesota, are at the southern periphery of the species’ north american range and within a low density range between two high moose density ranges to the northeast and northwest (fuller 1986). in the last 10 years, the northwestern moose population has dropped dramatically and now only numbers a few hundred animals (m. lenarz, minnesota department of natural resources, personal communication). white-tailed deer (wtd) expanded their distribution into northern minnesota around 1900 and were common in the area by the 1920s (petraborg and burcalow 1965). wtd densities in northeastern minnesota were estimated at between 6 and 8/km2 in the late 1930s (olson 1938, petraborg and burcalow 1965). moose and wtd are sympatric across a relatively narrow band of north america, and the species’ habitat use patterns within this band are not well understood. the two species are thought to have occurred sympatrically in the area that is now vnp at least since the early 1930s (cole 1987, gogan et al. 1997). fluctuations in wtd numbers might be due to changes in habitat and changes in moose population levels have been attributed to changes in vegetative types and seral stages (mech and karns 1978, cole 1987). following the 1971 little sioux fire in adjacent superior national forest, minnesota, moose densities increased to five times their previous number (neu et al. 1974, peek et al. 1976). both moose and wtd were found to consume similar browse after a fire in northern minnesota (irwin 1975). active fire suppression within vnp has limited recent natural disturbances, and together with logging restrictions, affected the vegetation composition. the low frequency of wildland fires since the establishment of vnp could be a factor contributing to relatively low densities of moose. in the absence of specific information on moose and wtd distributions and habitat use patterns at vnp, the relationship of each species to vegetative conditions remains unclear. parasite mediated competition between moose and wtd might be responsible for recent declines in moose numbers. meningeal worm, a parasite that is characteristically benign in wtd but fatal in moose, has been attributed for moose declines in minnesota and elsewhere (karns 1967, prescott 1974). the extent to which meningeal worm impacts moose abundance at vnp is an unresolved issue. spatial separation and differential habitat selection between moose and wtd may allow moose to persist in the presence of infected wtd (gilbert 1974). wtd in nova scotia were excluded from some habitats at high elevation by snow depth, providing moose with refuges from wtd during the winter season (telfer 1967). a “refugium” between moose and wtd in ontario was identified as a possible factor allowing moose to persist in the presence of sympatric populations of meningeal worminfected wtd (kearney and gilbert 1976). questions however have been raised concerning the validity of the refugia hypothesis. the purported benefits of seasonal refugia for moose in warmer months when the potential infection rate is highest might not exist because moose and wtd habitat use overlapped during other times of the year (nudds 1990). even partial refugia from infected wtd however may provide moose with a relative advantage (whitlaw and lankester 1994). in the absence of specific information on moose and wtd distributions and habitat use patterns at vnp, the relationship of the abundance of each species to vegetative conditions and the potential for moose refugia from meningeal worm infection remains unclear. this study was initiated to determine the relative spatial distribution and home range characteristics of moose and wtd at vnp, and to examine the alces vol. 40, 2004 cobb et al. sympatric moose and deer 171 influence of habitat types on these distributions, and to assess overlapping use patterns of the two species. study area vnp encompasses 882 km2 on the southern portion of the canadian shield along the u.s.-canada border. vnp is made up of a central landmass largely surrounded by lakes, called the kabetogama peninsula, and adjacent lands. approximately 40% of vnp is covered by 4 large lakes. there is little overall elevation change with a maximum topographic relief of 80-90 m (johnson and sales 1995). adjacent areas in minnesota include lands administered by the state (kabetogama state forest), the federal government (superior national forest), and privately owned lands. adjacent areas of ontario are mainly provincial crown lands. the study area boundary was defined as the area encompassed within the gis vegetation coverage (usgs 2001) of the vnp region (fig. 1). # international falls # rainy lake # namakan lake # crane lake # kabetogama lake # sandy point lakeunited states canada # kabetogama peninsula n ew s 5 0 5 10 kilometers # # # minnesota intern ational falls duluth twin cities water vnp boundary u.s.canada border habitat cover boundary fig. 1. location of voyageurs national park, northern minnesota. climate the climate is characterized as cold winters and cool summers. temperature extremes during the study period ranged from 35oc (august 1, 1989) to -39oc (december 30, 1990, national weather service, international falls, mn). average annual snowfall is 160 cm, with the most snowfall occurring during january (31 cm). the winter of 1988 – 1989 was a high snow year with 266 cm. snowfall in the winter of 1989 – 1990 was 155 cm, close to the longterm mean, while snowfall in the winter 1991 – 1992 was 247 cm, considerably higher than the mean. the first significant winter snowfall usually occurs in early november, and the last significant snow usually occurs in early april (national weather service, international falls, mn). the north atlantic oscillation (nao) index (lamb and peppler 1987, hurrell 1995) showed that winter temperatures were colder than average during the study period, with the winters of 1988 – 1989 and 1989 – 1990 sympatric moose and deer – cobb et al. alces vol. 40, 2004 172 being particularly cold. vegetation vnp lies on the boundary between southern boreal forest and northern hardwood forest types (pastor and mladenoff 1992). northern hardwood forests are dominated by red pine (pinus resinosa), white pine (p. strobus), red maple (acer rubrum), and black ash (fraxinus nigra) (kurmis et al. 1986). southern boreal forest types are characterized by a mosaic of secondary growth jack pine (p. banksiana), white spruce (picea glauca), quaking aspen (populus tremuloides), paper birch (betula p a p y r i f e r a ) , a n d b a l s a m f i r ( a b i e s balsamea) (kurmis et al. 1986). the soil in the region is thin and sandy (ohmann and ream 1971). varying sources and levels of disturbance have created spatial heterogeneity in the vegetation of the vnp region. logging has been an important influence on the current spatial distribution of vegetation across much of vnp. parts of vnp, including approximately 25% of the kabetogama peninsula, were extensively logged between 1910 and 1930 (crowley and cole 1995). the combined impacts of these harvests decreased the abundances of white spruce, balsam fir, white pine, and red pine on the kabetogama peninsula. the relative abundance of aspen consequently increased to higher levels post-harvest. while logging practices within vnp ceased with the inception of the park in 1975, the majority of forested lands adjacent to the park have continued to be managed for timber harvest. fire suppression efforts began in 1911 and have since limited major fires in the park region to 1917 – 1918, 1923, and 1936. fires burned substantial portions of the kabetogama peninsula in 1923 and 1936, adding to the mosaic of vegetative cover in vnp (fig. 2). human and naturally caused wildfires within the park since 1936 have been relatively small (<2.0 km2). wildlife woodland caribou (rangifer tarandus) and moose are thought to have been the most common ungulates in the vnp region in pre-historic times (cole 1987). wtd expanded northward into the region in the late 1890s and were reported to be common in the region by the 1920s (petraborg and burcalow 1965). woodland caribou were extirpated from the region by the 1940s (gogan et al. 1997). moose possibly declined in numbers from the establishment of the park in the mid – 1970s through the mid – 1980s (cole 1987). approximately 60 – 100 moose were estimated to inhabit vnp at a mean density of 0.23/km2 in the early 1990s (whitlaw and lankester 1994, gogan et al. 1997). estimated densities of wtd in and immediately adjacent to vnp ranged from 1.5/km2 to 11.5/km2 from 1975 (r. o. peterson, michigan technological university, unpublished report, 1976) through 1992 (whitlaw and lankester 1994, gogan et al. 1997). the reasons for the recent variations in moose and wtd population levels are unknown. methods gis vegetation coverage habitat availability of the study area was determined using a geographic information system (gis) vegetation coverage created by interpreting 1:15,840-scale color infrared (cir) aerial photographs taken in 1995 and 1996 (usgs 2001). the entire coverage consisted of 156,886 ha, of which vnp comprised 88,244 ha (56%) of the total. a total of 40 vegetation cover types defined the ground features within the project area. each vegetation cover type was further classified by canopy height and canopy density. for this study, we consolidated the original 40 gis vegetation cover alces vol. 40, 2004 cobb et al. sympatric moose and deer 173 types into 8 classes based on functional groups to facilitate analysis and alleviate the problem of missing habitat types during compositional analysis (table 1). pellet group transects we applied a 1 – km2 grid to a 1:50,000scale map of the park and randomly selected 32 cells as pellet group transect sampling units. two parallel transect lines were established in a north-south orientation with at least 100 m separation between transect lines within most sampling units. sampling units containing >50% water cover (11 of 32, or 34%) were limited to one transect line resulting in a total of 53 transect lines within the 32 sampling units. each transect line was 800 m long and consisted of 4 plots (22 m by 3.6 m) at 200 m intervals. a survey chain (20 m) was used to measure distance traveled while surveying. transect lines were sampled once in late may of 1989 and again in the late may 5 0 5 10 kilometers n ew s burned (1923 and 1936) logging (1950s and 1960s) fig. 2. historical fire and logging locations within voyageurs national park, minnesota. of 1991, after snow melt and prior to the onset of new vegetative growth. all pellet groups within each plot above the previous fall’s leaf litter were identified to species and tallied. we used the total number of moose and wtd pellet groups observed along each transect line in our analysis. we sampled 16 paired and 14 single transect lines in 1989, and 19 paired and 12 single lines in 1991. some sampling units (4 of 32, or 12.5%) were only visited one year due to a lack of personnel and access problems (private land, terrain). to insure that the detection probability was even between sampling units containing 1 and 2 transect lines, we randomly selected a single transect line from sampling units containing 2 transect lines for analysis. numbers of moose and wtd pellet groups detected within each sampling unit by sampling year were entered into a gis for interpretation. for moose, sampling units were stratified into those with pellet sympatric moose and deer – cobb et al. alces vol. 40, 2004 174 table 1. description of vegetation classes used in compositional analysis. consolidated vegetation classes were created by combining original gis habitat classifications from the usgs/nps vegetation coverage of the voyageurs national park region. cons olidat ed clas ses o riginal g is h abit at class ificat ion shrubland a lliance beaked h az el/serviceberry shrubland alliance bog birch/willow sat urat ed shrubland alliance leat herleaf s at urat ed dw arf shrubland alliance red o s ier d ogw ood/willow seas onally flooded s hrubland alliance sp eckled a lder s easonally flooded shrubland alliance n ort hern whit e cedar/red m ap le s at urat ed forest alliance n ort hern whit e cedar/yellow birch fores t alliance n ort hern whit e cedar fores t alliance n ort hern whit e cedar sat urat ed fores t alliance t amarack s at urat ed forest alliance black a sh/red m ap le black a sh/red m ap le sat urat ed fores t alliance jack p ine jack p ine/lichen nonvascular alliance jack p ine forest alliance jack p ine, red p ine w oodland alliance m osaic (jack p ine forest alliance and q uaking a s p en/p ap er birch fores t alliance) red/w hit e p ine red p ine forest alliance whit e/red p ine and q uaking a s p en fores t alliance whit e p ine forest alliance sp ruce/bals am f ir black sp ruce/q uaking a sp en fores t alliance a n d /o r w hit e sp ruce/bals am f ir/a s p en fores t alliance black sp ruce forest alliance black sp ruce s at urat ed forest alliance whit e sp ruce/bals am f ir forest alliance c ons olidat ed c las s es o riginal g is h abit at c las s ificat ion b ur o ak b ur o ak/o ak (w hit e, n ort hern p in, b lack) w oodland alliance b ur o ak fores t alliance a s p en/b irch p ap er b irch fores t alliance q uaking a s p en/p ap er b irch fores t alliance t rembling a s p en t emp orarily flooded fores t alliance q uaking a s p en w oodland alliance h erbaceous a lliance c anada b luejoint s eas onally flooded herbaceous alliance c at t ail/b ulrus h s em ip ermanent ly flooded herbaceous alliance c ommon r eed s emip ermanent ly flooded herbaceous alliance f ew -s eeded/w iregras s sedge s at urat ed herbaceous alliance h ards t em/soft s t em b ulrus h s emip ermanent ly flooded herbaceous alliance m os aic (1 s at urat ed dw arf s hrubland alliance and 3 w et land herbaceous alliances ) m os aic/c omp lex (5 w et land herbaceous alliances ) m os aic/c omp lex (7 w et land herbaceous alliances ) p ondw eed/h ornw ort /w at erw eed p ermanent ly flooded herbaceous alliance p overt y g ras s herbaceous alliance yellow /w hit e w at er lily p ermanent ly flooded herbaceous alliance w ild r ice s emip erm anent ly flooded herbaceous alliance alces vol. 40, 2004 cobb et al. sympatric moose and deer 175 groups “present” and those without pellet groups “absent”. for wtd, sampling units were stratified on the basis of abundance into low (1 – 21 pellet groups) and high (>21 – 42 groups) use areas. we determined the average percent composition of vegetation types, canopy densities, and canopy heights for sampling units within each stratum of moose and wtd from the modified gis vegetation map of vnp. we performed ttests to examine differences between habitat proportions in absent vs. present stratum of moose sampling units and high vs. low stratum of wtd sampling units. capture and radio telemetry we fitted 10 moose (3 bulls, 7 cows) with motion-sensing radio telemetry collars on the kabetogama peninsula between february 26 and march 2, 1989. each moose was immobilized and sedated with a mixture of carfentantil and xylazine hydrochloride via a barbed syringe fired from a helicopter. the immobilizing drugs were reversed with a hand injection of naltrexone. we fitted motion-sensing radio telemetry collars on 20 white-tailed deer (9 bucks, 11 does) within vnp between january 24 and march 9, 1989. thirteen wtd were captured within the moose bay-black bay region, 5 along the daley brook snowmobile trail, and 2 on or adjacent to cutover island. wtd were captured in collapsible clover traps (clover 1956) and immobilized using a pole-mounted syringe with a mixture of ketamine hydrochloride (ketaset) and xylazine hydrochloride. the immobilizing drugs were reversed with an intravenous hand injection of talozoline. instrumented moose and wtd were relocated via aerial radio telemetry at approximately 10-day intervals from january 24, 1989 to may 16, 1991. relocations were attempted on all animals throughout the study period unless there was a mechanical failure in the radio collar or the animal was confirmed dead. we calculated moose and wtd home ranges rather than use individual point relocations in habitat analyses to account for errors associated with radio telemetry point relocations (kernohan et al. 1998). we created 90% adaptive kernel home ranges for individual moose and wtd using home range extension (rodgers and carr 1998) in arcview 3.2 (esri 2000). the data were standardized by dividing each value of x and y by its respective standard deviation (seaman and powell 1996). we calculated the smoothing factor (h) individually for each animal using the biased cross-validation (bcv) method (sain et al. 1994). we limited our calculation of annual home ranges to those animals for which we secured a minimum of 30 relocations since kernel home range estimates suffer from inaccuracies and inflated sizes when small numbers of animal locations are used (seaman et al. 1999). all moose and 15 wtd were used in the home range analysis. one collared wtd fawn, approximately 9 months old at capture, was initially associated with a collared doe. this animal was included in the analysis because it established its own individual home range soon after capture, and therefore its locations were independent from its dam. we tested for significant differences between the mean sizes of moose and wtd home ranges and between male and female wtd mean home range sizes in vnp. samples of male moose (n = 3) were inadequate to test for sexual differences in moose home range size. habitat availability we determined available habitats separately for moose and wtd as 2 extended 100% minimum convex polygons (mcp) containing either all moose or all whitetailed deer radio-telemetry locations within our study area (fig. 3). we widened each mcp by 1.2 km for moose and 0.5 km for sympatric moose and deer – cobb et al. alces vol. 40, 2004 176 deer buffered mcp moose buffered mcp habitat cover boundary 5 0 5 10 kilometers n ew s fig. 3. buffered minimum convex polygons of moose and wtd radiotelemetry locations, depicting areas used to delineate habitat availability for compositional analysis. wtd to encompass the entire adaptive kernel home ranges of all study animals and calculated the percent composition of each habitat type available to moose and wtd using the extended mcps. we clipped the modified gis habitat coverage to individual moose and wtd adaptive kernel home ranges using arcview 3.2 patch analyst extension (rempel and carr 2003). lakes, ponds, and streams were excluded from habitat use or habitat selection calculations, however, vegetation adjacent to bodies of water were considered in the analysis. in addition to the 8 vegetation types, we examined the height and density of canopy covers within areas available to moose and wtd. habitat selection we compared habitat use to habitat availability using compositional analysis (aebischer et al. 1993). analysis was performed using the bycomp program (ott and hovey 1997) within the sas working environment (sas institute inc. 2000). this program first determined whether habitat use differed from random using wilks’ lambda ( ) statistics in multivariate analysis of variance (manova). if habitat use was nonrandom, habitats were ranked in order of preference and levels of significance between ranks were determined using a t-test. only home ranges located entirely within the study area (gis vegetation coverage extent) were included in habitat analyses. one migratory moose and one migratory wtd did not meet this criterion and were not used. small sample sizes precluded seasonal and sexual habitat selection analysis. a minimum of 10 animals per group (season or sex) is needed to produce reliable results using compositional analysis (aebischer et al. 1993). our data would not have met these standards when partitioned into groups by season or sex. compositional λ alces vol. 40, 2004 cobb et al. sympatric moose and deer 177 analysis required that each animal use all habitat types (aebischer et al. 1993). when proportional habitat use was estimated to be zero for moose and wtd, we replaced these values with 0.001. substituting a value smaller than the smallest recorded nonzero value produced results that were robust relative to the substituted value (aebischer et al. 1993). results pellet group counts a total of 1,674 deer pellet groups (820 in 1989, 854 in 1991) and 45 moose pellet groups (30 in 1989, 15 in 1991) were enumerated over all line transect surveys. wtd pellet groups were more abundant than moose pellet groups in all sampling units. twenty-two of the 32 total sampling units (68.8%) contained no moose pellet groups. all sampling units contained wtd pellet groups at varying abundances ( x = 16.8, sd = 12.0). moose and wtd pellet groups occurrence varied spatially across vnp (fig. 4). moose pellet groups were present in low (0 13) abundances ( x = 1.9, sd = 1.6) in sampling units in the central and eastern regions of the kabetogama peninsula and absent from all other sampling units. wtd pellet groups occurred at high (>21 42 pellet groups) abundances in sampling units in the central and western regions of kabetogama peninsula and in the southeastern corner of the park, and low densities (0 20 pellet groups) in sampling units on the eastern end of the kabetogama peninsula. only 3 sampling units (9.4%) contained moose pellet groups and high numbers of wtd pellet groups. these sampling units were located in the central region of the kabetogama peninsula and on the western periphery of moose pellet group distribution. the abundance of moose and wtd pellet groups varied with the average percent composition of sample unit habitat types. sampling units with high abundances of wtd pellet groups contained significantly more spruce/balsam fir habitat than did sampling units with low abundances of wtd pellet groups (t = 2.79, p = 0.02, table 2a). sampling units with moose pellet groups contained significantly less closed/continuous canopy cover (t = 2.25, p = 0.04) and significantly more open/discontinuous (t = 2.21, p = 0.04) than did sampling units lacking moose pellet groups (table 2b). sampling units with moose pellets also contained significantly less 12 – 20 m canopy cover (t = 2.37, p = 0.04) than did sampling units lacking moose pellets (table 2c). home range moose and wtd were relocated by fixed-wing aircraft on a 10-day mean interval (min = 1, max = 119, sd = 14) for a period of 842 days. with outliers removed, 10 moose were relocated 786 times and 20 wtd were relocated 1,032 times. each moose was relocated an average of 79 times (min = 30, max = 96, sd = 22). each wtd was relocated an average of 52 times (min = 6, max = 76, sd = 27). three (33%) radio-collared moose and 8 (40%) radiocollared wtd died during the study. moose home range averaged 48 km2 (min = 29, max = 141, sd = 33.5). one male moose had an especially large home range of 141 km2 because of seasonal migratory behavior. excluding this animal, the average annual moose home range size was 37 km2. the average annual wtd home range was 9 km2 (min = 2, max = 49, sd = 4.32). one female wtd that exhibited seasonal migratory behavior had an especially large home range (49 km2). excluding this animal, the average annual wtd home range was 6 km2. the average moose home range was significantly larger than the average wtd home range (t = 4.16, p < 0.01, fig. 5). sympatric moose and deer – cobb et al. alces vol. 40, 2004 178 moose pellet groups absent present n ew s 5 0 5 kilometers w td pellet groups low (1 21) high (>21 42) 5 0 5 kilometers n ew s fig. 4. (a) presence or absence of moose pellet groups and (b) abundance of wtd pellet groups in sampling units at voyageurs national park, minnesota, based on pellet count transects conducted in may 1989 and 1991. (a) (b) alces vol. 40, 2004 cobb et al. sympatric moose and deer 179 at least one other wtd, and all moose home ranges overlapped with at least one other moose. five of 10 moose home ranges overlapped with wtd home ranges, although the overlapping areas were relatively small. the total area of overlapping home ranges of instrumented moose and wtd was 6 km2. this area of overlapping home ranges encompassed 2.5% of all moose home range area and 6.0% of all wtd home range area, and was located in the central kabetogama peninsula (fig. 6). habitat availability and use available moose and wtd habitats were largely similar in terms of vegetation types, canopy height, and canopy density (table 3). however, 58% of combined moose home ranges vs. 9% of wtd combined home ranges had been burned or table 2. percent composition of (a) vegetation types, (b) canopy densities, and (c) canopy heights found in moose sampling units that contained pellet groups (present) and that contained no pellets (absent), and white-tailed deer (wtd) sampling units that contained high (27-40) and low (0-12)abundances of pellet groups. present high low (a) vegetation type aspen/birch 23 20.1 26.7 black ash/red maple 2.7 2 1.4 spruce/balsam fir 14.1 34.4* 12.0* bur oak 19.1 6.1 9.6 herbaceous alliance 8 6.3 8.8 jack pine 11.9 12.8 20.9 red/white pine 15.8 16.5 17.3 shrubland alliance 5.4 2 3.3 (b) canopy density dispersed/sparse (10-25%) 0.4 0 0.6 open/discontinuous (25-60%) 41.2* 28.3 29.7 closed/continuous (60-100%) 58.4* 71.7 70.2 (c) canopy heights open 7.7 5.9 9 <0.5 m 3.2 1.3 1.8 0.5 – 5 m 6.3 5 3.2 5 – 12 m 38.4 29.9 23.7 12 – 20 m 34.6* 49.1 50.6 20 – 30 m 9.8 8.8 11.610.7 11.1 26.5 21 47.6 55.6* 1.6 1.1 4.4 2.9 9.3 8.3 72.5 74.8* 0.4 0 27.1 25.2* 17.6 17.5 3.5 2 9.1 8.1 15.5 20.8 20.4 20 8.9 4.9 23 25.4 2 1.2 available absent moose wtd 0 10 20 30 40 50 60 70 80 moose deer h o m e r an g e a re a (k m 2) fig. 5. moose and wtd home range areas (km2), with 95% confidence intervals, at voyageurs national park, minnesota. the average male wtd home range (8 km2) was larger than the average female home range (4 km2), however, the difference was not statistically significant (t = 0.30, p = 0.77), even with the single female migratory wtd removed (t = 1.33, p = 0.21). all wtd home ranges overlapped with * indicates a significant difference ( = 0.05).α sympatric moose and deer – cobb et al. alces vol. 40, 2004 180 logged within the last 55 years. individual moose and wtd 90% kernel home ranges were largely similar in vegetation types with spruce/balsam fir and aspen/ birch types making up > 50% of the vegetation type for both ungulates (table 4). jack pine was slightly more abundant than the herbaceous alliance in moose home ranges, whereas herbaceous alliance was the third most abundant vegetation type in wtd kernel home ranges. moose and wtd showed similar rankings of abundance of canopy densities in their home ranges, however moose home ranges contained less closed/continuous and more open/discontinuous canopy densities than wtd. the 2 species differed in canopy height use with almost 50% of the 5 – 12 m height class available to moose and over 50% of the 12 – 20 m height class available to wtd. habitat selection vegetation types within moose home ranges differed significantly from available n ew s species overlap deer home ranges moose home ranges 5 0 5 10 kilometers fig. 6. home range overlap of instrumented moose and wtd in and adjacent to voyageurs national park, minnesota. vegetation types ( = 0.01, p = 0.04). moose showed a significant preference for spruce/ balsam fir over all other types except the shrubland alliance (t = 2.12, p = 0.07) and bur oak (t = 2.22, p = 0.07) (table 5a). the shrubland alliance, aspen/birch, herbaceous alliance, bur oak, and red/white pine types all tied for second in preference and did not differ significantly in preference from one another. moose exhibited significant nonrandom use of canopy densities ( = 0.30, p = 0.02). moose significantly selected open/discontinuous canopy cover over all others, but exhibited no significant difference in preference between closed/continuous and dispersed/sparse canopies (table 5b). moose exhibited nonrandom use of canopy heights ( = 0.09, p = 0.03) and showed a significant preference for 5 – 12 m canopy cover over open habitat and 12 – 20 m canopy cover (table 5c). there was no significant difference in preference between 5 – 12 m and 20 – 30 m canopy height (t = 1.02, p = 0.34), or between 5 – 12 m and λ λ λ alces vol. 40, 2004 cobb et al. sympatric moose and deer 181 (t = 5.61, p < 0.01) and open/discontinuous canopies (t = 4.86, p < 0.01) over dispersed/ sparse canopy (table 6b). selection for closed/continuous canopy over open/discontinuous canopy was not significant at p < 0.05 (t = 1.60, p = 0.12). wtd exhibited nonrandom use of canopy heights ( = 0.11, p < 0.01) with a highly significant (p < 0.02) preference for 12 – 20 m canopy over all others (table 6c). there was no evidence of significant preference for any other canopy height. discussion distribution pellet group sampling provided evidence that wtd were more widespread than moose at vnp. wtd pellet groups occurred in the high stratum toward the western and central portions of the kabetogama peninsula 0.5 – 5 m (t = 2.39, p = 0.06) or <0.5 m (t = 2.25, p = 0.06) canopy heights. vegetation types within wtd home ranges differed from available habitats but not significantly at p-value < 0.05 ( = 0.25, p = 0.09). wtd significantly selected spruce/balsam fir over all other vegetation types except aspen/birch (t = 2.03, p = 0.06) (table 6a). aspen/birch was significantly selected over all remaining vegetation types except herbaceous alliance (t = 1.83, p = 0.10). jack pine and bur oak tied for lowest in wtd preference at the home range scale. wtd exhibited significant nonrandom use of canopy densities ( = 0.27, p < 0.01), significantly selecting closed/continuous table 3. percent composition of available moose and white-tailed deer (wtd) (a) vegetation types, (b) canopy densities, and (c) canopy heights based on expanded minimum convex polygons of all moose and all wtd locations, respectively. m oose wt d (a) veget at ion t y p e a sp en/birch 24.1 28 black a sh/red m ap le 0.9 3.2 sp ruce/balsam f ir 26.8 26.6 bur o ak 7.6 2.7 h erbaceous a lliance 11.2 13.3 jack p ine 18.1 6.4 red/whit e p ine 5.1 6.1 shrubland a lliance 6.2 13.8 (b) canop y d ensit y d isp ersed/sp arse (10-25% ) 0.4 0.3 o p en/d iscont inuous (25-60% ) 29.3 22.7 closed/cont inuous (60-100%) 70.3 77 (c) canop y h eight s o p en 11.4 13.6 <0.5 m 2.2 2 0.5 – 5 m 6.3 9.8 5 – 12 m 34.3 33.1 12 – 20 m 44 39.3 20 – 30 m 1.8 2.1 moose wtd (a) vegetation type aspen/birch 22.6 28.2 black ash/red maple 0.5 2.5 spruce/balsam fir 36.9 40.5 bur oak 7.2 2.1 herbaceous alliance 10 11 jack pine 11.6 3.1 red/white pine 4.4 5.5 shrubland alliance 6.8 7.3 (b) canopy density dispersed/sparse (10-25%) 0.3 0.1 open/discontinuous (25-60%) 37.4 20.4 closed/continuous (60-100%) 62.3 79.5 (c) canopy heights open 10.1 11.1 <0.5 m 2.1 1.5 0.5 – 5 m 6.9 5.8 5 – 12 m 45.6 24 12 – 20 m 33.3 54.8 20 – 30 m 2.1 2.9 table 4. average percentage habitat use within moose and white-tailed deer (wtd) 90% adaptive kernel home ranges. λ λ λ sympatric moose and deer – cobb et al. alces vol. 40, 2004 182 and along the periphery of the southeastern portion of vnp while moose pellet groups were restricted to the central/eastern region of the kabetogama peninsula. the distribution of moose based upon our pellet group sampling and home range calculations were similar, and both agreed with the distribution of moose determined during aerial censuses (gogan et al. 1997). the table 5. simplified ranking matrices for moose based on comparing proportional (a) vegetation type, (b) canopy density, and (c) canopy height use within 90% adaptive kernel home range to proportions available within the available area (extended mcps). habitat classes are ranked from most preferred to least preferred. habitat classes that differ significantly in preference from random at p = 0.05 are indicated by either a “+++” or “—”. habitat classes that differ in preference from random at p = 0.10 are indicated by either a “++” or “—”. habitat classes that differ in preference from random at p > 0.10 are indicated by either a “+” or “-”. (a) vegetation type r a n k spruce/b alsam fir shrubland a lliance a spen/b irch h erbaceous a lliance b ur o ak r ed/w hite pine jack pine b lack a sh/r ed m aple spruce/balsam fir 1 . ++ +++ +++ ++ +++ +++ +++ shrubland alliance 2 -. + + + ++ +++ +++ aspen/birch 3 --. + + + +++ +++ herbaceous alliance 4 --. + + +++ +++ bur oak 5 -. + ++ ++ red/white pine 6 ---. +++ +++ jack pine 7 -----------. + black ash/red maple 8 -----------. (b) canopy density r a n k o pen/ d iscontinuous (25-60% ) c losed/ c ontinuous (60-100% ) d ispersed/ sparse (10-25% ) open/discontinuous (25-60%) 1 . +++ +++ closed/continuous (60-100%) 2 --. ++ dispersed/sparse (10-25%) 3 ---. (c) canopy heights r a n k 5-12 m 20-30 m 0.5-5 m o pen <0.5 m 12-20 m 5 – 12 m 1 . + ++ +++ ++ +++ 20 – 30 m 2 . + + + +++ 0.5 – 5 m 3 -. + + + open 4 --. + + <0.5 m 5 -. + 12 – 20 m 6 ----. capture and instrumenting of moose and wtd was completed prior to the establishment of the pellet group sampling transects and was therefore not dependent on our sampling of pellet groups. our calculated home ranges of instrumented wtd were largely coincident with the distribution of pellet group units that we assigned to the high wtd stratum. trapping locations for alces vol. 40, 2004 cobb et al. sympatric moose and deer 183 the species in the contiguous united states. annual home ranges of wtd at vnp (x = 5.74 km2, n = 15) were much larger than those in northeastern minnesota (mcp, 0.8 km2 summer, 0.4 km2 winter) (nelson and mech 1981). in general, the home ranges of wtd at the northern limits of the species distribution are larger than those in the s o u t h e r n p e r i p h e r y o f t h e i r r a n g e (severinghaus and cheatum 1956). there are a number of causes for varying home range sizes between locations. home range size might be dictated directly by an animal’s energetics (mcnab 1963). following this theory, animals of the same species in more productive habitats have smaller home ranges than those in poor habitats, as the latter require greater areas to secure the resources required for survival. other factors possibly influencing moose and wtd home range sizes at vnp include reproductive activity, relative distribution and diversity of suitable habitats, and species density (leptich and gilbert 1989, beier and mccullough 1990, ballard et al. 1991). sexual differences in home range size have been reported for moose and wtd. males of both species usually occupy larger home ranges than females (carlsen and farmes 1957, ballard et al. 1991) although no difference between male and female home range sizes has been observed in some areas (phillips et al. 1973, taylor and ballard 1979, hauge and keith 1981). there was no significant difference between male and female wtd home range sizes at vnp. our relatively small wtd sample caused large confidence intervals and strong outlier effects. moose and wtd in northern regions typically undergo significant seasonal home range shifts (messier and barrette 1985, van deelen et al. 1998). yarding behavior by wtd is common in northern regions (telfer 1967, rongstad and tester 1969, wtd were based upon our observations of high concentrations of deer in winter and not on the distribution of high densities of pellet groups based upon our pellet group sampling in may. while wtd pellet groups occurred in all sampling units, only 3 sampling units assigned to the high wtd stratum also contained moose pellet groups. this pattern is indicative of differing habitat use patterns between the 2 species. the inverse relationship between moose and whitetailed deer distributions at vnp is consistent with observations in adjacent ontario, where moose reached their highest densities in areas where white-tailed deer were <4/km2 (whitlaw and lankester 1994). moose densities in ontario were inversely related to the mean intensity of meningeal worm larvae in white-tailed deer pellet groups (whitlaw and lankester 1994). there is currently no information on the relative spatial distribution of meningeal worm larvae in wtd pellet groups at vnp. home range moose 90% adaptive kernel home ranges in vnp (x = 47.7 km2, n = 10) were among the largest recorded in the contiguous united states, and were much larger than those found in northwestern minnesota (x male = 3.1 km 2, x female = 3.6 km 2, n = 26) (phillips et al.1973) and northwestern ontario (x = 14.0 km2, n = 1) (addison et al. 1980). the differences may be greater than these comparisons of size alone suggest since the other studies used methods that generally produce larger home range estimates than does the adaptive kernel method used here. moose home ranges at vnp were substantially smaller than those in alaska, where home range sizes (mcp) are between 120 km2 and 350 km2 (gravogel 1984). wtd adaptive kernel home ranges at vnp were larger than most recorded for sympatric moose and deer – cobb et al. alces vol. 40, 2004 184 table 6. simplified ranking matrices for white-tailed deer (wtd) based on comparing proportional (a) vegetation type, (b) canopy density, and (c) canopy height use within 90% adaptive kernel home range to proportions available within the available area (extended mcps). habitat classes are ranked from most preferred to least preferred. habitat classes that differ significantly in preference from random at p = 0.05 are indicated by either a “+++” or “—”. habitat classes that differ in preference from random at p = 0.10 are indicated by either a “++” or “—”. habitat classes that differ in preference from random at p > 0.10 are indicated by either a “+” or “-”. (a) vegetation type r a n k spruce/b alsam fir a spen/b irch h erbaceous a lliance b lack a sh/r ed m aple shrubland a lliance r ed/w hite pine jack pine b ur o ak spruce/balsam fir 1 . ++ +++ +++ +++ +++ +++ +++ aspen/birch 2 -. ++ +++ +++ +++ +++ +++ herbaceous alliance 3 ---. + +++ ++ +++ +++ black ash/red maple 4 ----. +++ ++ +++ +++ shrubland alliance 5 --------. + + ++ red/white pine 6 ------. + + jack pine 7 --------. + bur oak 8 ---------. (b) canopy density r a n k c losed/ c ontinuous (60-100% ) o pen/ d iscontinuous (25-60% ) d ispersed/ sparse (10-25% ) closed/continuous (60-100%) 1 . + +++ open/discontinuous (25-60%) 2 . +++ dispersed/sparse (10-25%) 3 ----. (c) canopy heights r a n k 12 – 20 m o pen 5 – 12 m 0.5 – 5 m 20 – 30 m <0.5 m 12 – 20 m 1 . +++ +++ +++ +++ +++ open 2 --. + ++ + ++ 5 – 12 m 3 --. + + + 0.5 – 5 m 4 ---. + + 20 – 30 m 5 --. + <0.5 m 6 ---. nelson 1998), and by moose in eastern canada (proulx 1983). approximately 80% of white-tailed deer in nearby superior national forest, minnesota, exhibit migratory behavior (nelson 1998). migratory behavior has traditionally been thought of as an adaptive response to the presence of snow (townsend and smith 1933) or an anti-predator response (nelson and mech 1991). none of the wtd on the kabetogama peninsula exhibited migratory behavior typical of yarding based on the size and shape of their annual home ranges. one male moose extended its range into adjacent alces vol. 40, 2004 cobb et al. sympatric moose and deer 185 ontario in summer. two white-tailed deer captured and radio marked in the vicinity of d a l e y b a y s e a s o n a l l y m i g r a t e d during warmer months beyond the boundaries of the study area and returned to daley bay during the winter. vegetation in the vicinity of daley bay is predominately northern white cedar swamp (classified as shrubland alliance in this study), which is typically associated with white-tailed deer wintering yards (crawford 1982). wtd may not exhibit migratory behavior in vnp because of an abundance of winter habitat. forest maturation in adjacent superior national forest during the early 1970s provided white-tailed deer with abundant winter cover, and may have allowed the population to disperse into smaller groups rather then exhibit common yarding behavior (wetzel et al. 1975). winter cover, such as balsam fir, was abundant in vnp during this study due to a long-term absence of major wildfires and limited timber harvest in vnp since the 1930s, particularly in the area utilized by instrumented wtd. the potential abundance of preferred winter cover may be a reason for the non-migratory behavior of wtd at vnp. there was minimal overlap between instrumented moose and white-tailed deer home ranges. home range overlap between the two species was limited to a small region on the central kabetogama peninsula. concurrent pellet group sampling showed the same region of the kabetogama peninsula to be the only area within vnp where moose pellet groups occurred within the high-density stratum for wtd pellet groups. pellet group sampling showed that wtd occur throughout vnp at varying densities and that on the kabetogama peninsula, moose are generally limited to the central and eastern area. habitat selection our pellet group surveys provided information on winter habitat use only, while the radio telemetry data provided information on the habitat characteristics of yearround home ranges. the pellet group surveys indicate that wtd preferred the spruce/balsam fir vegetation type over all others but did not select for canopy height or density. in contrast, moose showed no preference for any vegetation type but did prefer the open/discontinuous canopy densities, and avoided the dispersed/sparse canopy densities and 12 – 20 m canopy height. the year-round telemetry data reveal a different pattern with wtd and moose selecting for vegetative type, canopy density, and canopy height. we attribute this difference to differences in sampling intensity and portions of the annual cycle embraced by each sampling technique. moose distributions appeared to be related by the distribution of canopy heights and canopy densities in vnp. lower canopy heights and canopy densities were more prevalent in sampling units containing moose pellets. the most common canopy height in sampling units containing moose was 5 – 12 m, but the most common canopy height in sampling units containing high abundances of white-tailed deer pellets was 12 – 20 m. lower discontinuous vegetation might provide moose with more accessible forage in the winter. both moose and wtd at vnp exhibited a strong preference for spruce/balsam fir habitat types. this type satisfies different needs of each herbivore species. moose habitat use generally is dictated by food abundance rather than shelter (telfer 1970) and balsam fir is an important source of forage for moose in boreal forests, especially during the winter season (irwin 1975, peek et al. 1976, ludewig and bowyer 1985, allen et al. 1987). however, balsam fir would also provide moose at vnp with a refuge from deep snow during winter: spruce and balsam fir were among the dominant sympatric moose and deer – cobb et al. alces vol. 40, 2004 186 overstory species in moose winter yards in southern quebec (proulx 1983). wtd are less adapted to harsh winter conditions than moose and select winter habitat based on thermal protection and shelter, rather than forage preference (telfer and kelsall 1979). spruce and balsam fir are considered poor quality forage for wtd (crawford 1982, blouch 1984) but offer wtd ideal winter protection. wtd in northeastern minnesota used balsam fir dominated stands frequently in late winter (wetzel et al. 1975). the unusually harsh winter conditions during our study might have caused whitetailed deer to utilize winter shelter habitats such as the spruce/balsam fir type for longer periods and at higher levels than usual. monthly snowfall exceeded 50 cm in three consecutive months during the winters of 1988 – 1989 and 1990 – 1991, and remained into early spring following these winters. wtd movements become restricted at snow depths of approximately 30 cm and confined at snow depths > 50 cm (telfer 1970). the aspen/birch vegetation type ranked second in wtd preference. both species are preferred wtd forage in boreal forests (cairns and telfer 1980, ludewig and bowyer 1985). wtd in adjacent ontario annually used aspen and birch more than other vegetation types (kearney and gilbert 1976). the aspen/birch type does not offer wtd as much thermal cover and snow protection as the spruce/balsam fir type. the shrubland alliance, aspen/birch, and herbaceous alliance vegetation types all ranked second in moose preference. these types offer moose important forage species. moose in northern minnesota forage extensively on aspen and birch (peek et al. 1976). the shrubland alliance includes northern red osier dogwood (cornus stolonifera) and willows (salix spp.) while the herbaceous alliance type contains hydrophilic plant species. moose eat both the woody and hydrophilic plants during the early spring and fall seasons in boreal forests (peek et al. 1976, jordan 1987). aquatic habitats also may provide moose with an escape from biting insects during the early summer season (ritcey and verbeek 1969). moose and wtd selected different canopy densities and heights at vnp. moose exhibited a significant preference for open/ discontinuous canopy density to all other canopy densities and for canopy heights of 5 – 12 m and 20 – 30 m. wtd preferred closed/discontinuous canopy to open/discontinuous canopy densities and significantly preferred 12 – 20 m canopy height to all others. differences between moose and wtd in body size may be an important factor in the differences in canopy preferences between these species. the amount of energy that ungulates expend in moving increases linearly with increasing snow depth, until breast height, at which it increases exponentially (parker et al. 1984). dense canopy cover displaces snow and causes structural changes to snow that influence the energy ungulates need to move and forage (kirchoff and schoen 1987). wtd are more restricted by snow depth and cold temperatures than are moose and therefore utilize vegetative types that allow movement during periods of deep snow (telfer 1970). moose are less affected by snow conditions than wtd and therefore would be more likely to select for forage availability rather than thermal and snow protection during winter. moose winter yards in southern quebec were in slightly closed canopy forest (41-80%) with tree heights of 9.1 – 18.3 m (proulx 1983). however, these yards tended to be on slopes that reduced the energetic cost of moving t h r o u g h d e e p s n o w ( p r o u l x 1 9 8 3 ) . moose selection of 20-30 m canopy cover was unexpected, however there are a few possible explanations for these observations. low abundances of 20-30 m canopy alces vol. 40, 2004 cobb et al. sympatric moose and deer 187 class were clustered throughout vnp, composed primarily of red/white pine overstory habitat. although moose typically do not exhibit a high preference for white and red pine (peek et al. 1976), as they did not in this study, understory vegetation in red/white pine habitat included aspen and birch that are palatable forage for moose. heat stress avoidance is another possible explanation for moose selecting high canopy cover (kelsall and telfer 1974). moose at vnp are at the southern periphery of the species’ distribution. southern populations of moose select forested upland sites during the summer season, possibly to reduce energy expenditure and enable them to forage for longer periods per day (miller and litvaitis 1992). during warm periods, the overhead canopy of a 20-30 m coniferous canopy could shade moose and reduce ambient temperatures better than other habitats. meningeal worm refuge we did not find a single sampling unit that was free of wtd pellets. it is therefore highly unlikely that moose had a complete refuge from meningeal worm-infected wtd within vnp. lower densities of wtd within moose range however might reduce the rate of meningeal worm transmission to moose, and thereby increase moose survival. this partial refuge could be enough to allow moose to survive in the presence of wtd. moose are able to maintain low population levels in the presence of meningeal worm if wtd numbers do not exceed 4.6/km2 (karns 1967). moose, wtd, and meningeal worm have existed sympatrically in ontario since the early 1980s, and their interaction does not appear to be negatively affecting moose population numbers (whitlaw and lankester 1994). the prevalence of meningeal worm in wtd in northwestern ontario was similar to vnp based upon the presence of larvae in feces and adults in the cranium (gogan et al. 1997). our pellet group surveys and radio telemetry data reveal evidence of spatial segregation and resource partitioning between moose and wtd at vnp. this spatial separation and differences in habitat preferences between moose and white-tailed deer at vnp may reduce the prevalence of meningeal worm infection in moose by providing moose with a partial refuge from meningeal worm-infected wtd (gilbert 1974). the high prevalence of meningeal worm in wtd at vnp (gogan et al. 1997) indicates that the parasite and its gastropod intermediate hosts are sufficiently abundant for transmission of the parasite to moose. the parasite is highly pathogenic to moose, and therefore moose are not expected to persist in its presence at vnp unless they are somehow isolated from infection. moose densities in adjacent ontario were lowest in areas with the highest mean intensity of meningeal worm larvae in wtd pellet groups (whitlaw and lankester 1994). the potential for meningeal worm transmission to moose may be even greater at vnp than ontario since wtd densities and the prevalence of meningeal worm are higher at vnp than in most regions in ontario (whitlaw and lankester 1994, gogan et al. 1997). partial habitat separation between moose and wtd at vnp may be a factor allowing moose to persist despite a high prevalence of meningeal worm. references aebischer n. j., p. a.robertson, and r. e. kenward. 1993. compositional analysis of habitat use from animal radiotracking data. ecology 74: 1313 – 1325. addison, r. b., j. c. williamson, b. p. sa u n d e r s , and d. f r a s e r. 1980. radiotracking of moose in the boreal forest of northwestern ontario. canadian field naturalist 94: 269 – 276. allen, a. w., p. a. jordan, and j. w. sympatric moose and deer – cobb et al. alces vol. 40, 2004 188 terrell. 1987. habitat suitability index models: moose, lake superior region. fish and wildlife service. biological report 82 (10.155). ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. beier, p., and d. r. mccullough. 1990. factors influencing white-tailed deer activity patterns and habitat use. wildlife monographs 54. blouch, r. i. 1984. northern great lakes states and ontario forests. pages 391410 in l.k. halls, editor. white-tailed d e e r : e c o l o g y a n d m a n a g e m e n t . stackpole books, harrisburg, pennsylvania, usa. cairns, a. l., and e. s. telfer. 1980. habitat use by 4 sympatric ungulates in boreal mixed wood forest. journal of wildlife management 44: 849 – 857. carlsen, j. c., and r. e. farmes. 1957. movements of white-tailed deer tagged in minnesota. journal of wildlife management 21: 397 – 401. clover, m. r. 1956. single-gate deer trap. california fish and game 42: 199-201. cole, g. f. 1987. changes in interacting species with disturbance. environmental management 11: 257 – 264. crawford, h. s. 1982. seasonal food selection and digestibility by tame whitetailed deer in central maine. journal of wildlife management 46: 974 – 982. crowley, k. f., and k. l. cole. 1995. patterns of temporal and spatial change in the vegetation of voyageurs national park. bulletin of the ecological society of america 77: 57. (esri) environmental systems research i n s t i t u t e . 2 0 0 0 . a r c v i e w 3 . 2 . redlands, california, usa. fuller, t. k. 1986. observations of moose, alces alces, in peripheral range in north-central minnesota. canadian field naturalist 100: 359 – 362. gilbert, f. f. 1974. parelaphostrongylus tenuis in maine: prevalence in moose. journal of wildlife management 30: 200 – 202. gogan, p. j. p., k. d. kozie, e. m. olexa, and n. s. duncan. 1997. ecological status of moose and white tailed deer at voyageurs national park, minnesota. alces 33: 187 – 201. grauvogel, c. a. 1984. seward peninsula moose population identity study. federal aid for wildlife restoration final report. alaska department of fish and game, juneau, alaska, usa. hauge, t. m., and l. b. keith. 1981. dynamics of moose populations in northeastern alberta. journal of wildlife management 45: 573 – 597. hurrell, j. w. 1995. decadal trends in the north atlantic oscillation: regional temperatures and precipitation. science 269: 676 – 679. irwin, l. l. 1975. deer-moose relationships on a burn in northeastern minnesota. journal of wildlife management 39: 653 – 662. johnson, c., and j. sales. 1995. influence of geomorphology on the vegetation of voyageurs national park. poster. tenth annual u.s. landscape ecology symposium. international association of landscape ecology, university of minnesota. jordan, p. a. 1987. aquatic foraging and the sodium ecology of moose: a review. swedish wildlife research supplement 1:119-137. karns, p. d. 1967. parelaphostrongylus tenuis in deer in minnesota and implications for moose. journal of wildlife management 31: 299 – 303. _____. 1998. population distribution, density and trends. pages 125 – 139 in a.w. franzmann and c.c. schwartz , editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . alces vol. 40, 2004 cobb et al. sympatric moose and deer 189 smithsonian institution press, washington d.c., usa. kearney, s. r., and f. f. gilbert. 1976. habitat use by white-tailed deer and moose on sympatric range. journal of wildlife management 40: 645 – 657. kelsall, j. p., and e. s. telfer. 1974. biogeography of moose with particular reference to western north america. naturaliste canadien 101: 117 – 130. kernohan, b. j., j. j. millspaugh, j. a. jenks, and d. e. naugle. 1998. use of an adaptive kernel home-range estimator in a gis environment to calculate habitat use. journal of environmental management 53: 83 – 89. kirchoff, m. d., and j. w. schoen. 1987. forest cover and snow: implications for deer habitat in southeast alaska. journal of wildlife management 51: 28 – 33. kurmis, v., s. l. webb, and l. c. merriam, jr. 1986. plant communities of voyageurs national park, minnesota, u.s.a. canadian journal of botany 64: 531 – 540. lamb, p. j., and r. a. peppler. 1987. north atlantic oscillation: concept and application. bulletin of the american meteorological society 68: 1218 – 1225. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use of moose in northern maine. journal of wildlife management 53: 880 – 884. ludewig, h. a., and t. bowyer. 1985. overlap in winter diets of sympatric moose and white-tailed deer in maine. journal of mammalogy 66: 390 – 392. mcnab, b. k. 1963. bioenergetics and the determination of home range. american naturalist 97: 130 – 140. mech, l. d., and p. d. karns. 1978. role of the wolf in a deer decline in the superior national forest. north central forest experimental station. u.s. forest service paper nc-148. st. paul, minnesota, usa. messier, f., and c. barrette. 1985. the efficiency of yarding behaviour by whitetailed deer as an anti-predator strategy. canadian journal of zoology 63: 785 – 789. miller, b. k., and j. a. litvaitis. 1992. habitat selection by moose in a boreal forest ecotone. acta theriologica 37: 41 – 50. nelson, m. e. 1998. development of migratory behavior in northern white-tailed deer. canadian journal of zoology 76: 426 – 432. _____, and l. d. mech. 1981. deer social organization and wolf predation in northeastern minnesota. wildlife monographs 77. _____, and _____. 1991. wolf predation risk associated with white-tailed deer movements. canadian journal of zoology 69: 2696 – 2699. neu, c. w., c. r. byers, and j. m. peek. 1974. a technique for analysis of utilization-availability data. journal of wildlife management 38: 541 – 545. nudds, t. d. 1990. retroductive logic in retrospect: the ecological effects of meningeal worm. journal of wildlife management 56: 617 – 619. ohmann, l. f., and r. r. ream. 1971. wilderness ecology: virgin plant communities of the boundary waters canoe area. north central forest experimental station. u.s. forest service resource paper nc – 63. st. paul, minnesota, usa. olson, h. f. 1938. deer tagging and population studies in minnesota. transactions of the north american wildlife and natural resources conference 3: 280 – 286. ott, p., and f. hovey. 1997. bycomp. sas version 1.0. british columbia forest service, victoria, british columbia, canada. sympatric moose and deer – cobb et al. alces vol. 40, 2004 190 parker, k. l., c. t. robbins, and t. a. hanley. 1984. energy expenditure for locomotion by mule deer and elk. journal of wildlife management 48: 474 – 488. pastor, j. r., and d. j. mladenoff. 1992. the southern boreal-northern hardwood forest border. pages 216-229 in h.h. shugart, r. leemans, and g.b. bonan, editors. a system analysis of the global boreal forest. cambridge university press, new york, usa. peek, j. m., p. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northwestern minnesota. wildlife monographs 48. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. petraborg, w. h., and d. w. burcalow. 1965. the white-tailed deer in minnesota. pages 11 – 48 in j.b. moyle, editor. big game in minnesota. minnesota department of conservation, technical bulletin no. 9. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37: 266 – 278. prescott, w. h. 1974. interrelationships of moose and deer of the genus odocoileus. naturaliste canadien 101: 493 – 504. proulx, g. 1983. characteristics of moose (alces alces) winter yards on different exposures and slopes in southern quebec. canadian journal of zoology 61: 112 – 118. rempel, r. s., and a. p. carr. 2003. patch analyst for arcview 3.2. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. ritcey, r. w., and n. a. m. verbeek. 1969.observations of moose feeding on aquatics in bowron lake park, british columbia. canadian field naturalist 83: 339 – 343. rodgers, a. r., and a. p. carr. 1998. hre: the home range extension for arcview. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. rongstad, o. j., and j. r. tester. 1969. movements and habitat use of whitetailed deer in minnesota. journal of wildlife management 33: 366 – 379. sain, s. r., k. a. baggerly, and d. w. scott. 1994. cross-validation of multivariate densities. journal of the american statistical association 89: 807 – 817. sas institute inc. 2000. sas institute inc. cary, north carolina, usa. seaman, d. e., j. j. millspaugh, b. j. ke r n o h a n, g. c. b r u n d i g e, k. j. raedeke, and r. a. gitzen. 1999. effects of sample size on kernel home range estimates. journal of wildlife management 63: 739 – 747. _____, and r.a. powell. 1996. an evaluation of the accuracy of kernel density estimators for home range analysis. ecology 77: 2075 – 2085. severinghaus, c. w., and e. l. cheatum. 1956. life and times of the white-tailed deer. pages 57 – 186 in w.p. taylor, editor. the deer of north america. the stackpole company. harrisburg, pennsylvania, usa. taylor, k. p., and w. b. ballard. 1979. moose movement and habitat use along the susitna river near devil’s canyon. proceedings of the north american moose conference and workshop 15: 169 – 186. telfer, e. s. 1967. comparison of a deer yard and a moose yard in nova scotia. alces vol. 40, 2004 cobb et al. sympatric moose and deer 191 canadian journal of zoology 45: 485 – 490. _____. 1970. winter habitat selection by moose and white-tailed deer. journal of wildlife management 34: 553 – 559. _____, and j. p. kelsall. 1979. studies of morphological parameters affecting ungulate locomotion in snow. canadian journal of zoology 57: 2153-2159. thompson, i. d., and d. l. euler. 1987. moose habitat in ontario: a decade of change in perception. swedish wildlife research supplement 1: 181 – 193. townsend, m. t., and m. w. smith. 1933. t h e w h i t e t a i l e d d e e r o f t h e adirondacks. roosevelt wildlife bulletin 6: 161 – 325. (usgs) u.s. geological survey. 2001. vegetation spatial database coverage for voyageurs national park. vegetation mapping project, upper midwest environmental sciences center, la crosse, wisconsin, usa. van deelen, t. r., h. campa, m. hamady, and j. b. haufler. 1998. migration and seasonal range dynamics of deer using adjacent deeryards in northern michigan. journal of wildlife management 62: 205 – 213. wetzel, j. f., j. r. wambaugh, and j. m. peek. 1975. appraisal of white-tailed deer winter habitats in northeastern minnesota. journal of wildlife management 39: 59 – 66. whitlaw, h. a., and m. w. lankester. 1994. the co-occurrence of moose, white-tailed deer, and parelaphostrongylus tenuis in ontario. canadian journal of zoology 72: 819 – 825. f:\alces\vol_38\pagema~1\3804.pdf alces vol. 38, 2002 van dyke et al. – ecosystem management and moose 5 5 ecosystem management and moose: creating a coherent concept with functional management strategies fred van dyke1, brian darby2, sarah e. van kley3, jamie d. schmeling4, and nathan r. dejager5 1department of biology, armending hall, wheaton college, wheaton, il 60187, usa; 2department of earth, ecological, and environmental sciences, university of toledo, toledo, oh 43606 , usa; 3department of botany and microbiology, george lynn cross hall, van vleet oval, university of oklahoma, norman, ok 73019-0245, usa; 4529 douglas avenue, apartment 16, holland, mi 49424-2701 usa; 5department of biology, university of minnesota duluth, duluth, mn 55812, usa abstract: ecosystem management is a popular but poorly defined concept in conservation biology. current vague, non-operational definitions provoke criticism of the concept and undermine credibility of its associated principles. we propose a definition of ecosystem management that emphasizes essential qualities of the concept rather than its accidental associations or properties, and that explains functional and operational attributes of ecosystem management rather than its descriptive characteristics. based on these criteria, we offer a definition of ecosystem management as “a pattern of prescribed, goal-oriented environmental manipulations that: (1) treat a specified ecological system of identifiable boundaries as the fundamental unit to be managed; (2) has, as its desired outcome, the achievement of a state or collection of states in the ecosystem such that historical components, structure, function, products, and services of the ecosystem persist within biologically normal ranges and with normal rates of change; (3) uses naturally occurring, landscapescale processes as the primary means of management; and (4) determines management objectives through cooperative decision-making of individuals and groups who reside in, administer, and/or have vested interests in the state of the ecosystem”. achieving workable ecosystem management is currently hindered by the lack of a unified vision and system of values for ecosystems, the absence of permanent inter-agency bodies with authority to manage ecosystems across multiple jurisdictions, and the lack of administrative mechanisms for the translation of ecosystem research findings into ecosystem management policies. we propose strategies to overcome these obstacles and examine moose (alces alces) as an example of a species that is both important to ecosystem management and may benefit from it. alces vol. 38: 55-72 (2002) key words: alces alces, ecosystem management, implementation, test case the concept of ecosystems dates to early in the twentieth century (tansley 1935), but the concept of managing ecosystems is more recent. of all modern efforts in the management and conservation of natural resources, none has proven more elusive in definition or more controversial in implementation than “ecosystem management.” speaking of the idea with unconcealed disdain, conservationist michael bean wrote, “rarely has a concept gone so directly from obscurity to meaninglessness without any intervening period of coherence” (bean 1997). less cynically, but not less optimistically, berry et al (1998) wrote, “no single operational definition of ecosystem management exists, although its basic principles are understood.” in the united states, 18 federal agencies have adopted or are considering adoption of programs based on ecosystem management concepts (congressional research ecosystem management and moose – van dyke et al. alces vol. 38, 2002 5 6 service 1994, christensen et al. 1996, haeuber 1996, haeuber and franklin 1996, prato 1999). representatives of 5 of these federal agencies participated in a signing of a joint agreement to proceed with ecosystem management at the ecological stewardship workshop held in tucson, arizona in 1995 (czech and krausman1997). to support such efforts, a wealth of attempted definitions of ecosystem management exists, many written by the agencies themselves (table 1). however, despite the intensity of effort and variety of expression, current ecosystem management has been described as “a loose collection of agency specific concept papers, policy guidance documents, and potential – or only partially implemented – administrative changes” (haeuber 1996). while some professionals view the concept of ecosystem management as an important paradigm shift, others see it is as the opposite, a vacuous phrase “desperately seeking a paradigm” (lackey 1998). various positions are: (1) that ecosystem management is not a new paradigm at all (slocombe 1993, taylor 1993, czech 1995, haeuber 1996); (2) that it is what managers have been doing all along (irland 1994, more 1996, berry et al. 1998); (3) that it is a dressed-up version of the u. s. forest service’s old “multiple use” management (czech and krausman 1997); (4) that it should be called “public lands management because public lands are ecosystems” (czech 1995); (5) that it is the same as conservation because it has the same goal and therefore should be renamed “ecosystem conservation” (czech 1995); (6) that it is a conspiracy to reduce the extractive use of natural resources and expel private citizens from public lands (christensen et al. 1996); and (7) that aldo leopold thought of it first (czech 1995, knight 1995, grumbine 1998). vague, non-distinctive definitions of ecosystem management encourage and justify criticisms of the concept (czech 1995). if we follow a classical authority such as aristotle, a useful definition of ecosystem management would be one that expresses the essence or nature of the entity and not merely its accidental properties (abelson 1967, more 1996). the ideal definition would be one that includes “all instances and only those instances” of the category we define, a definition specifying both the essence of ecosystem management and its boundaries. equipped with such a definition, we would be able to determine immediately if something is or is not ecosystem management (more 1996). but to make progress in our understanding we must determine the essence or distinctive nature of ecosystem management compared to other management strategies. valuable as an operational definition of ecosystem management might be, the definition alone is insufficient. mechanisms to enforce ecosystem management practices, and to overcome inherent and systemic obstacles to an ecosystem management approach must be constructed. in addition, the concept and practice of ecosystem management also raise legitimate concerns for those with vested interests in a particular species or resource. specifically, is ecosystem management such a broad concept that it will prove insensitive to the values of individual species, such as moose (alces alces)? for example, crichton et al. (1998) warn that “moose management is not counter to conservation biology or most other administrative program orientations (such as ecosystem or habitat management), [however] a danger exists to the resource – moose – if management attempts to be too inclusive and if readjustment occurs at the sacrifice or compromise of programs that have been the mainstay of professional management all along.” thus, to be effective, ecosystem management alces vol. 38, 2002 van dyke et al. – ecosystem management and moose 5 7 table 1. definitions of “ecosystem management” in 10 federal agencies in the united states (congressional research service 1994). agency definition department of agriculture the integration of ecological principles and social factors t o m a n a g e e c o s y s t e m s t o s a f e g u a r d e c o l o g i c a l sustainability, biodiversity and productivity. department of commerce national oceanic and activities that seek to restore and maintain the health, atmospheric administration integrity, and functional values of natural ecosystems that are the cornerstone of productive, sustainable economics. department of defense the identification of target areas, including department of defense lands, and the implementation of a “holistic approach" instead of a “species-by-species approach” in order to enhance biodiversity. department of energy a consensual process, based on the best available science that specifically includes human interactions and management; and uses natural instead of political boundaries in order to restore and enhance environmental quality. department of the interior bureau of land management the integration of ecological, economic, and social principles to manage biological and physical systems in a manner safeguarding the long-term ecological sustainability, natural diversity, and productivity of the landscape. fish and wildlife service protecting or restoring the function, structure, and species composition of an ecosystem, recognizing that all components are interrelated. national park service a philosophical approach that respects all living things and seeks to sustain natural processes and the dignity of all species and to ensure that common interests flourish. u.s. geological survey ecosystem management emphasizes natural boundaries, s u c h a s w a t e r s h e d s , b i o l o g i c a l c o m m u n i t i e s , a n d physiographic provinces, and bases resource management decisions on an integrated scientific understanding of how the whole ecosystem works. environmental protection agency to maintain overall ecological integrity of the environment while ensuring that ecosystem outputs meet human needs on a sustainable level. national science foundation an integrative approach to the maintenance of land and water resources as functional habitat for an array of organisms and the provision of goods and services to society. ecosystem management and moose – van dyke et al. alces vol. 38, 2002 5 8 must demonstrate that it is not only an operational concept, but that it can successfully meet needs of individual species of special importance in ecosystem function or of particular value to individual stakeholder groups. the role of individual species in ecosystem management must be stressed because some species are “drivers” and some are “passengers” in ecosystem processes. the drivers are active determinants of the characteristics of the ecosystem in which they live because of ecological functions that they perform in the system. the passengers “ride along” on the effects created by the drivers. moose are unquestionably “driver” species, or, in more familiar terms, “keystone” species, in every ecosystem in which they have been carefully studied. they have disproportionate effects on community or ecosystem processes and, as a result, disproportionately affect the abundance of other plant and animal species, as well as habitat composition in the landscape. we evaluate the problems of defining ecosystem management operationally, suggest mechanisms for its implementation and offer moose as a “test case” regarding its effects on an individual species. paradigm development of ecosystem management since the 1960s, managers of public lands and academics in applied sciences like wildlife management, range management, and forestry have written about “ecosystem concepts in management” (major 1969; van dyne 1969; wagner 1969, 1977). by the 1970s, the term “ecosystem management” was in common use (czech and krausman 1997). however, authors from this period almost always used such terms to describe either the management of populations as commodities or the manipulation of processes, structures, and functions of ecosystems in order to produce desired levels of animal populations or plant biomass (major 1969; wagner 1969, 1977). if this is all that “ecosystem management” means, then the concept would certainly not meet the criteria for a scientific paradigm, nor would it represent a genuine “paradigm shift” to any new concepts or ideas. ultimately, paradigms come to incorporate and express the values, theories, methods, and tools that a professional community prescribes and believes to achieve a desired condition (kuhn 1970, czech 1995). although the modern concept of ecosystem management still struggles with the problem of poor definition, its connotative attributes are nevertheless very different from concepts about “ecosystems and management” that were expressed in the 1960s and 1970s. today the increasing adoption of what is called “ecosystem management” does represent a genuine transfer of popular and professional loyalty from one group of ideas and values to another. this shift reflects a transfer of loyalty from the traditional “resource management” paradigm to values associated with ecosystem management. distinctions of the ecosystem management paradigm in the united states, major federal agencies have always had “jurisdiction” over ecosystems, but they have, until recently, never attempted to manage their jurisdictions “as ecosystems”. governed by a paradigm of resource management (rm), the entity of value was a particular “resource,” either an individual species or an abiotic component of the system, such as water, soil, or a mineral. the resource was seen as a commodity and its value was “use.” biologically, this meant that, in a rm paradigm, the value units of management were the species or abiotic components and the spatial units of management were the sites on which they occurred. the outcome of rm at the biological level is singlespecies management, either as commodialces vol. 38, 2002 van dyke et al. – ecosystem management and moose 5 9 ties (harvestable species of plants and animals) or as units of rarity to be preserved (endangered species). in rm, the mechanisms of management are site-specific activities performed by humans, usually through direct intervention. time scales are relatively short-term, and jurisdictional authority and management decisions are within the boundaries of individual agencies. the long-term goal is optimal, renewable, and sustainable production of natural resources as commodities for multiple uses, and, within this larger aim, individual management objectives are set and determined by demand for commodities that the system can supply. ecosystem management has emerged as a meaningful alternative to the rm paradigm and to more local, site specific management approaches, largely through 4 recent scientific and technical achievements: (1) the estimation of minimum viable populations (mvps) and population viability analysis (pva), leading to the scientific consensus that small populations of individuals in isolated reserves will not persist in the long term; (2) the development of remote sensing data collection techniques and geographic information systems (gis), which make the collection and analysis of landscape-scale data manageable; (3) increasing scale and complexity of environmental problems and associated threats that frustrate conservation efforts for individual species and habitats at local scales; and (4) a shift in public attitudes away from valuing the commodities produced by ecosystems for human use to valuing experience and appreciation of the functioning ecosystem itself. thus, ecosystem management owes its emergence not only to shifting public values, but also to increased technical opportunity. entity of value and sustainability. _ the ecosystem management (em) paradigm has gained support because of its ability to deal with changing biological and sociopolitical structures that frustrate the rm paradigm. what gives the em paradigm this advantage is a fundamental shift in the entity of value. in contrast to a former emphasis on the value of resources as commodities, em assumes that the entity of value is the ecosystem itself. that is, the ecosystem, on its own, is perceived as an object worthy of respect and admiration, valued for its beauty, complexity, history, and cultural significance. further, what is valued in the em paradigm is a state or collection of states of the ecosystem that permit long-term delivery of overall ecosystem services, the stability and persistence of ecosystem components (resident populations and communities), and the continuing, long-term stability of transfers of matter and energy within the system. the purpose of achieving such a state is to ensure the persistence of the ecosystem and its functions. specific management goals of how much can be taken from or used in the system are set by the capacity of the system to deliver the desired goods and services, not by the demand for the goods and services. the value of the ecosystem then rightfully entails human obligation to see that the ecosystem persists, and, although goods and services may be outputs of the ecosystem, the ultimate goal is the sustainability of the system, not the deliverability of resource commodities. biodiversity. _ in the em paradigm, the significance of individual elements of the ecosystem, whether communities, habitats, species, or abiotic features of climate, landscape, topography, soil, water, or elements are understood and determined in relation to their role in the stability and functioning of the system. all such components may have roles in the ongoing structure and function of the system, and as such they are considered and conserved at appropriate levels in management. thus, em is more attractive ecosystem management and moose – van dyke et al. alces vol. 38, 2002 6 0 to current managers and conservationists than rm. faced with continued increases in endangered and threatened species, managers are learning that management of ecosystems and landscapes often represents the only way to save both endangered species and overall biodiversity (wilson 1986, franklin 1993), as well as a means to preserve rare or poorly known habitats and ecological subsystems (franklin 1993). management mechanisms and decisions. — in em, the mechanisms of management are primarily through identification and manipulation of landscape-scale processes. management time scales are long-range, and units of management are large areas that may not fall within the jurisdiction of a single agency or government control, but may include jurisdictions of various levels of government as well as private ownership. thus, management decisions must incorporate decision-making strategies that involve all agencies with jurisdiction over lands or processes in the ecosystem, private landowners within or adjacent to the system who depend on outputs and services from the system, and nonresidents who have an interest in the state of the system. with these concepts in mind, we offer a definition that distinguishes ecosystem management from other types of land and resource management. we propose that ecosystem management be defined as “a pattern of prescribed, goal-oriented environmental manipulations that: (1) treat a specified ecological system as the fundamental unit to be managed; (2) has, as its desired outcome, the achievement of a state or a collection of states in the ecosystem such that historical components, structure, function, products, and services of the ecosystem will persist within biological and historical ranges and rates of change over long time periods; (3) uses naturally occurring, landscape-scale processes as the primary means of achieving management objectives; and (4) determine management objectives through cooperative decision-making of individuals and groups who reside in, administer, and/or have vested interests in the state of the ecosystem”. grumbine’s 10 themes of ecosystem management (grumbine 1994), the esas 8 primary characteristics of ecosystem management (christensen et al. 1996), and more’s (1996) 5 dimensions of ecosystem management can be seen as parallel expressions of similar values and concepts. these dimensions include; (1) the long-term sustainability of the ecosystem; (2) the maintenance of viable populations of all native species; (3) the representation of native ecosystem types across their natural range of variation within protected areas; (4) management through ecological processes; and (5) the accommodation of human use and occupancy within management constraints. the development of ecosystem management has consistently included and stressed 3 premises. first, the ecosystem, rather than individual organisms, populations, species, or habitats, is considered the appropriate management unit. second, emphasis is placed on the development and use of adaptive management models, which treat the ecosystem as the subject of study and research, and treat management activities as experimental and uncertain. this means that in ecosystem management, management decisions represent hypotheses about how ecosystems work. management actions that implement such decisions are therefore to be viewed as experimental tests of such hypotheses, and the outcomes of such management actions represent the results of these tests. thus, in ecosystem management, management decisions should carefully consider all reasonable alternatives. management actions should, whenalces vol. 38, 2002 van dyke et al. – ecosystem management and moose 6 1 ever possible, follow careful experimental design, include environmental controls (untreated sites or subjects), and be carefully monitored over time. the outcome of the management action should be viewed in terms of whether it supported or refuted the hypothesis of the management decision, and future decisions should be considered accordingly, under the full light of professional scientific scrutiny and public accountability. if the experimental design is sound, the results of the management action should be relatively unambiguous, but must still be interpreted stochastically (within a range of outcomes with differing probabilities), rather than as deterministic outcomes generated by simple cause-and-effect relationships. finally, ecosystem management is characterized by processes in which those with vested interests in the health and services of the ecosystem (stakeholders) participate in management decisions. defining the value and function of ecosystem management more (1999) defined public parks (such as national parks or national forests) as “organizations of natural and social resources that have been set aside in the public domain to accomplish a function or set of functions. generally these functions concern the preservation of a unique or …scarce resource in the service of the public good, both for present and future generations.” in the modern concept of ecosystem management, the ecosystem replaces the more limited concept of “park” as the entity of organization, but the need to define the function of the entity remains essential. for ecosystem management to succeed, managers must begin by defining explicitly what function or functions are to be accomplished by the ecosystem and its management. specific functions will vary according to individual ecosystems. most functions will require management for provision of essential ecosystem services, including climate and water regulation, conservation of native species, protection of interests of stakeholder groups, efficient use of ecosystem production that can be used as commodities, and long-term persistence of the ecosystem to ensure continuance of these services. a clear articulation of the function and values of ecosystem management is required to develop a unified set of values and objectives with the public, other government agencies and legislative bodies, and w i t h s p e c i f i c s t a k e h o l d e r g r o u p s . stakeholders are operationally defined as “persons or groups that have, or claim, ownership, rights, or interest” in the ecosystem and its management, past, present, or future (clarkson 1995). following values analysis, a management agency should subsequently conduct a “stakeholder analysis,” identifying various stakeholder groups and the nature of the vested interest of each in the management of the system (stead and stead 2000). stakeholders generally claim interest, ownership, or rights (legal, moral, individual, or collective) in a system as a result of past transactions between themselves and at least one of the managing agencies (clarkson 1995). if managers can successfully define and articulate the function of the ecosystem and its management in a manner that addresses the legitimate c l a i m s , r i g h t s , a n d o w n e r s h i p o f stakeholders, they can then evaluate ongoing and proposed management activities by objective guidelines. in turn, managers are then accountable for those guidelines. m a n a g i n g e c o s y s t e m c o m p o n e n t s , structure, and ecological function historically, a problem in managing ecosystem components, such as species, at the ecosystem level has been the lack of a generally accepted classification system of what constitutes a fundamental biological ecosystem management and moose – van dyke et al. alces vol. 38, 2002 6 2 conservation unit in ecosystems. a first step is to categorize ecosystems at regional, coarse, intermediate, and local geographic scales (fig. 1, poiani et al. 2000). once identified, ecosystem conservation requires identification and protection of focal ecosystems and the ecological processes that sustain them (pickett et al. 1992, meyer 1997). one methodology to meet these requirements is through the identification of “functional conservation areas,” defined as geographic domains that “maintain focal ecosystems, species, and supporting ecological processes within their ranges of variability” over the long-term (100-500 years) (fig. 2, poiani et al. 2000). three scales for functional conservation areas are recognized: sites, landscapes, and networks. sites conserve a small number of ecosystems or species at scales below landscape levels, while landscapes conserve many ecosystems and species at scales below regional levels. networks are integrated sets of sites and landscapes designed to protect regional scale species (poiani et al. 2000). an ecosystem can be judged to be “functional” according to 4 criteria: (1) it possesses the historic composition and structure of the ecosystem and its species within a natural range of variability; (2) its dominant environmental regimes are controlled by natural processes; (3) it is sufficiently large to possess at least one minimum dynamic area (50 times the size of the average disturbance patch) (pickett and thompson 1978); and (4) it is connected to other essential landscape elements and its species are free to move among those elements. managing stakeholder groups, ecosystem jurisdiction and political process: the obstacles to ecosystem management in the united states, one of the intentions of the elaborate planning and complex procedural requirements inherent in the forest and rangeland renewable refig. 1. a method of categorizing ecosystems at regional, coarse, intermediate, and local geographic scales. modified from poiani et al. (2000) ©2001 american institute of biological sciences. coarse alces vol. 38, 2002 van dyke et al. – ecosystem management and moose 6 3 sources planning act of 1974 (rpa) and the national forest management act (nfma) of 1976 was to improve decisionmaking and reduce conflict over use of the nation’s forest and range resources. in fact, the opposite has occurred. the rpa has become an object of ridicule within and outside of government agencies because it mandates long-range assessment and planning, but contains no funding mechanisms to achieve it (j. kie, personal communication). forest planning under the nfma has not fared much better. by 1990, 14 years after the passage of the nfma, 92 of 94 completed forest plans were under appeal. five were in court and one was declared illegal. further, 332 active appeals were pending against these plans, brought by conservation groups, commodity interests, off-road vehicle enthusiasts, state and local governments, native american tribes, and private citizens (behan 1990). such litigation manifests, in part, the long and difficult process of achieving a consensus about values in ecosystem management. the nfma directed the forest service toward an ecosystem management approach, and engagement in that process requires a longterm commitment to achieve lasting agreement over ecosystem values and management actions. however, the necessary conditions of public cooperation and trust remain unrealized. such failures in public cooperation and trust will be a fatal obstacle to ecosystem management if they are allowed to persist. successful ecosystem management will require the creation of permanent interagency committees, boards, and working groups in which all agencies and all primary stakeholders from the private sector are presented. such bodies must then move the concept of ecosystem management from the “discussion agenda” (visions discussed or defined by individual agencies) to the “decision agenda” (ideas that are submitted for decisive agency action). we identify 3 obstacles that such groups must overcome to achieve functional ecosystem management and ultimately produce effective procedures that translate ecosystem management into policy. fig. 2. definitions of functional sites, landscapes, and networks and their relationships to biodiversity at various spatial scales. modified from poiani et al. (2000). © 2001 american institute of biological sciences. coarse, ecosystem management and moose – van dyke et al. alces vol. 38, 2002 6 4 the need for unified vision and values for ecosystems and ecosystem management. — the resource management paradigm invested all decision making within the individual agency judged responsible for a particular resource, or for a particular land unit within the agency’s jurisdiction. in contrast, ecosystem management is actually a management system of stakeholder groups and management agencies. therefore stakeholders and agencies must have unifying values and purposes to participate constructively in managing ecosystems. the present reality is one of polarized, fragmented groups of stakeholders and agencies with different and conflicting values and visions of ecosystems, leading to separate agendas that foster distrust and conflict. for example, one research tool used to e v a l u a t e t h e i n t e r e s t s o f c o r p o r a t e stakeholders is the “survey inventory,” a list of up to 50 different issues important to stakeholders in corporate cultures (clarkson 1995). in a personal interview or written response, stakeholders rank listed issues according to their perceived importance. managers then use these responses as a first step to identify stakeholder interests, and begin to build value systems and management strategies informed by these rankings. the rankings collected directly from the stakeholders not only better inform managers of stakeholder interests, but also help managers to distinguish between “social issues” of ecosystem management (matters of importance to society at large, often already regulated by existing laws and regulations) and “stakeholder issues” (matters of importance to particular groups, often unregulated and not addressed by existing laws and regulations). tools such as survey inventories could help managing agencies and stakeholders to more clearly define and reach agreement on the management agencies’ “responsibilities to stakeholder groups”. if responsibilities for management are made explicit, they help to define what the prescribed outcomes of ecosystem management ought to be. if these responsibilities are fulfilled, agency-stakeholder relations grow in trust, move toward effective cooperation, and improve prospects for long-term success in ecosystem management. the performance of the agency can then be better evaluated by the stakeholders, and serve as a basis for discussion of future management strategies. the need for unified sources of information and analysis. — to cooperate effectively in ecosystem management, div e r s e a g e n c i e s n e e d a c o m m o n clearinghouse of information and analysis regarding ecosystem processes and their responses to management systems. however, current information is dispersed among scientific literature, proceedings of professional conferences, and agency reports. the information varies in quality, reliability, focus, format, and accessibility, and there are currently no uniformly accepted standards for data collection among agencies. we propose that such a clearinghouse be created, with appropriate oversight by agencies with jurisdiction over the ecosystem, before a management plan be developed for any particular ecosystem. the need to translate research into policy — if ecosystem management is to have a basis in science and a foundation of professional credibility, it must have the means to smoothly translate reliable research findings into informed policy. this condition requires established and ongoing channels of communication and high levels of trust among researchers, managers, and lawmakers. berry et al. (1998), for example, call for a radical restructuring of the ecosystem management effort. their proalces vol. 38, 2002 van dyke et al. – ecosystem management and moose 6 5 posal includes: (1) a federal, legislative mandate to achieve ecosystem management in all federal land management agencies; (2) establishing regional “boards of ecosystem management research” with representatives of all major stakeholders; (3) a common information clearinghouse to set clear and consistent standards for ecosystem research and serve as a single source for getting results of past studies; (4) an independent science oversight group, responsible to the board and appointed independently of any one agency or interest group to provide direction and review of current research and management efforts; and (5) a project management team responsible to the board that would collect research, development, and operational funds from agencies and stakeholder groups and allocate them to appropriate research efforts. the team would be advised of the merits of proposals and outcomes by outside researchers through independent peer review. implications of ecosystem management for moose currently, most jurisdictions with large moose populations manage moose as a featured species because of recreation, aest h e t i c , a n d e c o n o m i c c o n s i d e r a t i o n s (thompson and stewart 1998). as noted earlier, some have voiced concern that ecosystem management represents such a broad approach that individual species, such as moose, might not be effectively managed (crichton et al. 1998). in contrast, we suggest that moose populations may benefit from ecosystem management. moose have an affinity for early successional vegetation that tends to increase under management that actively employs ecosystem processes such as fire and flooding. moose also have high value among multiple stakeholder groups. finally, moose are often associated with habitats of high species richness. we also explore the potentially negative effects of large predators on moose populations. these effects may increase under ecosystem management practices that encourage the persistence, and even growth, of such predator populations. moose and ecosystem processes in both aquatic and terrestrial environments, moose can exert profound influence on the plant species composition, habitat distribution, and nutrient cycling of ecosystems. for example, in northern boreal forests, moose prevent saplings of preferred species from growing into the tree canopy, resulting in a forest with fewer canopy trees and a well-developed understory of shrubs and herbs (mcinnes et al. 1992). at light to moderate levels, browsing leads to increased production efficiencies (higher rates of production per biomass) in shrubs and saplings. through browsing, moose also reduce the quantity and quality of litter and soil nutrients, driving a complex set of ecological interactions between browse, litter quality, and soil nutrients (mcinnes et al 1992). similar effects are seen in mixed deciduous-coniferous forests, where moose typically browse preferentially on deciduous hardwoods. this pattern of feeding not only changes forest composition, but, more generally, reduces nitrogen mineralization, nitrogen inputs, and overall primary productivity of the forest because the browsing reduces the quantity and quality of litter returned to the soil (pastor et al. 1993). moose, in conjunction with snowshoe hare (lepus americanus), also can reduce fine root production in plants as a result of their herbivory on aerial biomass (ruess et al. 1998). in lakes and ponds, moose may consume up to 95% of submerged aquatic vegetation, particularly various species of pond lilies, which can trigger significant declines in such plant populations and induce major changes in plant species compoecosystem management and moose – van dyke et al. alces vol. 38, 2002 6 6 sition in the pond (belovsky 1981a). habitat relationships the intermediate disturbance hypothesis predicts that maximum species diversity, particularly plant species diversity, is most likely to occur in habitats experiencing intermediate or moderate levels of disturbance (loucks 1970, connell 1978, petraitis et al. 1989), because disturbance removes a subset of pre-existing species, making a portion of the area available for colonization. too little disturbance reduces areas available for colonization, and too much eliminates too many pre-established species, creating a “species debt” that new colonist species cannot fill in a short time. thus, ecosystem management must seek to incorporate both natural and prescribed patterns of environmental disturbance at intermediate levels to achieve its goal of enhancing the persistence of native species and the overall species richness of the ecosystem. while the terms “intermediate” and “moderate” are not always well-defined, they are often used to refer either to the magnitude of the disturbance or to its frequency or both (bendix 1997). on historic range, moose have typically occupied habitats associated with intermediate levels of disturbance, specifically habitats where vegetation is dominated by relatively short-lived species that are adapted to disturbances of intermediate strength and frequency, such as fire and flooding. in south-central montana, for example, shiras moose (a. a. shirasi) preferred aspen (populus tremuloides) habitats in all seasons compared to all other available habitat types table 2, van dyke et al. 1995). aspen is a short-lived deciduous tree whose presence in the surrounding landscape of a coniferous forest is strongly dependent on recurrent fire of intermediate frequency and magnitude. fire does not automatically ensure prolific growth and regeneration, but, on suitable sites, mature aspen with sufficient pre-burn root biomass will produce a strong suckering response with densities of up to 110,000 shoots per ha (renkin table 2. seasonal habitat selection by 3 male (m) and 10 female (f) moose in the fiddler and fishtail creek drainages, south-central montana, 1989-93. numbers indicate percentages. symbols in parentheses indicate selection for (+), selection against (-), or no selection (0). p (p = probability that difference between use and availability is due to random variation)< 0.01 for all cases of selection and for differences between sexes, except where noted. after van dyke et al. (1995). used by permission. moose locations % annual cover type available winter spring summer autumn m use f use pattern aspen 17.5 43.0 (+) 40.2 (+) 56.5 (+) 36.0 (+) 60.0 (+) 40.2 (+) m>f shrub-dominated wetland 8.1 23.9 (+) 20.7 (+) 8.7 (0) 17.41 (0) 17.11 (0) 16.5 (+) m=f lodgepole 55.0 21.8 (-) 20.7 (-) 17.4 (-) 31.1 (-) 12.6 (-) 25.4 (-) m (accessed november 2006). moen, r., j. pastor, and y. cohen. 1996. interpreting behavior from activity counters in gps collars on moose. alces 32: 101-108. _____, _____, _____. 2001. effects of animal activity on gps telemetry location attempts. alces 37: 207-216. _____, _____, _____, and c. c schwartz. 1996. effects of moose movement and habitat use on gps collar performance. journal of wildlife management 60: 659-668. mysterud, a., and e. ostbye. 1999. cover as a habitat element for temperate ungulates: effects on habitat selection and demography. wildlife society bulletin 27: 385-394. oosenbrug, s. m., r. w. mcneily, e. w. mercer, and j. f. folinsbee. 1986. some alces vol. 44, 2008 klassen and rea – nocturnal activity of moose 109 aspects of moose-vehicular collisions in eastern newfoundland, 1973-1985. alces 22: 377-393. peterson, r. l. 1955. north american moose. university of toronto press, toronto, canada. rea, r. v., d. p. hodder, and k. n. child. 2004. considerations for natural mineral licks used by moose in land use planning and development. alces 40: 161-167. risenhoover, k. l. 1986. winter activity patterns of moose in interior alaska. journal of wildlife management 50: 727-734. roshchevsky, m. p., n. a. chermnykh, and j. e. azarov. 1999. cardiac reactions in the behaviour of young moose. alces 35: 143-150. silverberg, j. k., p. j. pekins, and r. a. robertson. 2002. impacts of wildlife viewing on moose use of a roadside salt lick. alces 38: 205-211. _____, _____, _____. 2003. moose responses to wildlife viewing and traffic stimuli. alces 39: 153-160. simkin, d. w. 1963. tagging moose by helicopter. journal of wildlife management 27: 136-139. tankersley, n. g., and w. c. gasaway. 1983. mineral lick use by moose in alaska. canadian journal of zoology 61: 2242-2248. timmerman, h. r. 1992. moose sociobiology and implications for harvest. alces 28: 59-77. vercauteren, k. c., and m. j. pipas. 2003. a review of color vision in white-tailed deer. wildlife society bulletin 31: 684-691. zaitsev, v. a. 2002. structure of moose (alces alces) populations in russia with special reference to communication distances. alces supplement 2: 137-141. zheleznov, n. k., and l. m. fox. 2001. reproductive life history and fertility of moose in north asia. alces 37: 189-200. 4203(25-31).pdf alces vol. 42, 2006 lynch first nations hunting 25 does first nation’s hunting impact moose productivity in alberta? gerry m. lynch 4158 petty creek road, alberton, mt 59820, usa abstract: wildlife biologists and members of the hunting public in alberta voiced concerns that unregulated hunting by first nations’ hunters was detrimental to some moose populations. moose population dynamics were examined in 3 study areas where first nations hunting occurred. provincially licensed sport hunters were only allowed to harvest antlered moose in all 3 areas, but numbers of permits were unlimited. moose populations in some management areas were characterized by strongly biased sex ratios in favor of females, high mean age of the female cohort, and reduced reproductive performance. in wildlife management unit (wmu) 358, where hunting by first nations’ hunters was considered “heavy”, the sex ratio was not strongly biased, moose numbers were sustained at a higher level, and both pregnancy and twinning rates were higher than in the other areas. contrary to the fears of wildlife managers and sport hunters, moose hunting by first nations’ hunters in wmu 358 did not appear to be detrimental, but may have actually enhanced moose productivity. the moose harvest there probably resembled a selective harvest system where females as well as males were included. are currently managed under a non-selective male-only harvest strategy. alces vol. 42: 25-31 (2006) key words: age structure, alces alces, first nations hunting, moose harvest, native hunting, productivity, selective harvest, sex ratio in many jurisdictions across north america, the moose (alces alces) resource is shared by indigenous people (first nations people in canada, native americans in the united states), who are exercising their treaty hunting rights, and sport hunters licensed to groups have been known to occur and wildlife managers must consider the harvest by both when formulating the annual hunting regulations. an annual telephone questionnaire in alberta (lynch and birkholz 2000) provided reasonable estimates of the annual harvest by licensed sport hunters, but the take by first nations’ hunters was never documented. in response to the uncertainty concerning the total harvest of moose each year, those who set hunting regulations adopted a conservative approach by forbidding the removal of any females by sport hunters and later by limiting numbers of bull moose permits. the stated goal of moose management in alberta was to increase moose numbers in most wildlife management units (wmus). history has shown that unregulated hunting can profoundly impact wildlife populations by reducing animal numbers or by altering population characteristics, including age structures of population cohorts and sex ratios. in alberta, the northern moose management program (nmmp) enabled biologists to examine moose population characteristics over a large portion of northern moose range. three study areas (wmus 346, 350, and 358) allowed a comparison of the characteristics of moose populations in an area considered to be heavily utilized by first nations’ hunters (wmu 358) and 2 other areas where first nations hunting was not considered to be first nations hunting – lynch alces vol. 42, 2006 26 examine the impact of unregulated hunting by first nations’ hunters on the moose population in wmu 358, an area considered by moose biologists to have been heavily hunted by that group for many years. study area the 3 study areas were typical boreal forest, covered by mosaics of open muskeg and forests dominated by mixed and pure stands of white spruce (picea glauca), black spruce (p. mariana), aspen (populus tremuloides), and lodgepole pine (pinus contorta). locally important tree species were tamarack (larix laricina), white birch (betula papyrifera), balsam poplar (p. balsamifera), and jack pine (p. banksiana). shrub communities were dominated by species of alder (alnus spp.) and willow (salix spp.), depending on drainage. cut logging or oil and gas exploration. the 3 study areas were located in westcentral alberta, roughly north of highway 16 and south of the peace river. wmu 358 was situated on the alberta-british columbia boundary. some poaching and hunting by first nations’ hunters occurred in all areas, but first nations hunting in wmu 358 was feared to be excessive. hunting by licensed sport hunters was considered heavy in all areas, except in wmu 350, a semi-wilderness area with limited hunter access. moose in all areas were preyed upon by gray wolves (canis lupus), black bears (ursus americanus), and grizzly bears (u. arctos), but grizzlies were only occasional visitors to wmu 358. all wmu 350), logging, and the petrochemical industry. the areas of wmus 346, 350, and 358 were 5,205 km2, 6,253 km2, and 2,886 km2, respectively. methods moose surveys were conducted using the aircraft were used for the pre-survey activity. using robertson 44 and bell jet ranger helireached 20 % or less. a global positioning system (gps) and geographic information system (gis) helped with the logistics of surveying large geographical areas (lynch and shumaker 1995). helicopter wildlife management was the company contracted to captured female moose by net gun from a hughes 500 helicopter. an additional 24 females were darted from a robertson 44 helicopter and immobilized with carfentanil citrate and xylazine. captured a radio-collar, and a tooth was extracted for aging. moose transmitters were in the 150megahertz frequency range and included a mortality switch that doubled the pulse rate when a mortality occurred. moose incisor bars were collected from hunter-killed moose for aging. all teeth were sectioned and aged according to the pimlott (1959). radio-collared moose were found once per their location and status (dead or alive). when a transmitter was found to be in “mortality mode”, a helicopter was dispatched to the site immediately so the cause of death could be determined. each spring a helicopter was used to relocate newborn calves belonging to radioarea ensured that newborns were found before they were lost to early calf mortality. moose population parameters were compared between the 3 study areas to try to detect any adverse effects related to the unregulated hunting by first nations’ hunters in wmu 358. moose densities, adult sex ratios, calf production and survival, adult survival, and twinning alces vol. 42, 2006 lynch first nations hunting 27 rates were examined for discrepancies. hunting pressure (success and effort) by licensed sport hunters was measured by the annual telephone questionnaire. hunting pressure by first nations’ hunters was not as “light”, “moderate”, or “heavy” in the 3 observations by fish and wildlife division results areas every year during the 5 years of the nmmp (table 1). in wmu 346, the population mean ranged from 2,308 (1997-98) to 3,369 (1996-97). the low count in 1997-98 was not considered accurate and was attributed to poor survey conditions in that wmu, when warm winds suddenly melted most of the snow cover. excluding 1997-98, population means in wmu 346 ranged from 2,669 to 3,369. bulls per 100 cows ranged from 9 in 1993-94 to 21 in 1996-97. increasing bull:cow ratios were in response to changes to the hunting regulations beginning in 1994 that reduced the harvest of antlered moose by sport hunters. calves per 100 cows ranged from 31 to 46 at the time of the surveys, about 6 months after birth. the estimated moose population means in wmu 350 ranged from 2,701 (1994-95) to 3,593 (1996-97). the lower mean in 1994-95 was attributed to a greater harvest of bulls that year when hunting pressure shifted to wmu 350 from adjacent areas where the general hunting season format (unlimited over-thecounter license sales) was changed to a limited entry draw. this effect also showed itself in the lower bull:cow ratio that year (10 bulls per 100 cows) in the post hunting season population. other than 1994-95, bulls per 100 cows ranged from 23 to 29 in wmu 350. calves per 100 cows at the time of the surveys ranged wmu year population 95% c.l. per 100 cows density mean +/% bulls calves /100 km2 /mi2 346 93/94 3,118 16 9 42 65 1.7 94/95 2,792 19 15 45 54 1.4 95/96 2,669 15 20 31 51 1.3 96/97 3,369 18 21 42 65 1.7 97/98 2,308 11 17 46 44 1.2 350 93/94 2,952 19 23 45 51 1.3 94/95 2,701 16 10 32 45 1.2 95/96 3,557 20 25 39 57 1.5 96/97 3,593 19 25 35 57 1.5 97/98 3,204 14 29 32 51 1.3 358 93/94 1,328 15 20 42 46 1.2 94/95 2,478 15 22 56 86 2.2 95/96 2,682 18 30 46 93 2.4 96/97 2,842 15 28 50 98 2.6 97/98 2,552 11 24 56 88 2.3 table 1. five years of aerial survey results from wmus 346, 350, and 358 in northern alberta. first nations hunting – lynch alces vol. 42, 2006 28 from a low of 32 to a high of 45. in wmu 358, the population mean ranged from 1,328 in 1993-94 to a high of 2,842 in 1996-97. in 1994-95 the mean was 2,478, 80% higher than the previous year’s count. this discrepancy caused us to question the accuracy of our survey technique, so the moose survey in wmu 358 was replicated 2 months later in february 1995 in order to check the technique. the second survey resulted in a mean of 2,429, survey technique. the low numbers in 199394 were attributed to movement of moose out of wmu 358 into adjacent agricultural areas in response to deep snow. movements of moose into adjacent areas during winter wmu 358. future moose surveys in all areas were scheduled for completion by the end of december, before moose were driven out of a survey area or into heavy cover by deep snow (lynch and shumaker 1995). bulls per 100 cows in wmu 358 ranged from 20 to 28 during the 5-year period, and calves per 100 cows at survey time ranged from 42 to 56. sources of mortality of radio-collared percent (n = 39) died during the course of the project. six of 7 deaths in wmu 346 were due to being shot and 1 to predation. in wmu 350, the wilderness area, a total of 11 radio-collared moose died during the study. n = 6) was attributed to predation by gray wolves. one died giving birth, 1 died from winter-related causes, and 3 were shot. in wmu 358, where hunting by first nations’ hunters was common, 21 of those that died were shot (n = 20). the only other death in wmu 358 was related to severe winter conditions. over the term of the study, 224 cumulative annual adult radio-collared females were eligible to bear calves (table 3). of these, 176 (79%) produced litters. percent cows bearing calves was about the same in wmus 346 and 350 (74 % and 73%, respectively), but in wmu 358, the area heavily used by first nations’ hunters, 73 of 85 (87%) of eligible radio-collared cows produced calves (table 3). in spring of each year a helicopter was used to obtain a “visual” on each radio-collared cow moose in order to determine whether it in june was used, but it soon became apparent that calves were being lost to early mortality factors and were therefore not included in calving statistics. beginning in 1996, 3 accurate assessment of calf production. table 4 examines reproductive performance of radio-collared females during 1996 and 1997, wmu collared source of mortality n predation birthing winter shot totals 346 34 1 (14%) 0 0 6 (86%) 7 (21%) 350 36 6 (55%) 1 (9%) 1 (9%) 3 (27%) 11 (31%) 358 44 0 0 1 (5%) 20 (95%) 21 (48%) totals 114 7 (18%) 1 (3%) 2 (6%) 29 (74%) 39 (34%) table 2. numbers of female moose captured and sources of their mortality during the 5-year term of the nmmp in northern alberta. wmu eligible cows litters % 346 78 58 74 350 62 45 73 358 85 73 87 overall 224 176 79 table 3. reproductive performance of radio-collared moose as determined at time of parturition during spring of 1996 and 1997 in northern alberta. alces vol. 42, 2006 lynch first nations hunting 29 wmu 346, there were 38 litters that included 6 sets of twins (16% of litters). in wmu 350, 28 litters included 2 sets of twins (7%), and in wmu 358, where first nations’ hunters were active, there were 39 litters that included 10 sets of twins (26%). percent twins observed during the winter aerial surveys were 4.5, 4.2, and 7.5 in wmus 346, 350, and 358, respectively (table 4). the months of life. annual survival rates of radio-collared cows in wmus 346, 350, and 358 were 0.928, 0.863, and 0.751, respectively (table 5). calf annual survival rates were 0.745, 0.575, and 0.614, respectively. discussion the nmmp in wmus where cow and calf moose mortality studies were underway or where moose hunting regulations were changed from a general season format to a limited entry draw in order to reduce the bull harvest and restore less biased adult sex ratios. we were particularly interested in population trends in wmu 358, where many believed that moose were being over-hunted by first change in population means in any of the 3 study areas during the 5 years of the project. numbers in wmu 358 actually suggested a trend toward increasing, not decreasing moose numbers. moose densities in wmu 358 sustained themselves at 50-75 % higher levels compared to the other 2 study areas, in spite of removal of many female moose by first nations’ hunters. numbers of bulls per 100 cows more than doubled in wmu 346 in response to hunting season changes that curtailed the antlered moose harvest by licensed sport hunters. in 1993, prior to the regulations change, approximately 64 % of available bulls were shot each year. by 1996 this rate had dropped to 24 % and the sex ratio had gone from 9 to 21 bulls per 100 cows in the post hunting season population. in wmu 350, the wilderness area, the bull: cow ratio declined dramatically in 1994 when approximately 64 % of bulls were removed that season by licensed hunters. the spike in hunting pressure that year was attributed to movement of hunters into wmu 350 from adjacent areas where a limited entry draw had been initiated. the following year, in 1995, wmu 350 was included with adjacent areas limited by a draw for antlered moose permits. this reduced the bull harvest in subsequent years and led to a recovery in the ratio of bulls wmu radio-collared females aerial surveys n litters n twins % n cows single calf twins % calving % twinning 346 38 6 16 886 317 15 37.5 4.5 350 28 2 7 1,185 413 18 36.4 4.2 358 39 10 26 1,457 608 49 45.1 7.5 table 4. twinning rates among radio-collared moose and observed1 during 5 years of aerial surveys in northern alberta. 1calving and twinning rates determined by telemetry occurred at birth. rates determined from aerial surveys occurred approximately 6 months after birth. wmu females calves survival mortality % died survival mortality 346 0.928 0.072 14.9 0.745 0.255 350 0.863 0.137 28.2 0.575 0.425 358 0.751 0.249 25.2 0.614 0.386 table 5. annual survival rates calculated of radiocollared female moose and calves of radio-collared females in northern alberta. first nations hunting – lynch alces vol. 42, 2006 30 to cows. the ratio of bulls to cows in the post hunting season population appeared to be highly sensitive to hunting pressure. in wmu 358, licensed moose hunters removed 39, 25, 27, and 31% of available bulls between 1994 and 1997, respectively. in addition, an unknown number of bulls were harvested by first nations’ hunters and poachers. in spite of the heavy antlered harvest, post hunting season numbers of bulls per 100 cows in wmu 358 were 22, 30, 28, and 24, respectively, during the same period. it was apparent that removal of antlerless moose by first nations’ hunters helped to offset the loss of bulls to hunting, thus preventing the bull: cow ratio from becoming severely skewed. the causes of mortality of radio-collared females in the 3 study areas were not unexpected. the greatest level of mortality occurred in wmu 358, where first nations’ hunters, and possibly poachers, removed 46 % in the annual survival rates calculated for the 3 areas. wmu 358 had the lowest female moose survival rate (0.751), compared to 0.928 and 0.863 in wmus 346 and 350, respectively. we were able to age 1,206 incisor bars from hunter killed male moose and a sample of female moose from wmu 346. the females were harvested as part of a special hunting season to obtain reproductive tracts. the mean ages of bulls in wmus 346 and 358 were 2.5 and 2.7 years, respectively. in wmu 350, where hunting pressure was less, the mean age of males was 3.5 years. wmu 350 also had the lowest percentage of yearlings in the harvest, 32.4%, compared to 50.0% in wmu 346 and 49.7% in wmu 358. bull moose annual mortality rates in wmus 346, 350, and 358 were 49.9%, 33.7%, and 46.2%, respectively. these results were not unexpected and they were in agreement with perceived levels of hunting pressure. the mean age of females in wmu 346, where hunting by first nations’ hunters was minimal, was 7.2 years. we suspected that the female cohort in wmu 350 was also old, while that in wmu 358 was younger, due to the harvest of females there by first nations’ hunters. however, we did not have age data from female moose in those 2 areas. studies of reproductive performance of our radio-collared females suggested that older females might have been less productive than younger, prime-aged animals. seven radio-collared females demonstrated a pattern of alternate year breeding. all were 9 years old or older. all other adult females produced a litter annually. we thought that a younger female cohort in wmu 358, the area used by first nations’ hunters, would be more productive than the old female cohorts in wmus 346 and 350. was a factor in reducing the numbers of twins observed during the winter moose surveys. calf survival seemed highest in wmu 346 where there were fewer predators and less hunting by first nations’ hunters. the lowest calf survival rate was in wmu 350 where both bear species and wolves were abundant. in wmu 358, calves were hunted by first nations’ hunters. these observations were made as subsets of the larger nmmp. other factors that were beyond the scope of the nmmp, such as local climatic events and range quality and condition, were probably relevant factors not considered here. these data were considered observational and not cause and effect. the harvest of cows and calves by first nations’ hunters in wmu 358 was thought to resemble a selective harvest system of management, except that numbers of cows and calves harvested could not be regulated. under this regime, and despite the fact that comparisons between the 3 study areas may not have been sizes) when compared to other areas studied, the moose population in the area “heavily” hunted by first nations’ hunters had some of the highest bull:cow ratios, had the greatest population density, had the highest ratio of alces vol. 42, 2006 lynch first nations hunting 31 calves to cows, had the highest pregnancy rate among radio-collared cows, and had the highest twinning rate recorded. this does not mean that unregulated either sex moose populations, nor does it mean that unregulated hunting by first nations’ hunters is not detrimental to some other moose populations. it does suggest that moose managers in alberta should abandon their non-selective male-only management strategy for moose and opt for a selective harvest system that includes limited cow harvest and increased opportunity to harvest calves. far northern wmus, where moose populations are currently limited by predation, should continue to be managed for a male-only harvest, as prescribed by van ballenberge and dart (1982). acknowledgements this project was paid for by the government of alberta, alberta sustainable development, fish and wildlife division, and by hunters through the alberta conservation association. references gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological paper number 22. university of alaska, fairbanks, alaska, usa. lynch, g. m., and s. birkholz. 2000. a telephone questionnaire to assess moose harvest. alces 36:105-109. _____, and g. e. shumaker. 1995. gps and gis assisted moose surveys. alces 31:145-151. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23:315-321. van ballenberge, v., and j. dart. 1982. harvest yields from moose populations subject to wolf and bear predation. alces 18:258-275. 91 extending body condition scoring beyond measureable rump fat to estimate full range of nutritional condition for moose rebecca l. levine1, rachel a. smiley2, brett r. jesmer3, brendan a. oates4, jacob r. goheen5, thomas r. stephenson6, matthew j. kauffman7, gary l. fralick8, and kevin l. monteith1,2 1haub school of environment and natural resources, university of wyoming, 804 fremont street, laramie, wyoming 82072, usa; 2wyoming cooperative fish and wildlife research unit, department of zoology and physiology, university of wyoming, 1000 east university avenue, laramie, wyoming 82071, usa; 3department of fish and wildlife conservation, virginia tech, 310 west campus drive, blacksburg, virginia 24061, usa; 4washington department of fish and wildlife, 1111 washington street southeast, olympia, washington 98501, usa; 5department of zoology and physiology, university of wyoming, 1000 east university avenue, laramie, wyoming 82071, usa; 6sierra nevada bighorn sheep recovery program, california department of fish and wildlife, 787 north main street, suite 220, bishop, california 93514, usa; 7u.s. geological survey, wyoming cooperative fish and wildlife research unit, department of zoology and physiology, university of wyoming, 1000 east university avenue, laramie, wyoming 82071, usa; 8wyoming game and fish department, p.o. box 1022, thayne, wyoming 83127, usa. abstract: moose (alces alces) populations along the southern extent of their range are largely declining, and there is growing evidence that nutritional condition — which influences several vital rates – is a contributing factor. moose body condition can presently be estimated only when there is measurable subcutaneous rump fat, which equates to animals with >6% ingesta-free body fat (ifbfat). there is need for a technique to allow body fat estimation of animals in poorer body condition (i.e., <6% body fat). we advance current methods for moose, following those used and validated with other ungulate species, by establishing a moose-specific body condition score (bcs) that can be used to estimate ifbfat in the lower range of condition. our modified bcs was related strongly (r2 = 0.89) to ifbfat estimates based on measurable rump fat. by extending the predicted relationship to individuals without measurable fat, the bcs equated severe emaciation with 0.67% ifbfat, supporting the accuracy of the method. the lower end of nutritional condition is important for identifying relationships involving life-history characteristics because most state-dependent changes occur at lower levels of condition. therefore, until the bcs can be validated with moose carcasses, we believe our method to estimate body fat across the full range of condition should yield better understanding of the drivers underlying declining moose populations. alces vol. 58: 91 – 99 (2022) key words: alces alces, ingesta-free body fat, body condition score, moose, nutrition, ultrasound, ultrasonography, validation the nutritional condition (i.e., percent ingesta-free body fat [ifbfat]) of an individual integrates nutrient gains and losses as it reflects previous life-history and habitat quality (cook et al. 2007, monteith et al. 2014). indeed, nutritional condition forms the foundation for life-history of individuals and affects nearly every demographic component of a population (parker et al. 2009, monteith et al. 2014, stephenson et al. 2020). across moose (alces alces) distribution, nutritional limitation underpins body size, moose body condition scoring – levine et al. alces vol. 58, 2022 92 reproductive success, and population growth rate (murray et al. 2006, monteith et al. 2015, hoy et al. 2017, schrempp et al. 2019, jesmer et al. 2021). several nutritional metrics, including iron levels and fat content, were related to probability of pregnancy in western montana (newby and decesare 2020). in utah, production and recruitment of young increased linearly with rump fat measurements (ruprecht et al. 2016). similarly, moose in minnesota were less likely to be pregnant when malnutrition was indicated by bone marrow fat, blood indices, and rump fat measurements (murray et al. 2006, delgiudice et al. 2011). further, in wyoming, body fat was a strong predictor of pregnancy, parturition, survival, and therefore population growth rate (i.e., lambda), thus linking nutritional condition to demography (oates et al. 2021). the role of nutrition in the life-history of moose necessitates a reliable and reproducible metric for determining nutritional condition of individuals and populations to help identify factors affecting population demograhics and enhance conservation and management efforts for this species. methods used to assess nutritional condition of ungulates employ both post-mortem and in vivo indices, including marrow fat (cheatum 1949), kidney fat (riney 1955), back fat (anderson et al. 1972), visual examination of organ fat (kistener et al. 1980), and physical descriptions (franzmann 1977). in vivo methods are preferable because they allow for repeated sampling of individuals which yields potential to connect nutritional condition to life-history and environmental characteristics while avoiding animal sacrifice. when coupled with a body condition score (bcs) acquired via palpation, thickness of rump fat measured via ultrasonography has become the gold standard to accurately estimate total body fat in vivo for ungulates (cook et al. 2001b, 2021a). predictive equations following a standardized approach have been developed and calibrated for mule deer (odocoileus hemionus; stephenson et al. 2002, cook et al. 2007), elk (cervus canadensis; cook et al. 2001a, 2001b), bighorn sheep (ovis canadensis; stephenson et al. 2020), and caribou (rangifer tarandus; cook et al. 2021a). in moose, predictive equations for estimating percent ifbfat based on ultrasonography measurements of maximum depth of rump fat are highly related (r2 = 0.96; stephenson et al. 1998), but ultrasonography alone does not allow estimation across the full range of body condition (<1 mm rump fat). as in other north american cervids, subcutaneous rump fat is depleted when moose reach 5.63% ifbfat (cook et al. 2010, 2021a); however, the bcs derived from palpation to estimate ifbfat below that threshold has not been developed for moose. consequently, quantifying relationships with nutritional condition in moose is hampered by a lack of resolution at lower levels of ifbfat when fitness or behavioral consequences should be most evident (ruprecht et al. 2016, newby and decesare 2020). efforts to address this gap in knowledge do exist, including a body scoring system which delineates individual moose by describing appearance, boniness, and gait (franzmann 1977); however, the scores of live-captured moose using this technique had a statistically significant but weak relationship with ifbfat determined via ultrasonography (r2 = 0.34; delgiudice et al. 2011). similarly, while scoring systems validated for other cervids have been applied to moose (cook et al. 2021b), a species-specific bcs would be more appropriate given the morphological differences among species. validating the relationship between bcs and rump fat for moose would be ideal given its usefulness in other species (cook et al. 2001a, 2007, 2010); however, challenges of sacrificing a sufficient number of moose to determine body alces vol. 58, 2022 moose body condition scoring – levine et al. 93 composition via homogenization and chemical analysis (e.g., stephenson et al. 1998, cook et al. 2001a) have precluded its development. in lieu of validating a bcs for moose via sacrifice, we used an ad hoc approach to develop a bcs for estimating ifbfat of moose with no measurable rump fat. given the established rela tionship between bcs and ifbfat developed with other ungulates (cook et al. 2001a, 2007, 2021a; stephenson et al. 2002, 2020), we developed a bcs for moose. for moose with measurable rump fat, we then regressed their ifbfat estimates and bcs to develop a predictive equation (stephenson et al. 1998). we subsequently extended this relationship to include moose below the threshold of measurable rump fat to estimate ifbfat across the full range of nutritional condition. study area we studied moose (a. a. shirasi) from the sublette herd in the green river basin of northwest wyoming, usa (42.8653˚n, 110.0708˚w) in february 2011, 2012, and 2013 (see jesmer et al. 2017, 2021, oates et al. 2021). winters were characterized by mean temperatures below 18°c and deep snow (annual mean snowfall 160 cm). riparian areas used by moose were dominated by booth’s (salix boothii) and geyer’s willow (s. geyeriana). surrounding areas consisted of either mixed coniferous forest (abies lasiocarpa, pinus contorta, picea engelmannii, pseudotsuga menziesii), aspen forest (populus tremuloides), mixed conifer-aspen forest, or sagebrush (artemisia spp.) steppe. this population of moose was considered stable for the duration of our study (wyoming game and fish department, unpublished data). methods we captured 48 adult female moose via helicopter net-gunning on 13–15 february 2012. moose were blindfolded, hobbled, and restrained in a sternal recumbent position on their left side. the right, incisiform canine was removed following the methods of swift et al. (2002), and age was determined via cementum annuli (matson’s laboratory, milltown, montana, usa). we measured body length from the dorsal margin of the planum nasale to the tip of the tail following the contour of the body using a cloth tape, and measured chest girth from the middle of the sternum to the spinous process while maintaining the tape immediately posterior to the scapula and perpendicular to the spine. subsequently, we predicted body weight using the relationship between body length and chest girth (hundertmark and schwartz 1998). to assess nutritional condition, we measured the maximum depth of rump fat (maxfat; stephenson et al. 1998) using a bantam ii portable ultrasound device (e.i. medical imaging, loveland, colorado, usa) with a 5-mhz linear-array transducer (stephenson et al. 1998). we accompanied ultrasound with palpation and developed a modified bcs (appendix a), analogous to that validated for elk (cook et al. 2001a) and mule deer (cook et al. 2007) and highly correlated with percent ifbfat (r2 ≥ 0.86). the university of wyoming institutional animal care and use committee approved capture and handling procedures (protocol #20140124jg00057). our initial set of analyses used linear regression to establish the relationship between ifbfat and bcs. we calculated percent ifbfat of moose with measurable rump fat using established equations (stephenson et al. 1998), with scaled estimates to correct maxfat to body size (cook et al. 2010). previous maxfat analyses considered animals with minimal rump fat (<3 mm) to have no measurable fat because these measurements represent the fascia thickness (cook et al. 2001a, 2007). nevertheless, our use of conduction moose body condition scoring – levine et al. alces vol. 58, 2022 94 ultrasonography and high-resolution equipment allowed us to avoid inclusion of fascia thickness as part of the rump fat measurement. we therefore distinguished true maxfat measurements from fascia and included those individuals with maxfat >0 mm and <3 mm as animals with measurable rump fat. we excluded moose with no measurable rump fat (maxfat = 0) from our regression of ifbfat and bcs because our aim was to use the derived relationship to predict the ifbfat of these individuals. we used linear regression to establish the relationship between bcs and percent ifbfat (fig. 1) within the known range of ifbfat (>5.63%; animals with measurable rump fat). we then extended the relationship between bcs and values of ifbfat below 5.63%, assuming the linear relationship between bcs and ifbfat would hold (stephenson et al. 2020). during capture, we handled 1 moose that was in extremely poor condition, characteristic of an animal suffering from severe malnutrition, and which ultimately died later that winter. the mortality occurred in early spring (29 april), which was typical of malnourished moose in the region as they were not exposed to predators. based on previous experience with quantifying nutritional condition of ungulates, we expected this individual to have minimal remaining ifbfat (i.e., <1%). we used the estimates of ifbfat from the regression equation as a test case for our derived relationship, anticipating that our scoring system and regression should accurately estimate a starving moose to have little to no body fat. results estimates (±se) of age ranged from 3 to 10 years old (4.5 ± 0.3 years); only 4 of 48 individuals were >7 years old. average body and metatarsus length were 270.1 ± 1.8 cm (range: 223–290 cm) and 56.5 ± 0.2 cm (range: 52–60 cm), respectively. estimated body weight ranged from 244 kg to 419 kg, averaging 367.8 ± 4.8 kg. the maxfat measurements averaged 0.61 ± 0.01 cm, ranging from 0 to 2.0 cm. there was a strong linear relationship between bcs and ifbfat for animals with measurable subcutaneous rump fat (r2 = 0.89, n = 32; fig. 1). extending the linear relationship to include moose with a bcs but without measurable rump fat yielded none with ifbfat >6.0%. all individuals with bcs ≤2.75 were predicted to have no measurable rump fat, and conversely, all individuals with measurable rump fat had bcs > 2.75. the ifbfat estimate was 5.48% (95% ci: 5.15−5.80%) for individuals with bcs of 2.75 which was similar to thresholds where rump fat is depleted (5.8%, stephenson et al. 1998; 5.63%, cook et al. 2010). the predicted ifbfat for the individual in poor condition (presumed <1% ifbfat) was 0.67%. the population average of ifbfat was 6.42 ± 0.34% (range: 0.67–10.57%, n = 32). fig. 1. ingesta-free body fat (ifbfat) relative to body condition score (bcs) of adult female shiras moose during mid-february 2012, sublette county, western wyoming, usa. solid circles represent individuals with measurable subcutaneous rump fat and open circles represent individuals without measurable rump fat. alces vol. 58, 2022 moose body condition scoring – levine et al. 95 discussion poor nutritional condition underlies moose decline at the southern extent of their range (murray et al. 2006, delgiudice et al. 2011, vartanian 2011), thereby calling for adequate tools to monitor their nutritional condition (jesmer et al. 2017, 2021). we established a body condition scoring system to estimate ifbfat in moose (appendix a) with depleted subcutaneous rump fat using bcs systems validated for other species as a foundation (cook et al. 2001a, 2007, 2021a, stephenson et al. 2020). we derived a linear relationship between our scoring system (bcs) and ifbfat using moose where ifbfat could be calculated with measurable maxfat (stephenson et al. 1998). if our bcs scoring system was reliable, we expected that moose without measurable rump fat would have scores that corresponded with ifbfat levels below that detectable via ultrasound. these predictions were consistent with our findings; all moose without measurable rump fat were predicted to have <5.63% ifbfat. further, the derived relationship predicted that a severely malnourished moose had <1% ifbfat. until validation is possible via chemical analyses from animal carcasses, our equation to estimate ifbfat for moose without measurable rump fat should provide meaningful inference when nutritional limitation affects moose populations. indeed, following our reported method herein, ifbfat was related strongly to pregnancy, overwinter adult survival, parturition, and ultimately, was a strong predictor of lambda in the same population (oates et al. 2021). by combining a validated equation for moose with measurable rump fat with a modified bcs, our approach extends the utility of existing methods to quantify nutritional condition of moose. although an established scoring method (franzmann 1977) identified a relationship between moose condition and pregnancy status (testa and adams 1998), it explained only a portion of the variation in ifbfat (r2 = 0.34; delgiudice et al. 2011). with our system, bcs scores were highly correlated with ifbfat (r2 = 0.89, fig. 1), and were comparable to bcs validated in elk (r2 = 0.86, cook et al. 2001a), mule deer (r2 = 0.88, cook et al. 2007), and bighorn sheep (r2 = 0.77, stephenson et al. 2020). accordingly, our bcs system represents nutritional condition more accurately than previous scoring methods in moose, and it is commensurable to bcs systems used extensively in other ungulate species to assess fat reserves of animals in poor condition (monteith et al. 2013, long et al. 2014, proffitt et al. 2021). the lower end of nutritional condition, where fat reserves are depleted, is often the threshold beyond which animals face tradeoffs among nutritional reserves, reproduction, and survival. moose can survive milder winters with body fat <5.63%, but pregnancy rate (newby and decesare 2020, jesmer et al. 2021, oates et al. 2021) and survival probability decline (oates et al. 2021) below this threshold. thus, the point at which animals have depleted subcutaneous fat reserves is critical for drawing connections between life-history and nutrition. relationships between life-history and fat reserves are likely to be overlooked without measurement at the lowest extent of nutritional condition. indeed, changes in probability of pregnancy, parturition, and overwinter survival of adults occurred when ifbfat was <6% (oates et al. 2021), or below the detection range of measurable rump fat. our bcs for moose provided broader characterization of nutritional condition, particularly for individuals at the lowest extent of nutritional condition. we note the importance of adequate training on numerous animals (often >60 but dependent upon user adeptness) moose body condition scoring – levine et al. alces vol. 58, 2022 96 across a range of nutritional condition (cook et al. 2021a) to accurately assess condition using a bcs. the bcs technique, when properly used, aids in identifying factors limiting population growth while linking behavioral and ecological characteristics to nutritional condition. accurate assessment of nutritional condition is critical to identi fy stressors and sources of depressed productivity and survival associated with declining moose populations, and consequently, management options to enhance popu lation performance. acknowledgements we thank many landowners in sublette county, wyoming for allowing us access to their property for moose captures. we thank collaborators, including retired bridger teton national forest biologist, g. hanvey, grand teton national park personnel, and the wyoming game and fish department for logistical support. any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the u.s. government. we acknowledge p. pekins, editor, e. bergman, associate editor, and our reviewers, c. anderson and c. bishop for their careful consideration and feedback on earlier drafts of this manuscript. references anderson, a. e., d. e. medin, and d. c. bowden. 1972. indices of carcass fat in a colorado mule deer population. journal of wildlife management 36: 579–594. doi: 10.2307/3799091 cheatum, e. l. 1949. bone marrow as an index of malnutrition in deer. new york state conservationist 3: 19–22. cook, r. c., j. g. cook, d. l. murray, p. zager, b. k. johnson, and m. w. gratson. 2001a. development of predictive models of nutritional condition for rocky mountain elk. journal of wildlife management 65: 973–987. doi: 10.2307/ 3803046 _____, _____, t. r. stephenson, w. l. myers, s. m. mccorquodale, d. j. valesales, l. l. irwin, p. b. hall, r. d. spenser, s. l. murphie, k. a. schoenecker, and p. j. miller. 2010. revisions of rump fat and body scoring indices for deer, elk, and moose. journal of wildlife management 74: 880–896. doi: 10.2193/2009-031 _____, j. a. crouse, j. g. cook, and t. r. stephenson. 2021a. evaluating indices of nutritional condition for caribou (rangifer tarandus): which are the most valuable and why? canadian journal of zoology 99: 596–613. doi: 10.1139/ cjz-2020-0149 _____, _____, d. l. murray, p. zager, b. k. johnson, and m. w. gratson. 2001b. nutritional condition models for elk: which are the most sensitive, accurate, and precise? journal of wildlife management 65: 988–997. doi: 10.2307/3803047 _____, j. oyster, k. mansfield, and r. b. harris. 2021b. evidence of summer nutritional limitations in a northeastern washington moose population. alces 57: 23–46. _____, t. r. stephenson, w. l. myers, j. g. cook, and l. a. shipley. 2007. validating predictive models of nutritional condition for mule deer. journal of wildlife management 71: 1934–1943. doi: 10.2193/2006-262 delgiudice, g. d., b. a. sampson, m. s. lenarz, m. w. schrage, and a. j. edwards. 2011. winter body condition of moose (alces alces) in a declining population. journal of wildlife diseases 47: 30–40. doi: 10.7589/0090-3558-47.1.30 franzmann, a. w. 1977. condition assessment of alaskan moose. alces 13: 119–127. hoy, s. r., r. o. peterson, and j. a. vucetich. 2017. climate warming is associated with smaller body size and shorter lifespans in alces vol. 58, 2022 moose body condition scoring – levine et al. 97 moose near their southern range limit. global change biology 24: 2488–2497. doi: 10.1111/gcb.14015 hundertmark, k. j., and c. c. schwartz. 1998. predicting body mass of alaskan moose (alces alces gigas) using body measurements and condition assessment. alces 34: 83–89. jesmer, b. r., j. r. goheen, k. l. monteith, and m. j. kauffmana. 2017. statedependent behavior alters endocrine energy relationship: implications for conservation and management. ecological applications 27: 2303–2312. doi: 10.1002/eap.1608 _____, m. j. kauffman, a. b. courtemanch, s. kilpatrick, t. thomas, j. yost, k. l. monteith, and j. r. goheen. 2021. lifehistory theory provides framework for detecting resource limitation: a test of the nutritional buffer hypothesis. ecological applications 31: e02299. kistener, t. p., c. e. trainer, and n. a. hartmann. 1980. a field technique for evaluating physical condition of deer. wildlife society bulletin 8: 11–17. long, r. a., r. t. bowyer, w. p. porter, p. mathewson, k. l. monteith, and j. g. kie. 2014. behavior and nutritional condition buffer a large-bodied endotherm against direct and indirect effects of climate. ecological monographs 84: 513– 532. doi: 10.1890/13-1273.1 monteith, k. l., v. c. bleich, t. r. stephenson, b. m. pierce, m. m. conner, j. g. kie, and r. t. bowyer. 2014. life-history characteristics of mule deer: effects of nutrition in a variable environment. wildlife monographs 186: 1–62. doi: 10.1002/wmon.1011 _____, r. w. klaver, k. r. hersey, a. a. holland, t. p. thomas, and m. j. kauffman. 2015. effects of climate and plant phenology on recruitment of moose at the southern extent of their range. oecologia 178: 1137–1148. doi: 10.1007/s00442-015-3296-4 _____, t. r. stephenson, v. c. bleich, m. m. conner, b. m. pierce, and r. t. bowyer. 2013. risk-sensitive allocation in seasonal dynamics of fat and protein reserves in a long-lived mammal. journal of animal ecology 82: 377–388. doi: 10.1111/1365-2656.12016 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences in a declining moose population. wildlife monographs 166: 1–30. doi: 10.2193/0084-0173(2006)166[1:pnd aci]2.0.co;2 newby, j. r., and n. j. decesare. 2020. multiple nutritional currencies shape pregnancy in a large herbivore. canadian journal of zoology 98: 307–315. doi: 10.1139/cjz-2019-0241 oates, b. a., k. l. monteith, j. r. goheen, j. a. merkle, g. l. fralick, and m. j. kauffman. 2021. detecting resource limiation in a large herbivore population is enhanced with measures of nutritional condition. frontiers in ecology and evolution 8: 522174. doi: 10.3389/ fevo.2020.522174 parker, k. l., p. s. barboza, and m. p. gillingham. 2009. nutrition integrates environmental responses of ungulates. functional ecology 23: 57–69. doi: 10.1111/j.1365-2435.2009.01528.x proffitt, k. m., a. b. courtemanch, s. r. dewey, b. lowrey, d. e. mccwhirter, k. l. monteith, j. r. paterson, j. rotella, p. j. white, and r. a. garrott. 2021. regional variability in pregnancy and survival rates of rocky mountain bighorn sheep. ecosphere 12: e03410. doi: 10.1002/ecs2.3410 riney, t. 1955. evaluating condition of free-ranging red deer (cervus elaphus) with special reference to new zealand. new zealand journal of science and technology, b. general research 36: 429–463. moose body condition scoring – levine et al. alces vol. 58, 2022 98 ruprecht, j. s., k. r. hersey, k.hafen, k. l. monteith, n. j. decesaree, m. j.kauffman, and d. r. macnultya. 2016. reproduction in moose at their southern range limit. journal of mammology 97: 1355–1365. doi: 10.1093/jmammal/gyw099 schrempp, t. v., j. l. rachlow, t. r. johnson, l. a. shipley, r. a. long, j. l. aycrigg, and m. a. hurley. 2019. linking forest management to moose population trends: the role of the nutritional landscape. plos one 14: e0219128. doi: 10.1371/ journal.pone.0219128 stephenson, t. r., v. c. bleich, b. m. pierce, and g. p. mulcahy. 2002. validation of mule deer body composition using in vivo and post-mortem indices of nutritional condition. wildlife society bulletin 30: 557–564. _____, g. w. german, e. f. cassirer, d. p. walsh, m. e. blum, m. cox, k. m. stewart, and k. l. monteith. 2020. linking population performance to nutritional condition in an alpine ungulate. journal of mammology 101: 1244–1256. doi: 10.1093/jmammal/gyaa091 _____, k. j. hundertmark, c. c. schwartz, and v. van ballenberghe. 1998. predicting body fat and body mass in moose with ultrasonography. canadian journal of zoology 76: 717–722. doi: 10.1139/z97-248 swift, p. k., v. c. bleich, t. r. stephenson, a. e. adams, b. j. gonzales, b. m. pierce, and p. j. marshal. 2002. tooth extraction from live-captured mule deer in the absence of chemical immobilization. wildlife society bulletin 30: 253–255. testa, j. w., and g. p. adams. 1998. body condition and adjustments to reproductive effort in female moose (alces alces). journal of mammalogy 79: 1345–1354. doi: 10.2307/1383026 vartanian, j. m. 2011. habitat condition and the nutritional quality of seasonal forage and diets: demographic implications for a declining moose population in northwest wyoming, usa. m.s. thesis, university of wyoming, laramie, wyoming, usa. alces vol. 58, 2022 moose body condition scoring – levine et al. 99 appendix a. shiras moose body condition score score sacro-sciatic ligament base of tail1 caudal vertebrae2 sacrum 7 ligament covered in fat indiscernible nearly indiscernible, w/ much fat not discernible 6 ligament virtually indiscernible nearly indiscernible barely discernible, w/ much fat not readily discernible 5 ligament discernible, fat evident barely discernible barely discernible, w/ fat barely discernible 4.25 ligament discernible, some fat discernible, fat evident discernible, w/ fat barely discernible 3.75 ligament discernible vertebrae rounded discernible, but fleshed w/ some fat rounded, barely discernible 3.25 can pinch 0.5” w/o flesh covering vertebrae discernible, rounded individually discernible, but fleshed discernible ¼ way to tail 2.75 can pinch 1.0” w/o flesh covering vertebrae discernible, rounded individually discernible rounded, discernible 2.5 can pinch 1.25” w/o flesh covering vertebrae clearly discernible skeletal, but rounded rounded, discernible 2.25 can pinch 1.5” w/o flesh covering vertebrae clearly discernible skeletal, but rounded rounded, prominent 2 can pinch 1.75” w/o flesh covering vertebrae prominent and concave skeletal, w/ gaps rounded, prominent 1.75 can pinch 2.0” w/o flesh covering vertebrae prominent and concave skeletal, w/ gaps rounded, prominent 1.5 can pinch 2.25” w/o flesh covering vertebrae prominent and concave skeletal, w/ gaps skeletal, ≥ 0.5” protrusion 1.25 can pinch 2.5” w/o flesh covering vertebrae sharp and concave skeletal, sharp w/ gaps skeletal, ≥ 0.5” protrusion 1 can pinch ≥ 2.75” w/o flesh covering vertebrae sharp and concave skeletal, sharp w/ gaps skeletal, ≥ 1” protrusion last modified by k. l. monteith in 2022. note: bcs 2.75–3.0 = subcutaneous fat depletion point. we emphasize the importance of proper and repeated training to establish competency in assessing nutritional condition (cook et al. 2021a). 1 caudal vertebrae 2–3. 2 caudal vertebrae 6–7. appendix alces36_269.pdf 4016.p65 alces vol. 40, 2004 baskin et al. moose escape behaviour 123 moose escape behaviour in areas of high hunting pressure leonid baskin1, john p. ball2,3, and kjell danell2 1institute of ecology and evolution, leninsky pr. 33, moscow, 119071, russia; 2department of animal ecology, swedish university of agricultural sciences, umeå, se-901 83, sweden; 3e-mail: john.ball@szooek.slu.se abstract: although hunters cause more than 80% of moose mortality in some geographic areas, quantitative studies of how moose attempt to escape humans are surprisingly rare. we experimentally disturbed radio-collared moose of known age and of both sexes to study escape behaviour from humans. we found that larger groups of moose made fewer stops between being disturbed and settling down, and that larger groups exhibited a longer path length before quieting. we detected no significant effect of age (a potential measure of survival rate) on escape behaviour. the escape path of males was significantly longer than females even though the linear distance from the site of disturbance to the location where the moose settled down was not significantly different between the sexes. overall, the escape path of males from the site of disturbance to where they settled down was significantly more tortuous than that of females. although males are the preferred prey of hunters, the differences in escape behaviour between the sexes also may contribute to why males are more frequently killed by hunters. thus, in areas with heavy hunting pressure, hunters may be acting as a selective force that favours animals that immediately run away after disturbance by humans. finally, published evaluations of the use of hunter observations to index moose populations have often reported that considering the size of a hunting group is necessary to improve the accuracy of those data; our analysis suggests an explanation – differences in escape behaviour between the sexes. alces vol. 40: 123-131 (2004) key words: escape behaviour, gender differences, group size, hunters, hunting, moose, movements, predator, reaction distance, selective force, sweden predation has long been considered one of the most important selective pressures on animals in the wild (treves 2000). it is thus not surprising that ecologists have long studied many different aspects of predator and prey behaviour. for example, the observation that the vigilance of individual prey declines as their group size increases is one of the most frequently reported relationships in the study of animal behaviour (roberts 1996). among prey species in general, various aspects of escape or anti-predator behaviour have been found to vary with a great number of factors including group size (hebblewhite and pletscher 2002), distance to cover (white and berger 2001), temperature (fernandez-juricic et al. 2002), degree of predation risk (hamilton and heithaus 2001), experience with predators (jachner 2001), and the sex of the prey (magurran and nowak 1991). not surprisingly, evolution often has led to prey that are sensitive to the costs and benefits of different anti-predator behaviours under various circumstances (frid and dill 2002). studies of moose (alces alces) also have revealed a variety of anti-predator behaviours, that include standing its ground against wolves (canis lupus, mech 1970), increasing group moose escape behaviour – baskin et al. alces vol. 40, 2004 124 size when foraging further from cover (molvar and bowyer 1994), selecting unpredictable sites to give birth (bowyer et al. 1999), grouping together if predators are present, and increasing vigilance when with active young or when further from protective refugia (white and berger 2001, white et al. 2001). franzmann and schwartz (1998) provide a further overview of antipredator (and many other) behaviours in north american moose. however, few published studies have quantified the spatial aspects of escape behaviour of moose when confronted by potential wild predators or humans (but see glushkov 1976 and andersen et al. 1996). hunting is the most important mortality factor of moose in fennoscandia (e.g., 8191% of adult mortality; ericsson and wallin 2001) because large predators (brown bears, urus arctos, and wolves, canis lupus) are absent over large areas or occur only at very low densities (swenson et al. 1994, persson and sand 1998, ericsson et al. 2001). anecdotal reports by hunters and others suggest that there may be individual variation in escape behaviour – some moose use different escape tactics than others. furthermore, at least in russia, moose have been reported to exhibit different escape behaviour in areas with low and high hunting pressure (glushkov 1976). more knowledge on this issue is needed because humans might be acting as a strong selective agent in those areas that have high hunting pressure, and thereby may alter moose behaviour in the long term. perhaps most importantly, we need to understand escape behaviour if we are to understand and manage the interaction between the human predator and its prey. here, we report what is apparently the first quantitative investigation to consider: (1) does escape behaviour from humans differ between the sexes in moose; (2) is there any correlation between age and the way a given moose attempts to escape; and (3) does group size affect escape behaviour in moose? methods the study area (~4,000 km2) is just north of umeå (63°48’ n, 20°17’e), in coastal northern sweden (fig. 1). norway spruce (picea abies), scots pine (pinus sylvestris), and birches (betula pendula and b. pubescens) are the dominant tree species, and the field layer vegetation is dominated by bilberry (vaccinium myrtillis), lingonberry (v. vitis-idea), and heather (calluna vulgaris). the length of the growing season is about 150 days, starting around mid-may; snow normally arrives in early november and persists to the end of april (sna 1995). moose density in the study area (as determined by helicopter census and pellet group counts) ranged between 0.7 and 0.9 moose km-2 (j.p. ball unpublished data). predators (other than humans) capable of taking moose essenfig. 1. map of the study area in northern sweden where moose were experimentally disturbed to quantitatively investigate the effects of age, sex, and group size on escape behaviour. umeå is at the southern edge of the study area. alces vol. 40, 2004 baskin et al. moose escape behaviour 125 tially are absent (swenson et al. 1994, persson and sand 1998). moose are hunted heavily in the area and every year about one-third of the population is harvested. more importantly, in this area, hunters account for 81% of all deaths of female moose born and 91% of all male deaths (ball et al. 1999, ericsson and wallin 2001). using a dart rifle fired from a helicopter, we immobilized adult and calf moose with a mixture of ethorphine and xylazine hydrochloride (sandegren et al. 1987) during february to mid-march every year from 1990. each moose was fitted with a numbered radio collar and small numbered ear tags. moose were aged using two complementary methods: at marking, we determined their ages using tooth wear and eruption (skuncke 1949), and if later harvested, we obtained the jaw, sectioned the first molar, and counted the cementum annuli under at least 20x magnification (sergeant and pimlott 1959, fancy 1980, bubenik 1998). when good conditions for snow tracking occurred between january and april during the 1996-97 and 1997-98 winters, we conducted controlled disturbance experiments with these known-aged radio-collared moose. for all tests, a single observer (for consistency) first located the moose by radio-telemetry and thereafter the moose (one individual or group of moose) was disturbed in a uniform way (i.e., a slow quiet approach on skis at the same speed from an initial distance of 70-100 m). after the disturbance, the observer followed the movements of the radio-collared moose by radio signals until it settled down (i.e., started to feed, ruminate, lie down, etc.). then the observer tracked the moose backward in the snow to quantify the escape path taken (and avoid influencing the escape behaviour other than the initial controlled disturbance). we quantified the group size, and 5 response variables which quantified the spatial aspects of the moose’s escape behaviour (for a group of moose, the centroid of the group was used). these were: (1) the distance the moose ran, trotted, or galloped after the initial disturbance (“gallop distance” hereafter); (2) the total distance the moose moved before settling down (“path length to quieting”); (3) the straight line distance from the place where the moose was disturbed to the point where the moose quieted down (“straight length to quieting”); (4) the number of stops the moose made before quieting (“number of stops before quieting”); and (5) we calculated a simple index of tortuosity (“tortuosity”) by dividing the “straight length to quieting” by the “path length to quieting”. thus, an animal with a perfectly linear escape path would have a tortuosity index of 1.0, whereas a moose which travelled in a nonlinear fashion after being disturbed would have an index with a much lower value (e.g., 0.3). see mårell et al. (2002) for additional approaches to quantifying movement. indicator variable regression (also known as “dummy variable regression”; kleinbaum et al. 1987, tabachnick and fidel 2001) was used to test for the effects of the independent variables (sex, group size, and age) on the dependent variables that described escape behaviour. indicator variable regression is a form of a general linear model that is appropriate to a mixture of categorical (e.g., sex) and continuous (e.g., tortuosity) variables (kleinbaum et al. 1987, tabachnick and fidel 2001). the significance factor reported is the additional explanatory value of adding a variable to a model already containing the other two independent variables (i.e., this statistically controls for any correlations among these independent variables; cohen and cohen 1983, kleinbaum et al. 1987, tabachnick and fidel 2001). in keeping with standard practices, if and only if the anova for the moose escape behaviour – baskin et al. alces vol. 40, 2004 126 overall indicator variable regression (i.e., containing all independent variables) was significant, did we then examine which individual effects (e.g., sex, group size, or age) accounted for this. statistical analyses were performed with jmp version 4.0.5 (sas institute 2000). we report means ± 1 standard deviation unless otherwise noted. results in total, 29 controlled disturbances were performed (13 males and 16 females, including one cow of unknown age). the average age of the males we studied was 7.5 ± 2.2 years, and that of females 10.9 ± 4.5 years. the heavy hunting pressure in the study area is perhaps reflected by the rather strong escape responses made by moose to our experimental disturbances. the gallop distance was 293 ± 222 m, the path length to quieting was 1,324 ± 699 m, and the straight length to quieting was 855 ± 419m. indicator variable regression revealed that the distance that moose ran after being disturbed (gallop distance) showed no relationship with the independent variables sex, age, or group size (whole model r2 = 0.08, p = 0.59), so we did not examine separate effects. in contrast, the overall model with the dependent variable path length to quieting was significant (r2 = 0.36, p = 0.01); table 1 reveals that age was not important, but sex and group size were predictors (larger groups travelled shorter distances before quieting) and females travelled less (1,022m) before quieting down than did males (1,671m). the overall model testing the explanatory effects of sex, group size, and age on straight length to quieting was not significant (whole model r2 = 0.23, p = 0.11), so again we do not consider separate effects. the overall model testing the effects of sex, age, and group size on the number of stops before quieting was significant (r2 = 0.33, p = 0.02). here the independent variable that was important was group size (table 1), with larger groups making fewer stops. finally, the model testing the effects of sex, age, and group size on tortuosity explained nearly half of the observed variation (r2 = 0.46, p = 0.002). here, however, group size and age were not predictors, but sex was a predictor (table 1). females exhibited more linear escape paths (mean tortuosity index = 0.78) than did males (mean tortuosity index = 0.57). our analysis did not indicate an effect of age on escape behaviour. in contrast, our analysis revealed that group size was correlated with the path length to quieting (larger groups settled down after travelling a shorter path). furthermore, as group size increased, the group tended to keep moving (i.e., they made fewer stops). finally, several important aspects of escape behaviour differed between the sexes. although when initially disturbed both sexes galloped the same distance (gallop distance), and stopped the same linear distance away from source df sum of squares f ratio prob > f sex 1 2030616.9 5.7058 0.0255 group size 1 1962848.5 5.5154 0.0278 age 1 139210.9 0.3912 0.5378 sex 1 18.336596 1.6672 0.2095 group size 1 55.928378 5.0851 0.034 age 1 42.879506 3.8986 0.0605 sex 1 0.40314635 19.282 0.0002 group size 1 0.00000142 0.0001 0.9935 age 1 0.04315248 2.0639 0.1643 tortuosity number of stops before quieting path length to quieting (m) table 1. anova effect tables from the indicator variable regression on moose (alces alces) escape behaviour in sweden. the results of the overall anovas that were performed first are given in the text. significant effects are highlighted in bold. alces vol. 40, 2004 baskin et al. moose escape behaviour 127 the initial disturbance point (straight length to quieting), males travelled a greater total distance (1,671 ± 227 m) than females (1,022 ± 104 m) before settling down (path length to quieting), and males did this by moving in a much more tortuous path than did females (tortuosity index for males 0.58 ± 0.06, females 0.78 ± 0.03). discussion ungulates use a variety of different strategies to escape hunters. hiding is a special behaviour exhibited by some forestdwelling ungulates like wild boar and moose under extremely high hunting pressure (baskin 1976, 1998). moose also sometimes escape by running to a safe distance and, from there, observe human activities (glushkov 1976). this type of behaviour has also been reported for musk deer moschus moschiferus (zaitsev 1983), roe deer capreolus capreolus (danilkin 1996), reindeer rangifer tarandus (baskin and hjältén 2001), and alpine ibex capra ibex (krämer and aeschbacher 1971). often, animals demonstrate apparently reckless flight when they barely have seen or heard an approaching human. this type of flight behaviour is common under high hunting pressure for reindeer (baskin and hjältén 2001) and moose (glushkov 2002). a more extreme escape behaviour is for the animal to run beyond its home range; e.g., wild camels camelus bactrianus in mongolia (przewalsky 1878), european bison bison bonasus in caucasus at the beginning of the 20th century (filatov 1912), and american bison bison bison in north america at the end of the 19th century (roe 1951). our finding that the path length to quieting was 1,671m for males and 1,022m for females is comparable to the overall mean distance of 1,147m travelled by moose fleeing skiers and pedestrians in another rather heavily hunted scandinavian moose population (andersen et al. 1996). we were unable to find any published study on a lightly hunted population for comparison. our analysis, that revealed the tendency of males to escape by a more tortuous path than females, may have implications for their likelihood of escaping hunters, but also for how moose counts should be performed, and of understanding the potential biases in hunter observations of moose. we suggest that the greater tortuosity of the escape path of males may mean that they are more likely to come into repeated contact with hunters, and thus more likely to be killed. however, compared to females, males are preferentially harvested by swedish hunters anyway (ball et al. 1999), so we are unable to separate the magnitude of the effects of these two factors. in northern sweden, the most common method of hunting is through the use of baying dogs, whereby the moose stands its ground while it is distracted by a barking dog (and is subsequently shot by a hunter sneaking in) or it flees (and hunters stationed around a particular parcel of land may be able to shoot the moose; ball et al. 1999). thus, hunting with dogs should provide a selective pressure for moose that flee, rather than hide, to escape harvest. in sweden, hunting grounds are typically private, with a group of hunters having the right to hunt only on their particular hunting grounds. as a result, moose practicing a strategy of fleeing may escape the chain of shooters and find themselves beyond harvest by a particular group of hunters. because moose hunting seasons are long (september 1 – december 31), escaping moose often will find themselves in a new area where there is no group of hunters active on the same day. compared to escaping in a linear fashion, remaining in the same area (and moving in a tortuous path) seems to be a poor strategy because of the scenting ability of the hunter’s dogs; nevertheless, our analysis reveals that this is how males moose escape behaviour – baskin et al. alces vol. 40, 2004 128 we suggest moose escape behaviour is also influenced by group cohesiveness: when any group member (which our analysis suggests would often be a female) flees from a predator, so does the entire group. we suggest that moose managers should be aware of possible interactions among group size, hunting party size, and apparent sex ratio if using hunter observations to estimate moose population characteristics. further research in this area should help clarify these relationships, and we suggest that incorporating escape behaviour of moose in a quantitative manner, as we have done, may be necessary to understand the mechanisms affecting population estimates. our analysis provided no evidence of any relationship between age and escape behaviour, but our test of this aspect is not powerful because of sample size limitations. we suggest that future studies on escape behaviours focus on the effects of group size and sex because the age of wild moose cannot be judged from a distance and few studies will have the luxury, as we did, of tranquilizing moose first to determine their age. having age data is ideal, but our analysis suggests that it is not essential to understand the escape behaviour of moose towards humans. the behavioural ecology literature suggests two main advantages for grouping in the presence of predators. first, vigilance: many eyes to detect predators (see williams et al. 2003 and bednekoff and lima 1998 for recent overviews). second, the other major advantage of being in a group comes from risk dilution: the risk to a given animal in a group of 2 is 50%, in a group of 100 animals it is 1% (hamilton 1971, treves 2000, carbone et al. 2003). in accordance with the risk dilution hypothesis, our analysis reveals that moose apparently do feel safer as group size increases, as indicated by shorter travel distances to quieting. in conclusion, the differential mortality behave (perhaps the result of past selection by native predators). we suggest that, compared to females, males either prefer to keep a pursuing human within their sensory range or males have a stronger tendency to stay within their home range. our analysis suggests that if the observations of hunters are used to estimate population parameters, the proportion of males will be overestimated relative to their true fraction in the population (see ericsson and wallin 2001, sylvén 2003 for additional analyses). furthermore, if the differences between the sexes that we detected are indeed general, these behavioural differences may help to explain why the size of a hunting group often emerges as an important factor in studies that use observations of moose seen per hour and per hunter, to estimate moose numbers (ericsson and wallin 2001, sylvén 2003). we suggest that this may be because relative to smaller groups, larger groups of hunters may overestimate males even more because the same bull may be sighted several times as it circles to remain in the area. we do not suggest that our controlled disturbance by a skier just after the hunting season precisely reflects the responses of moose to hunters during the hunt. rather, our aim was to test if age, sex, or group size in moose is associated with differences in how moose attempt to escape from humans, and to do so in an area where humans are the main mortality source (and thus likely are acting as a selective force on moose behaviour). thus, quantitative variables like the distance to quieting, for example, may well differ between the hunting season and the immediate post-hunting period. on the other hand, the qualitative relationships revealed by our analysis (e.g., the differential responses between the sexes and the effect of group size) seem likely to be less temporally variable, although further research is needed to test this proposition. alces vol. 40, 2004 baskin et al. moose escape behaviour 129 exerted by hunters may be selecting for moose which employ a strategy of immediate linear flight, or at least run out of sensory contact with a human after disturbance. we suggest that escape behaviour is an under-studied aspect of moose ecology. this is surprising since humans account for over 80% of all moose deaths in sweden (ericsson and wallin 2001), and are thus expected to be a rather strong selective force. furthermore, understanding escape behaviour may be critical in improving our ability to use hunter observations to estimate moose populations. specifically, differences between the sexes in escape behaviour may lead to an apparently greater proportion of males being detected in hunter observations. the variation in escape behaviour with group size we detected may explain why the size of a hunting party often emerges as a significant factor in previous analyses that use hunter observations (e.g., ericsson and wallin 2001, sylvén 2003). based on our analysis, we suggest that quantifying the escape behaviour of moose may lead to a better understanding of the selective force human hunters are exerting on moose populations, and may lead to improvements in using hunter observations to manage moose populations. acknowledgements we thank the swedish university of agricultural sciences (grant to kjell danell) for financial support for leonid baskin during his stay in umeå, and the swedish environment protection agency (grants to kjell danell and john p. ball) for financial support. we also gratefully acknowledge the tireless and expert work by our field technicians (particularly åke nordström, e r i c a n d e r s s o n , a n d n i l s g u n n a r andersson) in the moose project, and k. wallin for his early work in building up the marked moose population for related studies. the enthusiastic co-operation of many hundreds of moose hunters over more than a decade allowed us to exactly age our moose, and made possible this and many other investigations. we thank t.r. stephenson, r. rea, and an anonymous reviewer for comments which greatly improved the manuscript. references andersen, r., j. d. c. linell, and r. langvatn. 1996. short term behavioural and physiological response of moose alces alces to military disturbance in norway. biological conservation 77:169-176. ball, j. p., g. ericsson, and k. wallin. 1999. climate change, moose and their human predators. ecological bulletins 47:178-187. baskin, l. m. 1976. behaviour of ungulates. nauka publishers, nauka, moscow, russia. _____. 1998. moose conservation in ecosystems of eastern europe. alces 34:395-407. _____, and j. hjälten. 2001. fright and flight behaviour of reindeer. alces 37:435-446. bednekoff, p. a., and s. l. lima. 1998. re-examining safety in numbers: interactions between risk dilution and collective detection depend upon predator targeting behaviour. proceedings of the royal society of london series bbiological sciences 265:2021-2026. bowyer r. t., v. van ballenberghe, j. g. kie, and j. a. k. maier. 1999. birthsite selection by alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80:1070-1083. bubenik, a. b. 1998. evolution, taxonomy and morphophysiology. pages 77-124 in a. w. franzmann and c.c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . moose escape behaviour – baskin et al. alces vol. 40, 2004 130 smithsonian institution press, washington, d.c., usa. carbone c., w. a. thompson, i. zadorina, and j. m. rowcliffe. 2003. competition, predation risk and patterns of flock expansion in barnacle geese (branta leucopsis). journal of zoology 259:301308. cohen, j., and p. cohen. 1983. applied multiple regression correlation analys i s f o r t h e b e h a v i o r a l s c i e n c e s . erlbaum, new york, new york, usa. danilkin, a. a. 1996. behavioral ecology of siberian and european roe deer. chapman and hall, london, u.k. ericsson, g., and k. wallin. 2001. agespecific moose (alces alces) mortality in a predator-free environment: evidence for senescence in females. ecoscience 8:157-163. _____, k. wallin, j. p. ball, and m. broberg. 2001. age-related reproductive effort and senescence in free-ranging moose alces alces. ecology 82:1613-1620. fancy, s. g. 1980. preparation of mammalian teeth for age determination by cementum layers: a review. wildlife society bulletin 8:242-248. fernandez-juricic, e., m. d. jimenez, and e. lucas. 2002. factors affecting intraand inter-specific variations in the difference between alert distances and flight distances for birds in forested habitats. canadian journal of zoology 80:1212-1220. filatov, d. 1912. caucasian bison. zapiski imperatorskoi akademii nauk, otdelenie phisiko-matematicheskoe, series 7, 30:140. (in russian). franzmann, a. w., and c. c. schwartz, editors. 1998. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. frid, a., and l. dill. 2002. humancaused disturbance stimuli as a form of predation risk. conservation ecology 6:11-16. glushkov, v. m. 1976. some peculiarities of moose behavior that determine hunter’s success. in: s.a. korytin, editor. behavior of hunting animals. s b o r n i k n a u c h n o t e k h n i c h e s k o y informatsii 51-52:51-58. (in russian). _____. 2002. ecological basis of moose population management in russia. thesis. vsesoyuznyi naychnoissledovatelsky institut okhotnichego khozyaistva i zverovodstva, kirov, russia. (in russian). hamilton, i. m., and m. r. heithaus. 2001. the effects of temporal variation in predation risk on anti-predator behaviour: an empirical test using marine snails. proceedings of the royal society of london series b-biological sciences 268:2585-2588. hamilton, w. d. 1971. geometry of the selfish herd. journal of theoretical biology 31:295-311. hebblewhite, m., and d. pletscher. 2002. effects of elk group size on predation by wolves. canadian journal of zoology 80:800-809. jachner, a. 2001. anti-predator behaviour of naive compared with experienced juvenile roach. journal of fish biology 59:1313-1322. kleinbaum, d. g., l. l. kupper, and k. e. muller. 1987. applied regression analysis and other multivariate methods. duxbury press, belmont, california, usa. krämer, a., and a. aeschbacher. 1971. zum slachterhalten des steinwildes (capra ibex) im oberengadin, schweiz. säugetierk mitteilungen 19:164-171. magurran, a. e., and m. a. nowak. 1991. another battle of the sexes the consequences of sexual asymmetry in mating costs and predation risk in the guppy, alces vol. 40, 2004 baskin et al. moose escape behaviour 131 poecilia reticulata. proceedings of the royal society of london series bbiological sciences 246:31-38. mårell, a., j. p. ball, and a. hofgaard. 2002. foraging and movement paths of female reindeer: insights from fractal analysis, correlated random walks and lévy flights. canadian journal of zoology 80:854-865. mech, l. d. 1970. the wolf: the ecology and behavior of an endangered species. natural history press, garden city, new york, usa. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate the alaskan moose. journal of mammalogy 75:621630. persson, j., and h. sand. 1998. vargen — viltet, ekologin och människan. svenska jägareforbundet, spånga, sweden. (in swedish). przewalsky, n. m. 1878. wild camel. priroda i okhota, 1-2. roberts, g. 1996. why individual vigilance behaviour declines as group size increases. animal behaviour 51:10771086. roe, f. g. 1951. the north american buffalo. a critical study of the species in its wild state. university of toronto press, toronto, ontario, canada. sandegren, f., l. pettersson, p. ahlqvist, and b. o. röken, 1987. immobilization of moose in sweden. swedish wildlife research supplement 1:785-791. sas institute. 2000. jmp statistics and graphics guide. sas institute, cary, north carolina, usa. sergeant, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23:315-321. skuncke, f. 1949. älgen, studier, jakt och vård. p. a. nordstedts och söners förlag, stockholm, sweden. (in swedish). (sna) sveriges national atlas. 1995. sveriges national atlas klimat, sjöar och vattendrag. bra böcker, höganäs, sweden. swenson, j. e., f. sandegren, a. bjärvall, a. söderberg, p. wabakken, and r. franzén. 1994. size, trend, distribution and conservation of the brown bear ursus arctos population in sweden. biological conservation 70:9-17. sylvén, s. 2003. management and regulated harvest of moose (alces alces) in sweden. agraria 371. doctoral thesis, department of conservation biology, swedish university of agricultural sciences, uppsala, sweden. tabachnick, b. g., and l. s. fidell. 2001. using multivariate statistics. fourth edition. allyn and bacon, needham heights, massachusetts, usa. treves, a. 2000. theory and method in studies of vigilance and aggregation. animal behaviour 60:711-722. wh i t e , k. s., and j. be r g e r. 2001. antipredator strategies of alaskan moose: are maternal trade-offs influenced by offspring activity? canadian journal of zoology 79:2055-2062. _____, j.w. testa, and j. berger. 2001. behavioral and ecologic effects of differential predation pressure on moose in alaska. journal of mammalogy 82:422-429. williams, c. k., r. s. lutz, and r. d. applegate. 2003. optimal group size and northern bobwhite coveys. animal behaviour 66:377-387. zaitsev, v. a. 1983. escape behaviour of musk deer (moschus moschiferus parviceps) in middle sikhote-alin. zoologichesky zhurnal 62:1718-1726. p9-24_4114.pdf alces vol. 41, 2005 lykke management of norwegian moose 9 selective harvest management of a norwegian moose population jon lykke værdalsbruket, 7660 vuku, norway abstract: the moose population at værdalsbruket in the county of nord-trøndelag, norwaythe moose population at værdalsbruket in the county of nord-trøndelag, norway has been studied since the 1930s. complete harvest and weight statistics for sex and age classes and detailed hunter observations have been collected since 1969 producing a data set of 2,667 harvested moose and 17,068 moose observations. these data were used to both manage and assess a selective harvest management system based upon annual hunter guidelines, contracts with sex-age quotas, and progressive pricing of hunting cost related to carcass weight. combined with a relatively high hunting pressure, the system has produced a controlled increase in the moose population, and an improved population structure with more prime bulls, higher mean age of cows, and an improved cow:bull ratio. long-term body weights and production have been stable, indicating a healthy moose population in balance with its resources. success of the harvest system depended largely on the level and progression of the hunting price-carcass weight relationship. alces vol. 41: 9-24 (2005) key words: alces alces, body weight, harvest, management, moose, norway, population dynamics, reproduction the history and development of the norwegian moose (alces alces) population is well documented and has shed light on interesting aspects of moose ecology. moose gradually occupied the country after the last ice age some 10,000 years ago, and were the hunters. as discussed by lykke (1960) and lykke and cowan (1968), pronounced of moose have occurred since. moose hides were exported from norway to britain as early as the 1100s, and moose were mentioned in older district laws. hunting regulations were enforced in the 15th century, and severe restrictions against killing moose in the 18th th century suggest scarcity at that time. moose were nearly extinct about 1800 and were rare outside south-central parts of norway and sweden. moose harvest statistics for norway (fig. 1) exist since 1889 and it population trends. moose population dynamics in norway have been addressed by many authors (e.g., skuncke 1949, lykke and cowan 1968, lykke 1974b, haagenrud 1986). by 1900 large predators were almost exterminated, moose hunting was better controlled, and forest pasturing of sheep and cattle declined, were probably due to variable hunting pressure. the rapid population increase from 1935-1960 was triggered by lower hunting pressure when forestry practices changed, and possibly by climatic factors. the dramatic population increase in the 1950s was due to increased browse production from and fairly low hunting pressure. around 1960 there were concerns about overpopulation, and moose were reduced through increased hunting pressure, even upon the most productive segments. forage producmanagement of norwegian moose lykke alces vol. 41, 2005 10 tion continued to increase at a high rate in response to logging for the rest of the century, factors of moose population dynamics. this paper addresses the moose population at værdalsbruket, which is one of the largest private properties in norway consisting of 900 km2 of forests, mountains, marshland, rivers, and lakes. the area lies in nord-trøndelag county in central norway and covers 60% of the municipality of verdal. most of the area is found in the upper, eastern part of verdal near the swedish border (fig. 2). my father, leif lykke, was managing director at værdalsbruket from 1931-1971 and i had the same position in 1971-2002. together, we were responsible for all forestry and wildlife management for > 70 continuous years, and generated a historical database of moose statistics, especially since 1969. analysis of these data is pertinent given the recent concerns about body size and reproduction, skewed population structure, overbrowsing and forest damage, and effects on biodiversity by moose in norway and throughout scandinavia (lavsund 1987, thompson 1990, angelstam et al. 2000, connor et al. 2000). the major objective of this study was to document the effect of a selective harvest system on the population structure, reproduction, body weight, and growth of the moose population at værdalsbruket, norway. study area the area (latitude 64° n) is a varied landscape of high and low productive forests, with short distances from sea to farmland to forests to mountains. there is a mixture of mountains and forested valleys with numerous rivers, streams, and waterfalls. altitude varies from 20-1100 m a.s.l., and 550 m near sweden. the area has a mix of coastal and inland climates. average july temperature is 12-15 °c, with maximum diurnal temperature of 30°c. average january temperature is –4 to –7°c, with a minimum of –35°c. average yearly precipitation is 800-1100 mm. snow depth is moderate in the lower western areas, while some eastern valleys may have > 2 m and are snow-covered 200 days a year. 0 5000 10000 15000 20000 25000 30000 35000 40000 45000 18 90 19 00 19 10 19 20 19 30 19 40 19 50 19 60 19 70 19 80 19 90 20 00 n u m b e r fig. 1. moose harvest in norway, 1889-2003 (bureau of statistics). alces vol. 41, 2005 lykke management of norwegian moose 11 the habitat was described previously by several authors (lykke 1974b, ahlen 1975, krefting and lykke 1976). the vegetation varies widely from highly productive farmland to alpine. there is a mixture of high and low productive forestland, bogs, tion of værdalsbruket shows 23% productive forestland with 3% highly productive, 12% other forestland, 10% marshland below treeline, 3% water, and 52% mountains. three main tree species occur with norway spruce (picea abies) dominant and representing 79% of the cubic mass; scots pine (pinus silvestris) is 12%, and birch (betula spp.) 9%. preferred moose browse species are sallow and willows (salix spp.), mountain ash (sorbus aucuparia), juniper (juniperus communis), aspen (populus tremula), scots pine, and birch. to some extent norway spruce, grey alder (alnus incana), and bird cherry (prunus padus) are browsed. the summer diet consists of a wide variety of plant species, and of special interest is browsing of blueberry (vaccinium myrtillus) in late fall. land use and management history forestry has been the main activity on the property since the 1600s, although portions were also used for grazing by laplanders (reindeer) and farmers (sheep and cattle). silvicultural practices have changed over time; before 1930 a selective cutting system was employed that was eventually replaced with moderate sized clearcuts. this change was of considerable importance to the moose population and was a primary factor in its moose included creation of a mixture of young and old forest stands, leaving forest vegetation as edge for marshland, water, and mountains, allowing broadleaved plants to grow with conifers in young stands, protection of marshland, and avoidance of herbicides. currently 40% of the productive forestland of værdalsbruket consists of young forests, most producing high volumes of moose browse. summer and winter areas, rutting locations, and calving grounds (lorentsen et al. 1991). home range size varies among animals, as does the distance between summer and winter ranges. the largest part of the population moves short distances from summer to winter range, relocating between higher and lower elevation (baskin 1987). the rest of the population is migratory with a traditional shift of home range along municipalities in norway and sweden. the moose population uses a larger area during summer than winter, notably in higher mountainous habitats. predators have minimal importance in the ecology of moose in verdal (lykke and cowan 1968). a few brown bears (ursus nnnno fig. 2. location of study area værdalsbruket, norway. norway sweden nord-trøndelag værdalsbruket management of norwegian moose lykke alces vol. 41, 2005 12 arctos) take the odd moose, but no wolf (canis lupus) predation exists. however, the bear population is slowly increasing, and a few wolf packs have established ranges in southeastern norway where they interact with the regional moose population (sand et al. 2004); both predators are protected by law. non-hunting losses are 10-15% of the legal harvest and vary annually due to snow depth, vehicular collisions, ice conditions (lykke 1952). under these circumstances, the harvest data presumably indicate reasonable trends in the moose population. the norwegian moose hunting system was described by l. lykke (1960), and j. lykke (1974a); essentially hunting rights belong to the landowner and 80% of the forest area is privately owned. since 1952, norwegian authorities have controlled harvest by area and the minimum area required to harvest moose relative to population density. in verdal it varies from 2-6 km2 per moose from lower to higher elevation. we divided værdalsbruket into approximately 20 hunting sections, each for exclusive use by a hunting team (usually 3-6 hunters) provided with a set quota of 2-6 moose. open season is set by the authorities, and has varied somewhat, although the majority of hunting and harvest occurs in late september and october. the season is closed during the main rutting period (2-9 october) and has recently been extended into november. a selective harvest system is important in moose management (rausch et al. 1974, mercer and strapp 1978) and was introduced by landowners and central and local authorities in norway. harvest composition was essentially determined by local authorities since the 1970s until recently, when more responsibility was granted to landowners (e.g., værdalsbruket) who develop 3-5 year management plans that are approved by local game boards. the objective of the selective harvest system is to protect the highest productive segment of the moose population. calves were protected in norway until 1963 when it became obvious they should be harvested to help manage the rapidly increasing population. historical protection of calves inhibited many hunters from shooting them; a situation that lasted into the 1980s and still prevails in some districts. an adaptable, selective harvest system has been used at værdalsbruket since 1945, but has played an important role in moose management only since the 1970s. important parts of the værdalsbruket moose management plan are protection of mature cows (maximum 15% of total harvest) and prime bulls, high harvest of calves and yearlings (minimum 65% combined of total harvest), maintenance of a balanced sex ratio, and with calf at heel have always been protected and sex composition of the adult kill has been predetermined. in 1960 we introduced a pricing system to protect older cows and bulls by making it more economically favourable to shoot young/small moose. the system was modipopulation goals by charging per kilogram of meat according to a progressive pricing formula. the formula in 2004 was: nok/kg = carcass weight (kg)/9 + 33 (5 for calves). for example, this formula produces a near doubling in price from 44 nok/kg (6.5 usd) for a 65 kg calf to 83 nok/kg (12.2 usd) for a 300 kg adult; both prices include 25% governmental tax. this approach was well received by hunters, and the formula is adapted to alter the progression to achieve methods the moose population estimates were based upon harvest statistics since the late alces vol. 41, 2005 lykke management of norwegian moose 13 1880s, yearly notes of moose abundance, non-hunting losses, poaching records, and hunter observations the last 3-4 decades (l. lykke 1962, 1968; j. lykke 1964). population characteristics were analyzed from annual harvest data and hunter observations from værdalsbruket and a small adjacent area. complete harvest data existed for the period 1945-2004 for 3,423 moose, with 92% of the kill after 1960. since 1969 each harvested moose (n = 2,667) was sexed, aged, and weighed (table 1). age was primarily judged by wear of incisors up to 6-7 years (heptner and nasimowitsch 1967). age was also determined from sectioned teeth in 1969-1972. the results were similar except for a few cases of misjudged 2-3 year olds. the carcasses were split into 8 pieces for weighing: 2 front legs, 2 hind legs, 2 sides, and the back and neck. since 1969 each hunting team provided bulls, cows without calves, cows with one calf, cows with two calves, and unknown. in total, 17,068 moose were observed on 26,010 man-days. it was assumed that the use of stable hunting teams in the same area each year provided reliable data (solberg and sæther 1999). such observations allowed comparison of relative moose density, sex and age composition of the herd, number of barren cows (schwartz 1998), reproduction, hunting pressure on various groups, and hunting success over time (baskin and lebedeva 1987, gaidar et al. 1990). hunting observed moose in each sex and age group that was harvested (harvest rate). results harvest statistics current harvest (2004) is about 10-fold higher than that in the late 1940s (fig. 4). since the minimum area requirement was introduced in 1952, one measure of hunter success is the percentage of predetermined moose harvested. the normal rate of success is 90% in værdalsbruket (fig. 4). the amount of harvested meat increased since 1970; likewise, the proportion of calf and yearling meat increased to > 50% of the total since 1990 (fig. 5). the peak in 1963 was a result of very high hunting pressure 0 20 40 60 80 100 50 100 150 200 250 300 kg. carcass w eight k r. /k g . 2003 1988 1977 1972 fig. 3. the temporal change in the relationship of price and carcass weight of harvested moose at værdalsbruket, norway. calves 1.5-year-olds 2.5-year-olds 3.5 years old total n % n % n % n % n % males 511 19 461 17 399 15 280 11 1,651 62 females 415 16 250 9 162 6 189 7 1,016 38 total 926 35 711 27 561 21 469 18 2,667 100 table 1. sex and age composition of the moose harvest, værdalsbruket, norway, 1969-2004. management of norwegian moose lykke alces vol. 41, 2005 14 (figs. 4 and 5). the overall sex composition of the harvest was dominated by males before 1955, declined in the early 1960s, and was relatively high in 1970-1985 (fig. 6). prior to 1980, only the odd (and large) calf was harvested, but the percentage of calves in the harvest increased to about 40% by 1990 (fig. 7). combined harvest of calves and yearlings rose to about 70%, and the percentage of harvested adult cows declined to about 12% (fig. 7). the harvest of mature moose declined overall since 1969. before 1980, about 40% of harvested adult moose were > 3 years old; current harvest is < 20%. excluding calves, 63% of the harvest was male since 1945, and 66% since 1969; after 1982, 54.5% of harvested calves were male. the average weight of all age and sex classes changed little over time; average weights of calf, yearling, and 2.5-year-olds were 61.3, 125.7, and 167.8 kg, respectively (fig. 8). average weights of male and female calves and yearlings were 63.2 and 58.6, and 128.5 and 120.7 kg, respectively. population characteristics hunter observations in 1969-2004 were used to estimate the annual sex and age composition of the moose population in late september and october (fig. 9). cows represented 40-45% of the population prior to 1985, and about 50% afterward. the percentage of bulls was 25-30% prior to 1985, declining to 20% afterward. calves were about 30% of the population during 1969-2004, ranging from 22.1-33.7%. before 1985, the sex ratio (cows:bulls) of observed adult moose averaged 1.50 or 60% cows, varying from 1.05-1.84 annually. since 1985, the ratio averaged 2.40 or 71% cows, but has dropped below 2.0 in recent years. reproduction was estimated from hunter observations since 1969; 35-40% of all cows had one calf at heel, and 10-15% had twins, with little variation since 1985. about 4550% of all cows were barren annually; a downward trend existed since 1994. the number of calves per calf-producing cow was relatively stable, about 1.20, as was the number of calves per total number of cows 0 20 40 60 80 100 120 140 1945 1950 1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 n u m b e r 0 10 20 30 40 50 60 70 80 90 100 p e rc e n t no of harvested moose harvest success,% fig. 4. annual moose harvest at værdalsbruket, norway, 1945-2004. alces vol. 41, 2005 lykke management of norwegian moose 15 < 50 kg carcass weight, represented 10-15% in 1985-1994 it was 15.2%, dropping to 10.5% in 1995-2004. the observed number of moose per man-day was a useful indicator of relative changes in moose density. an average of 0.6 moose (0.5-0.7) was observed per man-day before 1980 and about 0.7 moose (0.5-0.9) afterward (fig. 11). it took approximately 12 man-days to kill a moose (fig. 11) and this changed little despite the increasing moose harvest system produced change in harvest rate of various sex and age groups. since 1969, the harvest rate for bulls declined from 30-35 to 20-25%, and from about 10 to 5% for adult cows (fig.12). the harvest rate of calves increased from about 2 to 15%. the harvest rate for all moose fell from about 15 to 10%. discussion population development because there is almost no predation of moose in verdal, and non-hunting loss is only 10-15% of the legal harvest (lykke 1952, haagenrud et al. 1975), the harvest statistics (fig. 4) provide a reasonable estimate of the moose population. population growth is best by forest harvesting, hunting pressure, and harvest composition. multiple factors led to increased food production, but most important was the use of clearcuts introduced in verdal in the 1930s and increasingly used after world war ii (lykke 1974b). a high percentage of bulls in the harvest is indicative of low hunting pressure (cumming 1974, haagenrud and lørdahl 1979, solberg et al. 2001). the combined effect of low hunting pressure (fig. 6) and increased food production in the 1950s resulted in rapid population growth (fig. 4). there was an eventual concern about overpopulation because even norway spruce, a non-preferred forage, was heavily browsed (lykke 1964). a decision was made to reduce the moose population and increase hunting pressure of cows (fig. 6) and all large moose (fig. 5), especially in 1960-1966 (fig. 4). in the period 1967-1985, the population increased again, although more slowly because of previous overbrowsing, harvest composition in the early 1960s, and relatively high hunting pressure on adult moose. 0 2000 4000 6000 8000 10000 12000 14000 1960 1965 1970 1975 1980 1985 1990 1995 2000 m o o se m e a t h a rv e st , kg . 0 10 20 30 40 50 60 70 80 90 100 p e rc e n t total % f rom calves and yearlings fig. 5. kilograms of moose meat harvested at værdalsbruket, norway, 1960-2004. management of norwegian moose lykke alces vol. 41, 2005 16 35 40 45 50 55 60 65 70 75 80 85 90 95 100 1945 1950 1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 p e rc e n t m a le s fig. 6. male fraction of moose harvested at værdalsbruket, norway, 1945-2004. 0 10 20 30 40 50 60 70 80 1970 1975 1980 1985 1990 1995 2000 p e rc e n t % of calves % of calves + yearlings % of cow s >= 2 1/2 years fig. 7. age composition of moose harvested at værdalsbruket, norway, 1969-2004. cows alces vol. 41, 2005 lykke management of norwegian moose 17 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 1970 1975 1980 1985 1990 1995 2000 k g . calves yearlings 2 1/2 years old fig. 8. mean carcass weights of calves, yearlings, and two and a half year old moose harvested at værdalsbruket, norway, 1969-2004. 10 20 30 40 50 1970 1975 1980 1985 1990 1995 2000 p e rc e n t calves cow s bullscows fig. 9. sex and age composition of moose observed by hunters during the hunting season at værdalsbruket, norway, 1969-2004. management of norwegian moose lykke alces vol. 41, 2005 18 hunting pressure was gradually transferred to younger and smaller moose to reduce the harvest of mature moose (fig. 12). this was accomplished by “education” of hunters, stipulations in their contracts, and a progressive pricing system. the hunter effort required to kill a moose, approximately 12 man-days per moose similar to that measured in russia (gaidar et al.1990), was stable despite the higher population since 1970 (fig. ers to harvest the correct moose. although population management has been realized through the selective harvest system. since 1985 a stable, relatively high population has been maintained with high hunting pressure and harvest of young moose, thereby sparing most of the productive component of the population. hunter observations (number of moose observed per man-day, fig. 11) to some extent describe the relative change in the moose population. however, it was obvious that after lengthening the season gradually after 1980, the mean number of moose observed declined in october and november because of the removal of 20-30% of the population; the majority early in the hunt. the long, late fall season also includes leaf fall and the in number of moose seen per man-day does tion. for the last 15 years nord-trøndelag county used cersim (cervidae simulation model), a computer-based model that uses harvest statistics, hunter observations, and various biological parameters as input, to better integrate our hunter observations into population predictions. the calculations appear to be reliable, and are used to pre-determine the size and composition of moose harvest in subsequent fall seasons in each municipality. moose density moose population density is primarily related to browse production and availability, snow conditions, and predation (bubenik et al. 1975), and hunting pressure and strategy (ritcey 1974, timmermann 1987, boer 0.4 0.6 0.8 1.0 1.2 1.4 1970 1975 1980 1985 1990 1995 2000 calves per cow w ith calves calves per cow ,total fig. 10. reproduction based on hunter observations during the hunting season at værdalsbruket, norway, 1969-2004. with , total alces vol. 41, 2005 lykke management of norwegian moose 19 1991). værdalsbruket has about 400 km2 of moose habitat including productive forests, other forestland, and marshland below tree line with a current harvest of 2.5 moose per 10 km2. population estimates based on harvests, observed population structure and reproduction, hunting pressure, and population growth indicated a prehunt (fall) density of 1.0-1.1 moose per km2, and a winter density of 0.7-0.8 that underestimates localized winter concentrations. current harvest in norway is 3.0 moose per 10 km2, with winter density of 0.8-0.9 moose per km2. in the local county of nord-trøndelag harvest is 4.2 and some low lying, highly productive municipalities have harvests of 10-15 moose per 10 km2. the population density at værdalsbruket is similar to averages in norway and sweden (pimlott 1959, cederlund and sand 1991), and higher than typical in north america (karns 1998) or russia (baskin and lebedeva 1987). population structure because moose at værdalsbruket are not harvested randomly, the harvest composition (table 1, figs. 6 and 7) does not reveal population structure of the herd. however, i believe that hunter observations provide a reasonable estimate of population structure in the hunting season (fig. 9). current harvest strategy and composition. the cow: bull ratio increased from approximately 1.50 to 2.50 after 1985, and recently declined to < 2.0. however, there are some biases due to moose behaviour, time and length of season, and differences in hunting pressure on various sex and age groups (fig. 12). thus, the estimates of 30% calves and 20-25% bulls are probably low. although prime bulls have been protected to some extent, and a few are observed each year, increase the adult bull component of the population given continuous high hunting pressure of antlered moose. two reasons why the harvest of 60-65% males is sustainable, when only 52-54% are recruited, are that natural mortality and road/ railway accidents are higher in cows. cows are the most vulnerable segment because fig. 11. observed moose per man-day, and man-days per harvested moose during the hunting season at værdalsbruket, norway, 1969-2004. 0.4 0.5 0.6 0.7 0.8 0.9 1 1970 1975 1980 1985 1990 1995 2000 se e n m o o se /m a n -d a y -1 1 3 5 7 9 11 13 15 17 m a n -d a ys /m o o se s h o t observed moose per man-day man-days per shot moose management of norwegian moose lykke alces vol. 41, 2005 20 they are the largest and oldest segment of the population, they reside at lower elevations than bulls along roads and railways during winter, and they are in front during movements across roads, bad ice, and high water. cows represent 80% of adult mortality from car collisions and 77% of all non-hunting losses in the area. body weight body weight of moose in norway is inand quality of food, snow and other climatic conditions, population density relative to food sources, and genetic factors (krafft 1956, haagenrud and lørdahl 1977, hjeljord et al. 2000). it is believed that sex-age class weights have declined the last 20 years in conjunction with the increased moose population. weight reductions might be expected in a high population density that reduces forage quantity and quality, dramatic change in environmental conditions (e.g., snow depth), and a skewed sex ratio causing delayed breeding and births. such relationships are important to investigate locally given the local management autonomy in norway. young moose are ideal for such evaluations because they undergo rapid growth between age classes, they are accurately aged, and their sample size is large. the average body weights of younger sex and age groups (calf-2.5 years) were reasonably stable at værdalsbruket since 1980 (fig. 8); obvious annual variation was probably due to winter severity. the stable weights of these young age classes indicate a moose population in balance with its food supply, and the ability of a strict harvest program to ensure such balance and a reasonable age and sex structure. reproduction the possibility of lowered reproduction because of high population density and skewed population structure is central fig. 12. hunting pressure (harvest rate, %) on various sex and age groups of moose, værdalsbruket, norway, 1969-2004. 0 5 10 15 20 25 30 35 40 45 50 1970 1975 1980 1985 1990 1995 2000 p er ce n t adult bulls adult cow s calves totalcows alces vol. 41, 2005 lykke management of norwegian moose 21 to moose management concerns in scandinavia. the productivity of 1.20 calves per reproductive cow, and 0.7 calves per cow (fig.10) in værdalsbruket, seems reasonable given the location. cederlund and sand (1991) found calf recruitment in south sweden was twice that in northern sweden. markgren (1969) found that 51% of yearling cows ovulated in highly productive coastal areas and only 8% in inland habitat, and an adult ovulation frequency of 1.56 in coastal cows and 1.11 in inland cows. further, reproduction at værdalsbruket has been relatively stable for the last 35 years (fig. 9 and 10), despite variation in sex ratio of 1.5 to 3.0 cows per bull. a high cow:bull sex ratio typically results in a high percentage of late born calves. taiga moose are serial maters (bubenik 1998) and each cow occupies the bull for 3-5 days. thus, a bull does not have the chance to mate with possibly fewer in low density populations. the percentage of barren cows (young cows and “resting” cows) at værdalsbruket cows. the age and weight of cows are important factors in reproduction (markgren 1969, sæther 1987), and no cow is expected to reproduce every year; all have resting years and twins require more rest than single calves (sand and bergstrøm 2004); barren cows may be of any age (schwartz 1998). in 1985-1995 the barren cow percentage at værdalsbruket (55%) was somewhat higher than normal. further, there was a concurrent 5% drop in the percentage of male calves in the harvest and the percentage of late born calves was high. if large bulls produce more male calves (sæther et al. 2001), a lack of prime bulls may have cows, the time of mating, and production of male calves at værdalsbruket (bubenik 1987, 1990, 1998; sæther et al. 2001). of interest is that the male fraction of harvested calves in norway dropped from 56 to 51% from 1980 to 2003, and the low number of prime bulls in regional populations is a management issue in scandinavia. a slight increase in reproduction, a decline in barren cows, and an increase of male calves in the harvest have occurred at værdalsbruket since 1995. the rise in reproduction is probably higher than indicated because of the extended hunting season and high hunting pressure on calves (fig. 12). this higher production is probably associated with the higher mean age of cows, more bulls per cow, and more prime bulls in the population, all results of the selective harvest system. future challenges the data presented here indicate that a selective harvest system is an important and valuable tool in moose management. at værdalsbruket it led to controlled population growth, improved age and sex structure, and stable production and body weights. further, the implementation of the system was well received by hunters. the effectiveness of the system was largely dependent upon the level and progression of hunting price with body weight, and adherence to harvest guidelines future challenges include balancing the moose population density relative to growth and physical parameters, monitoring forest damage and effects on forest biodiversity, and implementing harvest strategies to manage population size relative to temporal changes in browse production. of major importance is to focus moose management curate harvest statistics of sex and age composition, body weights, hunting pressure, and population characteristics including sex and age composition and reproduction. these data and continuous hunting pressure allow annual adjustments in harvest strategy that avoid abrupt, periodic changes in the management of norwegian moose lykke alces vol. 41, 2005 22 moose population. although the preferred browse species of rowan, aspen, and sallow/willows are heavily browsed in the primary wintering areas, forest damage and impact on biodiversity at værdalsbruket is tolerable. under these circumstances, population parameters are a better guide than forest damage to decide an appropriate population density, especially in spruce-dominated habitat. regardless of overall winter density, there will always be over-browsing in localized areas of the winter range. given that hunting values are increasing and timber values have declined steadily since 1960, increased moose harvest and population density are possible. the population should be increased slowly, and controlled by a sensitive, selective harvest system. should the habitat situation change, for instance by reduced clear-cutting and browse production, the moose population may require reduction. in that case, an increased harvest strategy should protect an adequate portion of adult cows and bulls. the important part of a selective harvest sex and age groups of the moose population, while maintaining high production and hunting opportunity. future management programs should be designed to protect most mature moose, stabilize the cow:bull ratio at 1.5-2.0, and continue forestry practices acknowledgements i wish to express my sincere thanks to anders børstad – my successor at værdalsbruket, and to paul harald pedersen – wildlife manager at nord-trøndelag county, for their help in many ways. furthermore, i am most grateful to all my north american moose biologist friends who have given me inspiration along the way since my stay at the university of british columbia in 1968-1969 and the north american moose conference and workshop in alaska, 1968. special thanks to dr. vince crichton and tim timmermann, north american moose biologists, for valuable advice and critical review of the initial manuscript. references ahlen, i. 1975. winter habitats of moose and deer in relation to land use in scandinavia. viltrevy 9:45-192. angelstam, p., p.-e. wikberg, p. danilov, and k. nygren. 2000. effects of moose density on timber quality and biodiversity restoration in sweden, finland and russian karelia. alces 36:133-144. baskin, l. m. 1987. behaviour of moose inbehaviour of moose in the ussr. swedish wildlife research supplement 1:377-387. _____, and n. l. lebedeva. 1987. moose management in ussr. swedish wildlife research supplement 1:619-634. boer, a. h. 1991. hunting: a producthunting: a product or a tool for wildlife managers? alces 27:74-78. bubenik, a. b. 1987. behaviour of moosebehaviour of moose (alces alces) of north america. swedish wildlife research supplement 1:333-366. _____. 1990. principles of sociobiological management based on the impact of maturation processes on population behaviour in moose. abstracts of the third international moose symposium,symposium, syktyvkar, ussr:181. _____. 1998. behaviour. pages 173-222 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington,washington, d.c., usa. _____, h. r. timmermann, and b. saunders. 1975. simulation of population structuresimulation of population structure and size in moose on behalf of age-structure of harvested animals. proceedings of the north american moose conference and workshop 11:391-463. alces vol. 41, 2005 lykke management of norwegian moose 23 cederlund, g. n., and h. k. g. sand. 1991. population dynamics and yield of a moose population without predators. alces 27:31-40. connor, k. j., w. b. ballard, t. dilworth, s. mahoney, and d. anisons. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 36:111-132. cumming, h. g. 1974. annual yield, sex and age of moose in ontario as indices to the effects of hunting. naturaliste canadien 101:539-558. gaidar, a. a., n. n. grakov, and b. m. zhitkov lective moose hunts in a forest-taiga zone of russia. abstracts of the third international moose symposium, syk-symposium, syktyvkar, ussr: 107. haagenrud, h. 1986. elgens livshistorie. pages 9-35 in p. hohle and j. lykke, editors. elg og elgjakt i norge. gyldendalelg og elgjakt i norge. gyldendal norsk forlag, oslo, norway. _____, m. håker, and l. lørdahl 1975. elgundersøkelsene i grane,vefsn og hattfjelldal 1967-1975. viltforskningen:1-41. _____, and l. lørdahl. 1977. vektutviklingvektutvikling om høsten hos elg i trøndelag. (car-(carcass weight in moose from trøndelag, norway). meddelelser fra norsk vilt-meddelelser fra norsk viltforskning 3(3):1-27.3(3):1-27. _____, and _____. 1979. sex differential in populations of norwegian moose alces alces (l.). meddelelser fra norskmeddelelser fra norsk viltforskning 3(6):1-19. heptner, w. g., and a. a. nasimowitsch. 1967. der elch. die neue brehmbucherei. wittenberg lutherstadt, germany. hjeljord,o., e. rønning, and t. histøl. 2000. yearling moose body mass: im-yearling moose body mass: imselective feeding. alces 36:53-59. karns, p. d. 1998. population distribution, density and trends. pages 125-140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. krafft, a. 1956. størrelsen av norsk elg. særtrykk av jeger og fisker 11 og 12:1-13. krefting, l. w., and j. lykke. 1976. aa comparison of moose habitat in north america and norway. proceedings xvi. iufro world congress: 731-742. lavsund, s. 1987. moose relationships to forestry in finland, norway and sweden. swedish wildlife research supplement 1:229-244. lorentsen, ø., b. wiseth, k. einvik, and p. h. pedersen. 1991. elg i nord-trøndelag. fylkesmannen i nord-trøndelag, rapport 1:1-208. lykke, j. 1964. elg og skog. studies of moose damage in a conifer forest area in norway. papers of the norwegian state game research institute 2(17):1-57. _____. 1974a. moose management in norway and sweden. naturaliste canadien 101:723-735. _____. 1974b. elgens økologi og skjøt-1974b. elgens økologi og skjøtsel. moose ecology and management.moose ecology and management. norwegian journal of forestry 82:235337. _____, and i. m. cowan. 1968. moose management and population dynamics on the scandinavian peninsula, with special reference to norway. proceedings of the north american moose conference and workshop 5:1-22. lykke, l. 1952. avgang i elgstammen utenom jakttiden. jeger og fisker 10:369-373. _____. 1960. elgen og elgjakten. pageselgen og elgjakten. pages 197-219 in p.hohle, editor. jakt og fiske i norge, bind jakt. norsk arkivforskning, oslo, norway. _____. 1962. elgtelling i verdal. jakt-elgtelling i verdal. jaktfiske-friluftsliv 91:206-207. management of norwegian moose lykke alces vol. 41, 2005 24 _____. 1968. forvaltningen av en elgstam-forvaltningen av en elgstamme. jakt-fiske-friluftsliv 7:310-313. markgren, g. 1969. reproduction of moose in sweden. viltrevy 6(3):127-299. mercer, w. e., and m. strapp. 1978. moose management in newfoundland 1972-1977. proceedings of the north american moose conference and workshop 14:227-233. pimlott, d. h. 1959. moose harvests in newfoundland and fennoscandian countries. transactions of the north american wildlife conference 24:422448. rausch, r. a., r. j. sommerville, and r. bishop. 1974. moose management inmoose management in alaska. naturaliste canadien 101:705721. ritcey, r. w. 1974. moose harvesting programs in canada. naturaliste canadien 101:631-642. sæther, b.-e. 1987. patterns and proc-patterns and processes in the population dynamics of the scandinavian moose (alces alces): some suggestions. swedish wildlife research supplement 1:525-537. _____, m. heim, e. j. solberg, k. jakobsen, r. olstad, j. stacy, and m. svilland. 2001. effekter av rettet avskyting påeffekter av rettet avskyting på elgbestanden på vega. effects of sex-effects of sex population on the island of vega. nina fagrapport 049:1-39. sand, h., and r. bergstrøm. 2004. kalvar kostar kon krafter. svensk jaktsvensk jakt 5:70-73. _____, o. liberg, p. ahlqvist, and p. wabakken. 2004. algjakten kan hotas i vargområden. svensk jakt 10:84-86.svensk jakt 10:84-86. schwartz, c. c. 1998. reproduction, natality and growth. pages 141-172 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington d.c., usa. skuncke, f. 1949. algen. studier, jaktstudier, jakt och vård. p.a.norstedt & sons forlag, stockholm, sweden. solberg, e. j., and b.-e. sæther. 1999. “sett elg”. elgen: 63-67. _____, v.grøtan, m. heim, and b.-e. sæther. 2001. velger vi skogens konge eller skogens hellige kyr? elgen 11:44-48. thompson, i. d. 1990. effects of moose on structure and composition of forests in north america. abstracts of the third international moose symposium, syktyvkar, ussr: 75. timmermann, h. r. 1987. moose harvestmoose harvest strategies in north america. swedish wildlife research supplement 1:565579. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice f:\alces\vol_39\p65\3913.pdf alces vol. 39, 2003 schwartz et al. predator management and moose 41 large carnivores, moose, and humans: a changing paradigm of predator management in the 21st century charles c. schwartz1, jon e. swenson2, and sterling d. miller3 1interagency grizzly bear study team, u.s. geological survey, biological resources division, northern rocky mountain science center, montana state university, bozeman, mt 59717, usa; 2department of biology and nature conservation, agricultural university of norway, box 5014, n1432 ås, norway; 3national wildlife federation, 240 north higgins, suite #2, missoula, mt 59802, usa abstract: we compare and contrast the evolution of human attitudes toward large carnivores between europe and north america. in general, persecution of large carnivores began much earlier in europe than north america. likewise, conservation programs directed at restoration and recovery appeared in european history well before they did in north america. together, the pattern suggests there has been an evolution in how humans perceive large predators. our early ancestors were physically vulnerable to large carnivores and developed corresponding attitudes of respect, avoidance, and acceptance. as civilization evolved and man developed weapons, the balance shifted. early civilizations, in particular those with pastoral ways, attempted to eliminate large carnivores as threats to life and property. brown bears (ursus arctos) and wolves (canis lupus) were consequently extirpated from much of their range in europe and in north america south of canada. efforts to protect brown bears began in the late 1880s in some european countries and population reintroductions and augmentations are ongoing. they are less controversial than in north america. on the other hand, there are no wolf introductions, as has occurred in north america, and europeans have a more negative attitude towards wolves. control of predators to enhance ungulate harvest varies. in western europe, landowners own the hunting rights to ungulates. in the formerly communistic eastern european countries and north america, hunting rights are held in common, although this is changing in some eastern european countries. wolf control to increase harvests of moose (alces alces) occurs in parts of north america and russia; bear control for similar reasons only occurs in parts of north america. surprisingly, bears and wolves are not controlled to increase ungulates where private landowners have the hunting rights in europe, although wolves were originally exterminated from these areas. both the inability of scientific research to adequately predict the effect of predator control on ungulate populations and a shift in public attitudes toward large carnivores have resulted in an accelerating number of challenges to predator management in places where it is still espoused. utilitarian attitudes towards wildlife are declining in western cultures and people now increasingly recognize the intrinsic value of wildlife, including large predators. in the future, agencies responsible for managing resident wildlife will face increased pressure to balance the needs of the hunting public with the desires of non-hunting publics. we suggest that in the next century we will witness a continued shift in how wildlife agencies manage both moose and large carnivores. more attention will be paid to maintaining and restoring intact ecosystems and less toward sustainable yield of meat. alces vol. 39: 41-63 (2003) key words: alces alces, brown bear, canis lupus, gray wolf, grizzly bear, moose management, predator control, ursus arctos predator management and moose schwartz et al. alces vol. 39, 2003 42 there is ample evidence that both wolves (canis spp.) and bears (ursus spp.) kill and eat moose. there is also empirical data suggesting that this predation can limit moose numbers under certain conditions (ballard and van ballenberghe 1998). wolf reduction can result in increased numbers of moose under some circumstances (gasaway et al. 1983). however, there are only a few empirical studies supporting the principle that bear reduction programs result in enhancement of moose numbers (stewart et al. 1985, ballard 1992). to our knowledge, no study has addressed the long-term effects of bear control on moose numbers. in only a few places in north america does control or reduction in wolf and bear numbers continue to be strongly advocated by some citizens groups and agencies responsible for wildlife management. a similar attitude regarding wolf control for livestock safety also exists in europe. however, in the past 2 decades, agencies have witnessed increasing levels of concern toward and criticism of predator control programs by a more vocal public, particularly the environmental community. these groups question both the scientific validity and the philosophical basis for carnivore control. concurrently, there is an apparent shift in how society in general values large predators, particularly bears and wolves (duda et al. 1998). in recent times, a larger contingent of the public and scientific community finds inherent intrinsic value in carnivores and perceives their role as necessary in ecosystem function (miller et al. 2001); they oppose predator control that favors the more utilitarian attitude of “game production” as the objective of wildlife management. these attitudinal shifts are prevalent in both north america and europe despite dissimilar legal systems of game management. here, we review the evolution of predator control in north america and europe, agency culture and big game management, the scientific basis of predator management, and an apparent shift in social values in the past decades away from predator control and toward large carnivore conservation and management. early human–predator relationships three stages between humans and their environment have been described: hunting, shepherding, and agricultural (boitani 1995). human attitudes toward large carnivores have been shaped by these relationships. hunting economies were centered on herbivores as an important source of food. large predators were perceived as competition, but not as a threat. hunters had respect for and kinship with predators. this was reflected in attitudes of aboriginal peoples. nomadic shepherds disdained wolves as threats to their livestock and basis for livelihood. in contrast, sedentary herders had more tolerant attitudes toward wolves because they had housing to protect their livestock. farmers, producing crops and limited livestock, had leeway to be more tolerant. however, agricultural people living in latin cultures, characterized by closed villages, were more tolerant than those living in germanic societies, characterized by more open settlements and solitary farms (breitenmoser 1998). bears and wolves played important roles in the legends, beliefs, and lives of prehistoric peoples. bears were potentially dangerous and fearsome creatures to hunt and kill and had physical similarities to man (rockwell 1991). like man, bears were omnivorous, generalists, intelligent animals w i t h b i n o c u l a r v i s i o n . t h e y l i v e d sympatrically with humans and ate many of the same foods; when they stood erect or were skinned out, they shared physical similarities to humans (shepard 1996). these characteristics may have contributed to many similar myths and legends involving alces vol. 39, 2003 schwartz et al. predator management and moose 43 creatures that were half human and half bear as a consequence of mating between humans and bears (rockwell 1991, shepard and sanders 1992). bears disappeared in the fall and reappeared in the spring giving rise to beliefs that they had a special 2-way route to the afterlife and could commute readily between worlds; an ability worthy of great respect. all these characteristics gave bears a special importance in early cultures closely bound to wild animals. throughout the shared ranges of bears and humans, people developed elaborate and detailed rituals to appease the spirits of bears they killed (rockwell 1991, shepard and sanders 1992, edsman 1994) and many of these rituals continue today in some native american and eurasian cultures (black 1998). wolves played a great mythic-religious role because they possessed many similar characteristics with humans; they were great hunters, members of a pack (tribe or clan), defended territories, and hunted cooperatively (lopez 1978). large carnivores were eliminated from much of their former range both in europe and north america. the extermination of wolves in europe started in the middle ages and continued well into the early part of the 20th century (mallinson 1978, boitani 1995). in great britain, wolves were considered a threat to livestock and exterminated by the 17th century. they were exterminated from northern europe by the beginning of the 20th century, but survived in lower numbers in southern and eastern europe (boitani 1995). the extermination of brown bears followed a similar pattern; bears were extirpated from denmark before the middle ages (jessen 1929) and from britain during the 10th century (corbet and harris 1991). bears survived in relict populations in northern and southern europe, and in greater populations in eastern europe (swenson et al. 2000). the eradication of wolves and bears, and the eurasian lynx (lynx lynx), a predator on small ungulates, was directly caused by persecution, including bounties, and indirectly through habitat destruction and elimination of prey. bounties on wolves were initiated in england in the 1500s, on wolves and bears in sweden in 1647, and norway in 1733 (myrberget 1990, swenson et al. 1995, elgmork 1996). the destruction of the forests, due to the expansion of cultivation and overgrazing by domestic livestock, was an important factor, as was the extermination of native herbivores (breitenmoser 1998). for example, by the year 1200, 40% of switzerland’s forests had been cleared (breitenmoser 1998). the hungarian landscape was converted from 87% forest and wetland around 900 to less than 11% forest by 1920 (csányi 1997). also, the napoleanic wars in the 1800s resulted in the spread of modern firearms. the result was the virtual elimination of the remaining big game species in much of europe (breitenmoser 1998), even though over–hunting had contributed to the extinction of the wild boar (sus scrofa) in england in the 1500s, the capercaille (tetrao urogallus) in britain around 1790, and the complete extinction of the auroch (bos primigenius) by 1627 (myrberget 1990). the famous swedish taxonomist, carl von linné (linneaus) probably never saw a wild moose, and his description of the species in 1746 was based on a captive individual. in 1789, a swedish law allowed landowners unrestricted hunting on their land, with the result that moose were almost totally exterminated in sweden by 1825, when the law was repealed (bergström et al. 1993). with little wild ungulate prey, the large carnivores attacked the abundant domestic animals, increasing conflicts and persecution by people. when europeans colonized the north american continent, they brought their old world culture and traditions with them, including a view of wilderness as “something predator management and moose schwartz et al. alces vol. 39, 2003 44 alien to man–an insecure and uncomfortable environment against which civilization had waged an unceasing struggle” (nash 1982:8). to early settlers, wilderness was villainous as were the wild animals and indians living in it. taming wilderness meant the extermination of large carnivores, particularly wolves and bears. the prevailing attitude of colonial america was summarized in a quote from john adams in 1756: “the whole continent was one continuing dismal wilderness, the haunt of wolves and bears and more savage men. now the forests are removed, the land covered with fields of corn, orchards bending with fruit and the magnificent habitations of rational and civilized people” (kellert 1996:104). the difference between european colonizers and the american indians in attitude to wild lands and wild places was eloquently phrased by sioux chief luther standing bear (1932, from deloria 2001): “we did not think of the great open plains, the beautiful rolling hills, and winding streams with tangled growth as ‘wild.’ only to the white man was nature a ‘wilderness’ and only to him was the land ‘infested’ with ‘wild’ animals and ‘savage’ people. to us it was tame. earth was bountiful and we were surrounded with the blessings of the great mystery.” as civilization moved westward, the pioneers viewed predators much like the nomadic shepherds described by boitani (1995). since the prevailing form of livestock husbandry was to allow large herds of cattle and sheep to graze freely over vast areas, carnivores, particularly wolves and grizzly bears, were considered an economic threat. the pervasive attitudes of the time were captured in 2 salient quotes (from nrc 1997:135). historian and trapper stanley young wrote: “there was sort of an unwritten law of the range that no cow man would knowingly pass by a carcass of any kind without inserting in it a goodly dose of strychnine sulfate, in the hope of killing one more wolf” (young 1946:27). an early director of the u.s. biological society, e. a. goldman, wrote, “large predatory animals destructive of livestock and game, no longer have a place in our advancing civilization” (dunlap 1988:51). an early american stockman had similar views: “the destruction of these grizzlies is absolutely necessary before the stock business…could be maintained on a profitable basis.” (bailey 1931 cited in usfws 1993). with the exception of extreme northern minnesota, wolves were eliminated from the conterminous 48 states by the 1900s (boitani 1995). between 1800 and 1975, grizzly bears were eliminated from nearly 98% of their historic range (usfws 1993, mattson et al. 1995). at the time of the lewis and clark expedition, grizzly bears inhabited most of the western united states and extended out into the great plains (servheen 1999). they flourished where pacific salmon (oncorhynchus spp.) were abundant far inland into eastern idaho. today, they exist as only 5 remnant populations south of canada. three of these populations contain <50 individuals and only 2 contain >350 individuals (servheen 1999). one population, in the north cascades along the pacific coast, is highly endangered in both the united states and canada. no bears have been verified on the united states side of the border in recent decades. bears and wolves faired better north of the 49th parallel. the chronology of wolf extirpation in southern canada followed the pattern of agricultural and industrial settlement (hayes and gunson 1995). wolves were extirpated from many areas in the eastern provinces by the early 1900s, and significantly reduced in the western provinces by the 1930s (carbyn 1987, hayes and gunson 1995). grizzly bears followed a similar pattern. they were extirpated from part of their historic range in manialces vol. 39, 2003 schwartz et al. predator management and moose 45 toba, saskatchewan, and alberta, primarily in the prairies and boreal plains and are scarce in southern alberta and british columbia, where human populations are concentrated (macey 1979, banci 1991, banci et al. 1994, mclellan and banci 1999). even in alaska, where wolves and brown and black bears are still abundant, there were several attempts to eliminate them. according to sherwood (1981:24) “periodically alaskan civic leaders advocated the extermination of the animals [brown bears] and make ursus the symbol of an alleged colonialism that they claimed was inspired by conservationist sentiment and directed by bureaucrats in washington d.c., working in concert with vested absentee interests. this colonialism, they believed, prevented resident alaskan entrepreneurs from exploiting the territory’s natural resources and prevented resident politicians from setting the terms of that exploitation.” brown bears were perceived as a direct threat to the fledgling cattle industry on kodiak island (van daele 2003). political pressure from the industry resulted in regulations in 1929 allowing kodiak cattlemen to kill bears at any time they were considered a menace to livestock or property. according to sherwood (1981:59) one rancher’s advice to anyone encountering a brown bear was, “shoot them in the guts, in the foot, any place, but get a bullet into them.” prior to statehood in 1959, wolves in alaska were targeted in a major predator control program led by the united states government. wolf control was pervasive with the intent to increase moose and caribou populations. strychnine and cyanide were commonly used and later aerial gunning was employed as a very effective technique. following statehood, state management eventually evolved to where the wolf was listed as a game animal in 1963 (nrc 1997). as civilization expanded and human densities increased, predators were either significantly reduced or eliminated from much of their range in both north america and europe (woodroffe 2000). as we shall see, these early attitudes toward large carnivores have persisted, to varying degrees, around the world. early development of wildlife management systems throughout most of the historical record, wildlife managers have targeted predators and have had a significant impact on their numbers, distribution, and more recently, conservation. because wildlife has value for recreation and food, there are significant economic incentives for individuals, such as landowners, to acquire property rights to wildlife. thus, the system of wildlife management in a country can contribute to the attitudes towards predation by large carnivores on big game (lueck 1995). europe in most of europe, the kings and their chieftains controlled most of the lands and hunting rights during this feudal period, which was at its peak around 1000. the oldest general game law, introduced by king knut of denmark in 1016, established that no one owned wild animals, but that the king had some hunting privileges. this system declined as a result of corruption and the black death, and around 1348-49, hunting privileges, especially for big game, were transferred to the large landowners. this was completed in europe by the 17th and 18th centuries. in some countries, such as finland, norway, and switzerland, small farmers and others held many hunting rights (myrberget 1990). the norwegian parliament transferred all hunting rights, except for killing large carnivores, to landowners in 1899 (søilen 1995). predator management and moose schwartz et al. alces vol. 39, 2003 46 landowners often lost exclusive hunting rights following changes in political systems, such as the revolutions in france and russia and the introduction of communism in eastern europe. interestingly, the decision that hunting rights in the united states would not be held by the landowner, as is the case in britain, was made following the american revolution. after world war ii, europe had essentially 2 systems, with landowners having the hunting rights in western europe and governments managing hunting in the communistic eastern europe. hunting remained open to all citizens of portugal, italy, greece, and turkey (myrberget 1990). since the fall of communism, countries of eastern europe have been in flux about whether to revert to the former system of landowners owning hunting rights, as hungary has, or to retain state control over wildlife management, independent of land ownership, as has poland, slovakia, ukraine, and romania (csányi 1997, salvatori et al. 2002). north america in north america, prior to european colonization, the american indian tribes claimed rights to wildlife by protecting hunting and fishing territories (carlos and lewis 1995). “the ownership of game among native americans had an uncanny resemblance to current united states institutions. indian tribal societies, like state agencies, controlled wildlife stocks by enforcing the rights to hunting and fishing territories and restricting the time and method of harvest by tribal members.” (lueck 1995:3). american game laws are rooted in the past history of english common law. however, today’s american and english wildlife laws are markedly different. in the united states, ownership of wildlife resides with the people and is administered on their behalf by government, primarily the state governments. in great britain, the law places nearly all control in the hands of private landowners. in canada, wildlife is managed by the provincial governments on behalf of the people similar to the united states, but ownership is vested in the crown until the wildlife is legally killed. at this point property rights transfer to the hunter. the states and provinces have retained control over most wildlife management, but they have lost some authority to the federal governments with international treaties (e.g., the migratory bird treaty act and marine mammal protection act) and, in the united states, with the lacey act (controls interstate transportation of game) and the endangered species act. but the states and provinces have vigorously fought to retain the authority to manage wildlife (peek 1986). conflicts surrounding predator control to improve harvest of wild ungulates in his classic book “game management,” leopold (1933:3) defined game management as “the art of making land produce sustained annual crops of wildlife for recreational use.” this definition espoused a utilitarian philosophy of game management that established the direction of wildlife management for the next half century. w i l d l i f e a g e n c i e s “ m a n a g e d ” g a m e populations for a “sustained yield.” their primary clients were the hunting public, and up until the 1970s virtually all state and provincial wildlife agencies operated primarily under the principle of sustained use. most universities that trained students in the field of wildlife management were land grant or agricultural colleges with a focus on production and emphasized sustainable yield concepts. peek (1986:25), discussed 2 groups of conservationists as defined by harry et al. (1969): those with a conservation-utilization emphasis, and those with a conservationalces vol. 39, 2003 schwartz et al. predator management and moose 47 preservation emphasis. both groups were concerned with the perpetuation of natural resources and therefore could be classed as conservationists. however, people with a utilization emphasis were oriented toward the goal of resource exploitation, such as hunting, with aims of producing sustained yields by cropping surpluses. “wise use” was the doctrine of those with conservation-utilization emphasis and their philosophy was adopted by most wildlife and natural resource management agencies. this was encouraged by the importance of fees paid by hunters that were vital to management activities including the salaries of the managers. conservationist-preservationists, by contrast, were not oriented towards “wise use” but rather espoused an appreciative interest in the resource, preferably in its “natural state” (harry et al. 1969). state wildlife agencies were slow to acknowledge this doctrine in part, perhaps, because there was no mechanism for conservation-preservationists to regularly support wildlife management efforts with their fees. the conservationist-preservationist movement in north america greatly increased in influence in the early 1970s and became a major part of the biopolitical scene during the 1980s. it was largely responsible for broadening activities of wildlife and land management agencies in nongame management and has evolved into what is termed environmentalism (peek 1986). conservation-preservationists were and still are largely responsible for challenging predator control programs. alaska has carried out a program of wolf control with the specific goal of increasing ungulate populations for hunters during the same period that conservation and enhancement programs were underway elsewhere in the united states. the history of predator control in alaska provides a good example of the conflicts between the wise use and the conservation-preservation groups. during the decade of the 1980s, controversy grew around the state’s wolf control programs (stephenson et al. 1995), and environmental groups filed several court cases against the states in an attempt to stop wolf control. these conflicts reached a peak in the early 1990s. by 1994, the governor of alaska suspended the state’s wolf control program because it was judged to be an unacceptable treatment of wolves (nrc 1997). the governor called for a scientific review and indicated that he would not reinstate predator control unless it met 3 criteria: (1) it was based on solid science; (2) a full cost-benefit analysis showed it made economic sense for alaskans; and (3) it had broad public support. similar confrontations over wolf control programs have also occurred in canada (hayes and gunson 1995). alaska began efforts to reduce grizzly bears in some areas in 1980 in response to indications that moose numbers in certain areas were being maintained at chronically low levels by bear predation on neonates. moose numbers declined in the 1970s as a consequence of a series of severe winters and overharvest by hunters, in addition to significant predation on moose calves by bears (ballard et al. 1981). unlike efforts to control wolves, grizzly bear control efforts in selected areas were done by liberalization of hunting regulations instead of by trapping and shooting by state employees. bear reductions started in 1980 and ultimately included elimination of requirements that residents buy a special tag to hunt bears and generous bag limits that allowed hunters to take bears more frequently in targeted areas than elsewhere in alaska. moose numbers increased during the 1980s and 1990s but calf survival remained low. available evidence indicated that increases in moose numbers were unrelated to the increased bear harvests (miller and ballard predator management and moose schwartz et al. alces vol. 39, 2003 48 1992). efforts are ongoing in targeted areas of alaska to reduce bear numbers in order to provide more moose for hunters. unlike efforts to reduce wolf abundance to accomplish the same objective, bear reduction efforts have generated less controversy, perhaps because control efforts are done gradually by legal sport hunters instead of by government agents using trapping and aerial gunning techniques. also, unlike the government sponsored control efforts in the united states south of canada in the early 20th century, bear reduction by hunters has not yet resulted in measurable declines in bear density in the portions of alaska where it is ongoing, although it has caused changes in population composition (miller 1997). control of wolves to enhance hunting opportunity is supported by 48% of alaskan voters and by 65% of alaskan hunters (miller et al. 1998). the most vehement opposition to wolf control efforts comes from conservation groups outside of alaska, but these groups have been effective at stopping or curtailing active wolf reduction programs. it is possible that the general public may be more amenable to reductions in bear abundance than to reductions in wolf abundance because, unlike wolves, bears occasionally attack humans, resulting in a fear of bears in a significant proportion of the population. a survey of alaska voters indicated that 34% had concerns about bears that sometimes kept them from going into the countryside (miller et al. 1998). similar to the grizzly bear case in alaska, black bears (ursus americanus) and mountain lions (puma concolor) were thought to be reducing recruitment of elk (cervus elaphus) in the clearwater area of central idaho. in 2000, a research program was proposed to evaluate the biology of this relationship through predator reduction efforts accomplished by liberalized hunting regulations as well as control by government officials. predator advocacy groups organized a national campaign against the proposed research similar to that conducted against the wolf reduction effort in alaska. idaho officials modified their research proposal to accomplish targeted reductions using only liberalized regulations for hunting of predators. subsequently, objections to the research have largely dissipated even though the liberalized hunting regulations have been expanded beyond the boundaries of the originally proposed research area. difficulties in convincingly documenting trends in black bear and mountain lion abundance are likely to confound interpretation of any changes in elk recruitment or abundance found in the idaho study. we know of no case of planned reduction in bear densities in europe to increase ungulate numbers. the european brown bear is much less aggressive towards humans than the north american brown/grizzly bear (swenson et al. 1996), which may also influence people’s attitudes. the persecution of wolves in recent times has been justified partially by the reduction of predation on ungulates in some areas. nevertheless, we are unaware of any case where this was the primary justification, as reductions of livestock losses seem to be the primary objective (promberger and schröder 1993). we expected that planned reductions of these predators would have occurred in the countries where landowners own the economic rights to hunting. however, in such countries, bears seem to be popular enough for hunting in their own right that owners of the hunting rights accept this tradeoff. no country with this system has a huntable population of wolves, which might not be a coincidence. however, wolves recently have been established in norway and sweden, where this system exists. in norway, the state pays compensation for the loss of hunting income due to wolf predation, but it is a country where conflicts surrounding alces vol. 39, 2003 schwartz et al. predator management and moose 49 depredation on free-grazing sheep are very intense. in sweden, where there are almost no free-grazing sheep, large forest companies are contributing to the funding of wolf research because they want to know if wolves can help reduce forest damage caused by high moose densities. studies of predator–prey relationships b a l l a r d a n d l a r s e n ( 1 9 8 7 ) , v a n ballenberghe (1987), boutin (1992), nrc (1997), and ballard and van ballenberghe (1998) all provide thorough reviews of the recent studies of predator-prey dynamics of moose in north america; only 1 intensive study has been conducted in europe (swenson et al. 2001). wolves, brown bears, and american black bears are the principal predators of moose. there is general agreement that predation is a limiting factor for moose populations, but there is controversy regarding the magnitude of this limitation and if the evidence supports the hypothesis that predation regulates moose numbers (boutin 1992). many of the predator-prey studies dealing with carnivore(wolves and bears) moose relationships were not designed to answer specific questions about the impacts of predator control on moose demographics. additionally, the full impact of wolf predation and wolf control on moose demographics has received considerably more attention than similar impacts from bears. few studies detail the influences of both bear and wolf predation on moose in the same system at the same time (table 1). even fewer studies have been conducted long enough to determine the long-term impacts of predator control on moose population dynamics. in a review of predator control in alaska, the nrc (1997) came to 3 conclusions: (1) predator control experiments provide only negative evidence for the existence of an alternative stable state with relatively high numbers of both predators and prey. only 2 studies were monitored long enough to reveal the existence of such a state, and the evidence from those studies was negative or equivocal. existing evidence suggests that if predator control is to be used as a tool to increase ungulate populations, control must be both intensive and relatively frequent. there is no factual basis for the assumption that a period of intensive control for a few years can result in long-term changes in ungulate population densities; (2) experiments that resulted in increases in moose populations were conducted where wolves were relatively numerous, where bears were relatively uncommon and were not preying heavily on ungulate calves, where habitat quality was high, and weather was relatively benign. the evidence is inconclusive, but there is reason to believe that an intensive control effort, during which wolf populations are greatly reduced for several years and other factors are favorable, can result in short-term increases in moose populations; and (3) control experiments that appeared to have had some success used methods, such as aerial shooting, that are not currently politically acceptable. during the 1970s and early 1980s, wolf removal had strong support among management agencies in alaska and the yukon as a means of increasing moose densities (gasaway et al. 1983, ballard and larsen 1987). however, some wolf control experiments met with limited success, and telemetry studies during the same period (franzmann et al. 1980, ballard et al. 1981, franzmann and schwartz 1986, boertje et al. 1988, ballard and miller 1990) implicated bears as a limiting factor in calf survival. consequently, emphasis shifted toward both wolf and bear control programs, yet there was little empirical data to support bear control as a long-term management tool to increase moose numbers (table 1). current science allows managpredator management and moose schwartz et al. alces vol. 39, 2003 50 t ab le 1 . p re da to r r ed uc ti on e xp er im en ts (f ro m n r c 1 99 7) . p re da to r re du ct io n m et ho d an d lo ca ti on w ol ve s b ea rs r es ul ts a ir -a ss is te d e as tce nt ra l w ol f po pu la ti on r ed uc ed n ot d on e a k to 5 580 % b el ow p re -c on tr ol (g m u 2 0a ) nu m be rs fo r 7 y rs (1 97 682 ). f in la ys on , 49 -8 5% o f w ol f p op ul at io n n ot d on e y uk on re m ov ed fo r 6 y rs (1 98 389 ); hu m an h ar ve st r at e of m oo se an d ca ri bo u re du ce d by 9 0% . s ou th w es t w ol f nu m be rs r ed uc ed b y b ea r p op ul at io n y uk on 40 -8 0% fo r 5 y ea rs re du ct io n es ti m at ed (r os e l ak e; 1 98 287 ). at 7 -9 % (1 98 287 ). a is hi hi k, a pp ro xi m at el y 76 % o f w ol f n ot d on e y uk on po pu la ti on s re m ov ed o ve r 4 yr s (1 99 396 ). m oo se a nd ca ri bo u hu nt in g cu rt ai le d. n or th er n b c 1, 00 0 w ol ve s r em ov ed in n ot d on e 10 y ea rs (1 97 887 ); a lm os t 80 0 of w hi ch w er e re m ov ed in th e la st 4 y rs o f r em ov al . q ué be c w ol f po pu la ti on r ed uc ed t o a to ta l o f 81 b ea rs r em ov ed 48 -6 2% o ve r 4 y ea rs (1 98 285 ). ov er a 3 y r pe ri od in a di ff er en t a re a ( 19 83 -1 98 5) . a ve ra ge a nn ua l r at e of in cr ea se o f m oo se p op ul at io ns w as 1 5% d ur in g w ol f co nt ro l, an d 5% f or 1 2 yr s af te r th e en d of w ol f co nt ro l. a ve ra ge an nu al r at e of in cr ea se o f ca ri bo u po pu la ti on s w as 1 6% d ur in g, a nd 6 % fo r 7 yr s af te r th e en d of w ol f co nt ro l. in cr ea se d su rv iv al o f ad ul t c ar ib ou ; i nc re as ed n um be rs o f ca lv es /1 00 co w s fo r bo th m oo se a nd c ar ib ou . a ve ra ge a nn ua l r at e of in cr ea se f or m oo se a nd c ar ib ou a bo ut 1 618 % . h un ti ng s uc ce ss in cr ea se d. s ev en ye ar s af te r w ol f co nt ro l e nd ed , m oo se a nd c ar ib ou n um be rs b eg an to de cl in e. n o su bs ta nt ia l i nc re as e in m oo se p op ul at io ns o r co w :c al f ra ti os du ri ng p re da to r r em ov al . n o su bs ta nt ia l i nc re as e in m oo se p op ul at io ns o r co w :c al f ra ti os d ur in g pr ed at or re m ov al . in cr ea se d nu m be rs o f ca ri bo u ca lv es /1 00 c ow s. r es po ns e of m oo se hi gh ly v ar ia bl e an d no t c le ar ly r el at ed to w ol f re du ct io n. c on tr ol e nd ed in 1 99 6, to o re ce nt ly to a ss es s lo ng -t er m tr en ds . c al f su rv iv al r at es a nd p op ul at io n si ze s ap pa re nt ly in cr ea se d fo r al l un gu la te s in th e ar ea . w he n w ol f co nt ro l e nd ed , w ol f nu m be rs in cr ea se d ra pi dl y an d ca lf s ur vi va l d ec re as ed to p re -c on tr ol le ve ls . n o ap pa re nt c ha ng e in m oo se c al f su rv iv al r at e in e it he r w ol f or b ea r re m ov al a re a. alces vol. 39, 2003 schwartz et al. predator management and moose 51 e as tce nt ra l w ol f po pu la ti on w as r ed uc ed n ot d on e a k b y 28 -5 8% f or 3 y rs (1 98 184 ). (g m u 2 0e ) s ou th -c en tr al e xt en si ve a er ia l s ho ot in g p oi so ni ng p ro ba bl y a k an d po is on in g re du ce d al so r ed uc ed b ea r (g m u 1 3) w ol f n um be rs d ra m at ic al ly nu m be rs . (1 94 819 54 ). w ol ve s w er e re du ce d by a ft er w ol f c on tr ol 42 -5 8% fo r 3 y rs (1 97 678 ). en de d, 6 0% o f th e be ar p op ul at io n w as t ra ns -l oc at ed o r re du ce d by li be ra li ze d hu nt in g re gu la ti on s. g ro u n d -b as ed k en ai r ec re at io na l ha rv es ts p en in su la , in cr ea se d fo r 4 yr s a k (1 97 679 ). v an co uv er w ol f de ns it y w as r ed uc ed n ot d on e is la nd , b c ov er a 4 -y r p er io d (1 98 386 ) to a bo ut 1 0% o f pr eco nt ro l le ve ls . s as ka tc he w an n ot d on e u nk no w n pr op or ti on o f bl ac k be ar p op ul at io n w as r em ov ed in s pr in g of 1 98 3, a nd fr om an ot he r a re a in s pr in g of 1 98 4. t ab le 1 . c on ti nu ed . p re da to r re du ct io n m et ho d an d lo ca ti on w ol ve s b ea rs r es ul ts w ol f co nt ro l h ad n o m ea su ra bl e ef fe ct o n ca lf s ur vi va l. d ur in g an d af te r th is p re da to r re du ct io n pe ri od , c ar ib ou n um be rs in cr ea se d an d ha d m or e th an d ou bl ed b y th e ea rl y 19 60 s. t hi s co in ci de d w it h fa vo ra bl e w ea th er a nd r an ge c on di ti on s an d lo w ha rv es t by h um an s. w ol f co nt ro l d id n ot r es ul t i n hi gh a nn ua l i nc re as es in th e m oo se po pu la ti on . c al f su rv iv al in cr ea se d af te r be ar r em ov al , b ut b ea rs re tu rn ed to th e ar ea a ft er s um m er a nd m oo se c al f su rv iv al r et ur ne d to le ve ls b ef or e be ar r em ov al . n o ch an ge in c al f nu m be rs c ou ld b e at tr ib ut ed t o in cr ea se d be ar h ar ve st s. w ol f re du ct io ns w er e as so ci at ed w it h in cr ea se d m oo se p op ul at io ns , bu t th e hi gh es t nu m be r of m oo se o cc ur re d in a re as t ha t ha d bu rn ed 10 -2 5 ye ar s ea rl ie r r eg ar dl es s of th e ex te nt o f w ol f r ed uc ti on s. d ee r po pu la ti on s in cr ea se d fo ll ow in g w ol f co nt ro l. h un te r ef fo rt ap pe ar ed to b e en ha nc ed b y w ol f co nt ro l. in cr ea se d ca lf :c ow ra ti os a ft er b ea r r em ov al . t he fo ll ow in g ye ar , pr op or ti on o f ye ar li ng s in t he m oo se p op ul at io n w as h ig he r th an be fo re p re da to r co nt ro l. predator management and moose schwartz et al. alces vol. 39, 2003 52 ers to predict with reasonable confidence that predator removal at low moose densities can improve calf survival, but increased recruitment and subsequent growth of the moose population may or may not occur. data are generally unavailable on longterms effects of predator reduction. times change: from predator extermination to predator conservation europe because of the degradation of forests and loss of indigenous large herbivores, the attitudes of europeans towards the environment were changing by the end of the 1800s. in switzerland, laws requiring reforestation, banning livestock from forests, restricting hunting, and establishing game s a n c t u a r i e s w e r e p a s s e d i n 1 8 7 6 (breitenmoser 1998). similar laws, and an increase in effort to reintroduce ungulates into former habitats, occurred throughout europe. the first modern law regulating hunting and protection of game in norway was passed in 1845 (søilen 1995). the swedish hunter’s association, the norwegian association of hunters and anglers, and the swiss league for the protection of nature were founded in 1830, 1877, and 1909, respectively. during this same time period when forests and ungulate populations were increasing, human numbers were declining in rural areas as people migrated to industrial cities or emigrated to north america. as a consequence, numbers of livestock also declined (breitenmoser 1998). efforts to increase ungulate populations have been spectacular, and at present, many european countries have high densities. for example, in the mid-1980s, nearly 250,000 moose were harvested annually in the nordic countries, compared with about 72,000 in all of north america (haagenrud et al. 1987, kelsall 1987). most of these great increases occurred in the absence of large carnivores. today, forest damage caused by abundant ungulates is a widespread problem in europe (bergström et al. 1993, breitenmoser 1998) and this over-browsing has reduced the biodiversity of plants and invertebrates (suominen et al. 1999). during this period, steps were taken in several countries to save the remaining brown bears. in sweden, for example, official requests to remove the bounty on bears were made in 1889 and 1891, the second by a chapter of the swedish hunter’s association. parliament approved these requests in 1893 (lönnberg 1929). in 1905 the royal swedish academy of sciences issued a statement saying, “it is a matter of honor for our country that this interesting animal be protected from complete extermination.” several measures were taken to protect bears, including complete protection in national parks, which were first established in 1910. this effort was successful, and hunting was reintroduced in 1943 (swenson et al. 1995). brown bears received protection in poland in 1932 and in italy in 1939 (1992 in the abruzzo area). but, the bear received protection much later in other countries: 1955 in france, 1967 in spain, and 1972 in norway (servheen et al. 1999). efforts to save and increase bear populations in europe have been successful in many areas, and there are now about 50,000 brown bears in europe (ca. 14,000 outside of russia) with increasing and expanding, or at least stable, populations f o u n d i n n o r t h e a s t e r n e u r o p e , t h e carpathian region, the dinaric mountain range in former yugoslavia, and scandinavia (zedrosser et al. 2001). in many instances, bears are returning to countries that exterminated them because of successful conservation efforts in neighboring countries. in addition, the population has been increasing in most of russia, with the greatest increase in the european part (chestin et al. 1992). however, 8 of the 12 alces vol. 39, 2003 schwartz et al. predator management and moose 53 european populations are less than 500 bears and are decreasing. efforts to save the brown bear have concentrated on protection, and the species is protected or is a game animal in all of europe (zedrosser et al. 2001). additionally, 2 reintroductions have been attempted, and 2 populations have been augmented. the first reintroduction attempt in the world was in poland in 1938-44 and was unsuccessful. a second occurred in the central pyrenees of france in 1996-97. augmentations have occurred in austria in 1989-93 and an ongoing project in italy that started in 1999. brown bears are an important and prized big game animal in many countries. this is probably why swedish hunters worked actively for the species’ protection in the 1800s. it is also an important game species in many eastern european countries (salvatori et al. 2002). zedrosser et al. (2001) concluded that communism in eastern europe was not nearly as destructive to bear populations as the political systems in western europe, possibly because bears were managed for hunting by a few hunters, including foreign hunters with convertible currency, and because gun ownership was strictly limited. the extreme consequence of this was the situation in romania, where the dictator nicolae ceausescu allowed the bear population to increase to the highest densities in europe to provide hunting opportunities, such as shooting 24 bears in one hunt, and trophies for himself only (crisan 1994). romania is the only european country that has decided to reduce its brown bear population, from 8,000 to 6,000 by hunting, in order to reduce loss of human life and livestock losses (servheen et al. 1999, swenson et al. 2000). the situation for the wolf is quite different, even though some russian scientists began writing about the importance of the wolf’s place in nature at the end of the 19th and beginning of the 20th centuries (bibikow 1990). protection came much later: 1966 in sweden, 1973 in norway, 1976 in italy, 1993 in france, and 1995 in croatia and greece (promberger and schröder 1993, boitani 2000). it is fully protected in 11 of 27 european countries and has no protection at all in 9 countries. the total european population outside of russia is estimated to be over 18,000, but only 6 countries have more than 1,000 wolves, 11 have more than 500, and 8 have less than 50 (boitani 2000). wolves are increasing in many european countries and, like bears, are expanding into countries where they were formerly extirpated. in the soviet union, wolves were managed by zones, with extermination in intensive agricultural and reindeer (rangifer tarandus) husbandry areas, controlled at low density in areas with fewer people and agriculture, management as a hunted species in the largest zone, and complete protection in reserves (bibikow 1990). we know of no efforts to reestablish wolves in europe by reintroductions. however, in addition to the reintroductions and augmentations of bear populations, lynx have been reintroduced into many parts of europe, and more are planned (breitenmoser et al. 2000). both the brown bear and the wolf are protected and managed according to national legislation. in addition, most european countries are signatories of the bern convention, undoubtedly the most important agreement protecting large carnivores in europe. the bern convention was ratified on 19 september 1979 in bern, switzerland. its goal is to preserve wild animal species and their natural habitats. member countries must pay special attention to endangered, and potentially endangered, species listed in different appendices, each representing a different stage of endangerment. the brown bear and wolf are listed in appendix ii (strictly protected fauna species) requiring that actions must be taken predator management and moose schwartz et al. alces vol. 39, 2003 54 to protect them; forbidden are the capture or killing; wilful disturbance, and possession and trade. in addition, the recolonization of indigenous species must be promoted if doing so will enhance the likelihood of preservation. member countries can make reservations to the bern convention regarding means or methods of killing, capture, or other exploitation of listed species. seven countries have made reservations regarding protection of the brown bear (bulgaria, czech republic, finland, slovenia, slovakia, ukraine, and turkey) and 10 regarding the wolf (bulgaria, czech republic, finland, latvia, lithuania, poland, slovenia, slovakia, spain, and turkey) (boitani 2000, swenson et al. 2000). in addition, council directive 92/43/ eec, conservation of natural and wild fauna and flora (abl l 206, 22.07.1992), binds member states of the european union (eu). the main goal of the so-called florafauna-habitat directive is to secure species diversity by protection of habitats and protection of wild flora and fauna. actions must be taken by member countries to preserve all species and their habitats. the brown bear is a priority species of the eu. it is listed in appendix ii (species needing specially protected areas, except the populations in finland and sweden) and appendix iv (strictly protected species; capture, killing, and wilful disturbance not permitted). possession, transport, and trade of appendix iv species are strictly prohibited. the wolf is listed in appendix ii and i v ( i n b o t h c a s e s e x c e p t f o r s o m e populations in spain, greece, and finland). exemptions are given when it can be established that there is no negative impact to species preservation, to prevent serious damage to culture and livestock, for public health, sanitary, and safety reasons, and for scientific, restocking, and re-colonization purposes. the european parliament requests that members of the eu consider their resolutions, although they are not legally binding. those relevant to large carnivores are: (1) european parliament resolution, 24 january 1989 (a2-0377/88, ser. a), which calls for immediate steps to favor wolf conservation in all european countries, adopts the international union for conservation of nature wolf manifesto, and invites the european commission to expand and provide financial means to support wolf conservation; (2) european parliament resolution, 17 february 1989 (a2-339/88, abl c 69/201, 20.3.1989), which states that the european commission should promote or continue programs to protect the brown bear in the eu. actions for socio-economic development should be promoted in return for communities with protective measures for the brown bear. systems for bear damage prevention and damage compensation should be developed. a network of connected reserves and specially protected areas should be established (called the “natura 2000 network”); and (3) european parliament resolution, 22 april 1994 (a2-0154/94, abl c 128/427, 09.05.1994), which states that the european commission should not support and finance development that would have a negative effect on bear populations. protected areas and corridors for genetic exchange should be established to correct actions that have had negative impact on bear populations. measures to prevent the killing and capture of bears and protect bear habitat should be undertaken. financial support for damage compensation, and compensation for economic restrictions due to bear conservation, should be provided. north america wolves and grizzly bears were among the first species to be protected under the endangered species act (esa) in the united states, signed in 1973. wolves were listed alces vol. 39, 2003 schwartz et al. predator management and moose 55 as endangered in 1974 and grizzly bears as threatened in 1975. the esa is probably the most significant law in any nation designed to preserve and maintain biodiversity. the esa establishes that preservation of animal and plant species is a national priority that takes precedence over local interests in wildlife management, over economic interests, and even over certain rights of owners of private property (czech and krausman 2001). although the intent of the esa to apply to private lands is clear, the legal basis for this remains controversial and unresolved to some degree (sax 2001). distinct population segments can be listed under the esa for species like wolves and bears that are reduced in significant proportions of their former range but remain abundant elsewhere (e.g., wolves and grizzly bears are not listed in alaska, only south of canada). this national priority for species recovery mandated by the esa has worked well in the united states to recover large predators, like wolves and grizzly bears, which sometimes conflict, or are thought to conflict, with local economic or hunter interests. the esa was amended in 1982, to include section 10(j) designed to reduce landowner opposition to restoration of controversial species, like wolves and bears, to portions of their former range from which they were extirpated. this is accomplished by designating such reintroduced populations as “experimental”. populations restored as “experimental” are permitted more management flexibility on issues such as taking of nuisance individuals, permitting multiple uses of habitat, and reducing the requirement for federal review of land management and use activities (such as logging) that could adversely affect the species in experimental populations. this “experimental” provision was successfully used to restore wolves in yellowstone national park and central idaho. it was also the key to the reintroduction of red wolves (canis rufus) to the southeast and of mexican wolves (c. l. baileyei) to the southwest of the united states. so far, however, section 10(j) provisions have been inadequate to accomplish grizzly bear restoration in the wilderness areas of central idaho, which, even with this management flexibility, was opposed by key politicians in idaho. this opposition was based largely on misguided concerns over the level of physical danger grizzly bears posed to humans; a reprise of misconceptions that existed a century ago. a number of programs began in the 1990s, directed at long-term conservation and enhancement of large carnivores. much of this activity in north america was directed at areas where both wolves and bears had either been extirpated by early colonization or significantly reduced and declared threatened or endangered under the esa. wolves were reintroduced into the yellowstone ecosystem and to central idaho in 1995 and have recovered through improved management and natural dispersal in the midwest (minnesota, wisconsin, michigan). in both places, recovery goals in terms of population number have been achieved and proposals are pending to downlist and, ultimately, delist the species. thanks to the emphasis placed on the species under the esa, grizzly bear populations have also increased in the yellowstone and northern continental divide ecosystems and proposals to delist the species in and around yellowstone national park are expected as numerical population objectives have been achieved. in much of north america, there appears to be widespread recognition of the value of restoring healthy populations of predators like wolves and grizzly bears (duda et al. 1998, 2001). however, these attitudes are not universal and some states, notably idaho and alaska, retain strong sentiments against predators in favor of aggressive predator management designed to reduce predator management and moose schwartz et al. alces vol. 39, 2003 56 or eliminate depredations on livestock and the wild ungulates favored by hunting interests. these attitudes mirror the anti-predator attitudes that resulted in the near extirpation of predators in the previous century. such local opposition from political leaders has, temporarily, blocked efforts to restore grizzly bears to the bitterroot ecosystem in central idaho. the 14,800-km2 reintroduction area is designated wilderness and represents the best and largest place to restore a significant new population of grizzly bears in north america. the reintroduction proposal provided an unprecedented level of local participation and experimental designation of the restored grizzly population (fischer and roy 1998, usfws 2000, schoen and miller 2002). it received widespread national and local support from the public (duda et al. 1998, roy 2001) and from all professional wildlife management groups who commented. we believe it is likely that public support will ultimately result in grizzly bears being restored to this habitat. restoration and recovery for controversial and environmentally sensitive species like grizzly bears requires a collaborative approach to build popular support (servheen 1998). however, as the idaho bitterroot example demonstrates, even such approaches are sometimes insufficient to achieve success because of local negative attitudes about bears. changing social values wildlife management evolved originally as a means to assure continuation of hunting opportunities. frequently, attitudes of wildlife managers as well as of hunters reflect these origins and result in policies, such as predator control, that conflict with the concerns and preferences of the general public over the conservation of predators. the changes in official attitudes towards large carnivores, from policies of extermination, to those of conservation and enhancement, reflect changes in the attitudes of the general populace (duda et al. 1998, 2001). obviously, conflict will arise if the managing authorities are seen to be out-of-step with the prevailing public attitudes. one example of this was the liberalization of black bear hunting regulations, including spring hunting, hunting with hounds, and baiting, in colorado. citizens objected to these regulations based on concerns they were inhumane, but an intransigent bureaucracy failed to respond. the result was an overwhelmingly approved citizens’ initiative that set aside these regulations (beck and gill 1995, beck et al. 1995). similar initiatives passed in oregon (boulay et al. 1999) and, for mountain lions, in california. in alaska, voters approved a citizens’ initiative that overturned a regulation that allowed persons to take wolves by landing aircraft and shooting them. in british columbia, opponents of hunting grizzly bears succeeded in getting a moratorium on grizzly bear hunting implemented in 2001. the moratorium on hunting has since been lifted but it is clear that citizens threaten the continuation of grizzly bear hunting with conservation-preservationist attitudes toward bears. citizens’ initiatives are not practiced in european countries, but there is an obvious trend towards more protection of large carnivores. also, the killing of a few wolves by state employees shooting them from a helicopter in norway in winter 2000-2001 resulted in enormous european media coverage and negative public reactions from other parts of europe. some in europe are questioning the wisdom of the high densities of ungulates, which cause traffic accidents, forest damage, and reductions in forest biodiversity. these examples, and those provided earlier about opposition to predator control programs in alaska, illustrate aspects of a trend that sociologists have been observing; alces vol. 39, 2003 schwartz et al. predator management and moose 57 that utilitarian attitudes towards wildlife are declining in western cultures (decker et al. 1992, 2001). a recent metaanalysis (williams et al. 2002) of 38 studies about people’s attitudes towards wolves showed that people were generally positive to wolves (61%), that age, residence, occupation, education, and income influenced one’s attitude, and that about one-fourth of the people are neutral. the analysis also found that wolves are less popular in europe than in north america, as is also suggested by our review, based on levels and dates of protection. in europe, this may be influenced by a historical fear of wolves as a carrier of rabies that dates back to the middle ages (bibikow 1990), or a fear of the wolf as a potential killer of humans, which has more historical support from europe than north america (linnell et al. 2002). also, attitudes towards wolves vary within europe (boitani 1995), even in adjacent areas, as norwegians have more negative attitudes towards wolves than swedes (bjerke et al. 2001). wolf biologists working in scandinavia have reported the impression of an increased support for wolves, although they could not substantiate it (wabakken et al. 2001). however, the metaanalysis, which covered studies from the period 1972 – 2000, did not find any trend in support for wolves. williams et al. (2002) predicted increasing support for wolves over time due to increasing education and urbanization, but stressed that positive attitudes towards wolves in the general public are often weak and have the potential to shift rapidly if linked to other stronger attitudes. summary human attitudes toward large carnivores have been shaped by centuries of coexistence. these attitudes have changed markedly as human civilizations matured and industrialized. early peoples lacked the appropriate weapons to effectively control large predators. these cultures adapted ways to live with carnivores. as man acquired modern weapons, large predators and native ungulates were exterminated. we witnessed this first in europe where modern civilization first developed. however a second wave of extermination followed with the european colonization of north america. old world values were transported to the new world. in north america, attitudes of predator control were also adopted as part of the evolving profession of wildlife management, where large carnivores were still abundant in northern environments. carnivores were perceived as competitive and a threat to ungulates harvested by hunters. through history, human values toward large carnivores seem to be inversely proportional to carnivore abundance. society tends to value them more when they become rare or endangered. because predators were largely eliminated or reduced to remnant population in europe, social attitudes away from extermination and towards protection evolved more quickly. these attitudinal shifts are reflected in the recolonization of carnivores back into historic habitats. many citizens and scientists view this reoccupation as a valuable contribution to society and ecological processes. a similar transition has come more slowly to north america and even today there is an apparent inability by some members of the public and some wildlife managers to engender the changing social dimension that values predators in a broader context of ecosystem function, rather than an impediment to ungulate management for the hunting public. agencies in north america responsible for the management of ungulates and carnivores are currently faced with conflicting values and differing demands. if we can use information about the wolf as a guide, it predator management and moose schwartz et al. alces vol. 39, 2003 58 is possible to make some general conclusions. on the one hand, old values prevail with a continued emphasis on predator control. however, there appears to be a gradual shift toward carnivore conservation, especially among the more highly educated, urban, youth (williams et al. 2002). extrapolation of these results into the future suggest that there will likely be a gradual shift away from negative attitudes toward more positive attitudes as the older population is replaced (williams et al. 2002). game management agencies will need to shift toward a more modern construct that recognizes the intrinsic value of wild ecosystems and the wildlife they contain, including large predators. today’s biologists and moose managers face a difficult challenge of balancing biological principles with a diverse array of social and economic values often in conflict with principles of optimum or maximum sustained yield harvest. acknowledgements we thank k. west and r. jachowski for reviews of the manuscript. h. reynolds for providing literature detailing the alaska experience, and organizers of the norway conference for kindly inviting the senior author to present this topic. references bailey, v. 1931. mammals of new mexico. north american fauna 53. u.s. department of agriculture, biological survey. washington, d.c., usa. ballard, w. b. 1992. bear predation on moose: a review of recent north american studies and their management implications. alces supplement 1:1-15. , and d. g. larsen. 1987. implications of predator-prey relationships to moose management. swedish wildlife research supplement 1:581-602. , and s. d. miller. 1990. effects of reducing brown bear density on moose calf survival in south-central alaska. alces 26:9-13. , t. h. spraker, and k. p. t aylor. 1981. causes of neonatal moose calf mortality in south central alaska. journal of wildlife management 45:335342. , and v. van ballenberghe. 1998. predator/prey relationships. pages 247272 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. banci, v. 1991. the status of the grizzly bear in canada in 1990. committee on the status of endangered wildlife in canada, status report. ottawa, canada. , d. a. de m a r c h i, and w. r. archibald. 1994. evaluation of the population status of grizzly bears in canada. international conference on bear research and management 9:129142. beck, t. d. i., and r. b. gill. 1995. colorado status report. pages 9-12 in j. auger and h. l. black, editors. proceedings of the fifth western black bear workshop, human-black bear interactions, provo, utah, usa, 22-25 february 1994. , d. s. moody, d. b. koch, j. j. beecham, g. r. olson, and t. burton. 1995. sociological and ethical considerations of black bear hunting, a panel discussion. pages 119-133 in j. auger and h. black, editors. proceedings of the fifth western black bear workshop, human-bear interactions, provo, utah, usa, 22-25 february 1994. bergström, r., h. huldt, and u. nilsson. 1993. sverige, jakten och eg. svenska jägareförbundet, almqvist & wiksell tryckeri, uppsala, sweden. (in swedish). bibikow, d. i. 1990. der wolf. a. ziemsen alces vol. 39, 2003 schwartz et al. predator management and moose 59 verlag, wittenberg lutherstadt, ddr. (in german). bjerke, t., b. p. kaltenborn, and c. thrane. 2001. sociodemographic correlates of fear-related attitudes toward the wolf (canis lupus); a survey in s o u t h e a s t e r n n o r w a y . f a u n a norvegica 21:25-33. black, l. t. 1998. bear in human imagination and in ritual. ursus 10:343-347. boertje, r. r., w. c. gasaway, d. v. grangaard, and d. g. kellyhouse. 1988. predation on moose and caribou by radio collared grizzly bears in eastcentral alaska. canadian journal of zoology 66:2492-2499. boitani, l. 1995. ecological and cultural diversities in the evolution of wolf-human relationships. pages 3-11 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, occasional publication no. 35, university of alberta, edmonton, alberta, canada. . 2000. action plan for the conservation of wolves (canis lupus) in europe. nature and environment, no. 113. council of europe publishing, strasbourg, france. boulay, m. c., d. h. jackson, and d. a. immell. 1999. preliminary assessment of a ballot initiative banning two methods of bear hunting in oregon: effects on bear harvest. ursus 11:179-184. boutin, s. 1992. predation and moose population dynamics: a critique. journal of wildlife management 56:116127. breitenmoser, u . 1998. large predators in the alps: the fall and rise of man’s competitors. biological conservation 83:279-289. , c. breitenmoser-würsten, h. okarma, t. kaphegyi, u. kaphygyiwallmann, and u. m. müller. 2000. action plan for the conservation of the eurasian lynx (lynx lynx) in europe. nature and environment, no. 112. council of europe publishing, strasbourg, france. carbyn, l. n. 1987. gray wolf and red wolf. pages 378-393 in m. novak, j. a. baker, m. e. obbard, and b. malloch, editors. wild furbearer management and conservation in north america. ontario trappers association, toronto, ontario, canada. carlos, a. m., and f. d. lewis. 1995. strategic pricing on the fur trade: the hudson’s bay company, 1700-1763. pages 61-88 in t. l. anderson and p. j. hill, editors. wildlife in the marketplace. rowman and littlefield publishers, incorporated, lanham, maryland, usa. chestin, i. e., y. p. gubar, v. e. sokolov, and v. s. lobachev. 1992. the brown bear (ursus arctos l.) in the ussr: numbers, hunting and systematics. annales zoologici fennici 29:57-68. corbet, g. b., and s. harris, editors. 1991. the handbook of british mammals. third edition. blackwell scientific publications, oxford, u.k. crisan, v. 1994. jäger? schlächter ceausescu. verlag dieter hoffmann, mainz-ebersheim, germany. (in german). csányi, s. 1997. challenges of wildlife management in a transforming society: examples from hungary. wildlife society bulletin 25:33-37. czech, b., and p. r. krausman. 2001. the endangered species act. history, conservation biology, and public policy. john hopkins university press, baltimore, maryland, usa. decker, d. j., t. l. brown, n. a. connelly, j. w. enck, g. a. pomerantz, k. g. purdy, and w. f. siemer. 1992. toward a comprehensive paradigm of predator management and moose schwartz et al. alces vol. 39, 2003 60 wildlife management: integrating the human and biological dimensions. pages 33-54 in w. r. mangun, editor. american fish and wildlife policy: the human dimension. southern illinois university press, carbondale, illinois, usa. , , and w. f. siemer, editors. 2001. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. deloria, v., jr. 2001 american indians and the wilderness. pages 25-35 in t. kerasote, editor. return of the wild, the future of our natural lands. island press, washington, d.c., usa. duda, m., s. j. bissell, and k. c. young. 1998. wildlife and the american mind: public opinion on and attitudes toward fish and wildlife management. responsive management, harrisonburg, virginia, usa. , p. e. de michele, s. j. bissell, p. wang, j. he r r i c k, a. lanier, w. testerman, j. youder, and c. zurawski. 2001. public attitudes toward grizzly bear management in wyoming. appendix 1 in l. kruckenberg, editor. special report – draft wyoming grizzly bear management plan. wyoming game and fish department, cheyenne, wyoming, usa. dunlap, t. 1988. saving america’s wildlife. princeton university press, princeton, new jersey, usa. edsman, c.-m. 1994. jägaren och makterna, samiska och finska björnceremonier. dialektoch folminnesarkivet, almqvist & wiksell tryckeri, uppsala, sweden. (in swedish). elgmork, k. 1996. historic review of brown bears and wolves in centralsouth norway 1733-1845. fauna 49:134-147. (in norwegian with english summary). fischer, h., and m. roy. 1998. new approaches to citizen participation in endangered species management: recovery in the bitterroot ecosystem. ursus 10:603-606. franzmann, a. w., and c. c. schwartz. 1986. black bear predation on moose calves in highly productive versus marginal moose habitats on the kenai peninsula. alces 22:139-153. , , and r. o. peterson. 1980. moose calf mortality in summer on the kenai peninsula, alaska. journal of wildlife management 44:764768. gasaway, w. c., r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. haagenrud, h., k. morow, k. nygrén, and f. stålfelt. 1987. management of moose in nordic countries. swedish wildlife research supplement 1:635642. harry, j., r. gale, and j. hendee. 1969. conservation: an upper-middle class social movement. journal of leisure research 1:246-254. hayes, r. d., and j. r. gunson. 1995. status and management of wolves in canada. pages 21-34 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a c h a n g i n g w o r l d . c a n a d i a n circumpolar institute, occasional publication no. 35, university of alberta, edmonton, alberta, canada. jessen, k. 1929. the bear (ursus arctos l.) in denmark. meddelelser fra dansk geologisk forening 7:273-286. (in danish with english summary). kellert, s. r. 1996. the value of life: biological diversity and human society. island press, washington, d.c., usa. kelsall, j. p. 1987. the distribution and alces vol. 39, 2003 schwartz et al. predator management and moose 61 status of moose (alces alces) in north america. swedish wildlife research supplement 1:1-10. leopold, a. 1933. game management. university of wisconsin press, madison, wisconsin, usa. l i n n e l l , j . d . c . , r . a n d e r s e n , ž. an d e r s o n e, l. balc iauskas , j. c. blanco, l. boitani, s. braninerd, u. breitenmoser, i. kojola, o. liberg, j. løe, h. okarma, h. c. pedersen, c. promberger, h. sand, e. j. solberg, h. valdmann, and p. wabakken. 2002. the fear of wolves: a review of wolf attacks on humans. oppdragsmelding 731, norwegian institute for nature research, trondheim, norway. lönnberg, e. 1929. björnen i sverige 1856-1928. almqvist & wiksell tryckeri, uppsala, sweden. (in swedish). lopez, b. h. 1978. of wolves and men. charles scribner’s and sons, new york, new york, usa. lueck, d. l. 1995. the economic organization of wildlife institutions. pages 124 in t. l. anderson and p. j. hill, editors. wildlife in the marketplace. rowman and littlefield publishers, incorporated, lanham, maryland, usa. macey, a. 1979. the status of the grizzly bear in canada. national museum of natural science, ottawa, ontario, canada. mallinson, j. 1978. the shadow of extinction: europe’s threatened wild mammals. macmillan, london, u.k. mattson, d. j., r. g. wright, k. c. kendall, and c. j. martinka. 1995. grizzly bears. pages 103–105 in e. t. laroe, g. s. farris, c. e. puckett, p. d. doran, and m. j. mac, editors. our living resources: a report to the nation on the distribution, abundance, and health of u.s. plants, animals, and ecosystems. u.s. department of the interior, national biological service, washington, d.c., usa. mclellan, b. n., and v. banci. 1999. status and management of the brown bear in canada. pages 46-50 in c. servheen, s. herrero, and b. peyton, compilers. bears: status survey and conservation action plan. iucn/ssc bear and polar bear specialist groups. iucn, gland, switzerland, and cambridge, u.k. miller, b., b. dugelby, d. fopreman, c. martinez del rio, r. noss, m. phillips, r. reading, m. e. soulé, j. terborgh, and l. willcox. 2001. the importance of large carnivores to healthy ecosystems. endangered species updates 18:202-210. miller, s. d. 1997. impacts of heavy hunting pressure on the density and demographics of brown bear populations in southcentral alaska. federal aid in wildlife restoration, research final report, grants w-24-2, w-24-3, w-244, study 4.26, alaska department of fish and game, juneau, alaska, usa. , and w. b. ballard. 1992. analysis of an effort to increase moose calf survivorship by increased hunting of brown bears in south-central alaska. wildlife society bulletin 20:445-454. miller, s. m., s. d. miller, and d. w. mccollum. 1998. attitudes toward and relative value of alaskan brown and black bears to resident voters, resident hunters, and nonresident hunters. ursus 10:357-376. myrberget, s. 1990. wildlife management in europe outside the soviet union. utredning 18, norwegian institute for nature research, trondheim, norway. nash, r. 1982. wilderness and the american mind. yale university press, new haven, connecticut, usa. (nrc) national research council. 1997. wolves, bears, and their prey in predator management and moose schwartz et al. alces vol. 39, 2003 62 alaska: biological and social challenges in wildlife management. national academy press, washington, d.c., usa. peek, j. m. 1986. a review of wildlife management. prentice-hall, englewood cliffs, new jersey, usa. promberger, c., and w. schröder. 1993. wolves in europe, status and perspectives. munich wildlife society, ettal, germany. rockwell, d. 1991. giving voice to the bear, north american indian myths, rituals and images of the bear. roberts rinehart publishers, niwot, colorado, usa. roy, j. 2001. bitterroot noi to reevaluate rod: summary of all comments received 6/21/2001-8/21/2001. tallied by state and whether they agree or disagree with selection of no action alternative. unpublished u.s. fish and wildlife service document dated 17 september 2001. missoula, montana, usa. salvatori, v., h. okarma, o. ionescu, y. dovhanych, s. finïo, and l. boitani. 2002. hunting legislation in the carpathian mountains: implications for the conservation and management of large carnivores. wildlife biology 8:310. sax, j. l. 2001. legal and policy challenges of environmental restoration. pages 109-127 in v.a. sharpe, b. norton, and s. donnelley, editors. wolves and human communities: biology, politics, and ethics. island press, washington, d.c., usa. schoen, j., and s. d. miller. 2002. new strategies for bear conservation: collaboration between resource agencies and environmental organizations. ursus 13:361-367. servheen, c. 1998. the grizzly bear recovery program: current status and future considerations. ursus 10:591-596. . 1999. status and management of the grizzly bear in the lower 48 united states. pages 50-54 in c. servheen, s. herrero, and b. peyton, compilers. bears, status survey and conservation action plan. iucn/ssc bear and polar bear specialist groups, gland, switzerland, and cambridge, u.k.. , s. herrero, and b. peyton, compilers. 1999. bears, status survey and conservation action plan. iucn/ssc bear and polar bear specialist groups, iucn, gland, switzerland, and cambridge, u.k. shepard, p. 1996. traces of an omnivore. island press, washington, d.c., usa. , and b. sanders. 1992. the sacred paw: the bear in nature, myth and literature. penguin, new york, new york, usa. sherwood, m. 1981. big game in alaska: a history of wildlife and people. yale university press, new haven, connecticut, usa. søilen, e. 1995. sportsmenn i veideland. norges jegerog fiskerforbund, asker og bærums budstikke, billingstad, norway. (in norwegian). stephenson, r. o., w. b. ballard, c. a. smith, and k. richardson. 1995. wolf biology and management in alaska, 1981-1992. pages 43-54 in l. n. carbyn, s. h. fritts, and d. r. seip, editors. ecology and conservation of wolves in a changing world. canadian circumpolar institute, occasional publication no. 35, university of alberta, edmonton, alberta, canada. stewart, r. r., e. h. kowal, r. beaulieu, and t. w. rock. 1985. the impact of black bear removal on moose calf survival in east-central saskatchewan. alces 21:403-418. suominen, o., k. danell, and r. bergström. 1999. moose, trees, and ground-living invertebrates: indirect interactions in alces vol. 39, 2003 schwartz et al. predator management and moose 63 swedish pine forests. oikos 84:215226. swenson, j. e., b. dahle, and f. sandegren. 2001. brown bear predation on moose. fagrapport 48, norwegian institute of nature research, trondheim, norway. (in norwegian with english summary). , n. gerstl, b. dahle, and a. zedrosser. 2000. action plan for the conservation of the brown bear in europe (ursus arctos). council of europe, nature and environment, no. 114. , f . s a n d e g r e n , m . h e i m , s . b r u n b e r g , o . j . s ø r e n s e n , a . söderberg, a. bjärvall, r. franzén, s . w i k a n , p. w a b a k k e n, and k . overskaug. 1996. is the scandinavian brown bear dangerous? oppdragsmelding 404, norwegian institute of nature research, trondheim, norway. (in norwegian with english summary). , p. wabakken, f. sandegren, a. b j ä r v a l l , r . f r a n z é n , a n d a . söderberg. 1995. the near extinction and recovery of brown bears in scandinavia in relation to the bear management policies of norway and sweden. wildlife biology 1:11-25. (usfws) u.s. fish and wildlife service. 1993. grizzly bear recovery plan. u.s. fish and wildlife service, missoula, montana, usa. . 2000. grizzly bear recovery in the bitterroot ecosystem. final environmental impact statement, u. s. department of the interior fish and wildlife service, washington, d.c., usa. van ballenberghe, v. 1987. effects of predation on moose numbers: a review of recent north american studies. swedish wildlife research supplement 1:431-460. van daele, l. 2003. the history of bears on the kodiak archipelago. alaska natural history association, anchorage, alaska, usa. wabakken, p., h. sand, o. liberg, and a. bjärvall. 2001. the recovery, distribution, and population dynamics of wolves on the scandinavian peninsula, 1978-1998. canadian journal of zoology 79:710-725. williams, c. k., g. ericsson, and t. a. heberlein. 2002. a quantitative summary of attitudes toward wolves and their reintroduction (1972-2000). wildlife society bulletin 29:1253-1259. woodroffe, r. 2000. predators and people: using human densities to interpret declines of large carnivores. animal conservation 3:165-173. young, s. 1946. the wolf in north american history. caxton printers, caldwell, idaho, usa. zedrosser, a., b. dahle, j. e. swenson, and n. gerstl. 2001. status and management of the brown bear in europe. ursus 12:9-20. f:\alces\vol_38\pagemaker\3816. alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 123 evidence of carrying capacity effects in newfoundland moose w. e. mercer 1 and b. e. mclaren 2 18 virginia place, st. john’s, nl, canada a1a 3g6; 2government of newfoundland & labrador, department of forest resources & agrifoods, p.o box 2222, gander, nl, canada a1v 2n9 abstract: newfoundland moose (alces alces americana) increased following 1904, the year of successful introduction, to peak numbers in 1958. the population subsequently decreased to record low numbers by 1973, when an area-quota management system was instituted throughout the island (112,000 km2) in 38 moose management areas, in part, to respond to issues related to habitat and accessibility for hunting. subsequent quota-management manipulations permitted the islandwide population to increase in accessible areas to record high numbers by 1986, after which populations again decreased, to a 1999 estimate of 125,000 animals (post-hunt). we hypothesise that, unlike most studied irruptions of cervid populations, moose populations in newfoundland, and subsequently habitat carrying capacity (k), decreased on inaccessible range following 1958 to very low density, from which both have never recovered. decreases in relative numbers of young moose seen while hunting and during winter classifications are consistent with increases in the number of moose seen during increase phases during 1966–99. these observations are less obvious for less accessible management areas. we explore other recruitment and density relationships as they have been developed in association with our estimate of k in moose for newfoundland. we illustrate that, although some decrease in moose numbers following 1958 and 1986 was the result of management, changes to population size and to k also resulted in reduction in productivity, such that density dependence explains > 10% and up to 76% of hunter-observed recruitment. alces vol. 38 : 123-141 (2002) key words: accessibility, alces alces, carrying capacity, hunter reports, moose, newfoundland, population dynamics, quota management, recruitment moose (alces alces) populations are often difficult to compare because of geographic differences in scale, habitat continuity, and the relative importance of various mortality factors, including hunting, which introduces issues of hunter accessibility and its effect on population dynamics (van ballenberghe and ballard 1998). fundamental questions remain about what limits or regulates moose populations as there is much geographic variation in the relative effects of predators, human hunting, and primary production (gasaway et al. 1992, crête and courtois 1997, saether 1997, crête and daigle 1999). habitat carrying capacity (k) is a concept that assists managers in area comparisons and in resolving when density-dependent effects are predicted to occur (mccullough 1999, person et al. 2001). approximation of k for cervids has been discussed in theory by caughley (1976), and as an experiment by mccullough (1979). crête (1989: 378) described k as a bounded rather than a constant value, varying with effects of winter snowfall, annual primary productivity, and forest succession. the usefulness of mccullough’s approach to calculate k in areas where habitat changes is a problem also discussed generally in the original text (mccullough 1979: 156). in this paper, we attempt to estimate k using hunter reports of several local populations of moose. we compare areas where moose management has had more carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 124 fig. 1. selected management areas in newfoundland in which trends observed by either-sex, resident moose hunters are compared. shaded region is referred to as central newfoundland in this paper. gros morne national park (gmnp), terra nova national park (tnnp), and other areas referred to in the text are indicated. and less success in newfoundland over the past 4 decades. we hypothesise that, unlike most studied cervid populations (mccullough 1997), moose populations in newfoundland, and subsequently k, decreased on inaccessible range to very low density following initial irruptions, from which both have never recovered. we also predict that with an approach of more accessible, hunted populations to our estimates of k for these areas, lower numbers of young moose would be seen during hunting trips. these predictions allow us to illustrate density dependence in newfoundland moose based on hunter reports during 1966–99. the relative importance of density-dependent versus density-independent factors for fluctuations in ungulate populations in the absence of predation can be tested by the following specific questions: (1) are indices related to moose abundance (i.e., moose observed by hunters), or to hunting (i.e., total days of hunting and total licences issued) more important in explaining recruitment observed by hunters in autumn; (2) do all management areas experience declines following observed peaks in moose density; and (3) are observations of abundance related to observations of recruitment by hunters? our contention that variation in densitydependent reproduction depends on variation in k in time and space (crête 1989) is contrasted to claims that regulation of moose density is dependent on hunter functional response, and that habitat or food supply may influence only the synchronicity of p o p u l a t i o n c y c l e s i n n e w f o u n d l a n d (ferguson and messier 1996). we show limitations in hunter functional response to moose density in newfoundland, and we offer guidelines that may assist moose management in the future. population and management history the island of newfoundland (fig. 1), which encompasses 112,000 km2, forms a test case of the question of population regulation, because of the absence of predators of adult moose other than hunters. four adult moose (a. a. americana), 2 females and 2 males, were successfully introduced to newfoundland from new brunswick, canada, in 1904 (pimlott 1953, broders et al. 1999). rapid dispersal and low densities characterized the first 25 years of population increase (pimlott 1953) and wolves (canis lupus) were extirpated during that period (mercer 1995). during 1953–56, an increase rate was estimated for insular newfoundland at 0.33, based on observations of young moose in mid-winter surveys (pimlott 1959a). keith (1983) later calculated an average intrinsic rate of increase of r = 0.23 for north america in situations of carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 126 change may have caused the decrease in licence sales from 8,660 that year to 6,523 in 1953 (pimlott 1959b). later, minor changes to season length did not appear to affect licence sales (mercer 1974). gradually, hunting seasons have increased in length, beginning in september or october and ending between december and february; in 2002 the moose season was 15 weeks long, including a two-week, bowhunting season. in 2002, 27,820 moose licences were available in insular newfoundland (mercer 1995). methods we define k as the maximum density of moose that can be supported at equilibrium, in a stable environment and in the absence of time lags, as mccullough (1979) defined kcc. more generally, our definition agrees with odum (1953), who defined k as the upper asymptote of the logistic or sigmoid curve describing unimpaired population growth. the inflection point of the sshaped curve describing such growth has been used to determine maximum sustained yield (msy), also termed inflection point carrying capacity (icc) by mccullough (1979). msy has been shown to occur near 0.6 k for ungulates (mccullough 1984, person et al. 2001). much of the information for determining changes to the issue of moose licences (“quotas”) in newfoundland remains the same as that reported in mercer and manuel (1974). annual response to a questionnaire (“return”) attached to every moose licence is usually > 75%, but only following 1–2 reminders mailed to nonrespondents in february or march after the season closes. initial returns, including a completed questionnaire and the lower mandible of any moose taken on a licence (individual licences have always been for 1 moose), are about 50% of licence sales to resident hunters without a reminder. the mailed reminders are the extent of enforcement, although wildlife regulations stipulate returns are mandatory within 7 days of a kill or by the end of the hunting season. data from completed questionnaires are coded digitally and archived by june of each year, and records in this form date back to the 1966 hunting season. hunter trends are calculated annually from the resulting time series up to the hunting season of the previous year. on this timetable, the most recent hunter information can help set quotas only for the following hunting season, because draw notices (indicating successful licence applicants for the next season) are mailed by june each year. among those questions answered by moose hunters (including co-licence holders in a party hunt), the most reliable information has been considered for the calculation of trends by moose management area, as follows: (1) average number of animals reported killed by licence type (either-sex or male-only hunters) as a percentage of all licences sold by type (“hunter success”); (2) average number of days spent hunting by all licence holders spending at least 1 and at most 24 days hunting by licence type (“days of hunting”); and (3) average number of moose seen by the licence holder divided by the average number of days of hunting reported (“moose seen / day of hunting”). in instances where management areas were subdivided only for a portion of our study period, 1966–99, we pooled data from subareas for our trend calculations. we adjusted hunter success from the calculation using initial questionnaire respondents by assuming success of all nonrespondent hunters was represented by reports from reminder respondents. hunter respondents given first reminders only reported no difference in success from second-, third-, or fourth-reminder respondents, although all nonrespondents reported lower success than initial respondents without reminders (wildalces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 127 life division, newfoundland and labrador, unpublished data). we did not exclude any either-sex, resident licence holders in the calculation of days of hunting and moose seen / day of hunting. we reported, in addition to hunter success, a fourth trend; kill rate; expressed as total estimated moose killed / 10 days of hunting, a factor of the division of hunter success by days of hunting. we tracked licence issue and kill estimates from 1945, the year of the first general season for all newfoundland excluding the avalon and burin peninsulas, to 1999, the last year of records available to us (pimlott 1953, mercer and manuel 1974, mercer 1995). from these figures, we calculated hunter success for either-sex licence holders. additional information on questionnaire returns has been collected and archived but not regularly used in management, including information regarding co-licence holders, area access, age, and sex of moose seen during days of hunting, number of moose seen by calendar day, and date and location of moose kill. herein, we calculate a fifth trend, using age and sex classification of moose reported seen by hunters as an index of recruitment in autumn. for all hunters reporting “calves seen,” we calculate the number of young as a ratio of the number of “cows seen” (young seen / 100 adult females). we excluded, in this instance, hunters who did not report seeing “calves,” because there exists considerable bias in the sighting and identification of young moose, as discussed by pimlott (1959a) and mercer (1974). we show trends for the either-sex, resident moose hunter in insular newfoundland (averages for the island portion of the province of newfoundland and labrador), in central newfoundland (averages for management areas 16, 17, 22, and 24), and in management area 26 on the south coast (fig. 1). a network of forest access roads makes areas 16, 17, 24, and 22 the more accessible hunting areas studied, with mean geographic distances to the nearest road in each of these areas 2.4, 1.1, 1.0, and 0.7 km, respectively (mercer 1995). area 26 is among the least accessible hunting areas in newfoundland, with a mean distance to the nearest road of 9.6 km. thus, area 26 offers a likely example of a moose population at k (mercer and manuel 1974). to correspond our estimates of moose seen / day of hunting from hunter reports to moose density, and to calibrate hunter estimates of recruitment in autumn with midwinter, aerial observations, we considered area 24, in which sufficient aerial surveys had been conducted to form a time series similar to hunter trends. we used 8 estimates of population size, between 1973 and 1997, obtained from helicopter counts in winter with stratified random surveys, in which we adjusted all counts by a factor of 2.7 to correct visibility bias, an average correction factor for forested areas in newfoundland (oosenbrug and ferguson 1992, gosse et al. 2002). counts were divided over the entire survey area (including unforested regions) of area 24 to obtain average density in animals per square kilometre. we reported all linear relationships between indices with adjusted r2. stepwise multiple regressions, from which we reported mallow’s statistic (cp), were used to determine significant predictors of recruitment observed by groups of hunters, from insular and central newfoundland, and from management areas 16, 17, 22, 24, and 26. in each stepwise procedure, we allowed variables to enter if they were significant at p < 0.10, and to stay in the model if they were significant at p < 0.05. we included among those variables, for the corresponding hunting area, moose seen / day of hunting for the same year and for the previous year, young / 100 adult females alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 131 table 1. mean values for moose seen / day of hunting (m), hunter reports of autumn recruitment in young seen / 100 adult females (r), and hunter success (s, %), during 1984–89 and 1966–99 (excluding 1984–89). theoretical habitat carrying capacity, k, is derived from the longer period, by extrapolating hunter reports, where possible, to an autumn recruitment rate of 20 young seen / 100 adult females using linear regression (see fig. 7). k is reported in moose seen / day of hunting, ± 95% confidence intervals (from linear regression), and as density ( / km2) using the correspondence from fig. 8. we also report slope and adjusted r2 for the regressions, where p < 0.05. 1984–89 1966–99 k hunting area m r s m r s slope r2 p moose seen ( / km2) insular nf 1.10 49.4 85.2 0.69 56.4 68.0 -30.4 0.80 0.000 1.88 ± 0.51 7.5 central nf 1.23 45.7 83.0 0.60 57.3 61.7 -25.1 0.42 0.000 2.00 ± 0.61 8.0 area 16 1.47 46.1 95.7 0.48 57.8 72.6 -30.9 0.17 0.016 1.60 ± 0.48 6.4 area 17 1.85 42.6 95.2 0.76 58.1 72.1 -26.0 0.63 0.000 2.24 ± 0.64 9.0 area 22 1.59 47.2 93.0 0.66 59.4 79.1 -39.1 0.26 0.007 1.66 ± 0.44 6.6 area 24 0.83 49.1 91.7 0.56 57.9 73.9 -28.1 0.31 0.001 1.94 ± 0.59 7.8 area 26 1.32 35.9 89.5 0.98 28.7 75.9 — — 0.131 — — reports were 14–36% higher, ranging from 56–59 seen / 100 adult females. during 1984–89, the relationship between young seen / 100 adult females and moose seen / day of hunting formed a consistently shallower slope than the rest of the series, making the graphed relationship for the entire study period appear slightly curvilinear (fig. 7). hunter reports from area 26 suggest that there was a lower threshold in young moose reported in autumn of about 20 seen / 100 adult females. when we extrapolated the relationship excluding the peak period in other areas to estimate potential moose seen / day of hunting at 20 young seen / 100 adult females (i.e., at a theoretical zero population increase, or k), based on area 26, all areas produced similar estimates of 1.6–2.2 moose seen / day of hunting (table 1). consistently in all areas, according to stepwise regression, moose seen / day of hunting and the autocorrelated series lagged 1 year (young seen / 100 adult females in previous years), explained the most variance in young seen / 100 adult females. the best model fit occurred in areas 17 and 22, and in insular newfoundland as an average (table 2). in no areas were indices related to hunting (i.e., total days of hunting and total licences issued) significant in explaining recruitment in autumn according to stepwise regression. the principal component regressions also showed our models to be significant in all cases except in area 22, and the first principal component, explaining nearly 100% of the covariance in the 2 series, indicated that moose seen / day of hunting was negatively associated with the autocorrelated series, as in fig. 7, and explained a similar amount of variance as the linear regressions, as in table 1, r2 = 0.11 to 0.76. the match resulting from comparing hunter trends to aerial surveys in area 24 appeared to be approximately 4x the number of moose seen / day of hunting to arrive at carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 132 fig. 7. young seen / 100 adult females plotted against moose seen / day of hunting by either-sex, resident moose hunters in insular newfoundland, central newfoundland, and area 26, 1966–99 (solid circles). a linear regression is fit through the period, excluding observations from 1984–89 (open circles), and is extended toward the x-axis to make predictions about habitat carrying capacity, k (see table 1). fig. 8. correspondence between moose seen / day of hunting by either-sex, resident moose hunters (left axis, solid circles) and moose density ( / km2 ) estimated from mid-winter, aerial surveys (right axis, open circles) in area 24, (a) as a time series and (b) as a regression for corresponding years. the 90% confidence interval for moose density is a vertical line over the 1997 aerial survey estimate in (a), used to calibrate the series. survey result in 1997 weighed heavier than the estimates from the 1980s in our comparison. expressed in both instances as young / 100 adult females, the hunter observations of recruitment in autumn and aerial surveys of mid-winter recruitment in area 24 produced a near match, although the relationship is also weak statistically, r2 = 0.03 (fig. 9). matching our predictions for moose seen / day of hunting at 20 young seen / 100 adult females with the correspondence to aerial survey data in area 24, we suggest that densities of 6–9 moose / density estimates in moose / km2 in this area, although the relationship between the 2 indices was weak, r2 = 0.08 (fig. 8). the 4 survey results from the 1970s and the alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 133 table 2. results of principal component regressions of autumn recruitment in young seen / 100 adult females in newfoundland moose populations by hunting area, using either-sex, resident hunter reports of “calves” and “cows” seen during their trips, 1966–99. we report eigenvalues and factor loadings for 2 principal components, pc 1 and pc 2 , of the covariance matrix between moose seen / day of hunting (m) and the autocorrelated series of young seen / 100 adult females in the previous year (r t–1 ). we report from the stepwise regression procedures producing m and r t–1 as the only significant predictive variables: mallow’s statistic (cp); we report from the regressions using pc 1 to predict autumn recruitment: f, p, adjusted r2, sum of squares (ss), and error sum of squares (sse). pc 1 pc 2 factor factor hunting eigenloadings eigenloadings statistics from regression models area values m r t–1 values m r t–1 cp f p r2 ss sse insular nf 67.59 -0.03 1.00 0.02 1.00 0.03 2.9 102.5 0.000 0.76 1,600.7 484.0 central nf 95.21 -0.02 1.00 0.05 1.00 0.02 2.1 39.6 0.000 0.55 1,587.6 40.1 area 16 130.41 -0.02 1.00 0.13 1.00 0.02 1.1 4.8 0.036 0.11 557.5 3,612.8 area 17 167.57 -0.02 1.00 0.10 1.00 0.02 3.5 17.3 0.000 0.34 1,815.2 3,244.9 area 22 233.40 -0.02 1.00 0.14 1.00 0.02 2.2 3.7 0.065 0.07 785.2 6,626.7 area 24 108.79 -0.01 1.00 0.04 1.00 0.01 1.8 16.9 0.000 0.33 1,123.9 2,058.7 area 26 280.21 -0.02 1.00 0.19 1.00 0.02 1.0 4.2 0.052 0.11 1,027.1 6,149.2 km2 approach k for moose in newfoundland (table 1). discussion habitat carrying capacity for moose mccullough (1979) extended the linear regression of rate of recruitment on posthunt population size to a theoretical zero population increase to obtain an estimate of k for white-tailed deer ( odocoileus virginianus) in the george reserve, michigan, usa. although mccullough’s (1979) regression suggested linearity, our data indicate curvilinearity in the relationship between recruitment observed in autumn and population size (fig. 7). linearity appears to be a reasonable interpretation of recruitment in autumn for newfoundland moose during the increase phase (1973–83), until peak densities were achieved. this difference from mccullough’s (1979) interpretation may be explained by the possibility that the george reserve deer herd, which peaked at 34 deer / km2 during 1952–71, before a hunting experiment, was not allowed to reach densities high enough to show a decline to k. post-hunt peak density was 19 deer / km2 and mean density was much lower. in 1935, estimated density for the george reserve deer herd reached 48 deer / km2. deer population densities in other areas with no hunting have been higher yet. for example, in saratoga national park, new york, usa, without hunting, densities ranged between 37–74 deer / km2 and averaged 53 deer / km2 during 1985–94 (underwood and porter 1997). this density may be similar to densities observed in other parks where hunting is prohibited or restricted. mccullough (1997) reported very high densities in a population of blacktailed deer (o. hemionus columbianus) on angel island (2.2 km2), california, usa. there, during a period of 20 years (1965– 84), 5 peak populations were recorded to average more than 100 deer / km2. these densities did not appear to be declining over time. carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 134 fig. 9. correspondence between young seen / 100 adult females by either-sex, resident hunters (open circles) and by observers during mid-winter, aerial moose surveys (solid circles) in area 24, (a) as a time series and (b) as a regression for corresponding years. we estimate that k in newfoundland boreal forests approximates 6–9 moose / km2, calculated from theoretical hunter reports of 1.6–2.2 moose seen / day of hunting at an extrapolated hunter-observed autumn recruitment of 20 young / 100 adult females (table 1). this estimate depends on the assumption that changes to young seen / 100 adult females did not occur with changes to moose density for area 26 and that moose seen / day of hunting varied presumably only with annual weather or other random effects (i.e., area 26 was at k throughout our period of observation, when autumn recruitment was 20–40 young seen / 100 adults; fig. 7). the related assumption of zero population growth means that mean mortality between autumn and the following spring is about 30 young moose / 100 adult females. our hunter trend involving moose seen / day of hunting produces a conservative estimate of real moose density at high values, because: (1) we used a more conservative correction factor for visibility bias than indicated specifically for area 24 (oosenbrug and ferguson 1992); (2) there are many kills that occur with < 1 day of hunting, but our index of hunting effort does not measure less than that period; and (3) on average hunters see more moose when days of hunting are fewer, and hunting trips are very short when hunter success is > 80%. thus, our investigation of k from extrapolating moose seen / day of hunting is likely to be an underestimate particularly at h i g h d e n s i t i e s ; a g a i n , t h i s c o n t r a s t s mccullough’s (1979) conclusion that linear regression overestimates k. combining observations along several hunting routes, we also generalize the effect of changing habitat on our estimate of k. however, our estimate of k for moose is higher than estimates elsewhere, especially in the presence of wolves (gasaway et al. 1992, van ballenberghe and ballard 1998). our estimate of k in newfoundland should be compared to past density estimates for this unique situation. as an example of estimates during the first peak of moose densities in the boreal forest, densities in area 17 in 1960 were observed at 4.6 moose / km2 (bergerud and manuel 1969). this figure was likely an underestimate of the real population size, because the equivalent of 5.0 moose / km2 were shot along roads in 1962 (bergerud et al. 1968). multiplying the 1960 estimate by our visibility correction factor for forest of 2.7 would result in 12 moose / km2 as a minimum density for area 17, such that this earlier peak exceeds our upper estimate of k. estimates of moose density calculated only alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 135 for areas of forest cover, in inaccessible parts of newfoundland (cf. area 26), were stable but highly variable since 1960 (as in fig. 4), and were often > 12 moose / km2 (mercer 1995). to further support our estimates of k, which we suggest may have been exceeded by the peak densities achieved by all dispersing and expanding populations in newfoundland, new densities recorded in lowland forests in gros morne national park, without legal hunting, were 3.4 moose / km2 (uncorrected) and 7.4–10.6 moose / km2 depending on the visibility correction factor (mclaren et al. 2000). in nearby area 40, where moose have recently reached their highest numbers, the post-hunt density ranged from 3– 4 moose / km2 (uncorrected) or 8–10 moose / km2 (using the average visibility correction factor of 2.7) during 1989–99 (mercer 1995). comparisons may also be made among the estimates of deer and moose densities above if food production in different vegetation types and biomass production differences of the ungulates are considered. crête and manseau (1996) and crête and daigle (1999) have performed reviews of this type, in which they suggested that variation in moose biomass depends on the presence of other deer species and on the existence of predators. moose on the south shore of the st. lawrence river in the absence of wolves reach 740 kg / km2, while even in the presence of wolves, they reach a biomass exceeding 1000 kg / km2 on isle royale (crête and daigle 1999). in the forage-limited area these authors studied, on the québec-labrador peninsula, where wolves are present, production estimates are 78 young / 100 adult females, and autumn recruitment is relatively high (90% survival to autumn) compared with our estimates for newfoundland. our calculation of k in moose in the forested areas of newfoundland may be as high as the equivalent of 3 young, 8 adult females, and 4 adult males (table 1), which, according to crête and daigle’s (1999) estimates, approaches 5,000 kg / km2. to compare, 100 whitetailed deer / km2, with the same sex and age ratio, approaches 5,200 kg / km2. these estimates can apparently only be achieved in the absence of additive predation (gasaway et al. 1992, van ballenberghe and ballard 1998, person et al. 2001). population dynamics in newfoundland moose 1904–99 differences in k and resulting differences in population dynamics reported in this paper can be explained by differences in hunting pressure / unit area relative to local density of moose. for example, if we compare central newfoundland areas, then area 24, with the smallest amplitudes in moose seen / day of hunting (not shown), maintains a higher kill rate and a consequently lower density at peak, as well as a higher hunter-observed young to adult female ratio (table 1). in area 24, licence issue also explains more of the variance in the number of young relative to adult females reported seen than in any of the other areas, suggesting a stronger influence of hunting. in contrast, an inaccessible area, such as area 26, has a much lower kill / unit area relative to its density (fig. 4), and, as a result, a consistently low young to adult female ratio (table 1). in general, we conclude that barren areas of newfoundland (40% of insular newfoundland and represented here by area 26), or other areas less accessible to hunting, following a population peak approaching k in the 1950s, moose populations decreased and remained very low thereafter (fig. 10). in terra nova national park (forested and accessible but with negligible, illegal hunting of moose), the moose population now behaves similar to inaccessible areas in that it also experienced no recovery after a decline carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 136 fig. 10. summary of observed and theoretical population dynamics for moose in insular newfoundland: (a) young seen / 100 adult females by either-sex, resident hunters (solid line) as a function of estimated moose density ( / km2), with our range estimate of maximum sustained yield (msy) for forested habitat as vertical dotted lines; (b) annual recruitment (solid line) as a function of moose density (msy, range as in (a); and (c) a representation of population changes during 1940–2000 in forested, accessible areas (longer dashed line), which included 2 peaks above msy (range as in a, now represented by horizontal dotted lines), in 1958 and in 1986, and in forested portions of less accessible areas (shorter dashed line), in which density-dependent effects near or beyond our estimate of k resulted in a permanent decline in moose after 1 9 5 8 . t h i s f i g u r e w a s a d a p t e d f r o m mccullough (1984). following a 1958 peak in density. we predict that in this area, the population will remain low for many years; current estimates of mid-winter recruitment from the last aerial survey are 20 young / 100 adult females in the park (gosse et al. 2002), consistent with our estimates of zero population increase. we also predict that the moose population in gros morne national park under present management will decrease from its present high density and follow a similar trend. this population dynamic is also different from that of other parks (e.g., isle royale national park, michigan, usa; mclaren and peterson 1994) and other areas with wolves. in areas of newfoundland more accessible to hunters, the population dynamic is much different. we describe 2 major peaks in moose density (1958 and 1986) both exceeding msy but not necessarily k (fig. 10). we maintain that the decline following the 1958 island-wide peak in moose resulted in part from hunting in some accessible areas, but also from a natural die-off caused by habitat destruction when populations grew > k in less accessible areas. mercer and manuel (1974) recorded low autumn recruitment in all areas they reviewed, where young observed as a percentage of wintersurveyed moose were 20–40% in accessible areas and 10% in inaccessible areas. those authors hypothesized that the difference in winter recruitment was a result of destruction of winter food resources. according to a study of areas of different forest productivity in norway, differences in spring recruitment in moose begin with a measurable change in fecundity (saether et al. 1996). in newfoundland, a similar change in fecundity was observed in the 1980s in area 24 – consistent with the decline in young / 100 adult females observed in midwinter, aerial surveys (fig. 9), there were 44% young moose observed as twins in 1983, 21% in 1984, 18% in 1985, and generalces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 137 ally < 5% in current classifications (mercer 1995). nutrition-induced changes in vulnerability to predation likely result in later changes to summer recruitment, owing to the importance of mostly compensatory predation by black bear (ursus americanus), estimated at 22% (area 24) to 38% (elsewhere in central newfoundland) of young moose during summer (mercer 1995). for central newfoundland moose populations, we further estimate that density-dependent effects were not evident until a density of about 2 animals / km2 was reached (fig. 10), corresponding to about 0.5 moose seen / day of hunting (fig. 7). this density is below our estimate of msy, taken at 0.6 k, or about 3.5–5.5 moose / km2. using 2 methods, the moose population in insular newfoundland was estimated in 1988 at 167,000 (post-hunt) and 217,000 (pre-hunt), and a population decline was predicted (mercer 1995). if we use the post-hunt estimate, along with our assessment of the recent population decline by 25% of the peak density in 1986 (fig. 4), then the current (1999) population of moose in newfoundland is about 125,000 (posthunt). local moose densities range from < 0.1 to > 8 animals / km2 (mercer 1995); thus, some populations are clearly < msy, whereas others are probably > k (fig. 10). our estimate of the highest peak in moose densities from the first peak in hunter success (fig. 3) indicates that, consistent with caughley’s (1976) interpretation for ungulates, the first irruption occurred in 1958 and has never been exceeded. management of moose without wolves managers were late in responding with quota increases to the increase in moose populations in accessible areas throughout newfoundland during the 1980s and 1990s, and were thus indirectly to blame for the latest observed declines in autumn recruitment. in the 1970s, increases in licence issue were primarily from the promotion of the male-only licence and the opening of new hunting areas; only by the late 1980s did total licence sales increase substantially as a response to increasing moose densities (fig. 2). since 1974, males continue to represent 65–75% of the legal (reported) kill. mccullough (1979) illustrated that a male-only harvest cannot move a population away from k, and that msy cannot be achieved without harvesting females. combined with relatively little change in density of moose (a 25% decline) during the 1990s (fig. 4), harvest of most moose populations in newfoundland has led to declines in sex ratio (mercer 1995). licence issue has not substantially changed throughout the 1990s, resulting in some new density-dependent declines in autumn recruitment in areas where moose have arrived more recently, such as the northern peninsula (mclaren et al. 2000). although this generalized example of passive or precautionary management in licence issue was a response to presumed stable or declining moose populations and declining hunter success (fig. 3), this management was inconsistent with the decline in autumn recruitment also reported by hunters, especially in central newfoundland (fig. 6). moose management in areas without wolves and a lack of rigid control of hunting is a difficult enterprise. we illustrate the use of hunter statistics to show a biological phenomenon. for management of moose, it is useful to have measures of abundance and recruitment that are more cheaply obtained than by aerial survey, and hunter indices have often been suggested as an approach (courtois and crête 1993, timmermann 1993). other measures of abundance and recruitment based on cohort analysis, through more detailed investigation of the age structure of the hunter-killed population, rely on estimates of both kill and hunter effort, resultcarrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 138 ing in high and sometimes misleading correlations between reconstructions and the hunter indices, on which they are based (fryxell et al. 1988, ferguson 1993). attempts to calibrate these reconstructions with hunter indices result in circular arguments that have nonetheless been suggested for use in management (fryxell et al. 1988, ferguson and messier 1996). for a discussion of biases in such methodology, see caughley (1976). we apply hunter reports in an uncorrected fashion, taking advantage of their annual recurrence, to measure relative moose abundance over time. we calibrate these reports to real population dynamics by including hunter observations of young moose as a measure of autumn recruitment. trends from areas with frequent misreporting by hunters (not shown), such as coastal areas 23 and 29 (fig. 1), do not include correlations between moose observations as we have shown here, but we suggest that correlations between young and adult moose would not be directly misreported from areas that do show consistent trends. our main conclusion is that, while some of the decrease in newfoundland moose numbers following 1958 and 1986 is the result of management, changes to population size relative to k, as well as changes to k in some areas, resulted in densitydependent reproduction effects explaining > 10% and up to 76% of the decrease (table 2). our conclusions are consistent with what saether (1997) terms a “general hypothesis” regarding the relative importance of density-dependent versus densityindependent factors for fluctuations in ungulate populations in the absence of predation. in contrast to our interpretation of moose population dynamics in newfoundland (fig. 10), the assumption in an argument forwarded by ferguson and messier (1996) is that newfoundland hunters (and their managers) have always kept moose < k. unfortunately, we note that the basis of their argument, their measures of functional response in hunters, whether as effort (number of days spent hunting) or as kill rate (number of kills / licence / day), are not independent of their measure of moose density, which itself is based on hunter effort, measured by the number of days spent hunting (ferguson 1993). ferguson and messier’s (1996) argument for cycling in moose based on cohort reconstruction of population size, prompted by the analysis conducted by fryxell et al. (1988), is subject to nonindependent validation of population estimates, against kills / day of hunting and moose seen / day of hunting. as we show (fig. 4), these indices are highly correlated. moreover, biases in behaviour of hunters affect kills / day of hunting, as discussed by hatter (2001). the 2-year delay in management response to information obtained from hunters (see methods) is the most obvious of the “time lags” referred to by ferguson and messier (1996) in their argument for delayed density dependence in newfoundland moose hunting. we do not agree that this delay is responsible for cycling in moose. msy densities have not been maintained in newfoundland; moreover, moose populations reached either 1 or 2 peaks, during which reproduction was observably, affected (fig. 10). such missed opportunities or mistakes have more to do with the past actions of moose managers or with the absence of predation and consequent delayed regulatory mechanisms in moose populations (saether 1997) than with “socio-political changes [or] political events” (ferguson and messier 1996: 156). careful interpretation is required to understand population changes from indices of moose abundance, because these indices invariably represent a wide variety of populations with different dynamics in different habitats. in our most precise estimate of k, from combined data for all of alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 139 insular newfoundland, we average the observations of hunters for moose in different habitats and in different stages of population increase. managers should note that our estimate of k is a range, and k varies naturally both in time and space (crête 1989). moreover, our estimate, particularly for forested regions, is affected by expansion of forestry operations and thereby continuous supply of new habitat areas during the period of data collection; the same reason pimlott (1953) accounted for the increase in moose during the early part of the 20th century. also, extrapolation by linear regression does not explicitly take into account density-dependent effects in moose that may become apparent only at higher densities than those observed over our management period (we report a decline in density dependence at a threshold of 20 young seen / 100 adult females). further, we have no indication from hunter returns whether both winter and summer habitat affect k, but we assume both are important. the structure, succession and composition of natural forest communities have continued to be altered in inaccessible areas like area 26, so that their ability to produce moose has been marginalized (fig. 6). assuming as we have that area 26 represents a population at k throughout our study period, we show how weather, as well as other random effects, can create variation in moose seen, young seen, and hunter success between years (figs. 4 and 6). our hunting indices reflect only areas that are hunted, mostly space that is < 2 km from a road. even in accessible areas, there is considerable moose range that is > 2 km from road, which forms refugia, in which habitat quality is reduced when density is > k from lack of hunting. throughout such areas, regenerating fir (abies balsamea), birch (betula papyrifera, b. cordifolia, a n d b . a l l e g h a n i e n s i s ) , a n d o t h e r hardwoods (e.g., cornus, prunus, sorbus) have been eliminated except for some heavily browsed, sparsely distributed saplings. on the south coast, moose depend on atypical foods, such as branches of blown down trees, lichens, and low shrub and herb communities (albright and keith 1987). eventually, population condition is affected under these circumstances (ferguson et al. 1989), as for gros morne national park (mclaren et al. 2000). during periods of increasing population size observed in the 1980s, yearling harvest as a proportion of the moose hunt increases, and following these increase phases, jaw size declines (mercer 1995). an example of a population introduction and subsequent crash following poor condition occurred when moose were experimentally transported to brunette island, between the burin peninsula and the south coast of newfoundland (mercer 1995). in contrast, as we considered for central newfoundland, accessible areas of moose in newfoundland support a population fluctuating around a “long-term equilibrium” density, unlike their erupting phase from 1904–58 (fig. 10). we speculate, consistent with predictions by saether (1997), that for moose in newfoundland in the absence of hunting, a stable equilibrium between the population and food resources is not possible. acknowledgements we are grateful to m. strapp, former statistician for the newfoundland wildlife division, for his careful attention to the hunter statistics over many years. production of this paper was supported by the department of forest resources and agrifoods, government of newfoundland and labrador. references albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of carrying capacity in newfoundland – mercer and mclaren alces vol. 38, 2002 140 newfoundland. canadian field-naturalist 10:373–387. bergerud, a. t., and f. manuel. 1969. aerial census of moose in central newfoundland. journal of wildlife management 33:910–916. , , and h. whalen. 1968. the harvest reduction of a moose population in newfoundland. journal of wildlife management 32:722–728. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8:1309–1315. caughley, g. 1976. wildlife management a n d t h e d y n a m i c s o f u n g u l a t e populations. pages 183–246 in t. h. coaker, editor. applied biology. volume 1. academic press, london, uk. courtois, r., and m. crête. 1993. predicting moose population parameters from hunting statistics. alces 29:75– 90. crête, m. 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67:373–380. , and r. courtois. 1997. limiting factors might obscure population regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal of zoology, london 242:765–781. , and c. daigle. 1999. management of indigenous north american deer at the end of the xxth century in relation to large predators and primary p r o d u c t i v i t y . a c t a v e t e r i n a r i a hungarica 47:1–16. , and m. manseau. 1996. natural regulation of cervidae along a 1000 km latitudinal gradient: change in trophic dominance. evolution and ecology 10:51–62. ferguson, s. h. 1993. use of cohort analysis to estimate abundance, recruitment and survivorship for newfoundland moose. alces 29:99–114. , w . e . m e r c e r , a n d s . m . oosenbrug. 1989. the relationship between hunter accessibility and moose condition in newfoundland. alces 25:36–47. , and f. messier. 1996. can human predation of moose cause population cycles? alces 32:149–161. fryxell, j. m., w. e. mercer, and r. b. gellately. 1988. population dynamics of newfoundland moose using cohort analysis. journal of wildlife management 52:14–21. gasaway, w. c., r. d. boertje, d. v. grandgard, k. g. kellyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. gosse, j., b. mclaren, and e. eberhardt. 2002. comparison of fixed-wing and helicopter searches for moose in a midwinter habitat-based survey. alces 38: 47-53. hatter, i. w. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces 37:71–77. keith, l. b. 1983. population dynamics of wolves. pages 66–77 in l. n. carbyn, editor. wolves in canada and alaska: their status, biology and management. report series 45, canadian wildlife service, ottawa, ontario, canada. mccullough, d. r. 1979. the george reserve deer herd: population selection of a k-selected species. university of michigan press, ann arbor, michigan, usa. . 1984. lessons from the george reserve, michigan. pages 211–242 in alces vol. 38, 2002 mercer and mclaren – carrying capacity in newfoundland 141 l. k. hall, editor. white-tailed deer: ecology and management. stackpole books, harrisburg, pennsylvania, usa. . 1997. irruptive behaviour in ungulates. pages 69–98 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance: deer ecology and population management. smithsonian institution press, washington, d.c., usa. . 1999. life-history strategies of ungulates. journal of mammalogy 79:1130–1146. mclaren, b. e, c. mccarthy, and s. p. mahoney. 2000. extreme moose demographics in gros morne national park, newfoundland. alces 36:217– 232. , and r. o. peterson. 1994. wolves, moose, and tree rings on isle royale. science 256:1555–1558. mercer, w. e. 1974. recruitment rates of newfoundland moose. proceedings of the north american moose conference and workshop 10:37–45. . 1995. moose management plan for newfoundland. report on file with wildlife division, newfoundland and labrador, st. john’s, newfoundland, canada. , and f. manuel. 1974. some aspects of moose management in newfoundland. naturaliste canadien 101:657–671. odum, e. p. 1953. fundamentals of ecology. saunders, philadelphia, pennsylvania, usa. oosenbrug, s. m., and s. h. ferguson. 1992. moose mark-recapture survey in newfoundland. alces 28:21–29. person, d. k., r. t. bowyer, and v. van ballenberghe. 2001. density dependence of ungulates and functional responses of wolves: effects on predatorprey ratios. alces 37:253–274. pimlott, d. h. 1953. newfoundland moose. transactions of the north american wildlife conference 18:563–581. . 1959a. reproduction and productivity of newfoundland moose. journal of wildlife management 23:381–401. . 1959b. moose harvests in newfoundland and fennoscandian countries. transactions of the north american wildlife conference 24:424–448. saether, b.-e. 1997. environmental stochasticity and population dynamics of large herbivores: a search for mechanisms. trends in ecology and evolution 12:143–149. , r. andersen, o. hjeljord, and m. heim. 1996. ecological correlates of regional variation in life history of the moose, alces alces. ecology 77:1493– 1500. sokal, r. r., and f. j. rohlf. 1995. biometry: the principles and practice of statistics in biological research. third edition. w.h. freeman and co., san francisco, california, usa. timmermann, h. r. 1993. use of aerial surveys for estimating and monitoring moose populations – a review. alces 29:35–46. underwood, h. b., and w. f. porter. 1997. reconsidering paradigms of overpopulation in ungulates: white-tailed deer at saratoga national historic park. pages 185–200 in w. j. mcshea, h. b. underwood, and j. h. rappole, editors. the science of overabundance: deer ecology and population management. smithsonian institution press, washington, d.c., usa. van ballenberghe, v., and w. b. ballard. 1998. population dynamics. pages 247– 274 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. alces35_41.pdf 49 tracking mooseand deer-vehicle collisions using gps and landmark inventory systems in british columbia caleb sample1, roy v. rea1, and gayle hesse2 1university of northern british columbia, 3333 university way, prince george, british columbia, canada v2n 4z9; 2wildlife collision prevention program, british columbia conservation foundation, 4431 enns road, prince george, british columbia, canada v2k 4x3. abstract: vehicle collisions with moose (alces alces) and deer (odocoileus spp.) pose a serious threat to all motorists travelling highways traversing habitats of these two ungulates. in british columbia, mitigation measures to reduce such collisions are based on spatially-accurate records of collisions involving moose and deer that are collected by the province’s highway maintenance contractors. to date, the british columbia ministry of transportation and infrastructure (bc moti) uses the paper-based wildlife accident reporting system (wars) established in 1978 to maintain carcass records. we compared carcass location data collected in 2010 to 2014 by bc moti using wars to that collected by northern health connections bus drivers using a newly developed gps-based system (otto® wildlife device). in total, 6,929 carcasses (1,231 moose, 5,698 deer) were recorded using wars and 474 (167 moose, 410 deer) using the otto® wildlife device. we compared data collected along 2,800 km on the same highways in the same seasons of the same years. we found more carcass locations were identified with the wars method, but that in certain geographic regions, the otto® wildlife system identified several unique locations. we contend that more complete and finer-scale carcass location data is possible using a gps-based system such as otto® wildlife, than currently collected solely with the paper-based wars method. alces vol. 56: 49–61 (2020) key words: alces, car, collision, deer, gps, moose, odocoileus, roadkill, survey, vehicle british columbia is a province richly inhabited by large mammals and where populations are abundant and highways traverse their habitat, wildlife vehicle collisions (wvc) are a management concern. between 2013 and 2017, annual collision records from the insurance corporation of british columbia (icbc) averaged 11,000 animal-related crashes, with ~700 human injuries and 3 fatalities (icbc 2018). in addition to the safety threat resulting in $41 million in icbc insurance claims, the annual value of wildlife-specific mortality is estimated at $466 million (sielecki 2010). spatially-accurate, comprehensive wvc records are critical to limit threats to motorists (huijser et al. 2007) and to provide valuable insights about spatial and temporal wvc patterns useful to implement specific wvc mitigation measures. the british columbia ministry of transportation and infrastructure (bc moti) uses the wildlife accident reporting system (wars) to document when and where wildlife carcasses occur and are collected throughout the province. highway maintenance contractors are required to remove carcasses from numbered highways and submit a monthly, paper-based report to bc moti through wars (sielecki 2010). maintenance contractors use date-ofcarcass retrieval and the landmark moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 50 kilometre inventory (lki) (bc moti 2018b) to record the removal locations. at the time of our study, the carcass reporting methodology in wars had remained largely unchanged since its inception in 1978. several weaknesses have been identified previously including imprecise collision locations and incorrect and incomplete reporting of wildlife species (sielecki 2010). to determine how wars data collected under the current system might differ from a gps-based electronic record-keeping system, we partnered with persen technologies inc. (persentech) in winnipeg, manitoba to develop the otto® wildlife gps unit (fig. 1; hesse et al. 2010). this device was specifically designed to capture real-time gps coordinates, and the time and date of sightings of dead and live moose (alces alces) and deer (odocoileus spp.; mule, white-tailed, and black-tailed deer all referred to as “deer”), which together comprised 83.1% of wvcs in 2003–2007 (sielecki 2010). units were designed for dash-mounting with push button controls to enter data (including sound replay) without requiring the operator of the device to stop their vehicle. we partnered with the northern health authority and diversified transport of prince george and installed otto® wildlife devices on the dashboards of northern health connections buses for drivers to collect data on deer and moose throughout the province. to determine the usefulness of gps technology in wvc mitigation planning, we compared the similarity of moose and deer wvc data collected with wars and otto wildlife on selected highways. our null hypothesis was that both methods produced similar temporal and spatial patterns for moose and deer wvcs. study area our study area was located in british columbia, canada, extending from abbotsford in the south to fort st. john in the north (hwys 1 and 97) and from prince rupert on the pacific west coast to valemount (hwys 5 and 16) in east-central british columbia near the alberta border, a total of 2,798 km of highway. it is recognized that variable widths in right-of-ways likely influenced detection rates at certain locations on all highways. the north and east sections of the study area are characterized by rugged, mountainous terrain with deeply incised valleys (child 1992), with terrain to the south and west flat to rolling with hundreds of small lakes and wetlands (heard et al. 1997). although mostly an homogeneous unit on a drumlinized till plateau surrounding periglacial lake deposits, it is dissected by many rivers, lakes and wetlands (child 1992) and divided by the rocky mountains in the north and east. the landscape is dominated by coniferous fig. 1. the otto® wildlife gps device used for capturing location data for live and dead deer/ moose. the unit is powered by two aa batteries or can be plugged into the vehicle’s accessory receptacle. alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 51 forests of hybrid white spruce (picea engelmannii x glauca) and subalpine fir (abies lasiocarpa). lodgepole pine (pinus contorta var. latifolia) and trembling aspen (populus tremuloides) pioneer secondary successional sites (meidinger and pojar 1991), as do many species of willows (salix spp.) and other woody browse plants used by moose and deer. moose densities in the core of our study area were estimated at 0.63–0.78 moose/km2 (cadsand et al. 2013); deer densities were unknown. methods otto® wildlife system the otto® wildlife device provided gps locations of carcasses and additional information about the animal. to record a live sighting, the appropriate species button (fig. 1) was pressed to activate a coloured led and a vocal playback of “deer” or “moose” to confirm that the correct species was recorded by the operator. to catalog a carcass, the “dead button” was pressed immediately after the species button. pushing the “dead button” 3 times allowed the driver to indicate that a record was in error. latitude, longitude, time of day, and date were recorded when any of three buttons designed to collect data were pushed. previous to our study, and to verify that the otto® wildlife devices were recording accurate locations of carcasses and live sighting points of interest (poi), hesse et al. (2010) compared otto® wildlife poi locations to existing government gis layers and found that only 1.5% of otto® wildlife location data were ≥10 m from the digital road atlas (dra) layer (https://www2.gov.bc.ca/ gov/content/data/geographic-data-services/ topographic-data/roads). this study also confirmed (via exit interviews), that otto® wildlife did not pose a safety concern to vehicle operators and that the units were performing as per their intended design. the university of northern british columbia partnered with the northern health authority to dash-mount otto® wildlife devices in 10 northern health connections buses. data were subsequently collected by bus drivers for live and dead moose/deer and reported to bus dispatchers from 10 june 2010 to 15 july 2014. we collected data from dispatchers every 3–6 months and converted otto® wildlife and wars data from the same time period to kmz (keyhole markup language zipped) map files to compare moose and deer carcass data from the two methods. because the wars data only contained carcass records, we made no comparison of live sightings. an inevitable source of error associated with this method of comparison was that bus drivers operated on set schedules, driving certain sections of highway once daily, while maintenance contractors patrolled highways and responded to wvc reports 24 h/d, 7 days a week. therefore, from the outset of the study, we acknowledged maintenance contractors patrolling routes driven by busses would record more carcasses than bus drivers. carcass data were sorted and organized using microsoft excel. records within 500 m of each other were identified using arcgis in a bc albers 1983 coordinate system and quantum gis desktop distance matrix tool, with a linear output matrix type (set within the qgis distance matrix tool; qgis 2.4 development team 2014) and were deemed to be possibly referencing the same carcass (hyrcha and rea, unpublished). all carcass records within 500 m of each other were further evaluated for duplication using the following criteria: carcasses recorded <2 min apart of each other by the same driver on the same date were considered individual sightings of distinct animals, and both records https://www2.gov.bc.ca/gov/content/data/geographic-data-services/topographic-data/roads https://www2.gov.bc.ca/gov/content/data/geographic-data-services/topographic-data/roads https://www2.gov.bc.ca/gov/content/data/geographic-data-services/topographic-data/roads moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 52 were retained. it was considered highly probable that records of the same species at the same location occurring between 2 min and 24 h of each other were recorded by different busses; therefore, these records were classified as the same carcass (recorded twice) and one record was discarded. we accounted for location and time to determine whether same-species records occurring 1–7 days apart were duplicate sightings. for example, two records in close proximity on a stretch of highway with a very low wvc rate were deemed likely as duplicates. a spatial map of carcass locations created with google fusion tables identified concentrations or hotspots of mvcs (fig. 2). wars we obtained wars data for the same period and along the same highways where otto® wildlife devices were used. these data were not georeferenced, but provided written records of carcass locations using a series of established landmarks along highways. therefore, these locations had to be georeferenced in google earth using the landmark kilometer inventory (lki; bc moti 2018b). we created kmz mapping files for both wars and otto® wildlife data and overlaid them for comparison. we then visually inspected and analyzed the area around each otto® wildlife and wars record to identify matches. comparison of otto® wildlife and wars we first considered the spatial proximity of the wars and otto® wildlife records. because the majority of lki landmarks occurred within 1 km (bc moti 2018b), our first sorting criteria was a maximum separation of 1 km. however, because it was possible that the wars spatial data included errors > 1 km, we evaluated matching of the datasets using maximum possible separation distances of 1, 3, and 5 km (s1, s3, s5). separation in recording dates between wars and otto® wildlife records was also used to identify matched records. for each of the 3 spatial sortings, we allowed maximum temporal separations of 1, 3, and 5 days (t1, t3, t5) between matching records from the two databases. we combined the s and t values to simultaneously specify a record’s spatial and temporal sorting. for example, a spatial and temporal range of 3 km and 5 days had the sorting criterion s3t5. comparative analyses were performed separately for the 9 different sorting criteria (s1t1–s5t5). moose and deer carcass records from the otto® wildlife devices were also categorized by 4 seasons of 3-month increments: spring (march, april, may), summer (june, july, august), fall (september, october, november), and winter (december, january, february). after comparing the otto® wildlife and wars data to identify matching records, the data were sorted by time of year, latitude, highway, and separation of carcass location and time. to supplement the numerical data, geographic illustrations of carcass and live sighting locations (for otto® wildlife only) were constructed using the online application “mapmaker” (mapmaker.com 2018). results a total of 167 moose and 410 deer carcasses were recorded using otto® wildlife devices, and 1,231 moose and 5,698 deer carcasses with wars. with the most relaxed sorting criterion of up to 5 km and a 5-day separation in reporting (s5t5), 20% and 27% of otto® wildlife moose and deer were classified as having a match with wars data; with the strictest criterion (s1t1) these matches were 15% and 10%, respectively. http://mapmaker.com alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 53 the averages across all sorting criteria were 16% and 21% of moose and deer records, respectively (table 1). moose carcasses were most commonly recorded during winter (n = 48) and fall (n = 45) with fewer in summer and spring. fig. 2. a mapped sample of data from the middle of the study area (north-central bc) using s5t5 (spatial separation and temporal separation of 5 km and 5 days) otto® wildlife deer carcass records, with circles indicating otto® wildlife records with a corresponding wars record (matching), and triangles indicating unique, unmatching otto® wildlife records (mapmaker.com, 2018). http://mapmaker.com moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 54 the number of otto® wildlife moose carcasses with a matching wars record varied by season. the average matching rate was much higher in winter (26%) and fall (19%) than in spring (3%) and summer (7%) (table 2). most deer carcasses were recorded in fall (n = 131), winter and spring had similar levels of reporting, and summer had the lowest level (n = 82) (table 3). as with moose carcasses, the rate of matching varied by season with highest matching in spring (26%) and lowest in winter (15%) (table 2). matching moose and deer otto® wildlife carcasses and proportions were summarized based on the highway traveled (tables 4 and 5). carcasses of both species were most frequently recorded along highway 97 (103 moose, 296 deer) and highway 16 table 1. the number of matched/unmatched records of deer and moose carcasses by sorting criteria as collected with the otto® wildlife and wars systems in 2010–2014, british columbia, canada. the sorting criteria indicate spatial (s; 1, 3, 5 km) and temporal separations (t; 1, 3, 5 d) used to determine matches. sorting criteria matched # otto® deer unmatched # otto® deer matched # otto® moose unmatched # otto® moose matching (%) with wars (deer, moose) s5t5 110 300 33 134 27, 20 s5t3 95 315 31 136 23, 19 s5t1 87 323 25 142 22, 15 s3t5 104 306 32 135 25, 19 s3t3 92 318 30 137 22, 18 s3t1 83 327 25 142 20, 15 s1t5 68 342 22 145 17, 13 s1t3 65 345 20 147 16, 12 s1t1 60 350 16 151 15, 10 ave. 21, 16 table 2. the number of matched/unmatched records of moose carcasses by sorting criteria and season as collected with the otto® wildlife and wars systems in 2010–2014, british columbia, canada. the sorting criteria indicate spatial (s; 1, 3, 5 km) and temporal separations (t; 1, 3, 5 d) used to determine matches. values are shown separately for records with and without matching wars records. sorting criteria matched (n = 16–33) unmatched (n = 134–151) spring summer fall winter spring summer fall winter s5t5 2 4 16 11 31 37 29 37 s5t3 2 4 14 11 31 37 31 37 s5t1 1 2 13 9 32 39 32 39 s3t5 2 4 15 11 31 37 30 37 s3t3 2 4 13 11 31 37 32 37 s3t1 1 2 13 9 32 39 32 39 s1t5 1 4 10 7 32 37 35 41 s1t3 1 4 8 7 32 37 37 41 s1t1 1 2 8 5 32 39 47 43 ave. 3 7 26 19 alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 55 (55 moose, 97 deer); lower numbers were recorded on highways 2, 5, and 27. the proportion of matching otto® wildlife records were similar on highways 16 and 97, with average matching rates of 13% and 17% for moose (table 4), and 18% and 21% for deer (table 5). certain areas along highway 97, such as between quesnel and williams lake had a large proportion of matched deer carcasses; conversely, other areas had low matching rates. a low proportion of otto® wildlife and wars records were matched on highway 16 (fig. 3), where – between table 3. the number of matched/unmatched records of deer carcasses by sorting criteria and season as collected with the otto® wildlife and wars systems in 2010–2014, british columbia, canada. the sorting criteria indicate spatial (s; 1, 3, 5 km) and temporal separations (t; 1, 3, 5 d) used to determine matches. values are shown separately for records with and without matching wars records. sorting criteria matched (n = 60–110) unmatched (n = 300–350) spring summer fall winter spring summer fall winter s5t5 34 22 38 16 69 60 93 78 s5t3 28 20 31 16 75 62 100 78 s5t1 27 16 29 15 76 66 102 79 s3t5 32 21 35 16 71 61 96 78 s3t3 28 19 29 16 75 63 102 78 s3t1 26 15 27 15 77 67 104 79 s1t5 25 13 18 12 78 69 113 82 s1t3 22 13 18 12 81 69 113 82 s1t1 21 11 17 11 82 71 114 83 ave. 26 21 21 15 table 4. the number of matched/unmatched records of moose carcasses by sorting criteria and highway (h) as collected with the otto® wildlife and wars systems in 2010–2014, british columbia, canada. the sorting criteria indicate spatial (s; 1, 3, 5 km) and temporal separations (t; 1, 3, 5 d) used to determine matches. values are shown separately for records with and without matching wars records. sorting criteria matched (n = 16–33) unmatched (n = 134–151) h 2 h 5 h 27 h 16 h 97 h 2 h 5 h 27 h 16 h 97 s5t5 1 1 0 8 23 1 5 1 47 80 s5t3 1 1 0 8 21 1 5 1 47 82 s5t1 1 1 0 7 16 1 5 1 48 87 s3t5 1 1 0 8 22 1 5 1 47 81 s3t3 1 1 0 8 20 1 5 1 47 83 s3t1 1 1 0 7 16 1 5 1 48 87 s1t5 0 1 0 6 15 2 5 1 49 88 s1t3 0 1 0 6 13 2 5 1 49 90 s1t1 0 1 0 5 10 2 5 1 50 93 ave. 50 17 13 17 moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 56 topley and vanderhoof – there were 37 unmatched records and none of deer. similarly, there were no matching records of 10 deer carcasses on highway 97 between cache creek and hope. multiple stretches of highways with clustered deer carcasses were identified: highway 16 from sinkut falls road to vanderhoof (0.84 deer/ km), highway 97 between chetwynd and taylor (0.83 deer/km), and highway 97 between quesnel and clinton (0.46 deer/ km) in the northern most parts of the province. the average matching rate of otto® wildlife and wars records for both table 5. the number of matched/unmatched records of deer carcasses by sorting criteria and highway (h) as collected with the otto® wildlife and wars systems in 2010–2014, british columbia, canada. the sorting criteria indicate spatial (s; 1, 3, 5 km) and temporal separations (t; 1, 3, 5 d) used to determine matches. values are shown separately for records with and without matching wars records. highway 27 had only one recorded carcass. sorting criteria matched (n = 60–110) unmatched (n = 300–350) h 5 h 16 h 27 h 97 h 5 h 16 h 27 h 97 s5t5 7 19 1 83 9 78 0 213 s5t3 7 18 1 69 9 79 0 237 s5t1 7 17 1 62 9 80 0 244 s3t5 6 18 1 79 10 79 0 217 s3t3 6 17 1 68 10 80 0 220 s3t1 6 15 1 56 10 82 0 232 s1t5 5 16 1 46 11 81 0 250 s1t3 5 15 1 44 11 82 0 252 s1t1 5 15 1 39 11 82 0 257 ave. 38 18 21 fig. 3. an example from north-central bc of a particularly high density of unmatching otto® wildlife deer carcass records is found on highway 16 between topley and vanderhoof. triangles and circles indicate unmatching and matching records, respectively (mapmaker.com, 2018). http://mapmaker.com alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 57 species ranged from 13 to 50% (tables 4 and 5). discussion our data indicate that the majority of otto® wildlife moose (80–90%) and deer carcass records (73–85%) had no matching wars record, with the relative level of discrepancy dependent on our sorting criteria (days and km separating carcass records). the greater than expected proportion of unmatched records can be partially explained by carcasses removed by other agencies (e.g., ministry of environment, rcmp), the public (such as trappers for baiting traps), or scavengers. furthermore, carcasses concealed by snow plowing before the animal was recorded by a maintenance contractor would influence both availability and matching frequency. conversely, the proportion of wars records with a matching otto® wildlife record was <3% for both moose and deer; albeit, a low percentage was not unexpected since buses travel stretches of highway only once or less daily, whereas maintenance contractors patrol the same roads several times daily, every day of the week. if maintenance contractors use the lki as intended to record carcass locations in wars, then s1 should be a sufficient criterion, allowing for a spatial recording error of ~1 km on either side of the carcass. depending on the highway classification, a carcass reported to, or detected by highway maintenance contractors must be removed as soon as possible or within 3 days (hesse and rea 2016). almost all otto® wildlife records collected for this study were on primary highways (bc moti 2018a), so presumably most carcasses would be removed quickly. therefore, a 5-day separation (t5) between matching records is a generous matching criterion, and a 3-day separation (t3) should be sufficient assuming protocols are followed. nevertheless, only 20% of moose and 27% of deer carcasses in otto® wildlife were classified as matching with the most relaxed sorting criteria (s5t5). these large discrepancies are unexplained and require further study. we found that wvcs with moose and deer are most likely to occur in fall and winter, similar to findings by icbc (rea 2006, o’keefe and rea 2012) and bc moti (sielecki 2010). these seasonal peaks in wvcs are similar to those identified in alaska (garrett and conway 1999) and northern sweden (neumann et al. 2012). an unexpected finding was the seasonal relationship in the proportion of matching between the otto® wildlife and wars records. the matching of moose carcasses between the two databases was higher in fall and winter than spring and summer. it is possible that maintenance contractors and bus drivers can more easily distinguish carcasses in fall and winter when the contrast between a dark-bodied moose and the lighter landscape makes a carcass easier to spot. most carcass matches for deer occurred in spring with fewest in winter. beyond the limited sample size, a possible explanation is that carcass reporting by maintenance contractors is less of a priority when crews are preoccupied with plowing and salting roads. further, plowed snow could reduce detection and/or bury deer carcasses and lower the probability of matching. the probability of an otto® wildlife carcass record having a corresponding wars record was influenced by region as certain areas had very low rates of matching. highways in british columbia are maintained by contractors in 28 service areas throughout the province (bc moti 2019), which suggests varied efficiency among maintenance contractors at reporting and removal of carcasses, or that predators or some other agent (e.g., conservation officers, trappers, or the general public; hesse moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 58 and rea 2016) may remove carcasses more frequently. furthermore, the variation in width of road shoulders and highway rightof-ways, ditch depth, and areas which require brush-cutting may also contribute to differences in detection and matching rates. the gps-based, otto® wildlife system records sightings of live moose and deer that can be mapped and displayed visually (fig. 4). unfortunately, we had no basis to compare between the systems because wars does not provide these data, but encourage mapping of similar data for mitigation planning. even though clusters of live animal locations may not necessarily correspond to locations of potential high wvc risk per se (neumann et al. 2012), these data are useful to road safety planners to fig. 4. map of study area in bc showing live moose sightings recorded by northern health authority bus drivers during the study period (mapmaker.com, 2018). http://mapmaker.com alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 59 determine what engineering or environmental factors might explain differences between highway segments with and without wvcs. another limitation was the higher probability of observing carcasses with the otto® wildlife devices in daylight hours. northern health buses operate on set weekly schedules, mainly between 0630 and 2100 hr (northern health connections 2018), whereas highway maintenance contractors are required to remove carcasses all hours of the day. for example, if a bus departs in the morning from a northern centre en route to southern british columbia, drivers are likely to observe and record mvcs from the previous night (most collisions are nocturnal) near the bus’s point of origin; conversely, maintenance contractors would likely have removed carcasses as the bus nears its terminus farther south. it would be useful to collect otto® wildlife records 24 h daily and incorporate daily and random route starttimes (hesse et al. 2010). in some cases, northern health bus drivers may have pressed otto® wildlife buttons too early or too late to pinpoint a carcass location or missed carcasses; both would increase the occurrence of unmatched records. although enthusiasm was high when this project launched, it is possible that drivers became less keen on spotting and recording carcasses as the novelty of the project diminished, as occurred during an earlier pilot study (hesse et al. 2010). it should be noted that most maintenance contractors are trained to “keep an eye out for carcasses” while bus drivers understandably may not have carcasses as their primary search image. having bus drivers ride with maintenance contractors (and vice versa) could provide for a standardized carcass-spotting protocol and reduce possible biases. as discussed by hesse et al. (2010), several modifications could increase the ease of utility of the otto® wildlife device. for instance, bigger buttons with different textures for different species would allow drivers to locate the desired button more quickly, increasing the locational accuracy of records. a button to erase the last keystroke would allow drivers to quickly and easily correct entry mistakes, and additional buttons to record animal behaviour might provide unique and valuable data (hesse et al. 2010). the use of a smartphone application like that developed in alberta (alberta ministry of transportation 2017) would help update the wvc record-keeping system in british columbia. a smartphone application can be easily designed to utilize gps services and capture the latitude/longitude of collision locations, eliminating the need to reference carcass locations to established roadside landmarks in the lki system. not only would a gps-based record-keeping system facilitate more accurate locations, it could also provide supplementary data including photos and videos if combined with dashcam technology. in summary, data from the otto® wildlife units provided for a useful comparison of wvcs collected with the traditional wars system. while temporal and spatial patterns between moose and deer-vehicle collisions were mostly similar for the two systems as expected, matching of wvc data was low. overall and as expected, more wvcs were recorded with wars, but many unique carcasses were recorded with the otto® wildlife system. we recommend the use of gps-enabled data collection devices by highway maintenance contractors to provide accurate location data for moose and deer wvcs to promote road safety for motorists and wildlife alike. postscript: in 2018, the british columbia ministry of transportation and moose and deer-vehicle collisions – sample et al. alces vol. 56, 2020 60 infrastructure began the staged rollout of a new policy that mandates the use of gps technology by all highway maintenance contractors in all service areas for the purposes of collecting more accurate wars data; all carcasses will be identified with a gps location by 2023. acknowledgements we thank dr. d. bowering, the northern health authority the ministry of transportation and infrastructure, and diversified transportation for partnering with us and installing the otto® wildlife devices in their buses. we especially thank the drivers who eagerly worked with us to collect these data. we would like to thank s. emmons, j. svendson, and s. o’keefe for helping with logistics and managing the data during the project and f. franczk with persentech industries for working with us to design the otto® wildlife. we thank l. sielecki for a review of a previous draft of this manuscript. lastly, we thank our two reviewers for valuable recommendations that led to improvements on an earlier draft of our manuscript. references alberta ministry of transportation. 2017. alberta wildlife watch program. (accessed january 2019). british columbia ministry of transp­ ortation and infrastructure. 2018a. highway classification. (accessed march 2019). _____. 2018b. landmark kilometre inventory july 2018 version (201807). (accessed march 2019). _____. 2019. highway maintenance. < h t t p s : / / w w w 2 . g o v . b c . c a / g o v / content/transportation/transportation i n f r a s t r u c t u r e / c o n t r a c t i n g t o t r a n s p o r t a t i o n / h i g h w a y b r i d g e maintenance/highway-maintenance.> (accessed august 2019). cadsand, b., d. c. heard, j. courtier, a. batho, and g. s. watts. 2013. moose density and composition in the southern omineca region, winter 2011–2012. regional report for the ministry of forest, lands, and natural resource operations. 35p. child, k. n. 1992. moose management in the omineca sub-region. in d. f. hatler, editor. proceedings of the 1991 moose harvest management workshop, november 5–7, 1991. kamloops, british columbia. wildlife branch, british columbia environment, victoria, british columbia, canada. garrett, l. c., and g. a. conway. 1999. characteristics of moose-vehicle collisions in anchorage, alaska, 1991–1995. journal of safety research 30: 219–223. doi: 10.1016/s0022-4375(99)00017-1 heard, d., s. barry, g. watts, and k. child. 1997. fertility of female moose (alces alces) in relation to age and body composition. alces 33: 165–176. hesse, g., and r. v. rea. 2016. quantifying wildlife vehicle underreporting on northern british columbia highways 2004–2013. report prepared for the british columbia ministry of transportation and infrastructure, northern region, prince george, british columbia, canada. _____, _____, n. klassen, s. emmons, and d. dickson. 2010. evaluating the potential of the otto® wildlife gps device to record roadside moose and deer locations for use in wildlife vehicle collision mitigation planning. wildlife biology in practice. 6: 1–13. doi: 10.2461/wbp.2010.6.1 http://www.transportation.alberta.ca/content/doctype253/production/alberta​wildlife​watchprogramplan.pdf http://www.transportation.alberta.ca/content/doctype253/production/alberta​wildlife​watchprogramplan.pdf http://www.transportation.alberta.ca/content/doctype253/production/alberta​wildlife​watchprogramplan.pdf http://www.transportation.alberta.ca/content/doctype253/production/alberta​wildlife​watchprogramplan.pdf https://www2.gov.bc.ca/gov/content/transportation/transportationinfrastructure/transportation-planning/highway-classification https://www2.gov.bc.ca/gov/content/transportation/transportationinfrastructure/transportation-planning/highway-classification https://www2.gov.bc.ca/gov/content/transportation/transportationinfrastructure/transportation-planning/highway-classification https://www2.gov.bc.ca/gov/content/transportation/transportationinfrastructure/transportation-planning/highway-classification https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/landmark-kilometre-inventory https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/landmark-kilometre-inventory https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/landmark-kilometre-inventory https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/landmark-kilometre-inventory https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/landmark-kilometre-inventory https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/contracting-to-transportation/highway-bridge-maintenance/highway-maintenance https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/contracting-to-transportation/highway-bridge-maintenance/highway-maintenance https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/contracting-to-transportation/highway-bridge-maintenance/highway-maintenance https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/contracting-to-transportation/highway-bridge-maintenance/highway-maintenance https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/contracting-to-transportation/highway-bridge-maintenance/highway-maintenance alces vol. 56, 2020 moose and deer-vehicle collisions – sample et al. 61 huijser, m. p., j. fuller, m. e. wagner, a. hardy, and a. p. clevenger. 2007. animal-vehicle collision data collection: a synthesis of highway practice. national cooperative highway research program (nchrp) synthesis 370. transportation research board, washington, d. c., usa. icbc. 2018. quick statistics for the media manual. (accessed january 2020). mapmaker.com. 2018. mapmaker. (accessed january 2018). meidinger, d., and j. pojar. 1991. ecosystems of british columbia. special report no. 6. ministry of forests research branch, victoria, british columbia, canada. neumann, w., g. ericsson, h. dettki, n. bunnefield, n. s. keuler, d. p. helmers and c. radeloff. 2012. difference in spatiotemporal patterns of wildlife road crossings and wildlife-vehicle collisions. biological conservation 145: 70–78. doi: 10.1016/j.biocon.2011.10. 011 northern health connections. 2018. northern health connections route schedule. (accessed january 2019). o’keefe, s., and r. v. rea. 2012. evaluating icbc animal–vehicle crash statistics (2006–2010) for purposes of collision mitigation in northern british columbia. report for the insurance corporation of british columbia, north vancouver, canada. rea, r. v. 2006. elucidating temporal and species-specific distinctions in patterns of animal-vehicle collisions in various communities and regions of northern british columbia. in using collision data, gps technology and expert opinion to develop strategic countermeasures recommendations for reducing animal–vehicle collisions in northern british columbia. research report, road health-university wildlife collision mitigation research team, prince george, british columbia, canada. sielecki, l. e. 2010. wars 1988–2007: wildlife accident reporting and mitigation in british columbia: special annual report. environmental management section, engineering branch, british columbia ministry of transportation and infrastructure, victoria, british columbia, canada. (accessed january 2018). https://www.icbc.com/about-icbc/newsroom/documents/crashes-involving.pdf https://www.icbc.com/about-icbc/newsroom/documents/crashes-involving.pdf https://www.icbc.com/about-icbc/newsroom/documents/crashes-involving.pdf http://mapmaker.com https://www.darrinward.com/lat-long/ https://www.darrinward.com/lat-long/ https://nhconnections.ca/portals/12/documents/bus-schedule-nh-connections.pdf https://nhconnections.ca/portals/12/documents/bus-schedule-nh-connections.pdf https://nhconnections.ca/portals/12/documents/bus-schedule-nh-connections.pdf https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 https://www2.gov.bc.ca/gov/content/transportation/transportation-infrastructure/engineering-standards-guidelines/environmental-management/wildlife-management/wildlife-accident-reporting-system/wars-1988-2007 alces35_151.pdf alces vol. 47, 2010 lankester and foreyt moose infected with giant liver fluke 9 moose experimentally infected with giant liver fluke (fascioloides magna) murray w. lankester1 and william j. foreyt2 1101-2001 blue jay place, courtenay, bc, canada, v9n 4a8; 2department of veterinary microbiology and pathology, washington state university, 413 bustad hall, pullman, wa 99164-7040, usa abstract: moose (alces alces) are abnormal, dead-end hosts of the giant liver fluke fascioloides magna. the worms migrate extensively in moose causing considerable hepatic tissue damage before eventually dying. few reach sexual maturity and eggs are seldom, if ever, passed in feces. occurrence of the parasite in moose depends on the presence of a competent definitive host and suitable aquatic snail intermediate hosts of the genus lymnaea. there is no clinical evidence that f. magna kills moose although the considerable tissue pathology seen in some heavily infected livers is suggestive that they do. in this study, 2 farm-reared moose calves (2 months old) and a yearling moose (15 months old) were given 50, 110, and 225 f. magna metacercariae, respectively, and observed for 12.5-16 months. no outward signs of disease were observed. the livers of the 2 animals infected as calves were swollen and contained bloody tracks, extensive fibrosis, and capsules; 1 and 11 immature flukes were recovered. the liver of the animal infected as a yearling had 3 large, thick-walled capsules but no flukes. weight gain and behaviour of all were similar to those of uninfected farm-reared moose. known aspects of the biology of this parasite and our experimental results suggest that f. magna is unlikely to have been a major factor in the recent moose decline in northwestern minnesota. alces vol. 47: 9-15 (2011) key words: alces alces, experimental infection, fascioloides magna, liver flukes, moose, odocoileus. the giant liver fluke (fascioloides magna) develops normally in white-tailed deer (odocoileus virginianus), wapiti (cervus elaphus canadensis), and caribou (rangifer tarandus caribou), but in moose (alces alces) migrates extensively causing considerable damage to liver tissue (pybus 2001). moose are an abnormal dead-end host with few worms reaching maturity, and eggs are likely prevented from leaving infected livers by the resulting fibrosis and thick-walled, closed capsules unconnected to bile ducts (lankester 1974). infection in moose depends on continuous cohabitation with normal cervid hosts and appropriate conditions for transmission, including persistent aquatic habitat for the required, intermediate snail hosts of the genus lymnaea. the extensive and noticeable tissue damage seen in some moose has led to speculation that infection may cause death, particularly in nutritionally stressed animals (pybus 2001, lankester and samuel 2007). based on counts of flukes and liver cysts in dead moose, murray et al. (2006) concluded that the giant liver fluke was a significant mortality factor responsible for a marked decline in the moose population in northwestern minnesota. however, no clinical evidence exists that links the presence of flukes in livers with death of moose. to better understand any such impact on moose, we administered metacercariae of the giant liver fluke to 3 captive, hand-reared moose, observed their behaviour and weight gain following infection, and performed subsequent necropsies. moose infected with giant liver fluke lankester and foreyt alces vol. 47, 2011 10 methods moose were acquired in 1977 and 1978 as orphaned calves (0.5-2 months old) in the vicinity of thunder bay, ontario where f. magna is unreported in deer. two calves were infected shortly after capture; a third, obtained the previous year, was infected when 15 months old (table 1). they were housed in a compound without access to any infected water body from the date of capture. initially they were bottle-fed a formulated milk diet (385 ml of carnation milk with an equal volume of whole milk and 2 egg yolks) supplemented with pelleted beet pulp, alfalfa hay, and commercial dairy ration (nutrena sweetflow-16; lankester et al. 1993). animals were weaned at 16-18 weeks of age and maintained thereafter on dairy ration and alfalfa hay supplemented sporadically with fresh browse. metacercariae were obtained from baldwin enterprises (monmouth, oregon, usa) and their viability on arrival was confirmed in a subsample by microscopic examination of flame cell movement. both free and stillencapsulated metacercariae were counted and immediately administered in water by stomach tube to moose lightly anaesthetized with xylazine hydrochloride (rompun, haverlockhart laboratories, mississauga, ontario, canada). moose were observed daily and any clinical signs noted. they were eventually euthanized using t-61 (hoechst canada inc., montreal, quebec, canada), bled by cutting neck vessels, and hung to obtain whole weights; visceral organs were removed and inspected grossly. the liver was weighed, its volume measured by water displacement, and sliced at thickness of 1-1.5 cm. migrating and encapsulated flukes were dislodged from liver tissue by gently agitating the slices in a pail of warm physiological saline, and were recovered by pouring the saline through a 2 mm mesh screen. flukes were pressed gently between 2 glass microscope slides and examined for eggs in the uterus as an indication of maturity. they were measured to the nearest mm after formalin fixation. protocols were approved by the animal care committee of lakehead university. results both moose infected as calves showed depressed appetite for 7-10 days after infection, but thereafter resumed normal feeding and behaviour until they were euthanized 12.5-14 months later. a female yearling moose raised from a calf, but not infected, died at 14 months of age of accidental trauma and was considered a control; its liver weighed 3.3 kg and was 3,050 cc in volume. calf #1 was given 50 metacercariae when 2 months old and euthanized 12.5 months later; its liver weighed 4.9 kg with volume of 4600 cc and had a rounded marginal edge with a single 3 cm long immature fluke (table 1). seven capsules were present (4 were 3.5-4.5 cm and 3 were 1.0-1.5 cm in diameter), most in the vicinity of the hilus, with one protruding from the surface just beneath the serosa. capsules had thickened, fibrous walls with incorporated black pigment. they were filled with brown-black, pasty material and the larger had a thin, stoney inner lining. small, scattered patches (1-3 mm diameter) of black pigment were visible on the surface of the liver and in the omentum adjacent to the liver. calf #2 was given 35 and 75 metacercariae at 2 and 3 months old, and was euthanized 14 months after the first infection. its liver appeared somewhat enlarged (6.1 kg and 5,650 cc) with rounded marginal edge. eleven immature flukes (1.5-2.0 cm long) were recovered; 2 were in narrow (3-4 mm diameter) blood-filled tracts while the precise location of the others could not be determined. about 30-40% of the liver volume was comprised of thick-walled capsules and diffuse fibrosis. ten capsules (3-4 cm in diameter) had thick, fibrous walls and 15 smaller capsules (1-1.5 cm) had a grey-black inner lining (not stoney) and were filled with grey-black (clay-coloured) alces vol. 47, 2010 lankester and foreyt moose infected with giant liver fluke 11 pasty material. meandering tracts with fibrous walls (1-2 mm thick) were filled with dark red blood. the antero-dorsal surface of the liver was covered with whitish fibrinous tags and adhered firmly to the diaphragm. diffuse accumulations of black pigment were visible on the surface of the liver, in omental fat, and in mesentery and lymph nodes around the lower colon. moose #3 was given 225 metacercariae when 15 months old and euthanized 16 months later. it exhibited vigorous rutting behaviour and had developed average-sized antlers in the fall of its 3rd year. there was a whitish fibrinous coating over 25% of the diaphragmatic surface of the liver, but no areas of adhesions were seen nor was black pigment visible in the mesenteries. the liver (3.8 kg, 3600 cc) had a sharp marginal edge. three whitish capsules (1.7-6.0 cm diameter) were visible as raised areas on the surface of the liver. they had fibrous walls 4-6 mm thick and were filled with pasty to hardened grey-black matter; no flukes were recovered. discussion interpreting the results of parasitic infections reproduced experimentally is often difficult. in nature, the impact of infection on individuals commonly depends on dosage, or numbers of infective forms acquired, and over what time period (samuel et al. 1992, lankester 2002). these natural acquisition rates usually are unknown and experimenters may be tempted to exaggerate doses to ensure infection. seldom is it practical to administer repeated small doses or “trickle infections” over a period of time, as would more closely approximate what probably occurs in nature (prestwood and nettles 1977). the outcome of infection may also depend on host age at first exposure and whether the initial exposure induces a degree of protection against further infection (i.e., concomitant immunity; lankester 2002). the varying effects of dose on the outcome of f. magna infection are evident from experimental infection of mule deer, a host not commonly infected in nature (butterworth and pybus 1993). mule deer fawns given 250-500 metacercariae all died within 163 days of infection, whereas 3 of 4 fawns given 50 metacercariae survived; flukes matured and eggs passed in their feces (foreyt 1992, 1996). in our experiment, f. magna was relatively efficient in reaching the liver of 2-monthold moose given doses presumed moderate. considerable liver damage resulted but no worms reached sexual maturity; the longest was 3 cm but mature flukes can be 8 cm long (pybus 2001). the parasite had lower success in reaching the liver of the moose infected at 15 months old and all flukes were dead 16 months after infection. having to traverse the large functioning rumen of an adult moose may, in part, explain the lower recovery. whether the diet provided to captive, experimental animals may have altered their response to fluke infections cannot be judged. cattle, like moose, are dead-end hosts of f. magna, yet commonly become infected in enzootic areas. infections are generally calf sex age (months) at infection # of metacercariae infection duration (months) body weight (kg) liver weight (kg) liver volume (cc) # flukes # capsules 1 ♂ 2 50 12.5 291 4.9 4600 1 7 2 ♀ 2, 3 35, 75 14 218 6.1 5650 11 25 3 ♂ 15 225 16 334 3.8 3600 0 3 table 1. description of captive moose infected experimentally with metacercariae of the giant liver fluke (fascioloides magna). all animals were captured as calves (0.5-2.0 months old; assumed born 15 may) in an area without f. magna and held in captivity without access to infected water until euthanasia. moose infected with giant liver fluke lankester and foreyt alces vol. 47, 2011 12 sub-clinical and go undetected until slaughter (wobeser et al. 1985, pybus 2001). for example, 12 domestic calves given 1000 metacercariae each and monitored for 26 weeks were described as healthy but had conspicuous liver damage. relatively few flukes were recovered (1-32) and weight gain of infected animals was similar to that of controls (conboy and stromberg 1991). weights of our infected calves were similar to those of calves raised at the facility in subsequent years (lankester et al. 1993) and to those reviewed by broadfoot et al. (1996) including weights of their 11-month-old farm-reared animals (206 kg for females and 228 kg for males); mean liver weight for both sexes was 4243 g (range = 3800-5500). whole weights of wild, 2.5-year-old moose in manitoba were higher (288 and 332 kg; crichton 1979) but calves reared in captivity often weigh less than wild animals due to several factors including captive diets and digestive disorders (addison et al. 1983, welch et al. 1985). in nature, calf and yearling moose are less likely to be found with giant liver fluke infection than older animals (karns 1972, pybus 1990, 2001). is this because they are particularly susceptible to infection and die unnoticed, or are they somewhat refractory, either physiologically or because their feeding habits reduce the likelihood of infection? our results suggest that calves are not unusually vulnerable to hepatic disease, but their feeding habits probably reduce exposure. moose encounter fluke metacercariae while feeding on aquatic plants, and lepitzki (1998) found that trematode metacercariae (presumably f. magna) appeared on aquatic vegetation in greatest numbers during june and in midaugust-early september in the marshes of vermilion lakes, alberta. adult moose consume submerged and floating aquatic plants in greatest amount from mid-june to mid-july (cobus 1972, fraser et al. 1982), but calves rarely forage in a similar manner at such a young age. the rumen fluke paramphistomum spp. (although non-pathogenic in moose) is acquired similarly by moose ingesting metacercariae encysted on aquatic vegetation; moose <1.4 years old had fewer rumen flukes than older animals and calves <2.5 months had none in a study in sibley provincial park, ontario (snider and lankester 1986). poor recruitment is a feature of moose declines occurring in the past 15-20 years in areas west of lake superior (i.e., northwestern ontario, southeastern manitoba, and northwestern minnesota; lankester 2009). however, our results indicate that moderate doses of f. magna do not kill young moose and suggest that poor recruitment of young is unlikely explained by undetected calf mortality due to f. magna. an inference (murray et al. 2006) that liver flukes elicit high moose calf mortality has been attributed to karns (1972), but we could not confirm this interpretation. on the contrary, karns (1972) reported that the net productivity of moose was greater in northwestern minnesota where the prevalence of f. magna was 87%, than in the northeast where only 17% of moose were infected; this difference in productivity would seem related to factors other than liver fluke. recently, maskey (2008, 2011) concluded that a low and apparently declining prevalence of f. magna (<20%) was probably not the cause of a moose decline in north dakota occurring simultaneously with a decline in adjacent northwestern minnesota (murray et al. 2006). it is noted that the number of flukes (intensity of infection) in some moose in northwestern minnesota was high (murray et al. 2006). moose classified as likely to have died of fluke infection were defined as those with “signs of severe pathological damage to tissue and organs and no other overt cause of death” or those where “liver flukes were abundant.” not surprisingly, this group had more flukes (48.1 ± 9.6, n = 23) than animals considered dying of non-fluke related causes (14.2 ± 2.2, n = 69). however, the high prevalence of infection (89%, n = 100) was similar to that alces vol. 47, 2010 lankester and foreyt moose infected with giant liver fluke 13 (87%, n = 128) found >30 years earlier when the population was robust and moose hunting was reinstated (albeit intensity data were not reported, karns 1972). lankester (1972) examined a much smaller sample of moose from neighbouring southeastern manitoba and found 64% of livers with signs of fluke infection; 1-5 immature flukes were recovered from 3 of 7 animals with liver damage. butterworth and pybus (1993) found flukes (16.7 ± 7.3, range = 5-30) in 52% of moose (n = 22) in banff national park in alberta, and 2.7 ± 0.3 flukes (range = 2-3) in 63% of moose (n = 9) in kootenay national park in british columbia. pybus (1990) found 4% of adult moose (n = 191) with 3 ± 1 (range = 2-5) flukes in the foothills region of alberta, and shury (1995) found flukes (x = 8) in 40% of moose >6 years old (n = 10) from banff national park. there are several other reasons why f. magna is unlikely a major factor in moose declines. this parasite has a disjunct distribution across north america and has never occurred in some areas where moose declines are known (lankester 2009). as well, the prevalence of giant liver fluke infection increases with age of the host and reaches a plateau in older animals (lankester and luttich 1988, pybus 2001). but the mean intensity of infection is similar within infected age classes, suggesting the development of an immunological resistance as infection accumulates (pybus 2001). results reported here, and the observations of pybus (1990) suggest that moose react strongly to worms in the liver and liver hypertrophy accompanying infection may eventually mitigate some of the hepatic tissue damage. lastly, liver fluke infections have an aggregated distribution in definitive host populations. most individuals have relatively low-moderate numbers of worms while a few heavily infected animals carry the majority of the parasite population (addison et al. 1988, lankester and luttich 1988, mulvey and aho 1993). thus, even if the heaviest fluke infections can be shown to cause the death of moose, the greatest impact of the disease would be expected in only a relatively small portion of the population. references addison, e. m., j. hoeve, d. g. joachim, and d. j. mclachlin. 1988. fascioloides magna (trematoda) and taenia hydatigena (cestoda) from white-tailed deer. canadian journal of zoology 66: 1359-1364. _____, r. f. mclaughlin, and d. j. h. fraser. 1983. raising moose calves in ontario. alces 18: 246-270. broadfoot, j. d., d. g. joachim, e. m. addison, and k. s. macdonald. 1996. weights and measurements of selected body parts, organs and long bones of 11-month-old moose. alces 32: 173-184. butterworth, e., and m. j. pybus. 1993. impact of the giant liver fluke (fascioloides magna) in elk and other ungulates in banff and kootenay national parks. parks canada, banff national park, banff, alberta, canada. cobus, m. 1972. moose as an aesthetic resource and their summer feeding behaviour. proceeding of the north american moose conference and workshop. 8: 244-275. conboy, g. a., and b. e. stromberg. 1991. hematology and clinical pathology of experimental fascioloides magna infection in cattle and guinea pigs. veterinary parasitology 40: 241-255. crichton, v. f. j. 1979. an experimental moose hunt on hecla island, manitoba. proceedings of the north american moose conference and workshop 15: 245-279. foreyt, w. j. 1992. experimental fascioloides magna infections of mule deer (ococoileus hemionus hemionus). journal of wildlife diseases 28: 183-187. _____. 1996. mule deer (odocoileus hemionus) and elk (cervus elaphus) as experimental definitive hosts for fascioloides moose infected with giant liver fluke lankester and foreyt alces vol. 47, 2011 14 magna. journal of wildlife diseases 32: 603-606. fraser, d., b. k. thompson, and d. arthur. 1982. aquatic feeding by moose: seasonal variation in relation to plant chemical composition and use of mineral licks. canadian journal of zoology 60: 31213126. karns, p. d. 1972. minnesota’s 1971 moose hunt: a preliminary report on the biological collections. proceeding of the north american moose conference and workshop. 8: 115-123. lankester, m. w. 1974. parelaphostrongylus tenuis (nematoda) and fascioloides magna (trematoda) in moose of southeastern manitoba. canadian journal of zoology 52: 235-239. _____. 2002. low-dose meningeal worm (parelaphostrongylus tenuis) infections in moose (alces alces). journal of wildlife diseases 38: 789-795. _____. 2009. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53-70. _____., and s. luttich. 1988. fasciolodes magna (trematoda) in woodland caribou (rangifer tarandus) of the george river herd, labrador. canadian journal of zoology 66: 475-479. _____., and w. m. samuel. 2007. pests, parasites and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose (2nd edition). university press of colorado, boulder, colorado, usa. _____., t. wheeler-smith, and s. dudzinski. 1993. care, growth and cost of captive moose calves. alces 29: 249-262. lepitzki, d. 1998. giant liver flukes and snails in banff national park. final report prepared for the friends of banff national park, heritage resource conservation (banff national park) and alberta sport recreation parks and wildlife foundation. banff national park warden office library, banff, alberta, canada. maskey, j. j. 2008. movements, resource selection, and risk analyses for parasitic disease in an expanding moose population in the northern great plains. ph.d. thesis, university of north dakota, grand forks, north dakota, usa. _____. 2011. giant liver fluke in north dakota moose. alces 47: 1-7. mulvey, m., and j. m. aho. 1993. parasitism and mate competition: liver flukes in white-tailed deer. oikos 66: 187-192. murray, d. l., w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1-30. prestwood, a. k., and v. f. nettles. 1977. repeated low level infection of whitetailed deer with parelaphostrongylus andersoni. the journal of parasitology 63: 974-978. pybus, m. j. 1990. survey of hepatic and pulmonary helminths of wild cervids in alberta, canada. journal of wildlife diseases 26: 453-459. _____. 2001. liver flukes. pages 121-149 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals (2nd edition). iowa state university press, ames, iowa, usa. samuel, w.m., m. j. pybus, d. j. welch, and c. j. wilkes. 1992. elk as a potential host for meningeal worm: implications for translocation. the journal of wildlife management 56: 629-639. shury, t. k. 1995. an update of giant liver fluke (fascioloides magna) infection in ungulates of banff national park 1989-1994. parks canada, western and northern service centre, calgary, alberta, canada. snider, j. b., and m. w. lankester. 1986. alces vol. 47, 2010 lankester and foreyt moose infected with giant liver fluke 15 rumen flukes (paramphostomum spp.) in moose of northwestern ontario. alces 22: 323-344. welch, d. a., m. l. drew, and w. m. samuel. 1985. techniques for rearing moose calves with resulting weight gains and survival. alces 21: 475-491. wobeser, g., a. a. gajadhar, and h. m. hunt. 1985. fascioloides magna: occurrence in saskatchewan and distribution in canada. canadian veterinary journal 26: 241-244. alces35_105.pdf alces37(1)_207.pdf 127 estimation of moose parturition dates in colorado: incorporating imperfect detections eric j. bergman1, forest p. hayes2, and kevin aagaard1 1colorado parks and wildlife, 317 w. prospect ave., fort collins, co 80521, usa; 2wildlife biology program, university of montana, forestry 108, 32 campus drive, missoula, mt 59812, usa. abstract: researchers and managers use productivity surveys to evaluate moose populations for harvest and population management purposes, yet such surveys are prone to bias. we incorporated detection probability estimates (p) into spring and summer ground surveys to reduce the influence of observer bias on the estimation of moose parturition dates in colorado. in our study, the cumulative parturition probability for moose was 0.50 by may 19, and the probability of parturition exceeded 0.9 by may 27. timing of moose calf parturition in colorado appears synchronous with parturition in more northern latitudes. our results can be used to plan ground surveys in a manner that will reduce bias stemming from unobservable and yet-born calves. alces vol. 56: 127–135 (2020) key words: alces, calf-at-heel, detection probability (p), ground surveys, parturition, recruitment throughout north america and europe, researchers and managers use surveys of moose productivity to evaluate populations for harvest management purposes (boertje et al. 2007, grøtan et al. 2009, milner et al. 2013); however, surveys are prone to bias (williams et al. 2001, white 2005). when surveying for newborn moose calves, one source of bias is associated with the detection probability (p) of moose calves-at-heel (bergman et al. 2020). more specifically, if a calf is observed, then p is conceptually 1 for that individual during that occasion. however, if a calf is not observed, then uncertainty about its presence exists (i.e., was the calf simply not observed, or was there no calf to be observed). if surveys are conducted near the peak time of parturition, this bias is confounded by the possibility that cows may have not yet given birth. calf-at-heel estimates are also prone to bias as calf mortality occurs. however, multiplying monthly (or daily) calf survival rates by calf-at-heel ratios provides a numerical correction for bias stemming from calf mortality (bergman et al. 2020). no simple multiplicative, numerical correction exists for pre-parturition observations. fortunately, accounting and accommodating for many types of bias is possible in both modelling and survey design. first, estimates of p can be modelled from repeated observations (bergman et al. 2020). once estimated, p is used to inflate calf-at-heel or calf:cow ratios to reduce bias in estimates. an example of such an approach was completed by bergman et al. (2020) who used occupancy modelling and 3 years of ground observation data from radio-collared cow moose to generate a summertime estimate of p = 0.80. we suggest that a supplemental approach to reducing bias stemming from unborn calves is to quickly and efficiently conduct calf-at-heel surveys after the bulk of parturition has occurred. parturition dates in colorado – bergman et al. alces vol. 56, 2020 128 under ideal conditions and with modern technology, timing of moose parturition can be estimated with minimal uncertainty. for instance, vaginal implant transmitters (vit) are used to alert researchers to the timing and location of a birthing event (patterson et al. 2013, 2016, mclaren et al. 2017). this approach is often used when the objective is to capture and collar newborn calves. however, it requires capturing adult females to assess pregnancy status and to deploy a vit. the recent development of satellite-based vits minimizes the previous need for daily ground or aerial monitoring, but the technology remains cost prohibitive for most routine management purposes. a second and increasingly tractable approach to estimate parturition dates and locations for large herbivores is also tied to satellite technology. satellite collars now allow researchers to shorten the duration between sequential locations of animals and achieve nearly real-time transmission of data. movement algorithms, or even close scrutiny of sequential data points can be used to identify clustered locations that are often indicative of birthing events (severud et al. 2015, mclaren et al. 2017, cameron et al. 2018). however, neither traditional vhf radio-collars, store-on-board gps collars, nor early generation satellite collars provide the frequency of locations and the nearly real-time transmission of data necessary to identify birth sites. our objectives for this research were twofold. first, using ground observation data, we estimated a parturition date curve for moose in colorado. managers in colorado and elsewhere will benefit from estimates of the timing of parturition made more precise by incorporation of p, such that they can implement recruitment surveys when a threshold (such as >90%) of birthing events is predicted. our second objective was to correct estimates of parturition timing for p in this modelling process, thereby improving the precision of parturition date estimates. we hypothesized that accounting for p would shift the date of cumulative births to an earlier date, thereby allowing managers to initiate calf surveys at an earlier date without pre-parturition bias. study area we conducted this research across 3 study areas in colorado. the 2 most northerly were located in jackson (north park) and larimer (laramie river) counties, with the southern study area (san juan mountains) in hinsdale and mineral counties (fig. 1). north park was a high elevation (2,400–2,750 m), wide (14–46 km) mountain valley surrounded on the west by the park range mountains, on the south by the rabbit ears mountain range, and on the east by the rawah and never summer mountain ranges. to the north of the study area was a mix of private and public lands managed primarily for agricultural and open rangeland purposes. moose habitat in north park followed small rivers and creeks comprised of a diversity of willow (salix spp.) communities. moose also used aspen (populus tremuloides), lodgepole pine (pinus contorta), and englemann spruce (picea engelmannii) forests. much of the pine and spruce forests in north park and throughout colorado experienced mountain pine beetle (dendroctonus ponderosae) outbreaks during the latter part of the 20th and first decade of the 21st century, placing these forests into an array of successional stages (hayes 2020). the laramie river study area was located ~ 40 km northeast of north park with the rawah mountain range (3,200–3,840 m) separating them. it was also a high elevation mountain valley (2,470–2,800 m), although the valley floor was not as wide as north park (3–9 km). diverse willow stands located along the rivers and creek corridors gave way alces vol. 56, 2020 parturition dates in colorado – bergman et al. 129 to more upland aspen, lodgepole pine, and englemann spruce forests. the san juan mountains study area in southern colorado at 2,750–3,130 m elevation was higher than the north park and laramie river study areas. it was comprised of narrow valleys (0.5–1.5 km wide) with vegetation communities similar to those in the northern study areas. management authority of moose belonged to colorado parks and wildlife (cpw) in all 3 study areas, and each sustained limited cow and bull harvest. predator assemblages were consistent across study areas with black bears (ursus americanus) and mountain lions (puma concolor) the primary predators of moose, although coyotes (canis latrans) could potentially kill newborn moose calves; wolves (canis lupus) and grizzly bears (ursus arctos horribilis) were absent. predation pressure was considered low in each study area. methods field methods we captured cow moose (>1 year-old) via helicopter darting for 4 winters between mid-december and the end of january, 2015–2018. we sedated moose using one of three different drug combinations: 1) bam (54.6 mg of butorphanol, 18.2 mg of azaperone, and 21.8 mg of medetomidine) in combination with ketamine (200 mg), 2) carfentanil (3 mg) in combination with xylazine (100 mg), or 3) thiafentanil (10 mg) in combination with xylazine (25 mg). once sedated, we blindfolded each animal and administered oxygen (via nasal canula) to minimize the risk of adult and fetal hypoxia. we fitted moose with either a vhf radiocollar (advanced telemetry systems, isanti, minnesota, usa; usa model: m2520b), a store-on-board gps/vhf collar ([advanced telemetry systems; usa model: g2110d], or a satellite/vhf telemetry collar fig. 1. map of colorado, usa (black rectangular perimeter) depicting 3 study areas in relation to nearby cities and communities. study units are depicted by gray filled polygons. parturition dates in colorado – bergman et al. alces vol. 56, 2020 130 [vectronics aerospace gmbh, berlin, germany; model: vertex plus, and advanced telemetry systems; usa model: g5-2d]). blood samples were taken to determine pregnancy status using pregnancy specific protein b (pspb, wood et al. 1986). after handling, capture drugs were antagonized with naltrexone (100 mg, antagonist for carfentanil and thiafentanil), tolazoline (500 mg, antagonist for azaperone and xylazine), and atipamezole (100–150 mg, antagonist for medetomidine and xylazine). all capture and handling methods were approved by the institutional animal care and use committees at colorado parks and wildlife (#08-2013) and the university of montana (#032-17cbwb-060517). during the first year, no moose had been previously captured or collared. in subsequent years, previously collared moose were neither targeted nor avoided by the capture crew. as a result, this random process meant that the pregnancy status of some collared moose was unknown. each spring and summer we conducted ground surveys to evaluate the calf-at-heel status of each collared cow. we began observations in early may and continued through august. pregnant moose, based on pspb results at the time of capture, were prioritized for observation. once these animals were observed, we completed observations of radio-collared animals with unknown pregnancy status. typically, ground observations were completed by a single observer by relying on previous known locations and using vhf signals for ground tracking. a second observer was used for individual moose that consistently evaded observation by a single observer. in cases with two observers, one homed in on the moose using the described techniques, with the second observer stationed along the expected exit route with the goal of observing the moose as it passed by. moose observations typically fell into 2 categories: stationary or moving. stationary observations were made of moose that were either bedded or standing idly while they foraged. stationary observations often lasted from 5 to 20 min and ended when a moose stood and moved or foraged out of sight. moving observations were those of moose displaced by an observer. to be considered a completed observation, observers needed to see the entire moose and the surrounding 2 m of space. repeat observations were made on animals throughout the summer to increase the detection probability of calves, and to help determine the fate of calves. we recorded date, time, and location of each observation. analytical methods our objective was to estimate the parturition date for moose in colorado. thus, only cow moose that were eventually observed with calves-at-heel were included in analyses. cow moose that were never observed with a calf helped inform calf-at-heel and calf:cow ratios (bergman et al. 2020), but did not inform estimates of birth dates. the date of each observation was standardized against the date of the earliest survey (27 april), allowing for simple numerical progression throughout the survey period. we used a hierarchical bayesian model to evaluate the probability of parturition during the study period (mccarthy 2007, gelman et al. 2009). the hierarchical component refers to the multiple levels included in the model, which are ultimately integrated to estimate posterior estimates for each parameter of interest (gelman et al. 2009). the base model for parturition probability included an estimate of the probability of detection and followed a logistic regression with a “logit” link. the model had the following form: logit (ρ i ) ~ α + (β × ϑ i ) + γδi + γτi, alces vol. 56, 2020 parturition dates in colorado – bergman et al. 131 where α is the global intercept, ϑ is the date of observation (with corresponding coefficient β), and random effects (γ) of year (δ) and cow (τ) for each cow, i. the predicted calf presence, ρ, was influenced by an estimated probability of detection 0.8 (bergman et al. 2020) and followed a bernoulli distribution, modeled as: y i ~ bern (φ i ); φ i = μ × ω i ; ω i ~ bern (ρ i ), where the observed calf detections (y) follow a bernoulli distribution with probability (φ) informed by the product of the detection probability (μ) and estimated calf detection (ω, per cow i). this is standard practice for including detection probability in bayesian models (mccarthy 2007). the random effects were given vague normal priors with uniform precision (inverse of variance): θ θ σ σ  =  n u 0, ; 1 ; ~ 0, 10 . 2 2 the global intercept and β coefficient were given vague normal priors: n [0, 1.0 × 10−6]. we ran the model using the “runjags” package (denwood 2016) in r (r core team 2019) including 3 chains, with 10,000 iterations per sample and a burn-in of 5,000 iterations and a thinning parameter of 5. we determined convergence when r-hat < ~1.1 for monitored parameters (gelman and rubin 1992). results we captured 46 individual cow moose that were observed with spring or summer calves-at-heel, providing for 86 unique animal-by-year observations (i.e., some cows were observed multiple years). we made a total of 213 unique observations of these individuals. within a single year, the minimum number of observations of an individual moose was 1 (when a cow was observed with a calf during the first observation and subsequent observations were not made), and the maximum was 5. we made an average of 1.72 (sd = 0.96) observations of each cow. our earliest survey was on 27 april 2016 (day 1), and our earliest observation of a calf was on 17 may 2016. based on raw observation data, the median annual parturition dates ranged from 8 june through 18 july; however, these dates reflect uncorrected adjustments. as expected, accounting for p shifted dates for the predicted probability of parturition to an earlier period. the cumulative parturition probability for moose was 0.50 by day 22 (19 may, with a 95% credible interval [ci] range from 19 to 20 may). in consideration of cumulative parturition among all moose, without correcting for detection probability, the probability of parturition exceeded 0.90 on day 64 (june 30, 95% ci = 20 june to 19 july; fig. 2). when detection probability was incorporated into the model, the cumulative probability of parturition exceeded 0.90 by 30 days after 27 april (27 may; 95% ci = 26 may to 28 may; fig. 2). discussion evolutionary theory suggests that for many large, northern ungulates, the peak and duration of parturition periods are shaped to occur within a narrow window of time (rutberg 1987). but because our ground-based field methods to estimate parturition dates were laborious and observation rates low (0–5/day), the date of first observation for many cows extended well into summer. however, analytical adjustments to the estimation of parturition dates parturition dates in colorado – bergman et al. alces vol. 56, 2020 132 (i.e., accounting for p) buffered the bias associated with our slower field methods and led to the prediction that 50% of parturition events had occurred by 19–20 may. this 2-day window aligned very closely with the range of median parturition date of 19–22 may reported by gasaway et al. (1983) and keech et al. (2000) for the interior of alaska. similarly, bowyer et al. (1998) reported a mean parturition date of 25 may for moose in denali national park, and concluded that 95% of births occurred during a 16-day window. the median parturition date for moose calves in southwest yukon was also 25 may (larsen et al. 1989). our results also aligned with parturition dates for moose in the eastern united states and scandinavia. in new hampshire, musante et al. (2010) and jones et al. (2017) reported a median date of 19 may with 78 and 90% of births occurring between 13 and 27 may, respectively. parturition dates in norway were also similar (23 may), but dates were sensitive to the number of mature bulls in the population (sæther et al. 2003). finally, severud et al. (2015) reported a slightly earlier mean parturition date (14 may) for moose in minnesota, but a 1-month range of parturition (2 may–2 june). this earlier mean parturition date aligned with that reported in ontario (13 may; patterson et al. 2016). we estimated that the cumulative probability of parturition increased from 0.50 to fig. 2. predicted probability of parturition by date (shown as days since 27 april), modeled with (solid) and without (dashed) including the probability of detection, colorado, usa. the horizontal dotted-dashed line indicates a 90% parturition probability. the vertical lines indicate the days on which that 90% parturition probability was estimated to have been achieved for each model (dotted lines represent 95% credible intervals). raw observational data are depicted as black dots. alces vol. 56, 2020 parturition dates in colorado – bergman et al. 133 0.90 between 19 and 27 may, indicating that colorado has a similarly narrow parturition period as reported across much of moose range. this narrow window may be shaped by the interaction of habitat and season (rutberg 1987, bowyer et al. 1998), as well as predation (bergerud 1975, testa 2002). perhaps less intuitive was that colorado’s moose appear to calve in synchrony with moose at more northern latitudes. in comparison, the onset of spring and summer is generally earlier and winter later in colorado. this variation in seasonality could potentially afford moose in colorado and other southern populations flexibility from tight parturition patterns identified in northern populations; however, no shifts in parturition date are apparent. while the seasonality of colorado’s southern latitude may be mediated by high elevations, the parturition synchrony within the species across latitudes may prove to be relevant and informative in the face of a generally warming environment. more specifically, moose occupy a wide geographical and latitudinal range, over which seasons are not perfectly synchronous. yet, they apparently maintain tight synchrony in the timing of parturition across this range. thus, concerns over the shifting of seasonality due to global warming (i.e., earlier spring and delayed winter) may not intrinsically, or negatively impact the timing of parturition. from a management perspective, the estimation of parturition dates in colorado was particularly useful to design field surveys. one goal of refining productivity surveys is to reduce bias, and as noted, one source of bias is p and its confounding effects when moose are transitioning between pregnancy and calf-at-heel. ideally, surveys should be implemented post-parturition, but early enough that neonatal mortality is minimal. after applying the probability of detection at which the probability of parturition reached 90%, we recommend that surveys in colorado be initiated on 27 may, or 34 days prior to the date predicted without considering probability of detection. in addition, incorporating probability of detection decreased the credible interval (by about 28 days) associated with the predicted date at which 90% of parturition events occurred. while ground surveys cannot fully replicate the results of aerial surveys, managers can use our results to improve and facilitate the timing of ground surveys. for example, a concerted ground survey effort at the end of may, conducted typically on foot or horseback with a large number of volunteers and field personnel, should produce a productivity estimate with minimal bias from pregnant females and calf mortality. importantly, the narrower credible interval indicates that earlier surveys need not compromise confidence, and applying our refined, optimal date to initiate earlier surveys will produce measurable savings in effort and agency resources. acknowledgents this research was funded in part by a united states fish and wildlife service federal aid research grant, cpw game cash fund, and auction and raffle grants administered by cpw. we are also indebted to the efforts of s. boyle, b. cimpher, r. cordova, a. howell, a. maclean, s. peterson, b. smith, and k. yeager who spent many hours tracking and observing moose. early drafts of this manuscript were improved by the comments provided by m.alldredge, d. tripp, m. lankester, and 2 anonymous reviewers. references bergerud, a. t. 1975. the reproductive season of newfoundland caribou. canadian journal of zoology 53: 1213–1221. doi: 10.1139/z75-145 parturition dates in colorado – bergman et al. alces vol. 56, 2020 134 bergman, e. j., f. p. hayes, p. m. lukacs, and c. j. bishop. 2020. moose calf detection probabilities: quantification and evaluation of a ground-based survey technique. wildlife biology. doi: 10.2981/wlb.00599 boertje, r. d., k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494–1506. doi: 10.2193/2006-159 bowyer, r. t., v. van ballenberghe, and j. g. kie. 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79: 1332–1344. doi: 10.2307/1383025 cameron, m. d., k. joly, g. a. breed, l. s. parrett, and k. kielland. 2018. movement-based methods to infer parturition events in migratory ungulates. canadian journal of zoology 96: 1187–1195. doi: 10.1139/cjz-2017-0314 denwood, m. j. 2016. runjags: an r package providing interface utilities, model templates, parallel computing methods and additional distributions for mcmc models in jags. journal of statistical software 71: 1–25. doi: 10.18637/jss. v071.i09 gasaway, w. c., r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84: 1–50. gelman, a., j. b. carlin, h. s. stern, and d. b. rubin. 2009. bayesian data analysis, 2nd edition. chapman & hall press, boca raton, florida, usa. _____, and d. b. rubin. 1992. inference from iterative simulation using multiple sequences. statistical science 7: 457–511. doi: 10.1214/ss/1177011136 grøtan, v., b.-e. sæther, m. lillegård, e. j. solberg, and s. engen. 2009. geographical variation in the influence of density dependence and climate on the recruitment of norwegian moose. oecologia 161: 685–695. doi: 10.1007/ s00442-009-1419-5 hayes, f. p. 2020. resource selection and calving success of moose in colorado. m. s. thesis, university of montana, missoula, montana, usa. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450–462. doi: 10.2307/3803243 larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rate of moose mortality in southwest yukon. journal of wildlife management 53: 548–557. doi: 10.2307/3809175 mccarthy, m. a. 2007. bayesian methods for ecology. cambridge university press, cambridge, united kingdom. mclaren, a. a. d., j. f. benson, and b. r. patterson. 2017. multiscale habitat selection by cow moose (alces alces) at calving sites in central ontario. canadian journal of zoology 95: 891–899. doi: 10.1139/cjz-2016-0290 milner, j. m., f. m. van beest, e. j . solberg, and t. storaas. 2013. reproductive success and failure: the role of winter body mass in reproductive allocation in norwegian moose. oecologia 172: 995–1005. doi: 10.1007/s00442-012 2547-x musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose alces alces population in northeastern united states. wildlife biology 16: 185–204. doi: 10.2981/09-014 patterson, b. r., j. f. benson, k. r. middel, k. j. mills, a. silver, and m. e. obbard. alces vol. 56, 2020 parturition dates in colorado – bergman et al. 135 2013. moose calf mortality in central ontario, canada. journal of wildlife management 77: 832–841. doi: 10.1002/ jwmg.516 _____, k. j. mills, k. r. middel, j. f. benson, and m. e. obbard. 2016. does predation influence the seasonal and diel timing of moose calving in central ontario, canada? plos one 11: e0150730. doi: 10.1371/ journal.pone.0150730 r core team. 2019. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. https://www.r-project. org (accessed april 2020). rutberg, a. t. 1987. adaptive hypotheses of birth synchrony in ruminants: an interspecific test. the american naturalist 130: 692–710. doi: 10.1086/284739 sæther, b.–e., e. j. solberg, and m. heim. 2003. effects of altering sex ratio structure on the demography of an isolated moose population. journal of wildlife management 67: 455–466. doi: 10.2307/ 3802703 severud, w. j., g. delgiudice, t. r. obermoller, t. a. enright, r. g. wright, and j. d. forester. 2015. using gps collars to determine parturition and cause-specific mortality of moose calves. wildlife society bulletin 39: 616–625. doi: 10.1002/wsb.558 testa, j. w. 2002. does predation on neonates inherently select for earlier births? journal of mammalogy 79: 1332–1344. white, g. c. 2005. correcting wildlife counts using detection probabilities. wildlife research 32: 211–216. doi: 10.1071/wr03123 williams, b. k., j. d. nichols, and m. j. conroy. 2001. analysis and management of animal populations. academic press, inc. san diego, california, usa. wood, a. k., r. e. short, a. darling, g. l. dusek, r. g. sasser, and c. a. ruder. 1986. serum assays for detecting pregnancy in mule and white-tailed deer. journal of wildlife management 50: 684–687. doi: 10.2307/3800981 https://www.r-project.org https://www.r-project.org alces35_191.pdf alces36_35.pdf f:\alces\vol_38\pagema~1\3801.pdf alces vol. 38, 2002 bottan et al. adaptive management of ontario moose 1 adaptive management of moose in ontario brian bottan1, dave euler2,3, and rob rempel4 1202 briar bay, thunder bay, on, canada p7c 1l8, brianbottan@hotmail.com; 2faculty of forestry and the forest environment, lakehead university, 955 oliver road, thunder bay, on, canada p7b 5e1, birchpt@sympatico.ca; 4ontario ministry of natural resources, centre for northern forest e c o s y s t e m r e s e a r c h , 9 5 5 o l i v e r r o a d , t h u n d e r b a y , o n , c a n a d a p 7 b 5 e 1 , rob.rempel@mnr.gov.on.ca abstract: early policy decisions affecting moose (alces alces) management in ontario were based on data that were not reliable, but were the only basis available for policy development. as data collection increased in accuracy and reliability, policy decisions have also improved. in the last decade of the 20th century, adaptive management has been discussed and advocated as the best approach to managing natural resources since it was first developed in the early 1970s. the ontario ministry of natural resources has instituted at least some of the characteristics of adaptive management in managing moose. the 1960s and 1970s were periods of extensive learning and maturation for biologists and wildlife managers with respect to ontario’s moose herd. the experience and knowledge gained from these periods were used to develop goals and objectives which would eventually become ontario’s 1980 moose policy and the first steps of adaptive management. the later phases of the adaptive approach, to evaluate the earlier objectives and learn from them, are reviewed and discussed. the goals established in 1980, probably cannot be achieved, however, the learning associated with the process is important in order to manage adaptively. alces vol. 38: 1-10 (2002) key words: adaptive management, conservation, moose, moose hunters, moose policy, research moose (alces alces) management in ontario began with r. l. peterson’s investigations during 1949-51 which provided the basis for early management policy and the starting point for subsequent investigations (cumming 1974). early management objectives of the 1940s grew from a philosophy of conserving and protecting wildlife via enforcement of regulations made under the game and fish act (cumming 1974). moose management at this time was in its infancy as biologists and wildlife managers strove to uncover the uncertainties of age and sex ratios, herd numbers, preferred habitats, diet, birth rates, range, influence of predators, and anthropogenic impacts. some 50 years have past since peterson’s first investigations, and it is time to review progress in managing ontario’s moose herd. using ontario as a case study, the objective of this paper is to evaluate past management decisions with respect to current theories of adaptive resource management. this discussion will examine, in chronological order, ontario’s use of adaptive management techniques, with specific attention paid to ontario’s 1980 moose policy. it should be noted that this paper is a commentary, which portrays the opinions and views of the authors with regards to ontario’s use of adaptive management to manage moose. adaptive management although there is no clear consensus on what does or does not constitute adaptive 3present address: birch point enterprises, box 6, site 4, rr #4, echo bay, on, canada p0s 1c0 adaptive management of ontario moose – bottan et al. alces vol. 38, 2002 2 management, it is generally known “as a formal process for continually improving management policies and practices by learning from their outcome” (taylor et al. 1997: 2). the most common difference between adaptive and traditional approaches to management is that traditional approaches typically lack reliable feedback mechanisms that encourage learning. furthermore, adaptive management differs from traditional approaches because it is a systematic, rigorous approach to learning by doing, rather than a haphazard, trial and error approach. the first critical step in adaptive management is to develop clear, defined management objectives in terms of ecosystem function. the adaptive process is a systematic, cumulative approach to learning, where without clearly defined objectives, learning cannot begin. thus, management objectives must contain measurable goals, specified over appropriate time frames and spatial scales from which to learn about the ecosystem. once management objectives have been stated, the next step is to identify questions and uncertainties about the ecosystem in order to develop the best policy (taylor et al. 1997). management must ask the “need to know questions” that will distinguish whether or not the objectives have been achieved. as well, recognizing uncertainty about the ecosystem is necessary to avoid asking “nice to know” questions, rather than those that contribute to learning. unfortunately recognizing uncertainty is difficult because it often leads to controversy and adverse reaction from peers or the public. the third step in adaptive management is to explore potential effects of alternative hypotheses on key response indicators (taylor et al. 1997). in the context of moose management, key response indicators may be changes in hunter harvest rates, sex or age distributions, or population counts. this exploration is achieved through the design of (experimental) management policies (or models) and monitoring schemes for reliable feedback. staff creativity and experience is of the utmost importance at this stage, especially when discrimination between alternative hypotheses becomes difficult, sometimes requiring new approaches that deviate from the norm (hilborn et al. 1979, walters 1986, mcallister and peterman 1992). for example, traditional approaches used in the natural sciences (biology, forestry, and ecology) may give way to new approaches developed in the social sciences (human dimensions) as a way of exploring alternative hypotheses (applegate and witter 1984, lautenschlager and bowyer 1985, decker and richmond 1994, decker and enck 1996, bottan 1999, bottan et al. 2001). perhaps one of the most important steps in the process of adaptive management is monitoring. gibbs et al. (1999) described monitoring as the collection and analysis of repeated observations or measurements to evaluate changes in condition and progress toward meeting management objectives. several proponents of the adaptive process (holling 1978, ringold et al. 1996) have advocated the role of monitoring in the adaptive management process. this is particularly important when evaluating the utility of several alternative hypotheses and the success of stated management objectives (gibbs et al. 1999). the last steps in the adaptive management process are feedback loops, where collected data are analyzed, management objectives are adjusted, and information is communicated to policy makers and the public. predetermined changes (qualitative or quantitative) in key indicators should trigger predetermined changes in management activities or objectives (taylor et al. 1997). although there are many other steps in the adaptive process where failure may alces vol. 38, 2002 bottan et al. adaptive management of ontario moose 3 occur, perhaps the feedback process is the most critical. to quote hilborn (1992: 12), “if you cannot respond to what you have learned, you really have not learned at all”. even the best management objectives, monitoring programs, and data analyses will go to waste if one cannot apply what one has learned. in conclusion, despite the intuitive appeal of the adaptive management concept, there are few examples in wildlife management where it has been applied successfully (gibbs et al. 1999). there are numerous pitfalls such as technical, economic, ecological, institutional, and social challenges that affect the implementation and effectiveness of adaptive management. adaptive management requires managers and decision makers who are willing to learn by doing, and who acknowledge that making mistakes is part of learning (taylor et al. 1997). case study the objective of this paper is to discuss ontario’s use of adaptive management in the context of managing moose. due to the vast number of management decisions that have occurred over the past 50 years, only a few points from selected decades will be used to illustrate ontario’s use of adaptive management or the lack thereof. special attention is paid to ontario’s 1980 provincial moose policy. the 1960s the 1960s witnessed for the first time in ontario’s moose management history a new division of fish and wildlife policy statement containing 4 management principles. the thrust of these principles concentrated on maximum sustainable yield, multiple and full uses of the resource, and recognition of public uses. the intent was to provide hunters with more hunting opportunities, to protect and increase the existing herd, and to provide others with opportunities to use the same land in which moose inhabit (cumming 1974). these policy statements, written at a time when adaptive management was not well developed, were typical for that time. for example, in 1967, one of the statements of purpose concerning moose management in ontario was: “to provide the most hunting and viewing of moose which can be sustained without interfering with other interests” cumming (1974 : 676). in 1969, the purpose was revised “to provide: (1) a moose population as large as can be reconciled with timber production and forest management in general, and (2) as much hunting and viewing as the populations will sustain.” cumming (1974 : 676). while these statements indicate a great purpose, they have little or no measurable attributes. thus, after the policy was implemented and several years elapsed, little or no opportunity was available to learn from the policy. adaptive management advocates that clear management objectives be established with measurable outcomes in order to foster learning. without that first critical step, subsequent learning is much more difficult, if not impossible. the 1970s the 1950s and 1960s were periods in which ontario’s moose herd was able to sustain the demands management placed upon it; but by the 1970s it was apparent that these demands had begun to take a toll on herd numbers. steadily increasing hunter population and increased access (eason et al. 1981; eason 1985, 1989; bisset 1991) attributable to changes in forestry practices (e.g., mechanization) (thompson and stewart 1998) resulted in high hunter success rates (timmermann and gollat 1983). moose managers realized that there were significant problems with the moose population, which made the 1970s a critical period adaptive management of ontario moose – bottan et al. alces vol. 38, 2002 4 for ontario’s moose (omnr 1990). to deal with the problem, the ontario ministry of natural resources (omnr) implemented a number of passive approaches. for example, control measures were instituted to reduce the number of moose harvested in the province. control measures such as the delay of opening season were used to prevent the rut coinciding with the opening day of the hunting season. a full-scale telemetry project was initiated in 1972 and aerial survey techniques developed by fowle and lumsden (1958) and cumming (1958) in the 1950s were standardized in 1973. wildlife management units (wmus) were established in 1975 to “allowed managers to organize wildlife population data in separate geographic areas on the basis of land form, forest types, and habitat potential” (omnr 1990: 28). the establishment of wmus reduced moose management from a provincial scale to a local scale. each of these initiatives were designed to gain a better understanding of moose behaviour, habitat preferences, range, and numbers. as well, the information gained from these initiatives was used to improve the effectiveness of moose management objectives and to develop solutions to increase ontario’s moose herd. although the approaches (initiatives) managers applied to the problem of a declining moose herd were well meant, these steps failed to increase the herd (bisset 1991) and truly lacked clear policy objectives that are conducive to adaptive learning. the research initiatives were positive developments and the management efforts all moved moose management towards a solution to a difficult problem. the 1980s continuing their efforts to rectify the problems in moose management discovered a decade earlier, ontario’s provincial government established a new management policy, in 1980, that set specific goals and objectives for the herd: (1) to increase the herd from 80,000 to 160,000 animals by 2000; (2) to harvest 25,000 moose annually by 2000; (3) to provide 875,000 hunter days annually by 2000; and (4) to create sites where 1 million people annually can observe moose by 2000 (omnr 1980, 1990; timmermann and buss 1998). with this policy, the ontario ministry of natural resources established its first clear goals and objectives for wildlife that one day could be measured. setting clear goals and objectives is one of the first steps to managing adaptively. for the most part, the fear of failure leads agencies to establish well-meaning statements of intent, that are so vague that an observer cannot judge if the goals have been accomplished (taylor et al. 1997). the 1980 policy was in contrast to the policies of the 1960s that advocated maximum sustained yield and “best use” ideas, but with no indication of exactly what was intended. whether they knew it or not, wildlife managers in ontario had taken the first step towards managing adaptively. the goal of 160,000 moose was based on the idea that ontario moose populations in areas with good habitat, with wolves and bears present, but no human hunting, (e.g., quetico provincial park and chapleau crown game reserve) had approximately 0.40 moose per square kilometer. moose density in these areas was supplied to one of us (euler) by field staff working in those areas at that time. subsequent publications, (crête et al. 1981, allen et al. 1987, crête 1989) demonstrated that moose populations in similar conditions were capable of attaining similar densities, although sometimes predation or human hunting reduced the population below these levels. with this population goal in mind, the herd was expected to sustain by 2000 an annual harvest alces vol. 38, 2002 bottan et al. adaptive management of ontario moose 5 of 25,000 moose, provide hunters with 875,000 hunting days afield, and 1 million viewing opportunities for people provincewide. these targets were selected because they seemed attainable, and if not attained, would constitute an opportunity to learn more about moose management. the most significant addition to the 1980 policy was the adoption of the selective harvest system in 1983 (euler 1983, heydon et al. 1992, timmermann and rempel 1998). selective harvest requires hunters to identify the age and sex of the moose before shooting, something some hunters found difficult (timmermann and gollat 1984). the system permits only a limited number of adult moose to be harvested, while shifting unlimited hunting pressure onto calves where the chance of overharvesting is lower. this was a difficult period of transition for hunters due to the fact that the old system of unlimited hunting, long seasons, and few restrictions had been in place for such a long period of time. however, these changes were necessary if ontario’s moose herd was to be protected and expected to grow. management also took steps towards manipulating habitat as a means of increasing ontario’s moose population. habitat guidelines were discussed and studied in the early 1980s (euler 1982), but were not formally released until 1988, when timber management guidelines for the provision of moose habitat were formally endorsed by the ministry (omnr 1988). the primary purpose of these guidelines was to assist resource managers in planning timber management activities with regard to forest access, harvest operations, site preparation, regeneration, and maintenance. while the habitat guidelines were supportive of the broad policy objectives, they were not laws and did not contain clear goals or objectives that could be measured at some later point. the 1990s in the 1990s, ontario took the next step on the road to adaptive management by instituting a long-term research project to test the effectiveness of the timber management guidelines for the provision of moose habitat (rodgers et al. 1996). the moose guidelines evaluation program was established in 1989 to evaluate the effectiveness of fine-scale components of the guidelines within an individual moose’s home-range, including protection of aquatic feeding areas, corridors, and leave-strips (rodgers et al. 1996). rempel et al. (1997) set out to test omnr policy, in particular, moose harvest regulations and the timber management guidelines. their study was designed as a mensurative, large scale experiment in which results suggested a strong interaction between hunter access to moose and habitat quality (rempel et al. 1997). furthermore, study results demonstrated that coordinated harvest and habitat management is required to successfully manage moose populations as simply managing habitat alone is insufficient to achieve 1980 policy objectives. mckenney et al. (1998) explored the power of spatial population models using geostatistical interpolation techniques to evaluate moose harvest policies. point survey data were used to generate a map of moose density in 5-year time periods, and wildlife management units (wmus) were subsequently overlaid on the map. this allowed for the identification of “spatial anomalies”, where wmu moose densities were unexpectedly low, or high, relative to densities in surrounding areas (mckenney et al. 1998). in conjunction with this work, some effort was also directed at evaluating moose demographic responses to the selective harvest system (timmermann and rempel 1998), and re-evaluating the provincial moose targets based on sub-regional environmenadaptive management of ontario moose – bottan et al. alces vol. 38, 2002 6 tal capability to support moose populations. to various degrees, these research projects involved both predictive modeling and empirical monitoring of moose response at the individual and population levels. the habitat research has an inherent feedback to moose management through the mandate to revise the moose habitat guidelines based on findings from the work. the population modeling research cycled back to management by providing science support for the project to reset moose population targets. although these research components contribute to an adaptive approach to moose management, they are essentially passive, retrospective studies evaluating policies that have already been implemented. a more direct and powerful approach would be to actively implement various management policies and guidelines on the landscape, such as altered cut block size and placement, or altered moose harvest rules, and then design a monitoring program to test the effects of these various management elements on moose populations. without this step it is difficult, if not impossible to attribute cause and effect to management options. consider, for example, the documented increase in moose density since implementation of the moose habitat guidelines (mckenney et al. 1998). spatially explicit maps of moose population increase clearly show that the herd has been increasing across ontario since the mid 1980s. however, at the same time habitat guidelines were implemented, the selective harvest system was also implemented, which dramatically decreased hunting pressure. which of the two policies has contributed the most to the observed increase in moose density? if we either created more edge habitat or further decreased hunting pressure, then which of these two actions would create the greatest population response? answers to these questions cannot be easily achieved through such a retrospective analysis. a controlled, large-scale management experiment is best suited, and this form of adaptive management is essential to quick, effective learning from management actions. none-the-less, if the study is well designed, factorial mensurative experiments using retrospective data can be conducted to statistically examine the effects and interactions of management actions. rempel et al. (1997) did this in their study on the effects and interaction of disturbance and road access on moose populations. but the knowledge gained by such work lacks the reliability of true manipulative experiments because of the inability to control for environmental variance and other factors that may be driving population response. discussion ontario took the first step in adaptive management for moose by establishing goals and objectives in 1980 that were clear and measurable. methods of controlling human hunting were introduced and habitat guidelines were developed for managers to help produce habitat conditions favorable for moose. these steps fulfill the first directive that emanates from adaptive management; managers must be explicit about what they expect. the subsequent steps of adaptive management, collecting information and comparing that information with expectations, has been recorded in timmermann and gollat (1986), heydon et al. (1992), timmermann and whitlaw (1992), rempel et al. (1997), mckenney et al. (1998), and timmermann and rempel (1998). these authors suggested that the goal of 160,000 moose as well as the other targets would likely not be reached by 2000. despite the fact that the stated population goal has not been achieved, it is clear that the moose population has increased since 1980. ontario’s moose population is now in the 100,000 alces vol. 38, 2002 bottan et al. adaptive management of ontario moose 7 to 120,000 range (simmons 1997, provincial auditor 1998, timmermann et al. 2002), considerably higher than the earlier estimate of about 80,000 in the early 1980s (bisset 1991). thus, the management approach was the appropriate one, however, the total goal of 160,000 moose in the province may not be possible given the current conditions of forest management, hunting management, and other ecological factors in the province. ontario’s moose harvests remain in the range of 10,000 to 12,000 animals, about half of the projected target of 25,000 by the year 2000 (simmons 1997, timmermann and buss 1998, timmermann et al. 2002). although these figures are below expectation, possibly a more important problem is that in ontario there currently lacks an accurate and timely way to measure the annual harvest. support for mandatory registration, including a variety of methods, has been documented by omnr (1980), hansen et al. (1995), and bottan (1999). as well, several others have advocated its use and importance to accurately assess annual harvests and adjust harvest quotas quickly (crichton 1992, timmermann and whitlaw 1992, timmermann et al. 1993, timmermann and rempel 1998). the remaining 1980 policy goals, 875,000 hunter days afield and 1 million viewing opportunities by 2000 are difficult to evaluate because there was little or no effort to integrate, and data that were collected are only reasonably accurate on a regional or provincial level (timmermann et al. 1993). timmermann et al. (1993) recommended that higher quality district mail survey data should be phased in to replace broad provincial statistics. thompson and stewart (1998) have reviewed habitat management strategies in the context of adaptive management and propose a more flexible approach to habitat for moose based on principles of natural disturbance. however, in ontario, rempel et al. (1997) suggest that the moose habitat guidelines designed to mimic natural disturbances solely will not increase moose densities. thus, if managers attempt to follow the idea of designing timber harvest to mimic natural disturbance patterns, management plans must also include restrictions on hunter access in order to increase moose densities (rempel et al. 1997). conclusions despite the intuitive appeal of the adaptive management concept, there are startlingly few examples in wildlife management in which the adaptive management loop has been completed (gibbs et al. 1999). adaptive management is an approach to management, not a single cookbook of steps that can be applied by rote to every management issue. cookbook approaches tend to stifle the creativity that is crucial for dealing effectively with uncertainty and change (taylor et al. 1997). unless a management agency adopts an adaptive approach to natural resource management, very little learning can take place and attempts to correct errors will not have a high success rate. while implementing adaptive management will not be easy, the alternative is to continue to learn slowly, repeating mistakes, reaching invalid conclusions, and missing opportunities to manage better (taylor et al. 1997). acknowledgements we wish to thank tim timmermann and al bisset for comments on early drafts. as well, special thanks go out to two anonymous referees for their critiques of the manuscript and suggestions for improvement. references allen, a., p. a. jordan, and j. terrrell. 1987. habitat suitability models: moose, lake superior region. biological readaptive management of ontario moose – bottan et al. alces vol. 38, 2002 8 port 82 (10.155). u.s. department of agriculture, fish and wildlife service, washington, d.c., usa. applegate, j. e., and d. j. witter. 1984. utility of socio-economic research in wildlife management. transactions of the north american wildlife and natural resources conference 49:43-53. bisset, a. r. 1991. the moose population of ontario revisited – a review of survey data, 1975–1991. ontario ministry of natural resources, toronto, ontario, canada. bottan, b. j. 1999. exploring the human dimension of thunder bay moose hunters with focus on choice behaviour and environmental preferences. m.sc.f. thesis, faculty of forestry and the forest environment, lakehead university, thunder bay, ontario, canada. , l. m. hunt, w. haider, and a. r. rodgers. 2001. thunder bay moose hunters: environmental characteristics and choice preferences. ontario ministry of natural resources, thunder bay, ontario. cnfer technical report tr-007. crête, m. 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67:373-380. , r. j. taylor, and p. a. jordan. 1981. optimization of moose harvest in southwestern quebec. journal of wildlife management 45:598-611. crichton, v. 1992. management of moose populations: which parameters are used? alces supplement 1:11-15. cumming, h. g. 1958. geraldton district plan for a statistically sound aerial moose survey. 22nd federal provincial wildlife conference, ottawa, ontario, canada. . 1974. moose management in ontario from 1948 to 1973. canadian field-naturalist 101:673-687. decker, d. j., and j. w. enck. 1996. human dimensions of wildlife management: knowledge for agency survival in the 21st century. human dimensions of wildlife 1:60-71. , and m. e. richmond. 1994. managing people in an urban deer environment: the human dimensions challenges for managers. pages 3-10 in j.b. mcaninch, editor. urban deer: a manageable resource? 55th midwest fish and wildlife conference, st. louis, missouri, usa. december 12-14, 1993. eason, g. 1985. overharvest and recovery of moose in a recently logged area. alces 21:55-75. . 1989. moose response to hunting and 1 km2 block cutting. alces 25:6374. , r. j. thomas, and k. oswald. 1981. moose hunting closure in a recently logged area. alces 17:111-125. euler, d. 1982. a moose habitat strategy for ontario. alces 18:180-192. . 1983. selective harvest, compensatory mortality and moose in ontario. alces 19:48-61. fowle, c. d., and h. g. lumsden. 1958. aerial censusing of big game with special reference to moose in ontario. presented at meeting of canadian wildlife biologists, ottawa, ontario, canada. gibbs, j. p., h. l. snell, and c. e. causton. 1999. effective monitoring for adaptive wildlife management: lessons from the galapagos islands. journal of wildlife management 63:1055-1065. hansen, s., w. j. dalton, and t. stevens. 1995. an overview of a hunter opinion survey of satisfaction with the ontario moose management system. alces 31:247-254. heydon, c., d. l. euler, h. smith, and a.r. bisset. 1992. modeling the selective moose harvest program in ontario. alces 28:111-121. alces vol. 38, 2002 bottan et al. adaptive management of ontario moose 9 hilborn, r. 1992. canadian fisheries agencies learn from experiences? fisheries 17:6-14. , c. s. holling, and c. j. walters. 1979. managing the unknown: approaches to ecological policy design. biological evaluation of environmental impacts. council on environmental quality and fish and wildlife service, u.s. department of the interior. fws/ obs-80/26. washington, d.c., usa. holling, c. s., editor. 1978. adaptive environmental assessment and management. john wiley & sons, new york, new york, usa. lautenschlager, r. a., and r. t. bowyer. 1985. wildlife management by referendum: when professionals fail to communicate. wildlife society bulletin 13:564-570. mcallister, m. k., and r. m. peterman. 1992. decision analysis of a largescale fishing experiment designed to test for genetic effect of size-selective fishing on british columbia pink salmon (onchorynchus gorbuscha). canadian journal of fisheries and aquatic science 49:1305-1314. mckenney, d. w., r. s. rempel, l. a. vernier, y. want, and a. r. bisset. 1998. development and application of a spatially explicit moose population model. canadian journal of zoology. 76:1922-1931. (omnr) ontario ministry of natural resources. 1980. moose management policy. wm.3.01.02. ontario ministry of natural resources. toronto, ontario, canada. . 1988. timber management guidelines for the provision of moose habitat. ontario ministry of natural resources, toronto, ontario, canada. . 1990. the moose in ontario. ontario ministry of natural resources, toronto, ontario, canada. provincial auditor. 1998. audit of the ministry of natural resources, fish and wildlife program. queen's printer. toronto, ontario, canada. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. ringold, p. l., j. alegria, r. l. czaplewski, b. s. mulder, t. tolle, and k. burnett. 1996. adaptive monitoring design for ecosystem management. ecological applications 6:745-747. rodgers, a. r., r. s. rempel, and k. f. abraham. 1996. a gps – based telemetry system. wildlife society bulletin 24:559-566. simmons, g. 1997. independent review of the moose and deer tag allocation for ontario: recommendations from ontario hunters. queens printer, toronto, ontario, canada. taylor, b., l. kremsater, and r. ellis. 1997. adaptive management of forests in british columbia. province of british columbia, victoria, british columbia, canada. thompson, i. d., and r. w. stewart. 1998. management of moose habitat. pages 377-401 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. timmermann, h. r., and m. buss. 1998. population and harvest management. pages 559-615 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , and r. gollat. 1983. age and sex structure of harvested moose related to season manipulation and acadaptive management of ontario moose – bottan et al. alces vol. 38, 2002 10 cess. alces 18:301-328. , and . 1984. sharing a moose in north central ontario. alces 20:161-183. , and . 1986. selective moose harvest in north central ontario a progress report. alces 22:395-417. , , and h. a. whitlaw. 2002. reviewing ontario’s moose management policy – 1980-2000 targets achieved, lessons learned. alces 38:11 45. , and r. s. rempel. 1998. age and sex structure of hunter harvested moose u n d e r t w o h a r v e s t s t r a t e g i e s i n northcentral ontario. alces 34:21-30. , and h. a. whitlaw. 1992. selective moose harvest in north central ontario a progress report. alces 28:137-163. , , and a. r. rodgers. 1993. testing the sensitivity of moose harvest data to changes in aerial population estimates in ontario. alces 29:47-53. walters, c. j. 1986. adaptive management of renewable resources. mcgrawhill, new york, new york, usa. 4210(65-74).pdf alces vol. 42, 2006 stimmelmayr et al. oral health of alaskan moose 65 incisor tooth breakage, enamel defects, and periodontitis in a declining alaskan moose population r. stimmelmayr1,2, j.a.k. maier3, k. persons4, and j.battig5 1community and natural resources department, tanana chiefs conference, 122 first ave., fairbanks, ak 99701, usa; 3institute of arctic biology, university of alaska fairbanks, fairbanks, ak 99709, usa; 4division of wildlife conservation, alaska department of fish & game, nome, ak 99762, usa; 5chena ridge veterinary clinic, fairbanks, ak 99709, usa abstract: we examined 56 anterior segments of mandibles from moose harvested from a declining population that was affected by tooth wear and breakage at higher rates than in moose elsewhere on the seward peninsula, alaska. incisor teeth were examined for extent of tooth wear and breakage, the degree and prevalence of surface enamel defects, and radiographic evidence of periodontitis. body size (incisor arcade width of adult moose) and body condition index (timing of tooth eruption in yearlings) of the seward peninsula population were compared to other alaskan moose populations. mean (± se) age of adult moose in the study was 4.6 ± 0.4 years. the age distribution of harvested moose was 32% yearling, 61% young adult (2-6 years old), 4% prime adult (7-11 years old), and 4% old moose the absence of older animals in the 2002 harvest. timing of tooth eruption in yearlings was within the range of other moose populations. mean tooth wear and breakage score was 2.1 ± 0.2. ninety-three percent of the teeth exhibited hypoplastic enamel defects (pits) and staining, while 59% exhibited vertical and horizontal fracture lines on both labial and lingual tooth surfaces. fifty-three percent of examined teeth showed radiographic signs consistent with periodontitis. evidence of osteoporosis was present in 74% of the examined jaws. we hypothesize that observed enamel defects exacerbate age-related tooth wear and breakage in this population thereby resulting in accelerated demise of older animals. the skewed age distribution, with very few animals > 7 years supports this conclusion. the etiology of the observed enamel defects is unclear and requires further investigation. alces vol. 42: 65-74 (2006) key words: alaskan moose, alces alces gigas, mandibular bone loss, moose, periodontitis, tooth wear and breakage many rural residents in alaska depend heavily on wild animals such as moose (alces alces) for their meat supply. in the late 1980s the alaska department of fish & game began collecting general harvest data on the moose population; hunters were required to report sex, age, and location of kills and submit the distal portion of jaws from harvested moose on the seward peninsula. department biologists found that the teeth were characterized by excessive breakage and wear (smith 1992). the seward peninsula moose population has been declining precipitously for more than 15 years, the cause(s) of which is unclear at productivity are dependent on oral health and tooth integrity such that a decline in chewing with age-related tooth wear in cervids (rangifer tarandus: tyler 1986, skogland 1988, kojola et al. 1998; cervus elaphus: perezbarberia and gordon 1998). certainly the productivity of an ungulate population may be negatively affected by excessive tooth wear 2present address: ross university school of veterinary medicine, department of structure and function; p.o. box 334, basseterre, st. kitts, west indies; rstimmelmayr@rossvet.edu.kn oral health of alaskan moose stimmelmayr et al. alces vol. 42, 2006 66 and breakage. we report the extent of incisor tooth wear and breakage, degree of surface enamel defects, and prevalence of periodontitis in 56 anterior segments of moose mandibles submitted by hunters on the seward peninsula in 2002. we also compare estimated body size of seward moose to moose of interior alaska using the distance between the buccal surfaces of canines (arcade width) in adult moose. in yearlings we used timing of tooth eruption as a proxy indicator of body condition. study area the seward peninsula, alaska, usa (nome latitude 64° 25’, longitude 165° 30’), is underlain by permafrost and is characterlowlands (usda forest service, http://www. fs.fed.us/colorimagemap/images/m125.html). soils are poorly drained, and wetlands comprise 53 % of the area. there are many inland and coastal lakes and ponds. vegetation includes moist-tundra, sedge-tussock meadows, sparse willows, birch, and isolated sprucehardwood forests, particularly along rivers; while alpine-tundra heath-meadows and barrens dominate the high mountains. climate is characterized by long, cold winters (average -24 to -19°c), and short, cool summers (+1 to +6°c), with heavy annual snowfall (1,000 to 2,000 mm) and rain (460 mm). occurrence june through august. human population is low and dispersed, with the largest settlement being nome. old and new mining enterprises exist throughout the region. methods hunters submitted 56 anterior segments of moose mandibles (adults: 36 males, 2 females; yearlings: 17 males, 1 female) during the 2002 hunting season (august-september). these specimens were frozen at -10°c within 24 hours of collection and kept frozen for 5 months before analysis. a lower incisor was extracted in adult animals (n = 38) for determination of age (wolfe 1969). age was estimated from tooth root annuli by matson’s laboratory (matson’s laboratory, lcc 2001; p.o. box 308, milltown, mt 59851, usa). body size and condition we used incisor arcade width (the distance between the buccal surfaces of the incisiform canines, i4) as a proxy of body size for adult moose. in moose, incisor arcade width is positively correlated to body mass (spaeth et al. 2001). arcade width was measured to the nearest 1 mm with a dial caliper (spaeth et al. 2001). lower jaws of adult female moose (n = 2) and jaws of juveniles with missing canines (n = 3) were not included in the analysis. status of tooth eruption was determined in moose < 2 years. timing of tooth eruption has been proposed to be linked to body condition in cervids (odocoileus hemionus: robinette et al. 1957; alces alces: peterson et al. 1983; cervus elaphus: loe et al. 2004). individual teeth were described as missing, deciduous, erupting, or permanent. tooth wear and breakage individual mandibular incisors were ments of adult moose mandibles were grouped according to tooth wear and breakage as deincisor teeth were scored on a scale of 1-5 for wear and breakage (class 1: < 15% of crown missing; class 2: 15-25% of crown missing; class 3: 25-35% crown missing; class 4: 3550% crown missing; class 5: > 50% crown gone, root canal exposed). macro-enamel defects both enamel tooth surfaces, lingual, the surface facing the tongue, and labial, the surface facing the lip, were macroscopically examined for staining, fracture lines, and differences in enamel thickness (enamel hypoplasia) such as pits, vertical and horizontal furrows, and plane defects (goodman and alces vol. 42, 2006 stimmelmayr et al. oral health of alaskan moose 67 normal, pitted, or sprinkled, and percentage of tooth area pitted and stained (< 10, 11-30, 31-50, 51-75, 76-100%) was recorded. a normal tooth showed smooth, translucent, glossy white enamel; a pitted tooth had white surface enamel with multiple, randomly distributed pits; a sprinkled tooth had a discrete and subtle appearance of uniformly scattered, brown-black staining and discrete pits; and class 10-100% described the extent of discrete total crown surface. periodontitis radiographs were taken of 37 adult jaws (36 males; 1 female) (transworld® 325 v) to determine prevalence of periodontitis (huumonen and orstavik 2002). mandibles were 35 x 43 cm) to maximize accuracy and sharpness of x-rays. right and oblique angle views were taken of each jaw (50 kv and 300 ma). changes of alveolar bone were recorded for each tooth. width and height of radiolucent lesions (areas of bone loss as indicated by decreased radiodensity) were measured with a ruler to the nearest 1 mm and then categorized end opening, alongside the root). occurrence of root caries was assessed and root length, distance between cemento-enamel junction and root tip, was measured (mm). mandibular bone loss was assessed qualitatively by scoring general radiolucency (absent, mild, moderate, and severe). we initially measured pocket depth and recession using a standard dental probe on individual teeth to detect periodontitis. notwithstanding, we eventually excluded these measurements from our analysis because the mandibles had gone through repeated freezing and thawing during the examination and this may have affected gum recession and pocket depth. we also did not assess tooth mobility (indicator of periodontal disease). restricted tooth movement in ruminants is normal due to the low alveolar crest height (nickel et al. 1973). finally, we documented the presence and absence of impacted food. statistical analysis one-tailed t-tests were used to compare differences between means. we assumed independence of contra-lateral teeth (perzigian 1977) and therefore did not apply a bonferroni correction for comparison of contra-lateral incisiform measurements. data were examined for unequal variance using f-test for unequal variances prior to conducting t-tests. statistical software used was analysis toolpack (®mireported as means ± se. results mean (± se) age of adult moose in the study was 4.6 ± 0.4 years. the age distribution of harvested moose was 32% yearling (n = 18), 61% young adult (2-6 years old; n = 34), 4% prime adult (7-11 years old; n = 2), and 4% old moose (> 11 years old; n = 2). no food impaction was present in the examined jaws. body condition and body size tooth eruption pattern of yearlings is summarized in table 1. left incisor 2 was fully erupted in 50% (n = 9) of the jaws examined, while the right incisor 2 was fully erupted in only 39% (n = 7). mean incisor arcade width for male adults (5.75 ± 0.04 cm) was greater than that for male yearlings (4.69 ± 0.06 cm; t = 15.43; 49 df, p < 0.0001). dentition incisor 1 incisor 2 incisor 3 incisor 4 permanent 18;18 9;7 0 0 erupting 0 3;6 1;0 0 deciduous 0 6;5 17;18 18;15 missing 0 0 0 0;3 table 1. eruption status of dentition (left; right) in yearling (n = 18) moose of the seward peninsula, alaska, 2002. oral health of alaskan moose stimmelmayr et al. alces vol. 42, 2006 68 tooth wear and breakage short, shovel-like crowns with distinct necks and thick roots characterized permanent incisor teeth. mean (± se) tooth wear and of examined jaws by tooth wear and breakage was unbroken (18%; n = 7), class 1 (18%; n = 7), class 2 (21%; n = 8), class 3 (26%; n = 10), class 4 (8%; n = 3), and class 5 (5%; n = 2). an example of irregular incisor tooth wear and breakage with multiple lingual and labial surface fracture lines, tooth staining, and partially missing crowns is depicted in figure 1. fracture lines and hypoplastic enamel defects data on labial and lingual surface fracture lines are summarized in table 2. both labial and lingual fracture lines on the same tooth were present in 59% (n = 201) of examined teeth (n = 339), with i4 having the least (table 2). fracture lines were absent in 18% (n = 61) of examined incisors. yearlings were the dominant age group without fracture lines (left (l) = 75%; n = 13; right (r) = 91%; n = 16) for i1; (l = 40%; n = 4; r = 91%; n = 6) for i2; the one yearling who had an i3 contained no fractures in that tooth; no yearlings had an i4, as their incisiform canines had not emerged. all teeth had varying degrees of staining and pitting (table 3). sixty percent of examined n = 199) pitting and discoloration of the total tooth surface enamel. no tooth discoloration and pitting was present in 4% (n = 15) of the teeth and 3% (n = 11) had pitting with no stains. periodontitis fifty-three percent of incisors (n = 140) had radiolucent lesions (table 4; fig.2), and 84 % (n = 117) of these were at the root end opening (table 4). mean (± se) radiolucent area ranged between 13.3 ± 2.5 and 58 ± 21 mm2 (table 5). total radiolucent area was not different between contralateral incisors (table 5) as were root length (table 5). root caries was present in 13% (n = 32) of examined incisors (table 4). seventy-four percent of examined jaws exhibited signs of osteoporosis. distribution of osteoporosis scores of examined jawbones (n = 34) was: absent (n = 9), mild (n = 11), moderate (n = 10), and severe (n = 4). discussion mean age of the 2002 harvested animals fell within the range of mean age of moose harvested on the seward peninsula in previous years (3.1 ± 0.3 to 5.1 ± 0.4; stimmelmayr and maier, unpublished data). the skewed age distribution of harvested moose however, location incisor 1 incisor 2 incisor 3 incisor 4 (l, r) (l, r) (l, r) (l, r) labial & lingual 41, 35 38, 36 22, 23 3, 3 lingual 1, 2 0,1 10,12 22, 21 labial 3, 4 0,1 0,0 0,0 none 8,11 5, 6 4, 3 12, 12 table 2. distribution of labial and lingual surface fracture lines of incisors in moose of the seward peninsula, alaska, 2002. tooth staining & pitting incisor 1 incisor 2 incisor 3 incisor 4 (l, r) (l, r) (l,r) (l,r) normal 0,0 2, 4 2,3 2,2 pitted1 0,0 2, 1 2,2 2,2 sprinkled2 17,16 4, 7 3,4 3,4 <10 4,5 1,0 1,1 2,2 11-30 5,5 4, 2 5,3 6, 6 31-50 22,21 11, 14 10,6 14,9 51-75 3,4 13,9 11,15 5,9 76-100 1,0 5, 7 2,4 2, 2 table 3. distribution of degree of staining and pitting of incisiform teeth in moose on the seward peninsula, alaska, 2002. 1 a pitted tooth demonstrated white surface enamel with multiple, randomly distributed pits. 2 a sprinkled tooth had a discrete and subtle appearance of uniformly scattered brown-black staining and discrete pits. alces vol. 42, 2006 stimmelmayr et al. oral health of alaskan moose 69 indicates a young population with animals > 7 years old missing, suggesting decreased survivorship of older animals. over the last 10 years (1991-2001), percentage of harvested prime moose has decreased by about half while percentage of harvested young adult moose has increased (maier, unpublished data), lending support to this hypothesis. mean, incisor-arcade width for adult males was 5.75 ± 0.04 cm and fell within the lower range of reported values for alaskan moose (5.7 – 6.1 cantly smaller incisor width than did adults. et al. (2001) where the incisor-arcade width sexes of alaskan moose at their ultimate size. arcade width in moose is strongly correlated to body mass; we hypothesize that the small young age distribution in the 2002 harvest, although reduction in body size and low recruitment are well-known density-related trade-offs in moose populations (ferguson et al. 2000). the small body size of moose in our study may be due to density dependence, although the moose population on the seward peninsula is thought to be below the current carrying capacity of the habitat. data regarding moose habitat on the seward peninsula is lacking. timing of the eruption of incisor 2 was highly variable (50%) but comparable to other free-ranging alaskan moose populations from the kenai peninsula and nelchina basin (peterson et al. 1983). hence, our data on fig.1. incisor teeth (lingual view) with lingual-labial enamel fracture lines extending netlike over the entire crown surface and individual irregularly shaped teeth with sharp angles and contours in a young adult moose on the seward peninsula, alaska, 2000. the mean tooth wear score (± se) for this animal is 2.25 ± 0.73. variable incisor 1 (l, r) incisor 2 (l, r) incisor 3 (l, r) incisor 4 (l, r) radiolucent lesion present 28,26 18,14 13,19 11,11 absent 6,7 17,19 20,14 21, 22 root end opening 18,22 16,12 11,17 10,11 alongside the root 10,4 2,2 2,2 1,0 root caries1 absent 17,18 18,20 19,20 25,28 present 8,10 3,2 3,4 2,0 fuzzy 7,6 13,11 9,6 4,3 opening and alongside the root) of radiolucent lesion and associated root caries in moose on the seward peninsula, alaska, 2002. 1five (l) and six (r) could not be evaluated for root caries. oral health of alaskan moose stimmelmayr et al. alces vol. 42, 2006 70 tooth eruption do not support the notion of a decline in body condition in yearling moose on the seward peninsula. nonetheless, the is considerable and may limit interpretation when comparing among moose populations (peterson et al. 1983, loe et al. 2004). mean tooth wear and breakage score 2.1 ± 0.2 was lower than scores from previous years (2.6 ± 0.3 to 3.9 ± 0.4; stimmelmayr and maier, unpublished data). fewer animals, however, had zero breakage with 18.4% in this study compared to 29% in 1999. density dependence, mineral content of soil, and typical moose feeding behavior close to the ground are all factors that have been implicated as possible causes in age-related tooth wear for different moose populations (smith 1992). incisiform tooth wear and breakage in our population differs from the age-related tooth wear observed in other alaskan moose populations (stimmelmayr and maier, unpublished data). moose incisor tooth wear on the seward peninsula is characterized by individual irregularly shaped teeth with sharp angles and contours (fig.1), with lingual-labial enamel fracture lines extending netlike over the entire crown surface. in contrast, incisors from alaskan moose from kalgin island are characterized by smooth labial and lingual surface enamel with even wear on all incisors (fig. 3), while incisors from tanana flats are characterized by smooth labial and lingual surface enamel with marked interproximal u-shaped wear (fig. 4.); lingual-labial surface fracture lines are singular and vertically oriented. tooth conditions similar to moose from the tanana teeth left right statistics incisor 1 radiolucent lesion (mm2) 34 + 8 58 + 21 t = -1.07, 31 df, p = 0.15 root length (mm) 35.6 + 0.7 35.9 + 0.6 t = -0.2, 68 df, p = 0.39 incisor 2 radiolucent lesion (mm2) 17.09 + 3.8 49.06 + 30.07 t = -1.02, 8 df, p = 0.17 root length (mm) 35.2 + 0.6 35.3 + 0.6 t = -0.16, 67 df, p = 0.44 incisor 3 radiolucent lesion (mm2) 19.7 + 3.5 28.4 + 13.2 t = -0.64, 17 df, p = 0.26 root length (mm) 33.3 + 0.6 33.4 + 0.7 t = -0. 14, 67 df, p = 0.45 incisor 4 radiolucent lesion (mm2) 13.3 + 2.5 13.7 + 5.5 t = -0. 07, 9 df, p = 0.47 root length (mm) 32.1 + 0.6 33.1 + 0.6 t = -1.14, 66 df, p = 0.13 table. 5. mean (± se) radiolucent lesion area (mm2)1 and root length (mm) of incisiform teeth categorized by side of mouth in moose on seward peninsula, alaska, 2002. 1 width and height of radiolucent lesions were measured with a ruler to the nearest 1 mm. fig.2. radiographic example of several periodontal lesions (arrows l: incisor 1, 2, and 3) in moose on the seward peninsula, alaska, 2002. alces vol. 42, 2006 stimmelmayr et al. oral health of alaskan moose 71 flats have been described in canadian moose from manitoba (young and marty 1986). we hypothesize that the degree and extent of enamel defects exacerbate age-related tooth wear and breakage in our population. reduction of hardness and elasticity of hypoplastic enamel in comparison to normal enamel has been observed in humans (mahoney et al. 2004). development of enamel defects has been related to exposure to a variety of physiological stresses (i.e., disease, malnutrition) during the enamel formation period (goodman and rose, 1990, 1991). the underlying etiology of enamel defects observed in our study is unclear and requires further investigation. the majority of examined teeth had both lingual and labial surface fracture lines (59%), while only 18% of all teeth had neither. lingual fractures probably precede labial ones based on the distribution pattern of fracture lines for all incisors (table 3). the observed difference in enamel thickness on the lingual surface (thinner) in ruminants may be responsible for our observation (miles and grigson 1990). brown tooth discoloration and pitting were prevalent with only 4% of teeth having no staining at all and 3% having no stains, but pitting (table 3). in contrast to domestic ruminants (nickel et al. 1973), tooth surface fracture lines and brown tooth staining are considered to be typical in adult moose incisors. the consumption of tannin rich forage is most likely the cause of the brown-black discoloration suggested for humans and laboratory animals (nordbo 1977, nordbo et al. 1982). periodontitis can play a key role in exacerbating tooth wear and premature tooth loss, as well as compromise animal health due to systemic illness (aitchison and spence 1984, debowes 1998, gorrel 1998, duncan et al. 2003). fifty-three percent of teeth had lesions consistent with periodontitis. periodontitis, affecting premolar and molar, is frequently reported in captive bovids and free-ranging moose (peterson et al. 1982), and other ruminants (miles and grigson 1990), and has been linked to food impaction and associated soft-tissue injury (miles and grigson 1990). no signs of food impaction were found in our study. interestingly, root caries were present in only 13% of teeth, despite the high prevalence of periodontitis. seventy-four percent of the examined jaws exhibited signs of osteoporosis of the mandibular bone. osteoporotic circular skulllesions and periodontal lesions associated with molars have been previously reported in moose on isle royale (hindelang and peterson 1993, 1996, 2000). osteoporotic lesions were absent in isle royale moose younger than 7 years, but increased with age. males in that study had a higher prevalence of lesions than fig. 3. incisor teeth (labial view) with smooth labial and lingual surface enamel and even wear on all incisors in moose on kalgin island, alaska, 2002. fig.4. incisor teeth (labial view) with smooth labial and lingual surface enamel and distinct interproximal u-shaped wear in moose from the tanana flats, alaska, 2002. oral health of alaskan moose stimmelmayr et al. alces vol. 42, 2006 72 did females. the prevalence rate of 32% in isle royale moose was considerably lower than we found (74%). metabolic stressors (e.g., antler growth, gestation, and lactation) have been implicated as potential causative factors for osteoporosis in isle royale moose (hindelang and peterson 1993, 1996, 2000), while environmental exposure to heavy metals (lead) is thought to play a role in the increased incidence of osteoporosis in moose in norway (bjora et al. 2001). alternative explanations for the increased rate of bone loss include inbreeding and a founder effect (recker and deng 2002, seeman 2003). tooth wear and breakage, periodontitis, hypoplastic enamel defects, and mandibular bone loss are common in moose of the seward peninsula. percentage of moose jaws affected by osteoporosis exceeds what is reported for moose on isle royale, michigan. it is unclear at this point whether periodontitis, mandibular bone loss, and tooth wear and breakage are related, but it is highly likely. independent of underlying etiological mechanisms (i.e., ciency, heavy metals) of the complex dental pathology in moose on the seward peninsula, we hypothesize that the degree and extent of enamel defects exacerbate their age-related tooth wear and breakage, thereby resulting in accelerated mortality in older animals > 7 years. the skewed age distribution in our harvest lends support to this hypothesis. acknowledgements this work was funded in part by a grant from the u.s. bureau of indian affairs (bia). we are indebted to the residents of the seward peninsula for their continued interest, involvement, and willingness to provide us with samples via the alaska department of fish & game. we thank drs. r.t. bowyer, k.j. hundertmark, v. van ballenbergh for reviewing an earlier draft of this manuscript, and the anonymous reviewers, who greatly improved m. sullivan for expert editing. references aitchison, g.u., and j.a. spence. 1984. dental disease in hill sheep: an abattoir survey. journal of comparative pathology 94: 285-300. bjora, r., j. a. falch, h. staaland, l. nordsletten, and e. gjengedal. 2001. osteoporosis in the norwegian moose. bone 29: 70-73. debowes, l. j. 1998. the effects of dental disease on systemic disease. veterinary clinic north america small animal practice 28: 1057-1062. duncan, w. j., g. r. persson, t. j. sims, p. braham, a. r. pack, and r. c. page. 2003. ovine periodontitis as a potential model for periodontal studies. cross-sectional analysis of clinical, microbiological, and serum immunological parameters. journal of clinical periodontology 30: 63-72. ferguson, s., a. r. bisset, and f. messier. and reproduction in moose alces alces. wildlife biology 6: 32-39. goodman, a. h., and j. c. rose. 1990. assessment of systemic physiological perturbations from dental enamel hypoplasias and associated structures. yearbook of physical anthropology 33: 59-110. _____, and _____. 1991. dental enamel hypoplasias as indicators of nutritional stress. pages 279-293 in m. a. kelley and c. s. larson, editors. advances in dental anthropology. wiley-liss incorporated, new york, new york, usa. gorrel, c. 1998. periodontal disease and diet in domestic pets. journal of nutrition 128(12 supplement): 2712s-2714s. hindelang, m., and r. o. peterson. 1993. relationship of mandibular tooth wear to gender, age and periodontal disease of isle royale moose. alces 29: 63-73. _____, and _____. 1996. osteoporotic skull alces vol. 42, 2006 stimmelmayr et al. oral health of alaskan moose 73 lesions in moose at isle royale national park. journal of wildlife diseases 32: 105-108. _____, and _____. 2000. skeletal integrity in moose at isle royale national park: bone mineral density and osteopathology related to senescence. alces 36: 61-68. huumonen, s., and d. orstavik. 2002. radiological aspects of apical periodontitis. endodontic topics i: 3-25. kojola, i., t. helle, e. huhta, and a. niva. 1998. foraging conditions, tooth wear and herbivore body reserves: a study of female reindeer. oecologica 117: 26-30. loe, l. e, e. l. meisingset, ø. a. mysterud, r. langvatn, and n. c. stenseth. 2004. phenotypic and environmental correlates of tooth eruption in red deer (cervus elaphus). journal of zoology, london 262: 83–89c. mahoney, e. k., f. s. m. ismail, n. kilpatrick, and m. swain. 2004. mechanical properties across hypomineralized/hypoplastic journal of oral science 112: 497-502. miles, a. e. w., and c. grigson. 1990. colyer’s variations and diseases of the teeth of animals. revised edition. university press, cambridge, u.k. nickel, r., a. schummer, e. seiferle, and w. o. sack. 1973. teeth. pages 75-97 in e. seiferle, editor. the viscera of the domestic mammals. verlag paul pareyspringer verlag, berlin, germany. nordbo, h. 1977. discoloration of dental pellicle by tannic acid. acta odontologica scandinavia 35: 305-310. _____, h. m. er i k s e n, g. ro l l a, a. attramadal, and h. solheim. 1982. iron staining of the acquired enamel pellicle after exposure to tannic acid or chlorhexidine: preliminary report. scandinavian journal of dental research 90: 117-123. perez-barberia, f. j., and i. j. gordon. 1998. area on the voluntary intake, digestion, chewing behavior and diet selection of red deer (cervus elaphus). journal of zoology 245: 307-326. perzigian, a. j. 1977. fluctuating dental asymmetry: variation among skeletal populations. american journal of physiological anthropology 47: 81-88. peterson, r. o., j. m. scheidler, and p. w. stephens. 1982. selected skeletal morphology and pathology of moose from the kenai peninsula, alaska and isle royale, michigan. canadian journal of zoology 60: 2812-2817. _____, c. c. schwartz, and w. b. ballard. 1983. eruption patterns of selected teeth in three north american moose populations. journal of wildlife management 47: 884-888. recker, r. r., and h. w. deng. 2002. role of genetics in osteoporosis. endocrine 17: 55-66. robinette,w. l., d. a. jones, g. rogers, and j. s. gashwiler. 1957. notes of tooth development and wear for rocky mountain mule deer. journal of wildlife management 21: 134–153. seeman, e. 2003. invited review: pathogenesis of osteoporosis. journal of applied physiology 95: 2142-2151. skogland, t. 1988. tooth wear by food limitation and its life history consequences in wild reindeer. oikos 51: 238-242. smith, t. e. 1992. incidence of incisorform tooth breakage among moose from the seward peninsula, alaska, usa. alces supplement 1: 207-212. spaeth, d. f., k. j. hundertmark, r. t. bowyer, p. s. barboza, t. r. stephenson, and r. o. peterson. 2001. incisor arcades of alaskan moose: is dimorphism related to sexual segregation? alces 37: 217-226. tyler, n. j. c. 1986. the relationship between the fat content of svalbard reindeer in autumn and their death from starvaoral health of alaskan moose stimmelmayr et al. alces vol. 42, 2006 74 tion in winter. rangifer special issue 1: 311–314. wolfe, m. l. 1969. age determination in moose from cementum layers of molar teeth. journal of wildlife management 33: 428-431. young, g. y., and t. m. marty. 1986. wear and microwear on the teeth of a moose (alces alces) population in manitoba, canada. canadian journal of zoology 64: 2467-2479. alces vol. 44, 2008 hickey re-colonization of moose in new york 117 assessing re-colonization of moose in new york with hsi models lisa hickey columbia university department of ecology, evolution, and environmental biology, 10th floor schermerhorn extension, 1200 amsterdam avenue, new york, new york, usa 10027 abstract: after nearly a century of decline and range contraction in the northeastern united states, moose (alces alces) have re-colonized adirondack park, new york due to improved habitat and adjacent source populations. in this paper i present the results of 2 habitat suitability index (hsi) models used to examine the pattern of moose recovery in adirondack park. sighting data collected in 1980-1999 by the new york state department of environmental conservation were used to compare moose locations with 3 suitability levels of moose habitat predicted by the hsi models. the 2 models indicated that most of adirondack park was a combination of suitable (49-73%) and most suitable habitat (10-35%) for moose; the majority (53-77%) of sightings occurred in suitable habitat. however, the distribution of moose locations derived from sighting data might have been influenced by where human recreational activity occurred because sighting locations were not well correlated with the most suitable habitat. the combined analysis of the sighting locations and the hsi models provided valuable insight into the current and potential occupation and distribution of moose in adirondack park. alces vol. 44: 117-126 (2008) key words: adirondack park, alces alces, habitat suitability index, habitat model, population. how species select and use habitat forms one of the central fields of inquiry in ecology (guisan and zimmerman 2000). determining how and when animals use different habitats provides information about their distribution and abundance relative to a given landscape (fielding and bell 1997, pearce and ferrier 2001, anderson et al. 2002). several methods have been used to investigate species-habitat relationships including statistical methods based on empirical data, expert-based knowledge, and a variety of modeling approaches (heglund 2002, anderson et al. 2003). species distribution models can be a useful tool to predict distribution of a species by relating records of species presence and absence to environmental factors (elith et al. 2006). predictive geographic modeling uses the species-environment relationship in an attempt to understand potential species distributions, and has been applied to habitat use by moose (alces alces) for nearly 2 decades in the form of habitat suitability index (hsi) models (allen et al. 1987, koitzsch 2002, snaith et al. 2002, dussault et al. 2006). it is important to identify factors that influence habitat selection by moose to understand patterns of habitat use and choose appropriate model parameters. habitat preferences are largely related to forage and cover requirements (dussault et al. 2006). habitat requirements and preferences change seasonally, but are comprised of a few basic elements including variable and patchy forests that contain young and old deciduous and coniferous cover, and wetlands and water environments (snaith et al. 2002). open or disturbed areas within mature hardwood forests provide early successional vegetation that is a diet staple in the growing season, while conifers provide food and shelter during winter (peek 1998, snaith et al. 2002). wetlands provide important sources of minerals and other nutrients, and also provide shelter from predators, insects, and high ambient temperature (peek 1998, snaith et al. 2002, dussault et al. 2006). re-colonization of moose in new york hickey alces vol. 44, 2008 118 historically moose were found throughout new york including the adirondack park (park) prior to european colonization (hicks 1986, karns 1998). european immigrants cleared and converted much of the forestland to farmland in new england and new york. this dramatic land use change and increasing human settlement led to higher hunting pressure and reduction of forest habitat that caused the decline and extinction of local moose populations (alexander 1993, bontaites and gustafson 1993, foster et al. 2002). the last record of a moose shot in the park was in 1861 (jenkins and keal 2004). populations in neighboring states of vermont and massachusetts were either extirpated or unable to serve as viable source populations due to habitat fragmentation (alexander 1993, foster et al. 2002). periodic, unsuccessful reintroductions of moose occurred in northern new york in the 1870s-early 1900s (hicks 1986). eventually, lack of hunting combined with recovery of forests and wetlands, the latter due to resurgent beaver (castor canadensis) populations, created improved moose habitat in the park by the late 1900s. moose dispersing from northern new hampshire began to occupy adjacent vermont in the 1960s (koitzsch 2002). subsequently, dispersing moose from vermont and southern canada provided the source animals that reestablished a moose population in the park and northern new york around 1980 (hicks 1986, hicks and mcgowan 1992); this population has expanded steadily (jenkins and keal 2004). in the first decade the population exhibited a highly skewed sex ratio of approximately 3 males:1 female, a ratio typical of newly established mammal populations (garner and porter 1990). it is estimated that ~400 moose reside in the park currently (ed reed, new york state department of environmental conservation, pers. comm.). as moose established a resident population in the park, the new york state department of environmental conservation (dec) monitored moose numbers and movements (hicks 1986). dec used public sightings of moose to monitor population trends in 1980-2004. sightings were solicited through newspaper articles and were accepted from individuals calling dec offices. the information included date of sighting, location or general area of sighting, length of sighting, behavior observed, age, sex, type of sign observed (animal, track, scat, carcass), additional comments, and the name and contact information of the observer. sightings were assigned latitude and longitude coordinates and an approximate elevation (increments of 100 m) based on the reported area of sighting. other data sources included aerial surveys and locations of radio-collared moose. these efforts slowed considerably after 1999, consequently, my analysis only included data prior to 1999 (fig. 1). hsi models were first developed by the united states fish and wildlife service (usfws) in the late 1970s and early 1980s (koitzsch 2002); they were derived from expert knowledge or process-based models (ray and burgman 2006). the models use habitat variables related to species presence, abundance, and distribution (e.g., habitat used for food and shelter) and represents them as quantifiable measures of suitability with values ranging from 1 (optimal) to 0 (not suitable) (koitzsch 2002, dettki et al. 2003, ray and burgman 2006). variables can be combined using different equations and assigned to discrete units of the landscape. they can be used to predict species distribution and effects of habitat change on species presence and distribution (dettki et al. 2003), and are useful for conservation planning and resource management (pearce and ferrier 2001, koitzsch 2002). allen et al. (1987) developed 2 models for moose based on suitable habitat in the lake superior area of northern minnesota. these models were based on expert knowledge and selection of habitat variables based on research of peek et al. (1976) in northern minnesota. alces vol. 44, 2008 hickey re-colonization of moose in new york 119 they have been widely applied and have provided the basis for the development of recent models to inventory moose habitat and document changes therein (koitzsch 2002, snaith et al. 2002, dussault et al. 2006). the hsi model i was based on a high resolution analysis of a small area (~600 ha) that included the annual habitat requirements of moose. the model was designed to analyze requisite food and cover and requires detailed information about browse availability. the hsi model ii was designed to rapidly assess the potential of larger areas to provide suitable browse and cover for moose; remote sensing data of coarse resolution can be used in this analysis (allen et al. 1987). it has been used recently to evaluate moose habitat in vermont (koitzsch 2002) and nova scotia (snaith et al. 2002). this study was designed to apply hsi model ii in adirondack park to assess the availability and suitability of habitat for moose. this habitat information and the moose sighting data collected by dec were evaluated to assess the relationship between sightings and relative habitat suitability. these exercises should provide information useful for management of moose in adirondack park. study area the study area was the park, 23,876 km2 of state and privately owned land in northern new york (fig. 2). the park was created in 1892 and its core zone is the forest preserve, an 11,331 km2 area incorporating old-growth forest that represents a distinct biological area (jenkins and keal 2004). it is characterized by mountains and highlands intersected by river valleys. elevation is consistently high reaching nearly 1600 m, with valley floors about 100-200 m (jenkins and keal 2004). the region has a cooler climate than the immediately surrounding area, with average winter snow depth >2.5 m. the park is dominated by forests and wetlands that exist in an ecological tension zone between new england forests/ appalachian zone and canadian/boreal forests (jenkins and keal 2004). past and present logging has played an important role in shaping the composition of the commercial forestland representing 75% of the park (jenkins and keal 2004). it is a predominantly coniferous forest interwoven patchily with hardwood stands, except at higher elevations where conifers dominate. predominant species in the temperate forests include red spruce (picea rubens), red pine (pinus resinosa), white cedar (thuja occidentalis), black ash (fraxinus nigra), aspen (populus tremuloides), beech (fagus grandifolia), hemlock (tsuga canadensis), sugar maple (acer saccharum), white pine (p. strobus), and yellow birch (betula lutea), while 0 50 100 150 200 250 300 19 80 19 81 19 82 19 83 19 84 19 85 19 86 19 87 19 88 19 89 19 90 19 91 19 92 19 93 19 94 19 95 19 96 19 97 19 98 19 99 year r ep o rt ed p u b lic s ig h ti n g s fig. 1. moose sightings in the adirondack park reported by the public to the new york state department of environmental conservation, 1980-1999. re-colonization of moose in new york hickey alces vol. 44, 2008 120 the core boreal area of the park is comprised of balsam fir (abies balsamea), black spruce (p. mariana), aspen, tamarack (larix laricina), white spruce (p. glauca), and white birch (b. papyrifera). the other major category of land cover is wetlands that are common and as large as 200 ha. there are 3 major types of wetlands including open river corridors, floating bogs, and large open bogs sometimes dominated by conifers. open wetlands are diverse and change species composition frequently due to beavers that create and maintain high species richness within wetlands (wright et al. 2002). predators are limited to coyotes (canis latrans) and black bears (ursus americanus) both of which prey on moose calves (ballard and van ballenberghe 1998). methods habitat suitability index models the hsi model ii is a gis based model that uses 4 variables or classes of land cover considered important in habitat use and selection by moose (table 1) (allen et al. 1987). this model was applied to the park and data for each class of land cover were extracted from raster layers provided by the adirondack park agency and adirondack gis users group, vector and raster layers provided by the wildlife conservation society, and the usgs nlcd 2001 raster for the adirondack park. moose sighting data from 1980-1999 were digitized in excel, a corresponding shape file was created, and this was projected into nad 1983 utm 18n to correspond with the projection of pre-existing layers used in the gis analysis. data were analyzed using arcgis 9.1, the spatial analyst extension, and hawths tools. two raster files of land cover were used; the 1982 land cover layer for the area provided by the adirondack park agency and adirondack gis users group (), and the 2001 nlcd layer for the area obtained from the usgs. the 1982 land cover map was projected in albers conical area and had a cell size of 63.615 m; the 2001 land cover map was also projected in albers conical area and had a cell size of 30 m. in order to make both maps compatible, the 2001 land cover map was degraded to 63.615 meters using spatial analyst tools in arcgis 9.1. both layers were then re-projected into nad 1983 utm 18n to render them compatible with other layers used in the analysis. both layers were clipped to the extent of a vector layer conveying the boundary of the park known as the blueline established in 2001 (fig. 2). cells were selected that corresponded to the following 4 habitat variables of hsi model ii: new/regenerating hardwood forest, wetlands, softwood, and old and mixed hardwood forest (table 1). all other land cover classifications in the selected layers were coded as no data. two vector layers depicting damage due to wind and ice storms in the late 1990s fig. 2. the location of adirondack park in northern new york state. source populations of moose came from southern canada and the neighboring state of vermont. alces vol. 44, 2008 hickey re-colonization of moose in new york 121 were converted to raster layers with the same cell size, and areas of high and medium level damage to hardwoods were removed from the older hardwood layer and converted to new/ regenerating hardwood forest. habitat variable 1 – new/regenerating hardwood forest the data used to classify cells representing new/regenerating hardwood forest were derived from multiple sources. cells classified as hardwood forest in both the 1982 and 2001 land cover maps were selected and the resulting layers were overlaid; raster calculator was used to identify which areas defined as hardwood forest in 2001 were not defined as such in 1982. after identifying these cells, i applied layers identifying damage from a 1995 wind storm and a 1997 ice storm (k. didier, wildlife conservation society) to the new hardwood forest layer. i assumed that cells classified as medium and high damage would contain mostly regenerating forest. i selected those cells and overlaid the resulting layers on the 2001 land cover map to determine the area of hardwood in the damaged areas, and subsequently added them to the new/regenerating hardwood forest category. raster calculator was used to calculate the number and percent of cells in each grid square represented by new/ regenerating hardwood forest on the finished land cover map. habitat variable 2 – wetlands cells classified as wetland areas were extracted from the 2001 land cover map because the area has not yet been completely classified by the usgs national wetland inventory or the adirondack park agency and adirondack gis users group. the 2001 land cover map had 2 categories of wetlands, woody wetlands and emerging herbaceous wetland. wetlands with woody vegetation were excluded because they don’t provide optimal food for moose (allen et al. 1987) consequently, the wetland layer only utilized cells categorized as emerging herbaceous wetland. the raster calculator was used to determine the number and percent of cells represented by wetlands in each grid square on the finished land cover map. habitat variable 3 – softwood cells classified as softwood forest were extracted from the 2001 land cover map. the raster calculator was used to determine the number and percent of cells represented by softwood forest in each grid square on the finished land cover map. habitat variable description (allen et al. 1987) description (nlcd 1982, 2001) optimum habitat suitability (% area) new/regenerating hardwood forest hardwood stands <20 yr old areas dominated by trees generally >5 m, and >20% of total vegetation cover. more than 75% of the tree pecies shed foliage simultaneously in response to seasonal change and area is <20 yr old. 40-50 emerging herbaceous wetlands emerging herbaceous wetlands emerging herbaceous wetlands 5-10 softwood forest conifer forest ≥20 years old with canopy cover >50% in evaluation area divided by total area areas dominated by trees generally >5 m tall, and >20% of total vegetation cover. more than 75% of the tree species maintain their leaves all year. canopy is never without green foliage. no age calculated. 5-15 old and mixed hardwood forests upland deciduous or mixed forests ≥20 years old. more than 25% of the canopy is older than ≥20 years old and composed of <50% canopy cover of conifers. areas dominated by trees generally >5 m tall, and >20% of total vegetation cover. more than 75% of the tree species shed foliage simultaneously in response to seasonal change and area is ≥20 years old. 35-55 table 1. description of the 4 habitat variables used in the hsi model ii (allen et al. 1987) and equivalent layers derived from national land cover datasets (1982, 2001), adirondack park. re-colonization of moose in new york hickey alces vol. 44, 2008 122 habitat variable 4 – old and mixed hardwood forest cells representing hardwood forest >20 years old and mixed hardwood forest were derived from multiple sources and combined to a single variable. cells classified as hardwood forest in both the 1982 and 2001 land cover maps were identified initially. these layers were overlaid and raster calculator was used to verify that areas defined as hardwood forest in 1982 remained as hardwood forest by definition in 2001. i applied layers indicating damage from the 1995 wind storm and 1997 ice storm to the old and mixed hardwood forest layers. i assumed that cells classified as medium and high damage should be classified as new/regenerating hardwood forest. i then reclassified (removed) the appropriate cells originally identified as hardwood forest on the 2001 land cover layer. raster calculator was used to determine the number and percent of cells represented by hardwood forest >20 years old and mixed hardwood forest in each grid square on the finished land cover map. habitat suitability i used the approach of koitzsch (2002) and calculated habitat suitability for moose in grid squares of 25 km2 that approximated the home range of moose in new england. the percent availability of each of the 4 variables was calculated for each grid square and results were applied to an equation that ranked the percentage of each variable according to hsi model ii (allen et al. 1987). weighted results were inserted into the equation: hsi = (si1 x si2 x si3 x si4) 1/4 grid squares were ranked for low, medium, and high suitability. optimal habitat was represented by cells with a value of 1.0 and less optimal habitat was represented by cells with value <1.0. results the hsi model ii indicated that suitable moose habitat occurred throughout the park (fig. 3 and 4). the spatial distribution of moose habitat was described in 3 gradations from “most suitable” to “least suitable.” two methods of classification were applied to the hsi data; an equal division of the grid squares into thirds based on hsi ranking (koitzsch 2002), and a division of the grid squares using a natural breaks (jenks) division. the natural breaks division groups data into classes that maximize the differences between classes; divisions are created in places where relatively big changes occur in the data. the percentages and distribution of “most suitable” and “suitable” habitat varied depending on the numerical analysis assigned to each category, but under 2 tested scenarios the area and percentage of “least suitable” habitat remained constant. when the 1026 hsi grid squares were divided into categories based on 3 equal divisions of the total possible values, there were179 grid squares (17%) of “least suitable” moose habitat (hsi = 0.0-0.31), 749 grid squares (73%) of “suitable” moose habitat (hsi = 0.32-0.66), and 98 grid squares (10%) of “most suitable” moose habitat (hsi = 0.67-1.0) (fig. 3). when the natural breaks division was used to categorize the hsi, there were159 grid squares (16%) of “least suitable” moose habitat, 503 grid squares (49%) of “suitable moose” habitat, and 361 grid squares (35%) of “most suitable” moose habitat (fig. 4). the relative proportions of the 3 gradations of habitat suitability (least suitable to most suitable) remained largely similar in area and spatially with both classifications. areas in the west, southwest, and northwest of the park were found to be “most suitable” when the grid squares were broken into thirds; these areas remained the same but expanded under the natural breaks division. areas in the northeast and the northwestern border of the park were found to be “least suitable” in alces vol. 44, 2008 hickey re-colonization of moose in new york 123 both classifications (fig. 3 and 4). moose sighting data (>1650 locations) were overlaid onto the hsi map and analyzed for presence in grid squares according to suitability under both classifications. more moose were sighted in the “most suitable” habitat under the natural breaks division (36.7%) than the other classification (10.7%); under both scenarios most moose were sighted in areas classified as “suitable” habitat (53.5 and 77.3%, respectively; fig. 3 and 4). sighting locations in “least suitable” habitat were 9.8 and 12%, or conversely, sighting locations in “suitable” and “most suitable” habitat combined were 90.2 and 88%, respectively. discussion although this exercise was done primarily to estimate suitable habitat for moose in the park, importantly, it also identified marginally suitable habitat that is important in management considerations of an expanding population. the “most suitable” habitat defined in hsi model ii is presumed to reasonably support 2 moose/km2 (allen et al. 1987). using this population density estimate, i extrapolated the “most suitable” habitat identified in the 2 classifications and calculated potential populations of approximately 4,900 (98 grid cells with 3 equal divisions) and 18,000 moose (361 grid cells with natural breaks). both of these estimates seem unreasonably high in comparison to well studied moose populations in vermont, new hampshire, and maine (5000-15,000 moose). nonetheless, the results indicate that there are large areas figure 3. habitat suitability index map of the adirondack park with moose sightings overlaid as points. grid squares are divided equal thirds based on hsi ranking. figure 4. habitat suitability index map of the adirondack park with moose sightings overlaid as points. grid squares are divided into thirds using a natural breaks (jenks) division. re-colonization of moose in new york hickey alces vol. 44, 2008 124 of suitable habitat for moose in the park and potential for continued population expansion and growth. there are no existing guidelines that pertain to choice of divisions for hsi models. obviously the highly varied results derived in this exercise could lead to different interpretations and management decisions. this variation could be related to a number of factors including differences in habitat relationships between minnesota and the park, classification discrepancies of suitable habitat, and flaws in the model and data. failure of the sighting location data to correlate well with the “most suitable” habitat may be partly explained by the confounding influence of human spatial patterns that biased moose sightings. all observation or presence/absence data were determined by human presence and hence are biased by sampling design. because the bias of human presence in the landscape dictates sightings, moose locations in certain areas may be grossly over or under sampled. there are several assumptions inherent in the hsi model that likely influenced the results. habitat models are often simplified to a limited number of factors that influence habitat selection by a given species leaving room for potential error by omission (dettki et al. 2003). several important elements influencing habitat use by moose were not incorporated in this hsi model including mineral licks, diversity of wetland cover types, winter cover, spatial patterns of interspersion of food and cover resources, and human disturbance (allen et al. 1987, koitzsch 2002). further, the model assumes that habitat is uniformly suitable on a year-round basis and that all 4 habitat variables are required throughout the year (allen et al. 1987, koitzsch 2002). however, moose are unlikely to use wetlands during winter, and do not necessarily require that 5-10% of their home range contain wetlands as long as suitable wetlands exist within a reasonable distance (peek 1998). these factors are likely to skew the model towards under-scoring habitat which may be highly suitable for use during all or part of the year, and could explain some of the discrepancy between sighting locations and areas of “most suitable” habitat. the nature of hsi models is that they are commonly applied to geographic areas where they were not developed and this could affect results and reliability (guisan and zimmerman 2000). further, correlation between sightings and suitable habitat based on the hsi may be based solely on pattern, rather than processes (dettki et al. 2003) that may be specific to an expanding population of moose, such as exists in the park. however, application of the hsi model in the park was unlikely to produce highly skewed results because forest composition, browse and cover species, and climate are reasonably similar in the park and the lake superior region of minnesota where the model was developed. two thirds of the moose sightings occurred within the 1892 blueline (fig. 3 and 4), the original boundary of the park that was also a biological boundary delineating a core ecological zone (jenkins and keal 2004). it is possible that the 1892 blueline is a good ecological delineation of core habitat for moose, and that the high number of sightings indicate that moose are re-colonizing and populating in this area faster. however, the majority of recreational trails in the park occur within the blueline, yet the main human population centers in the park lie outside this zone. therefore, sightings could have been influenced by the differences in location and habitual travel patterns of park residents versus those of recreational tourists who may have been more likely to report unique moose sightings. the sighting data collected by dec provided a 19 year picture of the increasing moose population in the park and was useful in developing citizen awareness and concern for its re-colonization. it indicated the trend in moose population and spatial and temporal patterns of moose re-colonization in the park alces vol. 44, 2008 hickey re-colonization of moose in new york 125 (jenkins and keal 2004). however, the lack of sampling design or measurable effort required to collect data means that the data should probably not be used to estimate population density and abundance or habitat use relationships (danielsen et al. 2005, mackenzie et al. 2006). although my analysis indicated that the sighting locations were not well correlated with the “most suitable” habitat as predicted by the hsi model, about 90% of sightings did occur in locations described as suitable or better habitat. further, the hsi model classified the majority of the park as suitable or better moose habitat (about 84%), and it also identified “least suitable” habitat, valuable information relative to population and habitat management. the results of this hsi model could be combined with a more rigorously designed sampling regime to ascertain more specific information about abundance and distribution of moose in the park. nevertheless, the combined analysis of the sighting locations and the hsi model provided valuable insight into the current and potential occupation and distribution of moose in the park. acknowledgements i would like to thank drs. joshua ginsberg, kate mcfadden and samantha strindberg, richard pearson, and yuri gorokhovich for their support in developing the models and designing this study. drs. michale glennon and heidi kretser of the wildlife conservation society’s adirondack program provided extensive guidance and assistance in acquiring data. the new york state department of environmental conservation and particularly al hicks graciously provided moose sighting data. funding for this work was provided by joshua ginsberg. references alexander, c. 1993. the status and management of moose in vermont. alces 29: 187-196. allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. biological report 82(10.155), u.s. fish and wildlife service, fort collins, colorado, usa. anderson, r. p., a. t. peterson, and m. gomez-laverde. 2002. using nichebased gis modeling to test geographic predictions of competitive exclusion and competitive release in south american pocket mice. oikos 98: 3-16. _____, d. lew, and a. t. peterson. 2003. evaluating predictive models of species’ distributions: criteria for selecting optimal models. ecological modeling 162: 211-232. ballard, w. b., and v. van ballenberghe. 1998. predator-prey relationships. pages 247-274 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. bontaites, k. m., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29: 163-168. danielsen, f., n. d. burgess, and a. balmford. 2005. monitoring matters: examining the potential of locally-based approaches. biodiversity and conservation 14: 2507-2542. dettki, h., r. lofstrand, and l. edenius. 2003. modeling habitat suitability for moose in coastal north sweden: empirical vs. process-oriented approaches. ambio 32: 549-555. dussault, c., r. courtois, and j. p. ouellet. 2006. a habitat suitability index model to assess moose habitat selection at multiple spatial scales. canadian journal of forest resources 26: 1097-1107. elith, j., c. h. graham, r. p. anderson, m. dudik, s. ferrier, a. guisan, r. j. hijmans, f. huettmann, j. r. leathwick, a. lehmann, j. li, l. g. lohmann, b. a. loiselle, g. manion, c. moritz, m. nare-colonization of moose in new york hickey alces vol. 44, 2008 126 kamura, y. nakazawa, j. m. overton, a. t. peterson, s. j. phillip, k. richardson, r. scachetti-pereira, r. e. schapire, j. soberon, s. williams, m. s. wisz, and n. e. zimmermann. 2006. novel methods improve prediction of species’ distributions from occurrence data. ecography 29: 129-151. fielding, a. h., and j. f. bell. 1997. a review of methods for the assessment of prediction errors in conservation presence/absence models. environmental conservation 24: 38-49 foster, d. r., g. motzkin, d. bernardos, and j. cardoza. 2002. wildlife dynamics in the changing new england landscape. journal of biogeography 10-11: 1337-1357. garner, d. l., and w. f. porter. 1990. movements and seasonal home ranges of bull moose in a pioneering adirondack population. alces 26: 80-85. guisan, a., and n. e. zimmermann. 2000. predictive habitat distribution models in ecology. ecological modelling 135: 147-186. heglund, p. 2002. foundations of speciesenvironment relations in predicting species occurrence: issues of accuracy and scale. pages 35-42 in j. m. scott, p.j. heglund and m.l. morrison, editors. predicting species occurrences: issues of accuracy and scale. island press, washington, d. c., usa. hicks, a. 1986. the history and current status of moose in new york state. alces 22: 245-252. _____, and e. mcgowan. 1992. restoration of moose in northern new york state. draft eis. new york state department of environmental conservation. albany, new york. jenkins, j. b., and a. keal. 2004. the adirondack atlas: a geographic portrait of the adirondack park. syracuse university press, syracuse, new york, usa. karns, p. 1998. population distribution, density and trends. pages 125-139 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. koitzsch, k. b. 2002. the application of a habitat suitability index model to vermont wildlife management units. alces 38: 89-107. mackenzie, d. i., j. d. nichols, j. a. royle, k. h. pollock, l. l. bailey, and j. e. hines. 2006. occupancy estimation and modeling: inferring patterns and dynamics of species occurrence. academic press, amsterdam, netherlands. national environmental protection act of 1969. 42 u.s.c. § 4321. pearce, j., and s. ferrier. 2001. the practical value of modeling relative abundance of species for regional conservation planning: a case study. biological conservation 98: 33-43. peek, j. m. 1998. habitat relationships. pages 275-285 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. peek, j. m, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48:1-65. ray, n., and m. a. burgman. 2006. subjective uncertainties in habitat suitability maps. ecological modeling 195: 172-186. snaith, t. v., k. f. beazley, f. mackinnon, and p. duinker. 2002. preliminary habitat suitability analysis for moose in mainland nova scotia, canada. alces 38: 73-88. 3916v2art1.p65 alces vol. 40, 2004 dodge et al. michigan moose 71 survival, reproduction, and movements of moose in the western upper peninsula of michigan william b. dodge, jr.1, scott r. winterstein1, dean e. beyer, jr.2, and henry campa iii1 1department of fisheries and wildlife, michigan state university, 13 natural resources building, east lansing, mi 48824-1222, usa; 2wildlife division, michigan department of natural resources, marquette, mi 49855, usa abstract: moose were extirpated from the lower peninsula of michigan by the late 1800s. although it is not clear if moose were extirpated from the upper peninsula (up), the population was at the very least, reduced to a low level by ca 1900. attempts to re-establish a population of moose in the up during the mid-1930s failed. the michigan department of natural resources made a second attempt to reestablish moose by translocating animals from canada to the western up in 1985 and 1987. based on optimistic estimates of survival and reproductive rates and habitat surveys, a population of 1,000 moose was expected by the year 2000. however, aerial surveys conducted in the winters of 1996 and 1997 produced population size estimates that were well below 1,000. to determine possible reasons for the slower than expected population growth, 84 moose were outfitted with radio-collars in the winters of 1999-2001. the survival, reproduction, and movements of these moose and 12 others radio-collared in 1995 were monitored from january 1999-june 2001. overall, 1999-2001 pregnancy rates averaged 75%. annual adult survival rate (0.88) was higher than yearling survival rate (0.82). first-year calf survival rate (0.71) was high, relative to highly preyed on populations. annually, approximately 6% of radio-collared moose, primarily yearlings, dispersed out of the study area. the size of moose home ranges was typical of those found in the deciduous/coniferous ecotone of the upper great lakes region. migratory adult moose had larger annual home ranges than did non-migratory adult moose. low productivity appears to be the likely cause of the slower than predicted population growth. data from this study can be utilized to facilitate management of moose in the upper great lakes region. alces vol. 40: 71-85 (2004) key words: alces alces, dispersal, home range, michigan, moose, radiotelemetry, reproduction, survival, wildlife translocations prior to extensive european settlement, the eastern sub-species or taiga moose (alces alces americana) ranged throughout the upper great lakes region as far south as the northern ohio state line (de vos 1964). in michigan, moose ranged throughout the state, except for the southwestern portion of the lower peninsula (lp) (wood 1914, baker 1983). currently moose are only found in the upper peninsula (up) of michigan. moose probably never reached high densities in the lp because habitat quality at the southern periphery of moose range was marginal. habitat quality probably improved in the northern lp in the mid1800s because the large forest openings created by widespread logging had begun to regenerate. this improvement was shortlived, however, because vast areas of wildlife habitat were destroyed by the catastrophic firestorms that raged throughout michigan following the removal of timber (brewer 1991). moreover, continued degradation of habitat from expanding human settlement (e.g., conversion of logged-over land to farmland [whitney 1987]) and unregumichigan moose dodge et al. alces vol. 40, 2004 72 lated hunting resulted in the extirpation of moose from the lp by the mid-1880s (wood and dice 1923). the last credible sighting in the lp may have been john roger's report of a moose at black lake in presque isle county in 1883 (wood and dice 1923, baker 1983). to protect the remaining moose in the up, the michigan legislature banned moose hunting in 1899. moose persisted longer in the up because extensive timber harvesting and human settlement did not occur until several decades later than it did in the lp (hudgins 1953). despite legal protection, moose numbers dwindled and by the end of the 19th century moose may have been briefly extirpated from the up. besides human influences, factors such as wolf (canis lupus) predation and disease, for example, were speculated as contributing to their demise (verme 1984). the reported poaching of a yearling female in mackinac county in 1899 is the last documented record of a moose in the up in the 19th century (hickie 1944). during the early decades of the 20th century, moose were often reported in the eastern up (wood and dice 1923), however, it is not clear whether these were from a small remnant population or moose that periodically immigrated from ontario, canada. two attempts were made to reestablish moose in the up. in the winters of 19351937 the michigan department of conservation (mdc) live-trapped and shipped 69 adult moose (32 m, 37 f) from isle royale in lake superior to several locations in the western up (wup) (hickie 1937, 1944). sightings of moose increased after these releases, but by the end of world war ii the population had again declined. poor physical condition of the released animals and increased poaching because of food rationing during the war likely contributed to the decline (verme 1984). the second attempt occurred in january-february 1985 and february 1987 when the michigan department of natural resources (mdnr) (formerly mdc), in collaboration with the ontario ministry of natural resources, translocated 58 adult (>2.5-years of age) and 3 yearling moose (25 m, 36 f; 57 of which survived > 1 week) from algonquin provincial park, ontario, canada, to western marquette county, michigan. based on habitat surveys (wilton 1982) and rudimentary population growth projections, the objective of these translocations was to produce a population of 1,000 moose by the year 2000 (mdnr 1991). fifteen years after the translocations, moose are still present in the wup, however the population has not increased as rapidly as expected. estimates of the population size from aerial surveys conducted in the winters of 1996 and 1997 were 107 and 120 moose, respectively. also, moose population size estimates for 1996 and 1997 derived from a deterministic population model were both < 500 (1996, n = 452; 1997, n = 494). our study was initiated to determine why the moose population in the wup has not increased as rapidly as anticipated. the objectives were to: (1) determine productivity of moose; (2) estimate sexand agespecific survival rates of moose; and (3) estimate home range sizes and monitor movements of moose. study area the study was conducted in a > 3,000 km2 area in the wup of michigan that included portions of baraga, dickinson, iron, and marquette counties (fig. 1). the area was selected because it surrounds the 1985 and 1987 translocation release sites and harbors the greatest known density of moose in the up. the continental climate of this area is less moderated by the great lakes than is the rest of the up because it is bounded to the south by a large landmass alces vol. 40, 2004 dodge et al. michigan moose 73 (wisconsin) rather than a large body of water (lake michigan). the result is a wider variation in seasonal temperatures, colder winter temperatures, and a greater chance of summer thunderstorms. annual snowfall ranges from 102 to 356 cm (eichenlaub 1990). the underlying precambrian bedrock, most of which is covered with glacial deposits, is part of the canadian shield. soils are mostly acidic because the parent material lacks free lime (mccann 1991). the area lies within the deciduous/coniferous ecotone and is 90% forested, primarily in secondary-growth. northern hardwood forests lacking american beech (fagus grandifolia), except along the lake superior shoreline, dominate upland areas. drier sites support scattered pines (pinus resinosa, p. strobus) and a s p e n ( p o p u l u s t r e m u l o i d e s , p . grandidentata). a variety of wetlands occur where bedrock is at or near the surface, including conifer bogs dominated by black spruce (picea mariana) and tamarack (larix laricina), hardwood swamps dominated by black ash (fraxinus nigra), red maple (acer rubrum), and yellow birch (betula alleghaniensis), conifer swamps dominated by northern white cedar (thuja occidentalis) and tamarack, and speckled alder (alnus incana) thickets. moose are classified as a game species (aho et al. 1995), but are currently not hunted. potential predators of moose, primarily calves, include black bears (ursus americana) and wolves, however their impact on the population is considered by biologists as negligible. modern land uses include iron mining, recreation, and timber production. iron marquette baraga dickinson n lake superior minnesota lp of michigan up of michigan ontario wisconsin la ke m ic hi ga n 10 0 10 20 30 kilometers core moose range highways counties legend fig. 1. moose study area in the western upper peninsula of michigan during jan 1999-jun 2001. michigan moose dodge et al. alces vol. 40, 2004 74 methods capturing and radio-collaring moose were captured via net-gunning from a helicopter (hughes 500 or bell long ranger l-3, hawkins & powers aviation, inc., greybull, wyoming, usa) and fitted with 4-hour motion sensitive radio-collars (vhf: telonics, inc., mesa, arizona, usa; gps: lotek wireless inc., newmarket, ontario, canada) in january-february, 19992001. we classified moose by sex and identified age (adult, > 2-years of age; yearling, 12-23 months of age; calf, < 12-months of age) based on body size. to minimize stress during capture, moose were blindfolded and their ears plugged with foam rubber. to avoid injury, moose were processed as quickly as possible (average handling time: x = 26 min, range = 15-50 min) and their vital signs and behavior were closely monitored. pregnancy determinations blood samples were taken from the jugular vein at time of capture and assays of blood serum for pregnancy-specific protein b (pspb) (haigh et al. 1993, stephenson et al. 1995) were used to determine the pregnancy status of cows. because pspb has been shown to reliably detect pregnancy in moose 40 days after conception (huang et al. 2000), we assumed that cows with detectable levels of pspb were pregnant. in years subsequent to initial capture, the pregnancy status of cows was determined through assays of fecal material, collected during winter, for fecal progesterone (fp4) levels (monfort et al. 1993, schwartz et al. 1995). the fp4 levels of non-pregnant cows were used to establish a 95% upper tolerance limit (fp4-95% utl) for pregnancy (messier et al. 1990). radio-tracking and monitoring radio-collared moose were monitored throughout the study period from a cessna172, -182, or -206 aircraft equipped with radio-telemetry tracking equipment (i.e., side-facing, 2-element, yagi antennas mounted to each wing strut, connected by coaxial cable to a switchbox in the cockpit). survival monitoring of radio-collared animals was conducted at least once a week and we attempted to obtain at least 2 relocations (radio-fixes) per moose per month. at each relocation we recorded gps coordinates, time of day, whether the moose was seen, and activity if seen. in addition, all radio-collared cows were approached on the ground in the winter, to collect fecal samples, and during the calving period (15 may-30 jun [verme 1984]). following calving, cows that gave birth were approached on the ground at monthly intervals to assess survival of their calves. survival annual (1 jun-31 may), summer (1 may-31 oct), and winter (1 nov-30 apr) survival rates (with 95% cis) were estimated for adults, yearlings, cows (adult females), and bulls (adult males) using micromort (heisey and fuller 1985), which incorporates the mayfield survival estimator (mayfield 1961, 1975; trent and rongstad 1974). survival monitoring of moose radio-collared in 1999 began on 1 may; thereafter animals entered the study on the day they were radio-collared (i.e., staggered entry [see pollock et al. 1989]). to accommodate staggered entry and meet the constant survival assumption of the mayfield estimator, the biological year was divided into monthly intervals with a constant weekly survival rate. period survival rates were then equal to the product of the monthly survival rates (e.g., kkkkkkkkkkkkk kkkkkkk). censored animals (i.e., those from which radio signal contact was lost) were included in survival analysis up to the point at which they were censored (vangilder and sheriff 1990). winter nov dec aprs s s s...= × × × ) alces vol. 40, 2004 dodge et al. michigan moose 75 date of death (or censoring) was estimated at halfway between the last recorded live signal and the date that the moose was first known to be dead (or censored). first-year, 0-6 month (~1 jun-30 nov), and 7-12 month (~1 dec-31 may) survival rate estimates were calculated jointly for radio-collared calves and un-collared calves of radio-collared cows. calves not seen in the spring that were subsequently radiocollared during their first winter were not included in calf survival analysis. survival monitoring of un-collared calves began the day they were first observed. because survival monitoring of un-collared calves occurred once a month, all calves were assumed to have a constant monthly survival rate. a calf was considered to have died if its cow had either died or was found alone for two consecutive months prior to the calf attaining 8-months of age (the earlierst age of known cow-calf separation). dead calves were assigned the date of death of their cow or the date halfway between the last date the calf was seen with its cow and the date the cow was first seen alone. home range and movements annual (1 may-30 apr), summer, and winter home ranges of adult moose were determined with the animal movement analyst extension (amae; hooge et al. 1999) to arcview® (environmental systems research, inc., redlands, california, usa) geographic information system (gis). the 95% utilization distribution (ud) of the fixed kernel (fk; worton 1989) was used to estimate home range sizes. the smoothing factor was calculated via least squares cross validation (lscv). home range size estimates were determined only when > 18 relocations annually and > 9 relocations per season were available. because a minimum of 30 relocations per animal is recommended to accurately estimate home range area with the fk method when lscv is used, our results likely overestimate moose home range sizes (seaman et al. 1999). radio-fixes that deviated from grouped relocation points were considered transitory, and were not included in home range estimates. moose were classified as migratory if < 25% of their seasonal home ranges overlapped and there was > 2 km between the center of the fk 25% ud of their seasonal home ranges. because the use of gps radio-collars was experimental, home range sizes were determined only for moose with vhf radio-collars. annual dispersal rates (with 95% cis) were estimated using micromort (heisey and fuller 1985). a moose was considered to have dispersed if it permanently emigrated > 30 km straight line from its capture location or previous home range. this distance was chosen because it was greater than the maximum distance (26 km) a migratory moose moved between seasonal home ranges. data analysis the mann-whitney u-test was used to make comparisons between fp4 concentrations of pregnant and non-pregnant cows and among home range size estimates. comparisons between survival estimates where made with a z-test statistic (pollock et al. 1989) when > 25 moose per treatment (e.g., age, sex) were available (winterstein et al. 2001). unless otherwise noted, significance level for all statistical analyses was α = 0.05. results capturing and radio-collaring thirty-four adults (32 f, 2 m), 4 yearlings (4 f), and 36 calves (13 f, 23 m) were captured and radio-collared with standard vhf collars. in addition, gps collars were placed on 4 adult moose (2 f, 2 m) in 2000 and 5 adult moose (2 m, 3 f) and michigan moose dodge et al. alces vol. 40, 2004 76 one yearling (1 m) in 2001. twelve adult moose (6 f, 6 m) that were radio-collared in 1995 were also part of the initial sample population. no moose died or were injured during capture and no signs of capture myopathy (e.g., muscle stiffness, lethargy) were observed following capture in any year. radio-tracking and monitoring during feb 1999-jun 2001, we conducted 195 radiotracking flights and recorded 2,384 relocations of radio-tagged moose. moose with vhf radio-collars were relocated an average of 1.75 times per month whereas those with gps radio-collars were relocated 1.50 times per month. more aerial observations of radio-tagged moose were made per flight in the winter ( x = 5.04, range = 0-30) than in the summer ( x = 1.15, range = 0-11). radio-tagged cows were relocated on the ground an additional 312 times, during which 281 observations of moose (183 cows, 98 calves) were made. cows were approached between 0900 and 2300 hours, however, 87% of approaches occurred after midday. pregnancy determinations sixty-nine percent (25 of 36) of captured cows from which useable blood serum samples were collected had detectable pspb levels indicating pregnancy. mean and median pspb concentrations pooled across 2000-2001 were 411.05 ± 69.38 (se) and 387.30 ng/ml, respectively (1999 pspb values were unavailable because pspb results were reported as positive or negative only). we collected 111 fecal samples (36 at capture, 75 post capture) from 41 cows. multiple fecal samples were collected from 18 cows in 2000 ( x = 2.18) and 19 cows in 2001 ( x = 2.05). fp4 concentrations fell into fairly distinct pregnant and non-pregnant groups each year, although the results were not unequivocal (fig. 2). pooled fig. 2. progesterone concentration in fecal material collected in winter from radio-collared cow moose during 1999-2001 in the western upper peninsula of michigan. each symbol represents the mean fecal progesterone concentration of a single moose. the 95% upper tolerance limit (95% utl; horizontal line) between pregnant ( , , ) and non-pregnant ( , , ) cows was 5.17 mg/g. 0 5 10 15 20 25 30 35 year f ec al p ro ge st er on e (u g/ g) 1999 2000 2001 5.17 ug/g mean ± se and median values for pregnant cows were 16.29 ± 0.98 and 14.46 mg/g, whereas for non-pregnant cows the values were 2.30 ± 0.32 and 2.66 mg/g, respectively. fp4 concentrations of pregnant cows were significantly different from those of non-pregnant cows (mann-whitney u test, zmwu = -7.1730, p < 0.0001). the fp4-95% utl for pregnancy was 5.17 μg/g of dried feces. retroactively applying the fp4-95% utl to the average fp4 levels of individual cows, 92% (44 of 48) of pregnant cows (i.e., those with positive pspb results and/ or that gave birth) and 94% (17 of 18) of non-pregnant cows would have been correctly identified. the average annual pregnancy rate (adjusted for cows for which pregnancy test results were not available, but that were observed with calves) was 74% (table 1). reproduction of the pregnant cows (i.e., those with positive pspb or positive fp4 results, or alces vol. 40, 2004 dodge et al. michigan moose 77 table 1. productivity of radio-collared adult cow moose studied during 1999-2001 in the western upper peninsula of michigan. year no. cows % cows % cows no. % spring year end pregnant1 reproducing calves twins calf : cow calf : cow produced 1999-00 18 78 78 19 36 1.06 : 1 0.76 : 1 2000-01 27 70 67 19 6 0.70 : 1 0.60 : 1 2001-02 41 76 59 29 21 0.71 : 1 x or total 86 74 65 67 19 0.78 : 1 0.72 : 1 1cows with positive pspb or fp4 results, plus cows for which pregnancy tests were not available, but were observed with calves in the spring. note: year-end calf:cow for 2001-02 unavailable; study ended 30 june 2001. that reproduced if pregnancy test results were not available), 78% (14 of 18) in 1999, 67% (18 of 27) in 2000, and 59% (24 of 41) in 2001 were observed with at least 1 calf in the spring (table 1). overall, adult cows produced 19 calves in 1999 and 2000, and 29 calves in 2001. the earliest visual confirmation of calving was 21 may in 1999, 24 may in 2000, and 15 may in 2001. frequency of twinning varied from 6% to 36% ( x = 19%). post-calving calf:cow ratios (table 1) decreased from 1.06 in 1999 to 0.70 in 2000 and remained relatively unchanged in 2001 (0.71). due to the loss of a greater number of calves than cows, calf:cow ratios decreased 28% during 1999-2000 and 14% during 2000-2001 (table 1). the year-end calf:cow ratio for 2001-02 was not available because the study ended. during the study, only 1 yearling female reproduced, giving birth to a single calf in 2000. among the remaining 9 yearling females, 3 had negative pspb results, 3 were observed alone during spring natality checks, and 3 were unknown as regards to calving. survival of 96 moose (60 f, 36 m) that were monitored, 72 (46 f, 26 m) were known to be alive at the end of the study, 4 (2 f, 2 m) shed their radio-collars, 4 (2 f, 2 m) had their gps radio-collars removed to collect the data stored in each collar (the gps radio-collar of 1 male was replaced with a vhf radio-collar), and 17 (11 f, 6 m) died. three (2 f, 1 m) deaths were attributed to cerebrospinal nematodiasis (parelaphostrongylus tenuis). one male was killed by a motor vehicle and wolves killed a yearling female. eight (47%) moose that died were also heavily parasitized by the large american liver fluke (fascioloides magna). three (1 f, 2 m) moose that died during the winter had < 20% femur marrow fat (dry weight) indicating severe malnutrition (peterson et al. 1984). however, all 3 had additional maladies (e.g., p. tenuis and/or f. magna) that likely contributed to their deaths. other causes of death included accidents, birthing complications, and old age. ages of dead moose, estimated by counting cementum annuli of sectioned first incisors, excluding calves, ranged from 1.0 to 13.0 years for females ( x = 4.75 ± 1.24, n = 10) and 1.5 to 7.5 years for males ( x = 3.0 ± 1.12, n = 5). annual survival of adults did not differ (z = 0.089, p = 0.465) between 1999-00 ( ŝ = 0.880) and 2000-01 ( ŝ = 0.873). also, although more adults died in the winter (n = 7) than in the summer (n = 4) no difference michigan moose dodge et al. alces vol. 40, 2004 78 difficulty of detecting calves that died shortly after birth. first-year survival of calves in 1999-00 was 20% lower than that in 200001 (table 2). sixty-seven percent (4 of 6) of calf mortalities in 1999-00 and 25% (1 of 4) in 2000-01 occurred within the first six months of life. calf survival from 0-6 months was 20% lower in 1999-00 than in 2000-01. thirty-three percent (2 of 6) of calf mortalities in 1999 and 75% (3 of 4) in 2000 occurred between 7 and 12 months of age. seven to twelve month calf survival did not differ from 1999 to 2000. home range and movements in 1999-00, 22% (4 of 18) of adults migrated between distinct summer and winter home ranges, while in 2000-01, 38% (14 of 37) of adults migrated seasonally. migration distance ranged from 2 to 26 km ( x = 11 km). the median date of arrival on summer home range was 22 may and on winter home range it was 13 october. home range sizes (table 3) of migratory adults did not differ between winter and summer (zmwu = -1.5615, p = 0.118). annual home ranges of migratory adult moose were larger (zmwu = 2.4664, p = 0.014) than those of non-migratory adult moose. no differences were detected between home range sizes of cows and bulls (zmwu = 0.0102, p = 0.992). although, cows attended by calves had was found between winter and summer survival rates of adults in either year (1999-00: winter ŝ = 0.934, summer ŝ = 0.966, z = -0.554, p = 0.301; 2000-01: winter ŝ = 0.913, summer ŝ = 0.933, z = -0.357, p = 0.364). in 199900, annual survival of bulls ( ŝ = 1.000) was 16% higher than that of cows ( ŝ = 0.840), whereas in 2000-01, annual survival of cows ( ŝ = 0.882) was 2% higher than that of bulls ( ŝ = 0.857). the small sample of radio-collared bulls in 1999-00 (n = 9) was probably the reason that survival of bulls was so high that year. survival rates of cows, between years (199900 ŝ = 0.840, 2000-01 ŝ = 0.882, z = -0.449, p = 0.335) and between winter and summer in 2000-01 (winter ŝ = 0.911, summer ŝ = 0.937, z = -0.398, p = 0.351) were not significantly different. annual survival of yearlings was lower than that of adults in 1999-00 (yearling ŝ = 0.840) and 2000-01 (yearling ŝ = 0.800). in 1999-00, survival rates of yearlings between summer and winter were only slightly different (0.914 vs. 0.919), whereas, in 2000-01 summer survival of yearlings was 6% greater than in winter (0.924 vs. 0.867). seventy percent (7 of 10) of calf deaths occurred in the winter. this seasonally skewed calf mortality pattern, however, could be a result of the frequency (monthly) at which calf survival was checked and the table 2. first-year, 0-6 month, and 7-12 month joint survival rates of radio-tagged and un-tagged calf moose of radio-tagged cows seen each spring during 1999-2001 in the western upper peninsula of michigan. 1mayfield estimator (mayfield 1961, 1975; trent and rongstad 1974). no. no. no. 95% 95% year category ŝ 1 calves months deaths lci uci 1999-00 first-year 0.634 17 157 6 0.439 0.915 0-6 month 0.754 17 87 4 0.568 0.988 7-12 month 0.840 13 70 2 0.657 1.000 2000-01 first-year 0.787 20 203 4 0.622 0.995 0-6 month 0.943 20 102 1 0.838 1.000 7-12 month 0.835 19 101 3 0.678 1.000 alces vol. 40, 2004 dodge et al. michigan moose 79 smaller home ranges than did solitary cows, the differences were not significant (zmwu = 0.6664, p = 0.505). five yearlings (3 m, 2 f) and 1 adult cow permanently dispersed a mean linear distance of 80 ± 16 km (se) (range = 30134 km) during the study. the 2 yearling females, 1 yearling male, and the cow dispersed during april-june. the other 2 yearling males dispersed in january and september, respectively. the estimated annual dispersal rate was 0.068 (ci, 0.000 < 0.068 < 0.139) in 1999-00 and 0.054 (ci, 0.000 < 0.054 < 0.122) in 2000-01. discussion pregnancy determination and productivity pspb and fp4 appeared to be accurate at determining pregnancy. however, 27% (6 of 22) of cows with positive pspb test results were not observed with calves during the spring. four of these cows also had fecal progesterone levels greater than the fp4-95% utl for pregnancy (5.17 μg/g). although each of these cows was approached 2-3 times, it is possible that we did not find a calf (calves) because it (they) died shortly after birth or within the interval between cow sightings. no cows with negative pspb results were observed with calves. in addition, using pspb results as a baseline, fp4 test results would have correctly identified 83% of pregnant and 90% of non-pregnant cows. adult pregnancy rates in the wup were relatively constant from year to year (cv = 4.72%), but were lower than the 84.2% average reported by boer (1992) for moose in north america. stenhouse et al. (1995) also reported higher pregnancy rates than were found in the wup. in western northwest territories where moose densities are low (0.14 0.16 moose/km2), the pregnancy rate of adult females was 96% (stenhouse et al. 1995). in contrast, adult pregnancy rates in the wup, where moose also occur at low density (0.29 moose/km2) (mdnr, unpublished data), were higher than those found by cox et al. (1997) in northwest minnesota where moose populations have been in decline. for example, at agassiz national wildlife refuge (anwr) where the population had decreased 62% during 1993-1994 (0.50 moose/km2 to 0.31 moose/ km2), the average adult pregnancy rate during 1995-1997 was 37.5% (cox et al. 1997). in addition, at beltrami island state forest (bisf), where moose had apparently been declining for decades (1971: 0.54 moose/ table 3. annual and seasonal home range sizes of migratory and non-migratory adult moose, bulls, cows, cows attended by calves, and cows not attended by calves during 1999-2001 in the western upper peninsula of michigan. no. no. mean median range season category moose locations (km2) (km2) se (km2) annual migratory 18 20-30 63 53 7.12 20-122 non-migratory 37 19-34 43 41 3.49 14-99 bulls 12 19-24 47 43 5.97 22-80 cows 43 20-34 50 44 4.19 14-122 cows w/calves 29 21-34 48 44 5.19 14-115 cows w/out calves 11 20-28 51 49 6.01 25-96 summer migratory 16 11-24 44 39 6.04 9-98 winter migratory 8 10-15 27 23 7.58 3-64 michigan moose dodge et al. alces vol. 40, 2004 80 km2 1996: 0.07 moose/km2), cox et al. (1997) reported that only 51% of cows were pregnant. our results and those from western northwest territories and northwest minnesota differ from that of boer (1992) who found that adult pregnancy rates were quite similar across a wide range of population densities, as well as geographic areas, winter severities, and habitats. however, pregnancy rates per se, may not be the best index of moose productivity. adult twinning rates and yearling pregnancy rates, the variable components of fecundity (boer 1992), are likely better indicators of moose productivity (aitken and child 1992) although few yearling females were radio-collared in any 1 year ( x = 3), the mean annual pregnancy rate for yearlings(< 9%) was low compared to the north american average of 48.7% reported by boer (1992). twinning rates were comparable to those found by blood (1974) in alberta, canada (range = 4-48%). however, the mean twinning rate was below that reported by mdnr biologists in the wup during 1985-1995 (aho et al. 1995; 36%) and by boer (1992; 33.3%). frequency of twinning has been shown to be a good indicator of cow health condition and habitat quality. on the kenai peninsula, alaska, 70% of cows living on high quality habitat gave birth to twins, whereas only 20% living on poor quality habitat did so (franzmann and schwartz 1985). furthermore, boer (1992) found a direct relationship between adult twinning rates and yearling pregnancy rates. therefore, these measures of productivity are likely influenced by the same habitat component. however, no study has yet quantitatively related habitat quality and availability to moose productivity (crête and courtois 1997). finally, it has been suggested that at low population density a low bull:cow ratio may affect breeding of cows and productivity (crête et al. 1981, albright and keith 1987). this does not appear to be a problem in the wup where bulls, on average, comprised 50% of the adult winter population during 1999-2002 (mdnr, unpublished data). survival adult and yearling survival rates were similar to those reported for other nonh u n t e d , l i g h t l y p r e y e d o n , m o o s e populations. for example, in alberta, canada, mytton and keith (1981), reported mean annual survival rates of 0.86 for adults and 0.83 for yearlings. also, in a newly established moose population in southwest colorado, olterman and kenvin (1998) reported slightly higher bull survival (0.94, >1%) and a slightly lower cow survival (0.83, <2%) than were found in this study. furthermore, mean annual cow survival was 28% and 19% higher than that reported by cox et al. (1997) at anwr (0.67) and bisf (0.72), respectively. the relatively high survival rate of calves in the wup suggests that predation is not a significant mortality factor. firstyear calf survival rates were similar to those in alberta, canada (0.67; mytton and keith 1981) and higher than those in northwest minnesota (0.56; cox et al. 1997), both areas where predation of neonates is low. this contrasts with studies in alaska and canada where bears and wolves often kill substantial numbers of calves. in south-central alaska, brown bear (ursus arctos) predation accounted for 73% of all calf deaths and survival of calves < 5-months of age was only 0.39 (ballard et al. 1991). also, in northeastern alberta, hauge and keith (1981) reported an annual calf mortality rate of 73%, of which 29% was attributed to wolf predation. alces vol. 40, 2004 dodge et al. michigan moose 81 home range and movements the proportion of seasonally migratory moose in the wup (20-40%) was similar to that reported by addison et al. (1980) in northwest ontario, canada (27%), and phillips et al. (1973) in northwest minnesota (20%). the distance between summer and winter home ranges was also comparable to those reported by addison et al. (1980) in northwest ontario ( x = 7 km, range = 2-13 km) and phillips et al. (1973) in northwest minnesota ( x = 16 km, range = 14-34 km). in contrast, migration distances were smaller than those found by ballard et al. (1991) in interior south-central alaska, ( x = 48 km, range = 10-68 km) and mauer (1998) southeast of the brooks range in alaska and canada ( x = 123 km, range = 18-196 km). because of the different methods (e.g., minimum convex polygon, probabilistic) used to estimate home range size and other difficulties (e.g., sample size, delineation of seasons, etc.), it is problematic to make comparisons among different studies. nevertheless, in general, moose home ranges were comparable in size to those found for moose in the upper great lakes region. for a more detailed comparison of home range sizes from different geographic locations i n n o r t h a m e r i c a s e e h u n d e r t m a r k (1998:316-317). five of twenty-six (19%) offspring permanently dispersed out of the core study area shortly after separation from their cows. in south-central alaska, ballard et al. (1991) reported that 33% of offspring fully dispersed from their natal home range, and that more males than females dispersed. although we did not observe a male biased dispersal, all moose that dispersed except 1, were yearlings. in central alberta, canada, lynch (1976) found that 50% of subadults (< 2-years of age) and 17% of adults dispersed. these values are probably overestimates however, because he considered moose from which radio contact had been lost to have dispersed. in the wup, aho et al. (1995) reported that a yearling female and a yearling male emigrated at least 160 km to wisconsin over a 9-month (mar 1989-dec 1990) and 7-month (mar-oct 1994) period, respectively. conclusions although moose in the wup appear to be well established, biologists believe that the population is below carrying capacity. based on a habitat suitability index (hsi) model (model 2; allen et al. 1987) covertype composition variables, patterson et al. (1995) estimated that suitable habitat in baraga county (1,073 km2) could support 1.72 moose/km2. by comparison, on 25 plots (16 of which wholly or partially fell within baraga county [total area = 1,549 km2]) designated as having high moose density, preliminary mark-resight population estimates for winter 2002 were 0.29 moose/ km2 (mdnr, unpublished data). this > 6fold difference between potential habitat carrying capacity and the estimated population size suggests that further growth of the moose herd in the wup is possible. the relatively high survival of all age categories indicates that moose are probably in good physical condition and that disease and predation are not limiting population growth. in addition, the size of moose home ranges and their seasonal movement patterns indicate that there is good interspersion and juxtaposition of suitable habitat types. low productivity appears to be the primary reason the moose population in the wup has not increased as rapidly as expected. because productivity is dependent on female body condition which, in turn, is directly related to food supply (franzmann and schwartz 1985), one possible explanation then, is that the quality, quantity, and availability of food is less than optimal for maximum productivity. michigan moose dodge et al. alces vol. 40, 2004 82 additionally, because moose in the wup are at the southern extent of their range in eastern north america, the environmental conditions (e.g., climate) that have prevented further range expansion have also likely played a role in limiting population growth. management implications our results suggest that low productivity, exhibited in below average adult pregnancy rates and low production of twins, coupled with nearly non-existent yearling reproduction, is an important reason the population has not increased as predicted. in retrospect, the original population objective of 1,000 moose 15 years after the 1985 translocation was overly optimistic. a reevaluation of this objective is warranted. in addition, a closer examination of the potential of moose habitat in the upper great lakes region through a quantitative assessment of the nutritional quality and availability of forage is suggested. furthermore, monitoring the survival and movements of moose, the collection and analysis of fecal samples for fp4, which has been shown to be a reliable indicator of pregnancy, and determination of spring calf production should be continued for several years. the information obtained by further study will assist wildlife managers of reintroduced and/or small moose herds to set realistic population objectives. acknowledgements we thank the michigan department of natural resources, michigan state university, and the involvement committee of safari club international for providing the primary funding for this research. we also appreciate the hard work of research technicians r. atkinson, e. heimerl, j. hill, d. jentoft, v. lane, b. martin, e. north, n. seward, and m. westbrock for their assistance in the field. r. aho, b. johnson, and b. roell of the mdnr wildlife division provided additional assistance. we also express our thanks to pilots s. adkins, n. harri, and d. minett of the mdnr forest, mineral, and fire management division. references addison, r. b., j. c. williamson, b. p. saunders, and d. fraser. 1980. radio-tracking of moose in the boreal forest of northwestern ontario. canadian field-naturalist 94:269-276. aho, r. w., s. m. schmitt, j. hendrickson, and t. r. minzey. 1995. michigan’s translocated moose population: 10 years later. wildlife division report number 3245, michigan department of natural resources, lansing, michigan, usa. aitken, d. a., and k. n. child. 1992. relationship between in utero productivity of moose and population sex ratio: an exploratory analysis. alces 28:175-187. albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of newfoundland. canadian field-naturalist 101:373-387. allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose. lake superior region. biological report 82. u.s. fish and wildlife service, fort collins, colorado, usa. baker, r. h. 1983. michigan mammals. michigan state university press. east lansing, michigan, usa. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114. blood, d. a. 1974. variation in reproduction and productivity of an enclosed herd of moose (alces alces). transactions of the international congress of game biologists 1:59-66. alces vol. 40, 2004 dodge et al. michigan moose 83 boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces supplement 1:1-10. brewer, r. 1991. original avifauna and postsettlement changes. pages 33-58 in r. brewer, g. a. mcpeek, and r. j. adams jr., editors. the atlas of breeding birds of michigan. michigan state university press, east lansing, michigan, usa. cox, e. w., d. l. murray, and t. k. fuller. 1997. moose population dynamics in northwestern minnesota – annual progress report and recommendations. pages 66-99 in minnesota department o f n a t u r a l r e s o u r c e s , w i l d l i f e populations and research unit 1997 report, minneapolis, minnesota, usa. crête, m., and r. courtois. 1997. limiting factors might obscure regulation of moose (cervidae: alces alces) in unproductive boreal forests. journal of zoology (london) 242:765-781. , r. j. taylor, and p. a. jordan. 1981. optimization of moose harvest in southwestern quebec. journal of wildlife management 45:598-611. de vos, a. a. 1964. range changes of mammals in the great lakes region. american midland naturalist 71:210231. eichenlaub, v. l. 1990. the climatic atlas of michigan. university of notre dame press, notre dame, indiana, usa. franzmann, a. w., and c. c. schwartz. 1985. moose twinning rates: a possible population condition assessment. journal of wildlife management 49:394396. haigh, j. c., w. j. dalton, c. a. ruder, and r. g. sasser. 1993. diagnosis of pregnancy in moose using a bovine assay for pregnancy-specific protein b. theriogenology 40:905-911. hauge, t. m., and l. b. keith. 1981. dynamics of moose populations in northeastern alberta. journal of wildlife management 45: 573-597. heisey, d. m., and t. k. fuller. 1985. evaluation of survival and cause-specific mortality rates using telemetry data. journal of wildlife management 49:688-674. hickie, p. f. 1937. a preliminary report on the past and present status of the moose, alces americana (clinton), in michigan. michigan academy of science, arts and letters 3:369-402. . 1944. michigan moose. michigan department of conservation, lansing, michigan, usa. hooge, p. n., w. m. eichenlaub, and e. k. solomon. 1999. using gis to analyze animal movements in the marine environment. united states geological survey, alaska biological science center, glacier bay field station, gustavus, alaska, usa. h u a n g , f . , d . c . c o c k r e l l , t . r . stephenson, j. h. noyes, and r. g. sasser. 2000. a serum pregnancy test with a specific radioimmunoassay for moose and elk pregnancy specific protein b. journal of wildlife management 64:492-499. hudgins, b. 1953. michigan: geographic backgrounds in the development of the commonwealth. privately printed, detroit, michigan, usa. hundertmark, k. j. 1998. home range, dispersal and migration. pages 303335 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. lynch, g. m. 1976. some long-range movements of radio-tagged moose in alberta. proceedings of the north american moose conference and workshop 12:220-235. mauer, f. j. 1998. moose migration: michigan moose dodge et al. alces vol. 40, 2004 84 northeastern alaska to northwestern yukon territory, canada. alces 34:7581. mayfield, h. 1961. nesting success calculated from exposure. wilson bulletin 73:255-261. . 1975. suggestions for calculating nest success. wilson bulletin 87:456-466. mccann, m. t. 1991. land, climate, and vegetation of michigan. pages 15-31 in r. brewer, g. a. mcpeek, and r. j. adams jr., editors. the atlas of breeding birds of michigan. michigan state university press, east lansing, michigan, usa. messier, f., d. m. desauliners, a. k. goff, r. nault, r. patenaude, and m. crête. 1990. caribou pregnancy diagnosis from immunoreactive progestins and estrogens excreted in feces. journal of wildlife management 54:279-283. (mdnr) michigan department of natural resources. 1991. draft moose management plan. michigan department of natural resources, wildlife division report. lansing, michigan, usa. monfort, s. l., c. c. schwartz, and s. k. wasser. 1993. monitoring reproduction in moose using urinary and fecal steroid metabolites. journal of wildlife management 57:400-407. mytton, w. r., and l. b. keith. 1981. dynamics of moose populations near rochester, alberta, 1975-1978. canadian field-naturalist 95:39-49. olterman, j. h., and d. w. kenvin. 1998. reproduction, survival, and occupied ranges of shiras moose transplanted to southwestern colorado. alces 34:4146. patterson, r. l., s. l. ockey, c. e. olson, and a brenner. 1995. analysis of population statistics (1985-1994) and habitat (1955-1991) for moose (alces alces) in the western upper peninsula of michigan. final research report submitted to the wildlife division, michigan department of natural resources, marquette, michigan, usa. peterson, r. o., j. d. woolington, and t. n. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88. phillips, r. l., w. e. berg, and d. b. sniff. 1973. moose movement patterns and range use in northeastern minnesota. j o u r n a l o f w i l d l i f e m a n a g e m e n t 37:266-278. pollock, k. h., s. r. winterstein, c. m. bunck, and p. d. curtis. 1989. survival analysis in telemetry studies: the staggered entry design. journal of wildlife management 53:7-15. schwartz, c. c., s. l. monfort, p. h. dennis, and k. j. hundertmark. 1995. fecal progestagen concentration as an indicator of the estrous cycle and pregnancy in moose. journal of wildlife management 59:580-583. seaman, d. e., j. j. millspaugh, b. j. ke r n o h a n, g. c. b r u n d i g e, k. j. raedeke, and r. a. gitzen. 1999. effects of sample size on kernel home range estimates. journal of wildlife management 63:739-747. stenhouse, g. b., p. b. latour, l. kutny, n. maclean, and g. glover. 1995. productivity, survival, and movements of female moose in a low-density population, northwest territories, canada. arctic 48:57-62. stephenson, t. r., j. w. testa, g. p. adams, r. g. sasser, c. c. schwartz, and k. j. hundertmark. 1995. diagnosis of pregnancy and twinning in moose by ultrasonography and serum assay. alces 31:167-172. trent, t. t., and o. j. rongstad. 1974. home range and survival of cottontail rabbits in southwestern wisconsin. alces vol. 40, 2004 dodge et al. michigan moose 85 j o u r n a l o f w i l d l i f e m a n a g e m e n t 38:459-472. vangilder, l. d., and s. l. sheriff. 1990. survival estimation when fates of some animals are unknown. transactions of the missouri academy of science 24:57-68. verme, l. j. 1984. some background on moose in upper michigan. michigan department of natural resources, w i l d l i f e d i v i s i o n r e p o r t 2 9 7 3 , lansing, michigan, usa. whitney, g. g. 1987. an ecological history of the great lakes forests of michigan. journal of ecology 75:667-684. wilton, m. l. 1982. report to the michigan department of natural resources concerning potential moose habitat. michigan department of natural resources, wildlife division, lansing, michigan, usa. winterstein, s. r., k. h. pollock, and c. m. bunck. 2001. analysis of survival data from radiotelemetry studies. pages 351-380 in j. j. millspaugh and j. m. marzluff, editors. radio tracking and animal populations. academic press, san diego, california, usa. wood, n. a. 1914. an annotated checklist of michigan mammals. university of michigan occasional papers no. 4. , and l. r. dice. 1923. records of the distribution of michigan mammals. michigan academy of science, arts and letters 3:425-469. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70:164168. alces36_77.pdf 67 availability and use of moose browse in response to post-fire succession on kanuti national wildlife refuge, alaska erin julianus1,2, teresa n. hollingsworth3, a. david mcguire4,5, and knut kielland6 1department of biology and wildlife, university of alaska fairbanks, 982 n. koyukuk dr., fairbanks, alaska usa 99775; 2present address: bureau of land management, 222 university ave, fairbanks, ak 99709, usa; 3us forest service pnw research station, po box 75680 university of alaska fairbanks, fairbanks, alaska, usa 99775; 4u.s. geological survey, alaska cooperative fish and wildlife research unit, po box 757020, university of alaska fairbanks, fairbanks, alaska, usa 99775; 5retired; 6institute of arctic biology, po box 757000, university of alaska, fairbanks, fairbanks, alaska, usa 99775 abstract: wildfire is a prominent landscape-level disturbance in interior alaska and associated vegetation changes affect quantity and quality of moose (alces alces) habitat. these changes are important to land and wildlife managers responsible for managing habitat and ensuring sustained yield of game species such as moose. considering the changing fire regime related to climate change, we explored post-fire dynamics of moose habitat to broaden understanding of local habitat characteristics associated with wildfire on the kanuti national wildlife refuge in interior alaska. we studied 34 sites in different aged stands (2005 burn, 1990 burn, 1972 burn, and unburned in the last 80 years) in august 2012 and 2013 to estimate summer browse density, biomass production, and browse use, and revisited each site the following march to estimate winter browse availability and offtake. we also used location data from 51 radio-collared moose to quantify use of burns on the kanuti national wildlife refuge. we found that summer density and biomass of preferred browse was highest at sites in the 1990 burn, although use of burns varied seasonally. despite high biomass in the most recent 2005 burn, radio collared moose avoided burns <11 years old in summer and had preference for older stands (>30 years old). winter browse offtake was highest in the 1990 and 1972 burns despite relatively high biomass available in the 2005 burn. the disparate use of burns, particularly low use of the 2005 burn, likely reflected a combination of influences including species composition and preference, predator avoidance strategies, a low density moose population, and historic moose distribution patterns. alces vol. 55: 67–89 (2019) key words: browse, habitat, habitat selection, interior alaska, moose, wildfire the fire regime in interior alaska is changing as a result of climate change. this shift is characterized by shorter fire intervals and an increase in late-season fires, frequency of large (>1000 km2) fires, and higher-severity fires (kasischke and turetsky 2006, kasischke et al. 2010) that influence post-fire vegetation patterns at local and landscape scales. specifically, higherseverity fires result in deeper burning of the surface organic layer that increases establishment of deciduous species while negatively impacting recruitment of black spruce (picea mariana) (johnstone 2006). increased prevalence of high-severity fires could cause a major vegetative shift from coniferous black spruce communities to those dominated by deciduous species (johnstone et al. 2010b). such 1present address: bureau of land management, 212 university ave., fairbanks, alaska, 99709, usa. moose post-fire habitat. – julianus et al. alces vol. 55, 2019 68 landscape-scale changes can impact wildlife habitat and, consequently, wildlife populations either in a positive or negative direction depending on species-specific habitat requirements. therefore, vegetation changes associated with a changing fire regime are essential to consider when developing future habitat management objectives. predicted change in the boreal fire regime is anticipated to be generally beneficial to moose (alces alces) because it is hypothesized that deciduous species will increase in and/or dominate certain plant communities (chapin et al. 2008, johnstone et al. 2010a). moose commonly consume willow (salix spp.), birch (betula neoalaskana), and aspen (populus tremuloides) regrowth maintained by natural disturbances such as wildfire. maier et al. (2005) found that in november moose preferentially use forest stands where fire occurred 11–30 years ago, and quantity and quality of browse is highest (oldemeyer 1974, oldemeyer et al. 1977, maccracken and viereck 1990, lord and kielland 2015). additionally, the physical structure of these stands provides moose year-round access to browse, whereas shorter vegetation in early seral stands (<11 years old) is often unavailable due to snow depth. likewise, mature birch and bebb’s willow (s. bebbiana) in late seral stands are often inaccessible given their height >3.0 m (wolff and zasada 1979, danell and ericson 1986). moose populations respond to disturbance and vegetative succession in a number of ways; for example, individuals actively immigrate into recently disturbed areas (peek 1974b) and moose density changes through time in response to habitat (loranger et al. 1991). wildfire and flooding are the primary natural disturbance agents on the kanuti national wildlife refuge (refuge). the varied fire history on the refuge has created many forest stands of diverse size and age, although it is dominated currently by black spruce communities highly susceptible to conversion to deciduous communities after severe wildfire. in addition, moose populations in the upper koyukuk river drainage, including the refuge, are primarily regulated by predation (stout 2010). the role of wildfire in areas with dense moose populations is well studied in alaska, specifically due to management concerns regarding habitat degradation and carrying capacity (boertje et al. 2000, 2009, lord and kielland 2015). conversely, habitat use is less explored in regions with lower density populations regulated by predation. although habitat is not believed to regulate the refuge moose population, it is important to understand the influence of a changing fire regime on the interactions between habitat dynamics and moose distribution and habitat use. we sought to examine habitat characteristics in stands at various stages of post-fire succession on the refuge to provide insight about these interactions. specifically, we evaluated browse availability and use in summer and late winter in multiple-aged burn scars within the refuge. we also used location data from radio collared moose to explore their use of burns. we predicted that summer and winter browse availability and use would be highest in 11–30 year-old stands, and that moose would exhibit a preference for these stands in winter. study area the study took place on the refuge which consists of ~3.2 million roadless ha (1.3 million acres) located between 65° 59’ to 66° 53’ n and 150° 58’ to 152° 58’ w in interior alaska (fig. 1). it is representative of the boreal forest biome characterized by plant diversity and vegetation patterns dictated by climate, hydrology, and wildfire. the climate is cold and continental, with alces vol. 55, 2019 julianus et al. – moose post-fire habitat 69 short hot summers and long cold winters. mean monthly temperature ranges from ~ −28°c in january to 20°c in july (western  region climate center 2014; http://www. wrcc.dri.edu/cgi-bin/climain.pl?ak0761). the growing season is short, generally beginning in late may and ending in august. the refuge occupies the broad lowland flats between the koyukuk and kanuti rivers. the kanuti basin is characterized by poor drainage and riparian wetlands created and maintained by seasonal flooding and the presence or absence of permafrost. vegetation patterns reflect drainage patterns, with lowland permafrost areas dominated by black spruce forests and tussock tundra. well-drained slopes are dominated by deciduous stands of aspen, birch, and upland shrubs such as willow and alder (alnus spp.). large white spruce (picea glauca) and riparian shrub species dominate permafrost-free riparian areas where secondary succession fig. 1. sample site locations and age of fire scars studied in the kanuti national wildlife refuge, alaska, usa (2012–2013). http://www.wrcc.dri.edu/cgi-bin/climain.pl?ak0761 http://www.wrcc.dri.edu/cgi-bin/climain.pl?ak0761 moose post-fire habitat. – julianus et al. alces vol. 55, 2019 70 is a consequence of flood patterns and frequency along river corridors (payette 1992, nowacki et al. 2001). moose density ranges from 0.07 to 0.18 moose/km2, with the kanuti population fluctuating between 551 and 759 moose since 1993 (julianus and longson 2018). the most recent refuge estimate was 1311 ± 252 (90% ci) in 2017 (julianus and longson 2018). hunting pressure is light and localized near villages and along navigable rivers. moose are considered large and healthy with high twinning rates (35–60%) indicative of good nutrition (franzmann and schwartz 1985, stout 2010) in the game management unit that includes the refuge. despite adequate bull:cow ratios (46–70 bulls:100 cows) and high pregnancy rates (96% from 2006 to 2009), fall recruitment is consistently low (33 calves:100 cows in november 2010; stout 2010) and purportedly due to high calf and yearling mortality from predation (saperstein et al. 2009, craig and stout 2011). the characteristics of adequate production yet low adult recruitment have been documented in other low density moose populations in alaska (bertram and vivion 2002, lake et al. 2013). methods site description we established 4 burn age strata across fire scars on the refuge based on seasonal landscape use patterns by moose (maier et al. 2005): 1) <11 year-old stands, 2) 11–30 year-old stands, 3) 30–80 year-old stands, and 4) stands that were unburned in the past 80 years of recorded fire history (hereafter unburned). we selected 3 different fire scars to represent burn age strata 1–3: a 2005 fire (f-05), a 1990 fire (f-90), and a 1972 fire (f-72) (fig. 1); unburned sites were visited to identify sites for burn stratum 4. we characterized abiotic factors across each burn stratum (f-05, f-90, f-72, and unburned). we used a digital elevation model (dem) to determine the mean, minimum, and maximum elevations, and arcmap 10.1 (esri, redlands, california, usa) spatial analyst extension to determine slope and aspect from the dem. slope was averaged across plots within each burn stratum and classified as flat, gentle (<10°), medium (10°–30°), or steep (>30°). the alaska landfire vegetation map (2008) was used to quantify vegetation types and stand height classes in each burn strata. because much of the refuge is dominated by black spruce communities considered low quality moose habitat, we excluded these during site selection. instead, we selected vegetation types that were more likely used by moose within each burn stratum (appendix a). we isolated vegetation types with >3 adjacent pixel groups (areas >30 m2) and generated different lists for randomly derived boat/float plane or helicopter accessible sites. in 2012, field work was restricted to areas accessible by boat/float plane from the kanuti river; in 2013, a helicopter was used to access more remote areas within a burn. for the boat/float plane accessible sites, a 200 m buffer was created around the kanuti river and tachlodaten lake (a lake ~12 miles north of the kanuti river) and random points were generated within 300 m outside the buffer. if necessary, a <200 m buffer was implemented to avoid sampling in the floodplain which was subject to flood disturbance dynamics. in total, 34 sites were sampled (8 in unburned, 9 in f-72, 8 in f-90, and 9 in f-05): 11 boat/ float plane sites in summer 2012 and spring 2013, and 23 helicopter sites in summer 2013 and spring 2014 (fig. 1). because the digital vegetation classification pre-dated f-05 and post-burn vegetation class information was lacking, we selected 6 of the 9 f-05 sites post-hoc while conducting fieldwork. we classified vegetation at these sites alces vol. 55, 2019 julianus et al. – moose post-fire habitat 71 using the scheme developed by viereck et al. (1992). summer field work and analyses a 30 m diameter plot was established at each site and flagged to facilitate relocation for winter browse surveys. the following were measured at each plot: vegetation community type, slope (°), aspect, elevation, average tree canopy height (m), and shrub height (m). additionally, we used photos to evaluate and classify fire severity at each plot as low, moderate, or high (kasischke et al. 2008). vascular and nonvascular plant species were inventoried and classified relative to moose browsing preference of deciduous trees and shrubs described in the literature (oldemeyer et al. 1977, wolff and zasada 1979, bryant and kuropat 1980; appendix a). we did not consider birch as preferred summer browse. two 30 m transect lines were established in each plot. we counted individual preferred plants within 1 m of the line (both sides or 120 m2) to estimate browse species density (individuals/ha) and evidence of past browsing (individuals browsed/ha) in the 120 m2 transect area. evidence of browsing by moose, snowshoe hare (lepus americanus), and other species was noted. browsing was identified from leaf stripping and the presence of dead stems. we counted stems at the general foraging height of moose between 0.5 and 3.0 m above ground level (wolff 1978, danell and ericson 1986). the extent of browsing was not described during summer, but architecture classes (unbrowsed, browsed, or broomed) were assigned to individual plants during winter field work (see winter field work and analyses). stems within 10 cm of each other were defined as one plant. at the center of the 30 m plot, we also established a second plot to measure browse biomass. the size of this sub-plot varied depending on browse plant density and vegetative homogeneity. within this sub-plot, the current annual growth (cag) of stems on preferred browse species was removed and oven dried at 110°c for 48 h. stem and leaf material were weighed separately, and leaf material was used to estimate summer biomass (kg/ha). we evaluated normality for all data sets prior to analysis; however, data were not normally distributed or easily transformed. therefore, we used the non-parametric kruskal–wallis one-way analysis of variance to detect differences in browse density, biomass, and browsed plant density among burn strata. we used the mann–whitney u test to detect pairwise differences between groups when the kruskal–wallis test indicated significance; alpha was set at 0.05 for all tests. median values are reported, as well as the first (25th) and third (75th) quartiles. winter field work and analyses we evaluated biomass availability and use of winter woody browse in the 4 burn strata (unburned, f-72, f-90, and f-05) following the methods of paragi et al. (2008) and seaton et al. (2011). sites established in 2012 were revisited in late march 2013, and sites established in summer 2013 were revisited in late march 2014. we re-established plot boundaries in the winter by delineating a 30 m diameter circle in the snow. within each plot, we recorded slope (°), aspect, and snow depth (m) and documented preferred and non-preferred browse species. although not considered preferred in summer, we classified birch as a preferred winter browse species (unpublished data, paragi et al. 2008). we counted the number of preferred plants present in the plot. in plots with high, relatively uniform densities of preferred browse species, we counted individuals in one quadrat of the 30 m circle and used these data to estimate the number of plants in the entire plot (707 m2). moose post-fire habitat. – julianus et al. alces vol. 55, 2019 72 in each plot we randomly selected 3 plants of each preferred browse species, or if <3 plants, all available specimens. we recorded the species, plant height, number of cag stems (0.5–3.0 m above ground level), and classified each plant as having 0%, <50%, or >50% dead cag stems. an architecture class was also assigned to each plant: unbrowsed (no evidence of browse), browsed (<50% of cag stems were from lateral stems produced from browsing), or broomed (>50% of cag stems were from lateral stems). we measured cag diameter (mm) on a random sample of 10 twigs/plant using dial calipers, and if a twig was browsed, the diameter at point of browsing (dpb). the winter sampling effort (stems/plot measured) is provided in appendix b. data were entered into a microsoft access database and processed using software written in r (microsoft corporation, redmond, washington, usa; r project for statistical computing, [accessed february 2015]). mass:diameter regression relationships for each browse species were previously developed (paragi et al. 2008) from sample twigs gathered on the refuge in 2007 and provided by the alaska department of fish and game (adfg) (t. paragi, adfg, personal communication). we calculated winter browse biomass availability and removal using these mass:diameter relationships, and our estimates of plant density (individuals/ha) and cag twigs/plant with the following formula: m m n n zb̂ j jk jk i ijk ijk h hijk= ∑ ∑ ∑ where b̂ denotes estimated plot biomass, twigs are denoted by h, plants i, species j, and sites k. m denotes total plants in each plot, m sampled plants, and n and n total and sampled twigs, respectively; z denotes individual twig biomass (g). the r output provided estimates of biomass production and removal at the plant, species, plot, and study area levels. we estimated proportional biomass removal rates (%) based on browse production and consumption for each area (kg/ha) per year. habitat use in 2008 the adfg, u.s. fish and wildlife service (usfws), national park service (nps), and bureau of land management (blm) initiated a radio telemetry study of moose in game management unit 24 which includes the refuge (joly et al. 2015). of the 120 moose, 51 (48 adult cows, 3 adult bulls) ranged at least partially within the refuge; the study targeted adult cows (97 of 123 captured animals). radio-collared moose were located monthly or as weather allowed during telemetry flights from 2008 to 2013. the average number of relocations per animal was 45, ranging from 31 to 56 per animal. radio-collared moose were observed when possible to, in part, document the vegetation type within which they were observed. capture efforts occurred throughout the refuge and were not confined to specific habitat types (e.g., burns); 25 moose were captured in unburned areas, 6 in >30 year-old burns, and 10 in both 11–30 and >11 year-old burns. we used their location data to evaluate use of the burn strata and assumed independence between locations (dunn and gipson 1977). the vhf data set was characterized by small (<50 locations) sample sizes for each marked animal. because appropriate methods for analyzing habitat use with these sample sizes are limited, we used methods described by neu et al. (1974) to examine general use of burn strata. habitat use by http://www.r-project.org http://www.r-project.org alces vol. 55, 2019 julianus et al. – moose post-fire habitat 73 individual moose was difficult to assess due to sample size; therefore, we combined all locations within the refuge for analysis. we used a chi-square goodness-of-fit test to determine whether moose exhibited seasonal patterns of habitat use that deviated from proportional habitat availability. we first determined proportional availability of burn strata by dividing the number of ha within each burn class by the total refuge area. we designated 2 seasons – “winter” (october–april) and “summer” (may– september) – and also a separate “calving” season (may 28–june 23; joly et al. 2015). we compared the observed number of seasonal locations in each stratum to the expected number based on each stratum’s proportional availability. if p < 0.05, we concluded that seasonal use did not occur in proportion to availability. where use of burn strata was not in proportion to availability (p < 0.05), we examined whether moose demonstrated preference (observed number of locations > expected proportion) or avoidance (observed < expected). we determined preference/ avoidance and the degree to which they were demonstrated using confidence intervals developed by neu et al. (1974). confidence intervals were constructed for the proportion of times an animal used each habitat type. the interval equaled: p z p p n p p z p p n 1 1 i k i i i i k i i 1 /2 1 /2 ) )( ( − − ≤ ≤ + − α α) )( (− − where pi is the proportion of moose locations in the ith burn stratum, n is the number of locations, and z(1−α/2k) is the lower standard normal variate corresponding to a probability tail area of α/2k where k is the number of burn strata (4). the 2k denominator was used because multiple confidence intervals were being computed simultaneously. we identified the degrees of freedom (df) as the number of available habitat types (k) minus 1. if the proportion of available habitat was included in the confidence interval, we concluded that preference for or avoidance of a burn stratum was not exhibited. if the lower bound of the confidence interval was greater than the proportion of available habitat, we concluded preference was exhibited; alternatively, if the upper bound was less than the proportion of available habitat, we concluded that avoidance was exhibited. results site description each burn used in this study was >80,000 ha and f-90 and f-05 occurred during 2 of the biggest fire seasons on record. abiotic characteristics of the 4 burn sites are summarized in appendix c. the f-72 burn perimeter contained both flat wetlands and uplands with gentle (<10°) south-facing slopes, with an elevation of 213 m. f-90 was also characterized by gentle slopes, although much of the burn scar was >300 m in elevation and dominated by upland vegetation types. the southern perimeter of f-05 abutted the foothills of the ray mountains with most of the burn consisting of wetlands and permafrost-rich soils; fire severity was classified as moderate-high based on multiple site assessments within the fire scar (appendix d). based on landfire (appendix e), f-72 was dominated by deciduous vegetation types, and f-90 consisted mostly of deciduous and tall shrub vegetation types (38% and 20% respectively). post-burn landfire data for f-05 were unavailable; however, prior to burning, f-05 was mostly deciduous (25%) and shrub vegetation types (35%). the unburned stratum contained a wide variety of vegetation types and was without a dominant cover type. vegetation types were further documented during site visits (table 1): the f-72 fire scar moose post-fire habitat. – julianus et al. alces vol. 55, 2019 74 was ~67% forest and 33% shrub; the f-90 fire scar was ~25% forest, 63% shrub, and 12% herbaceous; the f-05 fire scar was ~11% forest, 78% shrub, and 11% herbaceous; and the unburned stratum was ~63% forest and 37% shrub. canopy height varied considerably among burn strata (appendix c). in f-72 the height of >80% of vegetation was >10 m, and 65% was >5 m in f-90; conversely, 33% of vegetation in f-90 and 18% in f-72 was classified as shrubs 0.5–1.5 m in height, and 50% in f-05 was classified as shrubs >1.5 m. vegetation >5 m tall was mostly concentrated in riparian areas. in the unburned, only 45% of trees were >5 m; heights <5 m reflected the preponderance of old growth black spruce stands throughout. availability and use of browse during summer we documented 3 preferred browse species in unburned and f-72, and 5 in f-90 and f-05 (table 2); the range was 1–5 species at a given site. density of summer browse (excluding birch) ranged from ~500 to 18,000 individuals/ha across the burn strata; the kruskal–wallis test indicated that the median (mdn) values were different. pairwise comparisons (mann–whitney test) among strata indicated that browse density in f-90 and f-05 (mdn = 10,084 and 6833 individuals/ha, respectively) was greater than that in f-72 (2000 individuals/ha) and unburned (5666 individuals/ha) (u = 6–31, p = 0.01–0.04; fig. 2). no differences were found in plant density between unburned and f-72 (p > 0.05) or f-90 and f-05 (p > 0.05). relative abundance (based on the number of individuals) of browse species and birch in summer varied among burn strata (fig. 3). of the 6 species identified, 2 (salix arbusculoides and populus tremuloides) contributed little to overall abundance (0% unburned, 0% f-72, 0% f-90, and 12% f-05). willow species (s. pulchra, s. glauca, and s. bebbiana) dominated unburned (87%), f-72 (99%), and f-90 (98%). of note, betula neoalaska was 48% of the relative abundance in f-05. browse use (individuals browsed/ha) in summer was highest in f-72 and f-90 (u = 10–13, p = 0.008–0.03; fig. 2). the proportion of browsed individuals with evidence of browsing did not differ among unburned, f-72, and f-05 (p > 0.05). the proportion of individuals with evidence of browsing did not differ among f-72 and f-90 (p > 0.05); however, f-90 had a significantly higher proportion of browsing than unburned and f-05 (u = 9, p = 0.03 and table 1. vegetation types studied in 4 burn strata in the kanuti national wildlife refuge, alaska, usa. using the alaska vegetation classification (viereck et al. 1992), types are ranked based on their frequency in each stratum. stratum vegetation code vegetation type # plots unburned i.a needleleaf forest 3 ii.c low shrub 3 i.b deciduous forest 1 i.c mixed forest 1 f-72 i.a needleleaf forest 3 i.b deciduous forest 2 i.c mixed forest 2 ii.d dwarf shrub 2 ii.c low shrub 1 f-90 ii.c low shrub 3 i.b deciduous forest 1 i.c mixed forest 1 ii.b tall shrub 1 ii.d dwarf shrub 1 iii.a graminoid herbaceous 1 f-05 ii.c low shrub 5 ii.d dwarf shrub 2 i.b deciduous forest 1 iii.a graminoid herbaceous 1 alces vol. 55, 2019 julianus et al. – moose post-fire habitat 75 u = 10, p = 0.02 respectively), and was similar to that in f-72 and f-90 (p > 0.05). leaf biomass (excluding birch) in summer ranged from ~40 to >400 kg/ha (fig. 4). biomass in f-90 and f-05 (mdn = 143 and mdn = 189 kg/ha, respectively) was higher than in unburned and f-72 (mdn = 16 and mdn = 9 kg/ha, respectively; u = 10–20, p = 0.001– 0.03), a pattern consistent with browse density measurements. availability and use of browse during winter available winter biomass ranged from ~2 to 30 kg/ha across study sites (fig. 5), and was highest in f-90 and lowest in unburned (mdn = 28 and mdn = 24 kg/ha respectively; u = 9–12, p = 0.02–0.04). f-05 was dominated by birch, whereas willow was predominant in the other burn strata; e.g., willow was 61% of available biomass in f-90 and only 10% in f-05 (fig. 6). the relative offtake of woody biomass across all burn strata was 5.4% (95% ci = 3.9–6.9%; fig. 5). the highest offtake was 6% in f-72 and the lowest 4.5% in unburned and f-05. use (2.2 kg/ha) was higher in f-72 and f-90 than in unburned and f-05 (p = 0.001). moose generally took larger bites of willow in f-90 and unburned (both willow-dominated) and smaller bites in f-05 (birch-dominated) (fig. 7); broomed plants were not observed (data not presented). these burns were dominated by willow, whereas f-05 was dominated by birch. browsing on birch was not observed despite its high availability as potential winter browse in f-90 and f-05 (fig. 5 and 7). table 2. preferred browse species (trees and shrubs) documented in 4 burn strata in the kanuti national wildlife refuge, alaska, usa. genera are betula (b.), salix (s.), populus (pop.), picea (p.), alnus (a.), and rosa (r.). unburned f-72 f-90 f-05 preferred b. neoalaskana1 b. neoalaskana1 b. neoalaskana1 b. neoalaskana1 s. bebbiana s. bebbiana s. arbusculoides s. arbusculoides s. glauca s. glauca s. bebbiana s. bebbiana s. pulchra s. pulchra s. glauca s. glauca s. pulchra s. pulchra pop. tremuloides s. scouleriana pop. tremuloides non-preferred a. crispa a. crispa a. crispa a. crispa a. tenufolia b. glandulosa b. glandulosa b. glandulosa b. glandulosa b. nana b. nana b. nana b. nana b. neoalaskana2 b. neoalaskana2 b. neoalaskana2 b. neoalaskana2 p. balsamifera2 p. balsamifera2 p. mariana s. bebbiana2 s. bebbiana2 s. bebbiana2 pop. tremuloides p. glauca p. glauca pop. tremuloides2 r. acicularis p. mariana p. mariana p. glauca s. beauverdiana p. mariana 1considered as browse species in winter only; 2mature individuals (>3 m height). moose post-fire habitat. – julianus et al. alces vol. 55, 2019 76 fig. 2. density of total available and browsed plants (individuals/ha) for preferred browse during summer (excluding betula neoalaskana) on the kanuti national wildlife refuge, alaska, usa. the lower bound represents the 1st (25%) quartile, center lines indicate median values, and the upper bound represents the 3rd (75%) quartile. letters denote significantly different groups based on kruskal–wallis analysis of variance and mann–whitney u post-hoc pairwise comparisons. total browsed 0 5000 10000 15000 20000 d en si ty (i nd iv id ua ls /h a) burn a a b b a ab b a unburned 1972 1990 2005 fig. 3. composition (based on number of individuals) of preferred browse species and birch during summer on the kanuti national wildlife refuge, alaska, usa. bene denotes betula neoalaskana, sapu denotes salix pulchra, sagl denotes s. glauca, and sabe denotes s. bebbiana. “other” denotes populus tremuloides and picea balsamifera. 0 20 40 60 80 100 unburned 1972 1990 2005 re la � ve a bu nd an ce burn other sabe sagl sapu bene alces vol. 55, 2019 julianus et al. – moose post-fire habitat 77 fig. 4. total leaf biomass (kg/ha) of preferred summer browse by burn strata on the kanuti national wildlife refuge, alaska, usa. the lower bound represents the 1st (25%) quartile, center lines indicate median values, and the upper bound represents the 3rd (75%) quartile. letters denote significantly different groups based on kruskal–wallis analysis of variance and mann–whitney u post-hoc pairwise comparisons. 0 50 100 150 200 250 300 350 400 450 unburned 1972 1990 2005 bi om as s (k g/ ha ) burn a a b b fig. 5. winter biomass and removal (kg/ha) of preferred browse by burn strata on the kanuti national wildlife refuge, alaska, usa. the lower bound represents the 1st (25%) quartile, center lines indicate median values, and the upper bound represents the 3rd (75%) quartile. letters denote significantly different groups based on kruskal–wallis analysis of variance and mann–whitney u post-hoc pairwise comparisons. values of removed biomass represent salix spp. exclusively; betula neoalaskana was not browsed. 0 5 10 15 20 25 30 35 40 bi om as s (k g/ ha ) burn a b c d a b b a unburned 1972 1990 2000 total removed moose post-fire habitat. – julianus et al. alces vol. 55, 2019 78 habitat use capture locations of radio-collared moose did not appear to influence or relate to habitat use because the relatively large number of moose (25) captured in unburned did not demonstrate exclusive preference for this habitat. further, only 6 animals were captured in burns >30 years old, yet moose demonstrated preference for this stratum in both summer and winter. as such, preference or avoidance was likely not an artifact of capture location. during the “summer” season, moose exhibited preferential use of burns >30 years old and avoidance of burns <11 years old ( χ2 = 17.675, p < 0.001; fig. 8). moose did not appear to actively select or avoid unburned or 11–30 year-old burns (p > 0.05). cows (n = 120) preferred unburned stands ( χ2 = 11.766, df = 3, p = 0.01) during calving (28 may–23 june; fig. 8). in winter, moose demonstrated preference for stands 11–30 years old and avoidance of stands <11 years old ( χ2 = 36.074, df = 3, p < 0.001; fig. 8). winter use of unburned areas and stands >30 years old was proportional to availability. discussion overall, our results are consistent with the general understanding that moose habitat quality peaks at 11–30 years post-fire (maier et al. 2005). we found that density and biomass of summer browse were highest in f-90, and that browse removal was highest in f-90 and f-72. although browse density and biomass in f-05 were also high, use in summer was low. similarly, marked moose avoided <11 year-old stands and preferred >30 year-old stands in summer suggesting fig. 6. relative abundance (biomass) of winter browse in 4 burn strata on the kanuti national wildlife refuge, alaska, usa. sa spp. denotes salix spp. and bene denotes betula neoalaskana. 0% 20% 40% 60% 80% 100% unburned 1972 1990 2005 re la � ve a bu nd an ce burns sa spp. bene alces vol. 55, 2019 julianus et al. – moose post-fire habitat 79 fig. 7. panel a: frequency distributions of cag (current annual growth) and dpb (diameter at point of browsing) of willow in 4 burn strata on the kanuti national wildlife refuge, alaska, usa. panel b: frequency distribution of cag (current annual growth) of betula neoalaskana in each burn stratum; browsing of birch was not observed. fr eq ue nc y 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 1972 sa spp. diameter (mm) cag dpb fr eq ue nc y sa spp. diameter (mm) 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 1990 cag dpb fr eq ue nc y 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 2005 sa spp. diameter (mm) cag dpb 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 fr eq ue nc y unburned sa spp. diameter (mm) cag dpb panel a 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 fr eq ue nc y unburned bene diameter (mm) cag 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 fr eq ue nc y 1990 bene diameter (mm) cag 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 fr eq ue nc y 2005 bene diameter (mm) cag 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 fr eq ue nc y 1972 bene diameter (mm) cag panel b moose post-fire habitat. – julianus et al. alces vol. 55, 2019 80 fig. 8. panel a: selection (use/availability) of burn age classes by radio-collared moose in summer (may–september), kanuti national wildlife refuge, alaska, 2008–2013. panel b: selection of burn age classes by radio-collared cow moose during calving (28 may–23 june), kanuti national wildlife refuge, alaska, 2008–2013. panel c: selection of burn age classes by radio-collared moose during winter (october–april), kanuti national wildlife refuge, alaska, 2008–2013. values indicate proportion of relocations observed in each stratum. confidence intervals (95%) >1 indicate preference, whereas values <1 indicate avoidance. confidence intervals overlapping 1 indicate that use of strata occurred in proportion to availability. 0.40 0.60 0.80 1.00 1.20 1.40 panel a 1.60 unburned > 30 11-30 < 11 se le c� on (u se /a va ila bi lit y) burn age class 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 2.00 unburned > 30 11-30 < 11 se le c� on (u se /a va ila bi lit y) burn age class panel b 0.40 0.60 0.80 1.00 1.20 1.40 1.60 unburned > 30 11-30 < 11 se le c� on (u se /a va ila bi lit y) burn age class panel c alces vol. 55, 2019 julianus et al. – moose post-fire habitat 81 that moose were not using plentiful forage available in young burns; rather, use was focused in burns >11 years old. available winter browse ranged from < 1 to ~26 kg/ha across the burn strata. while consistent with a 2007 browse survey (22 kg/ha) on the refuge, these values are low compared to other areas in interior alaska. for example, estimates from similar ecological regions in interior alaska frequently average >200 kg/ha, with local estimates >400 kg/ha (paragi et al. 2008). while browse use in these regions vary, they are typically much higher than the <5% use that we measured; for example, use was >20% and as high as 49% in areas where biomass abundance was >200 kg/ha (paragi et al. 2008). but importantly, low use and consistently high twinning rates in our study area suggest that individually, moose are not negatively impacted by low browse availability (craig and stout 2014). available winter browse and summer biomass were highest in f-90 which supports our original hypothesis that this burn (1130 years post-fire) likely provides the best overall habitat of the 4 burn strata. in further support of this hypothesis was that winter offtake was highest in f-90 and f-72, and while considerable food resources (primarily birch) were available in f-05, the majority of winter browsing occurred in older stands. these results were corroborated by habitat use of the marked moose in summer and winter. selective feeding on higher quality forage is evident across all results. while winter biomass in f-05 was high relative to f-72 and unburned, it is important to note that estimated browse removal in this stratum was low (<0.5 kg/ha). we also observed that the relative abundance of birch to willow in f-05 was much higher than in other burns. despite its predominance in f-05, use was not observed, suggesting that although accessible and relatively plentiful in this burn, moose did not measurably use birch as winter forage. rather, they preferentially used willow species that are nutritionally superior to birch (hjeljord et al. 1982). we also found that the dpb of willow twigs was smaller in f-05 than the other burns (data not presented), suggesting that moose maximized browse consumption in older stands by taking larger bites, but possibly at the expense of nutritive value because digestibility declines as twig diameter increases. while the results generally support our hypothesis that 11–30 year-old burns would have high biomass, browse use did not occur strictly in proportion to availability. areas <11 years old had relatively high biomass, but browse use was minimal in these areas, and marked moose spent little time in recent burns. it is likely that vegetation/browse composition contributed to the patterns we observed, but historic moose distribution patterns (craig and stout 2011), the spatial distribution of collaring efforts (g. stout, adf&g, personal communication), and predation and predator avoidance strategies (ballard and van ballenberghe 1998) also influence relative habitat use. we found that moose in the refuge exhibited selective feeding behavior by consuming a higher relative proportion of willow than birch. they appeared to forego birch even in winter when available food resources were restricted to a few species of deciduous trees and shrubs, and avoided recent burns despite measurable food resources that were available in these areas. when they did feed in recent burns, they took smaller bites. these patterns in foraging behavior were likely a consequence of interactions between population density and habitat availability. in our study area, moose densities were moderate (craig and stout 2014), and as such, browsing pressure and competition for habitat and resources were moose post-fire habitat. – julianus et al. alces vol. 55, 2019 82 low, and browse pressure on food resources was minimal. thus, moose could afford to be selective not only as they foraged, but as they used the wide array of habitat types within their home range. our data indicate that moose were not using areas burned in the last decade, despite readily available food resources. gasaway et al. (1989) found that immigration rates are low in low to moderate density populations, as these populations are generally not constrained by limited space or food resources. similarly, schwartz and franzmann (1989) documented delayed and moderated density responses to disturbance in populations limited by predation. high density populations have undergone local density changes in as little as 2 years postfire (peek 1974a), but moose density in f-05 has remained low (craig and stout 2014). we hypothesize that this delayed population response will persist because moose in the refuge are less pressured to occupy recently burned areas because they are not habitat or forage-limited. assuming that forage availability is relatively unconstrained, on a relative scale it may be that behaviors that reduce predation risk or offspring establishing home ranges overlapping or adjacent to the cow’s home range (gasaway et al. 1985, ballard et al. 1991) are more influential on habitat use/selection. these results are particularly interesting in light of evidence suggesting a changing fire regime with larger, and more severe and frequent fires in interior alaska (kasischke et al. 2010). we found no differences in browse use in the fire severity categories in f-05 (unpublished data), although sample size was limited (n = 7). however, if a higher proportion of landscape shifted to “younger” successional stages, habitat use and preference may shift considerably; albeit, calculated preferences in habitat and forage use are often a quantitative function of relative availability and not biological importance. regardless, the effects of fire severity on post-fire vegetation will become an increasingly important factor in areas of moderate moose density. the relative effects of high-severity fires on browse quality, and how moose respond to such, are dependent on the species that regenerate/recolonize post-burn as illustrated by our disparate consumption data of willow and birch. predation on calves and yearlings in the upper koyukuk river drainage is high. calf mortality is estimated as 74% from spring parturition to population surveys in november, with 22% annual predation of yearlings, mostly by wolves (canis lupus) (adfg 2012). previous studies indicate that moose, particularly cows with calves, preferentially inhabit forest stands dominated by conifers that provide more protection from wolves and other predators (mech 1966, peterson 1977, poole et al. 2007). similarly, the marked cows showed preference for unburned stands during the calving season and >30 year-old stands throughout summer. vegetation in f-05 was characterized by homogeneous stands of early seral vegetation, and avoidance of burns <11 years old was presumably due to lack of vegetative cover and increased predation risk. although the characteristics of vegetation in f-05 will change considerably in the coming years, given the population characteristics of moose in the region, it may be a number of years before moose regularly use and establish core home ranges within f-05 and other recent burns. semi-annual moose surveys will continue on the refuge to quantify temporal changes in population and distribution. these surveys will help land and wildlife managers understand the nuances of reestablishment in recent burns, and to document changes in moose population dynamics and address broader management issues. alces vol. 55, 2019 julianus et al. – moose post-fire habitat 83 continued study of habitat and population change through time is particularly relevant in light of climate change. as deciduous forest succession becomes dominant in uplands of interior alaska, implications for moose and other species must continue to be explored. while research suggests that increase in deciduous species will benefit moose, the nutritive value of deciduous trees and shrubs varies, and other factors also influence habitat use. it is necessary to study successional change at both the landscape and individual burn scales, specifically as it relates to moose distribution and habitat use, to improve our understanding of habitat dynamics under a changing fire regime. acknowledgments this project was funded by the us fish and wildlife service region seven fire program and kanuti national wildlife refuge. we thank p. butteri, b. cogley, t. craig, l. dillard, b. haugen, r. lane, and t. st. clair for their valued field assistance. we thank pilots t. cambier, q. slade, m. spindler, and the pilots and staff of brooks range aviation. c. hamfler and j. rose provided assistance with r. any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the u.s. government. references alaska department of fish and game (adfg). 2012. operational plan for intensive management in game management unit 24b during regulatory years 2012–2017. department of wildlife conservation, juneau, alsaka, usa. ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114: 3–49. _____, and v. van ballenberghe. 1998. moose-predator relationships: research and management needs. alces 34: 91–105. bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior alaska. journal of wildlife management 66: 747–756. doi: 10.2307/3803140 boertje, r. d., m. a. keech, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314–327. doi: 10.2193/2007-591 _____, c. t. seaton, d. d. young, m. a. keech, and b. w. dale. 2000. factors limiting moose at high densities in unit 20a. federal aid in wildlife restoration research performance report, grant w-27-3, project 1.51. alaska department of fish and game, juneau, alaska, usa. bryant, j. p., and p. j. kuropat. 1980. selection of winter forage by subarctic browsing vertebrates: the role of plant chemistry. annual review of ecology and systematics 11: 261–285. doi: 10.1146/annurev.es.11.110180.001401 chapin, f. s. iii, s. f. tranor, o. huntington, a. l. lovecraft, e. zavaleta, d. c. natcher, a. d. mcguire, j. l. nelson, l. ray, m. calef, n. fresco, h. huntington, t. s. rupp, l. dewilde, and r. l. naylor. 2008. increasing wildfire in alaska’s boreal forest: pathways to potential solutions of a wicked problem. bioscience 58: 531– 540. doi: 10.1641/b580609 craig, t., and g. w. stout. 2011. aerial moose survey on and around kanuti national wildlife refuge. u.s. fish and wildlife service, kanuti national wildlife refuge, fairbanks, alaska, usa _____, and _____. 2014. aerial moose survey on and around kanuti national wildlife refuge. u.s. fish and wildlife service, kanuti national wildlife refuge, fairbanks, alsaka, usa danell, k., and l. ericson. 1986. foraging by moose on two species of birch when these occur in different proportions. moose post-fire habitat. – julianus et al. alces vol. 55, 2019 84 holarctic ecology 9: 79–84. doi: 10.1111/j.1600-0587.1986.tb01194.x dunn, j. e., and p. s. gipson. 1977. analysis of radio telemetry data in studies of home range. biometrics 33: 85–101. doi: 10.2307/2529305 franzmann, a. w., and c. c. schwartz. 1985. moose twinning rates: a possible population condition assessment. journal of wildlife management 49: 394–396. doi: 10.2307/3801540 gasaway, w. c., s. d. dubois, r. d. boertje, d. j. reed, and d. t. simpson. 1989. response of radio-collared moose to a large burn in central alaska. canadian journal of zoology 67: 325–329. doi: 10.1139/z89-047 _______, _______, d. j. preston, and d. j. reed. 1985. home range formation and dispersal of subadult moose in interior alaska. federal aid for wildlife restoration final report. alaska department of fish and game, juneau, alaska. hjeljord, o., f. sundstol, and h. haagenrud. 1982. the nutritional value of browse to moose. journal of wildlife management 46: 333–343. doi: 10.2307/3808644 johnstone, j. f. 2006. response of boreal plant communities to variations in previous fire-free interval. international journal of wildland fire 15: 497–508. doi: 10.1071/wf06012 _______, f. s. chapin iii, t. n. hollingsworth, m. c. mack, v. romanovsky, and m. turetsky. 2010a. fire, climate change, and forest resilience in interior alaska. canadian journal of forest research 40: 1302–1312. doi: 10.1139/x10-061 _______, t. n. hollingsworth, f. s. chapin iii, and m. c. mack. 2010b. changes in fire regime break the legacy lock on successional trajectories in alaska boreal forest. global change biology 16: 1281–1295. doi: 10.1111/j.1365-2486.2009.02051.x joly, k., t. craig, m. s. sorum, j. s. mcmillan, and m. a. spindler. 2015. moose movement patterns in the upper koyukuk river drainage, northcentral alaska. alces 51: 97–105. julianus, e. l., and s. longson. 2018. aerial moose survey on and around kanuti national wildlife refuge, november, 2017. u.s. fish and wildlife service, kanuti national wildlife refuge, fairbanks, alaska, usa. kasischke, e. s., and m. r. turetsky. 2006. recent changes in the fire regime across the north american boreal region – spatial and temporal patterns of burning across canada and alaska. geophysical research letters 33: 1–5. doi: 10.1029/2006gl025677 _____, _____, r. d. ottmar, n. h. f. french, e. e. hoy, and e. s. kane. 2008. evaluation of the composite burn index for assessing fire severity in alaskan black spruce forests. international journal of wildland fire 17: 515–526. _____, d. l. verbyla, t. s. rupp, a. d. mcguire, k. a. murphy, r. jandt, j. l. barnes, e. e. hoy, p. a. duffy, m. calef, and m. r. turetsky. 2010. alaska’s changing fire regime – implications for the vulnerability of its boreal forests. canadian journal of forest research 40: 1313–1324. doi: 10.1139/ x10-098 lake, b. c., m. r. bertram, n. guldager, j. r. caikoski, and r. o. stephenson. 2013. wolf kill rates across winter in a low-density moose system in alaska. journal of wildlife management 77: 1512–1522. doi: 10.1002/jwmg.603 landfire. 2008. existing vegetation type layer. landfire 1.1.0. u.s. department of the interior, u.s. geological survey. https://www.landfire.gov/index.php (accessed december 2015). loranger, a. j., t. n bailey, and w. w. larned. 1991. effects of forest alces vol. 55, 2019 julianus et al. – moose post-fire habitat 85 succession after fire in moose wintering habitats on the kenai peninsula, alaska. alces 27: 100–109. lord, r. e., and k. kielland. 2015. effects of variable fire severity on forage production and foraging behavior of moose in winter. alces 51: 23–34. maccracken, j. g., and l. a. viereck. 1990. browse regrowth and use by moose after fire in interior alaska. northwest science 64: 11–18. maier, j. a. k., j. m. verhoef, a. d. mcguire, r. t. bowyer, l. saperstein, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics: effects of scale. canadian journal of forest research 35: 2233– 2243. doi: 10.1139/x05-123 mech, l. d. 1966. the wolves of isle royale. fauna series no. 7. u. s. national park service. u.s. printing office, washington, dc, usa. neu, c. w., c. r. byers, and j. m. peek. 1974. a technique for analysis of utilization-availability data. journal of wildlife management 38: 541–545. doi: 10.2307/3800887 nowacki, g., p. spencer, m. flemming, t. brock, and t. jorgenson. 2001. ecoregions of alaska: 2001. open file report 02-297. u. s. geological survey, anchorage, alaska, usa. oldemeyer, j. l. 1974. nutritive value of moose forage. naturaliste canadien 101: 217–226. _______, a. w. franzmann, a. l. brundage, p. h. arneson, and a. flynn. 1977. browse quality and the kenai moose population. journal of wildlife management 41: 533–542. doi: 10.2307/3800528 paragi, t. f., c. t. seaton, and k. a. kellie. 2008. identifying and evaluating techniques for wildlife habitat management in interior alaska: moose range assessment. alaska department of fish and game, division of wildlife conservation, juneau, alsaka, usa. payette, s. 1992. fire as a controlling process in the north american boreal forest. pages 144–169 in h. h. shugart, r. leemans, and g. b. bonan, editors. a systems analysis of the global boreal forest. cambridge university press, cambridge, england. peek, j. m. 1974a. moose-snow relationships in northeastern minnesota. pages 39–49 in a. o. haugen, editor. proceedings of the snow and ice symposium. iowa cooperative wildlife research unit, ames, ia. _____. 1974b. initial response of moose to a forest fire in northern minnesota. naturaliste canadien 101: 131–174. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. u. s. national park service science monograph 11. u. s. government printing office, washington, dc, usa. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammalogy 88: 139–150. doi: 10.1644/06-mamm-a-127r1.1 saperstein, l., g. w. stout, and t. craig. 2009. aerial moose survey on kanuti national wildlife refuge, november 2007. u. s. fish and wildlife service, kanuti national wildlife refuge, fairbanks, alaska, usa. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose, and forest succession, some management considerations on the kenai peninsula, alaska. alces 25: 1–10. seaton, c. t., t. f. paragi, r. d. boertje, k. kielland, s. dubois, and c. l. fleener. 2011. browse biomass removal and nutritional condition of alaska moose (alces alces). wildlife biology 17: 1–12. doi: 10.2981/10-010 stout, g. w. 2010. unit 24 moose. pages 572–610 in p. harper, editor. moose management report of survey and inventory activities, 1 july 2007–30 june 2009. alaska department of fish moose post-fire habitat. – julianus et al. alces vol. 55, 2019 86 and game, division of wildlife conservation, juneau, alsaka, usa. viereck, l. a., c. t. dyrness, a. r. batten, and k. j. wenzlick. 1992. the alaska vegetation classification. general technical report pnw-gtr-286. department of agriculture, u. s. forest service, pacific northwest research station, portland, oregon. western region climate center. 2014. record monthly climate summary for bettles, alaska. noaa national climatic data center. http://www.wrcc. dri.edu/cgi-bin/climain.pl?ak0761 (accessed october 2014). wolff, j. o. 1978. burning and browsing effects on willow growth in interior alaska. journal of wildlife management 41: 135–140. doi: 10.2307/3800700 _____, and j. zasada. 1979. moose habitat and forest succession on the tanana river floodplain and yukon-tanana upland. alces 15: 213–244. alces vol. 55, 2019 julianus et al. – moose post-fire habitat 87 appendix a. preferred and non-preferred browse species classifications (based on literature review) established a priori in 4 burn strata on the kanuti national wildlife refuge, alaska, usa. preferred browse species salix alaxensis salix pulchra salix arbusculoides salix bebbiana populus. balsamifera populus tremuloides betula neoalaskana (winter only) non-preferred browse species picea mariana picea glauca alnus spp. betula glandulosa betula nana populus tremuloides1 populus balsamifera1 betula neoalaskana1 1mature individuals (>3 m tall). appendix b. sampling effort on winter browse surveys in 4 burn strata on the kanuti national wildlife refuge, alaska, usa, 2013–2014. stratum # plots # plants # twigs unburned 8 37 372 f-72 9 39 386 f-90 11 76 747 f-05 9 43 430 total 37 195 1935 moose post-fire habitat. – julianus et al. alces vol. 55, 2019 88 appendix c. elevation, slope, dominant aspect, and vegetation characteristics based on a digital elevation model (dem) and landfire data in 4 burn strata on the kanuti national wildlife refuge, alaska, usa. stratum unburned f-72 f-90 f-05 elevation (m) mean 224 213 332 261 min 121 116 160 151 max 1068 459 809 889 slope (°) mean 2.06 2 4 3 slope class (%)1 flat 12 31 16 46 gentle (<10°) 24 66 76 46 medium (10–30°) 32 2 8 9 steep (>30°) 32 0 0 0 dominant aspect southeast south southwest south canopy height (m) mean 9 8 3 1 tree height (m) max 12 10 5 4 1% of burn in each slope c. appendix d. fire severity classification in the f-05 burn stratum, kanuti national wildlife refuge, alaska, usa. classification was determined from photographs. site severity f-05 – 1 low f-05 – 2 low f-05 – 3 moderate/high f-05 – 4 moderate/low f-05 – 5 high f-05 – 6 moderate f-05 – 7 high f-05 – 8 moderate alces vol. 55, 2019 julianus et al. – moose post-fire habitat 89 appendix e. landfire classification of vegetation types in 4 burn strata on the kanuti national wildlife refuge, alaska, usa. note that vegetation types for f-05 reflect composition prior to burning. % class unburned f-72 f-90 f-05 closed tree canopy 28 50 55 15 dwarf shrubland 3 2 1 3 herbaceous – grassland 11 9 5 14 non-vegetated 7 2 2 4 open tree canopy 21 13 4 10 shrubland 30 23 34 53 sparse tree canopy 0 0 0 0 sparsely vegetated 1 0 0 1 % sub-class aquatic 2 1 0 1 deciduous 16 19 38 25 deciduous dwarf-shrubland 1 2 0 2 deciduous shrubland 23 21 20 35 evergreen 18 18 31 9 evergreen open tree canopy 17 8 3 7 mixed 2 16 0 0 mixed evergreen-deciduous open tree canopy 3 5 1 3 non-vegetated 7 3 2 4 perennial graminoid 10 8 5 11 perennial graminoid or annual 0 0 0 2 sparsely vegetated 1 1 0 1 % height class sparse 1 0 0 0 shrub > 1.5 m 2 8 21 47 shrub 0.5–1.5 m 13 2 8 0 shrub 0–0.5 m 6 8 4 0 herb >0.5 m 10 1 1 14 herb 0–0.5 m 1 0 0 0 forest >50 m 0 0 0 0 forest 5–10 m 23 15 9 25 forest 25–50 m 0 45 31 2 forest 10–25 m 22 21 25 10 forest 0–5 m 16 0 0 1 f:\alces\vol_39\p65\3926.pdf alces vol. 39, 2003 arnemo et al. – immobilization of moose 243 chemical immobilization of free-ranging moose jon m. arnemo1,2, terry j. kreeger3, and timo soveri4 1department of arctic veterinary medicine, the norwegian school of veterinary science, no-9292 tromsø, norway; 2department of forestry and wilderness management, hedmark university college, evenstad, no-2480 koppang, norway; 3wyoming game and fish department, 2362 highway 34, wheatland, wyoming 82201, usa; 4department of clinical veterinary medicine, faculty of veterinary medicine, university of helsinki, fi-04920 saarentaus, finland abstract: a wide range of drugs and drug combinations have been used to capture free-ranging moose (alces alces). currently, potent opioids are considered the drugs of choice for capture of free-ranging moose. recommended doses are carfentanil at 0.01 mg/kg or etorphine at 7.5 mg/adult. combining an opioid with a sedative drug like xylazine will increase the risk of bloat, regurgitation, and aspiration of rumen contents. extreme toxicity for humans and lost darts are major concerns when using potent opioids under field conditions. the best non-opioid alternative is medetomidine at 40-50 mg/adult combined with ketamine at 600 mg/adult. carfentanil, etorphine, and medetomidineketamine have wide safety margins in moose and the risk of severe anesthetic side effects in healthy animals is minimal. chemical immobilization from a helicopter in winter is considered the best capture method for moose. due to animal welfare considerations and a low therapeutic index, neuromuscular blocking agents should not be used in moose. a mortality rate greater than 2% during immobilization and a one month post capture period is not acceptable for routine moose captures. alces vol. 39: 243-253 (2003) key words: alces alces, anesthesia, capture, carfentanil, etorphine, immobilization, ketamine, medetomidine, xylazine free-ranging moose (alces alces) are chemically immobilized for various research and management purposes: radiotransmitter deployment, collection of biological materials, morphometry, health examination, and translocation. most moose are approached with a helicopter or occasionally by snowmobile, all-terrain vehicle, car, boat, or on foot, and drugs are administered by projectile darts fired from a dart gun. the first chemical immobilization of free-ranging moose was carried out in alaska in 1957-58 with nicotine, a neuromuscular blocking (nmb) agent (rausch and ritcey 1961). since then a wide range of drugs and drug combinations have been used to capture free-ranging moose in north america and europe, including other nmb agents, tranquilizers, sedatives, and anesthetics. franzmann (1982, 1998) has published excellent reviews of moose chemical immobilization. here we present an update on recommended drugs, doses, and methods for chemical capture of free-ranging moose. chemical capture versus net-gunning although helicopter net-gunning has been successfully used on moose, with an immediate capture mortality rate of less than 1% (carpenter and innes 1995), mortality rates as high as 14% have been reported from other projects using this method (olterman et al. 1994). there is no doubt that helicopter net-gunning is a useful method for capture of free-ranging ungulates, and in some species it is even considered to be better than chemical immobilization (kock et al. 1987a,b,c). in moose, however, we are not aware of a single immobilization of moose – arnemo et al. alces vol. 39, 2003 244 publication on stress physiology or possible long-term negative effects (e.g., exertional myopathy, increased risk of predation, reduced calving success, and reduced survival of offspring) after capture by helicopter net-gunning. whether net-gunning is an acceptable method for moose capture remains to be documented. immobilizing drugs there are three major groups of drugs currently used for wildlife capture: alpha-2 adrenoceptor agonists, opioid agonists, and cyclohexanes (kreeger et al. 2002). the nmb agents are a fourth group that was extensively used during the pioneer days of chemical immobilization. nmb drugs cause muscular paralysis but the animal is conscious, aware of its surroundings and fully sensory, and can feel pain and experience psychogenic stress. due to a very narrow range between effective immobilizing doses and lethal doses, mortality rates as high as 70% have occurred with nmb agents (kreeger et al. 2002). although inferior to modern immobilizing drugs, the nmb agent succinylcholine has been used for moose capture in recent years (delvaux et al. 1999). however, the reported mortality rate due to respiratory paralysis was 7% and only 63% of the immobilization attempts were successful. due to animal welfare considerations and the low therapeutic index (effective dose:lethal dose), succinylcholine or other nmb agents should not be used for moose immobilization. alpha-2 adrenoceptor agonists include xylazine, romifidine, detomidine, and medetomidine. these agents induce dosedependent sedation and analgesia and they have anxiolytic and muscle relaxing properties. the difference in potency between the alpha-2 agonists is species dependent, but no controlled studies have been done in moose or other wildlife species. in sheep, the equipotent sedative doses (mg/kg) for xylazine, romifidine, detomidine, and medetomidine are 0.15, 0.05, 0.03, and 0.01, respectively (kreeger et al. 2002). although these drugs may induce deep sedation and immobilization in large doses, sudden arousal may occur. in highly excited animals, induction times are usually prolonged and immobilization may be impossible regardless of the dose administered. alpha-2 agonists should therefore never be used as the sole agent for capture of freeranging moose. they are, however, very useful in combination with opioids or cyclohexanes. alpha-2 agonists have the ability to potentiate other cns-drugs; e.g., if ketamine is combined with medetomidine the effective anesthetic dose of ketamine is reduced by as much as 75% in some species (jalanka and roken 1990). the effects of alpha-2 agonists can be rapidly and permanently reversed by atipamezole, a potent and specific alpha-2 adrenoceptor antagonist (kreeger et al. 2002). other less specific reversal agents, such as yohimbine and tolazoline, can be used to antagonize xylazine, the least potent of the alpha-2 agonists. opioid agonists used for wildlife immobilization include carfentanil, etorphine, fentanyl, and, to some extent, thiafentanil and sufentanil (kreeger et al. 2002). in moose and other cervids, carfentanil (north america) and etorphine (europe) have been the primary opioids, either alone or in combination with xylazine (kreeger et al. 2002). carfentanil and etorphine both have high therapeutic indices in moose; i.e., the same dose can be used in most adults regardless of body weight. underdosing with opioids may cause excitement and hyperthermia and overdosing is therefore considered to be better than underdosing. although not a “new” agent for wildlife captures (stanley et al. 1988, 1989), thiafentanil (formerly identified as a-3080) is still an investigational drug for wild animal capture (citino et al. alces vol. 39, 2003 arnemo et al. – immobilization of moose 245 2001, grobler et al. 2001, kreeger et al. 2001, citino et al. 2002). the relative potencies of carfentanil, etorphine, and thiafentanil in moose are approximately 2:1:1 (mcjames et al. 1994, kreeger et al. 2002). the effects of opioids can be reversed by several opioid antagonists. naltrexone is the preferred agent due to its potency and l o n g d u r a t i o n ( i . e . , l e s s r i s k o f renarcotization). other opioid antagonists include naloxone, nalmefene, nalbuphine, and diprenorphine. cyclohexanes (also known as nmda antagonists) include ketamine and tiletamine. these drugs are general anesthetics; i.e., they induce unconsciousness and amnesia. however, due to severe side effects like muscle rigidity, frequent convulsions, and rough recoveries, these agents should only be used in combination with an alpha-2 agonist or another tranquilizing or sedative drug (kreeger et al. 2002). the relative potency between tiletamine and ketamine is approximately 2.5:1 and the duration of action of tiletamine is about three times longer than with ketamine. tiletamine is not available as a single product and is marketed in a 1:1 combination with the benzodiazepine agonist zolazepam. there is no reversal agent to the cyclohexane drugs. too early administration of an alpha2 antagonist in animals immobilized with an alpha-2 agonist in combination with ketamine or tiletamine, may uncover residual side effects of the cyclohexane component and can cause uncontrolled recoveries, hyperthermia, trauma, and even death (kreeger et al. 2002). in general, antagonists should be administered intramuscularly. intravenous injection of the reversal agent will cause complete recovery in less than one minute in animals immobilized with opioids alone. such rapid recoveries may be stressful to the animals and may jeopardize the safety of both animals and people. intravenous administration of reversal agents should therefore only be considered in an emergency situation. carfentanil a large number of free-ranging moose have been immobilized with either carfentanil alone or carfentanil combined with xylazine (franzmann 1982, 1998; roffe et al. 2001; kreeger et al. 2002). recommended doses of carfentanil alone are 0.01 mg/kg or 3-6 mg/adult. carfentanil is marketed as a 3 mg/ml solution (wildnil®, wildlife pharmaceuticals inc., ft. collins, colorado, usa) and the dose for an adult moose will fit into a standard dart of most remote drug delivery systems. for reversal, naltrexone at 100 mg per mg carfentanil should be administered (kreeger et al. 2002). in several studies carfentanil at 3-4 mg/ adult has been used in combination with xylazine (e.g., cervizine® 10 mg/ml, wildlife pharmaceuticals inc.) at 25-175 mg/ adult to improve muscle relaxation and to potentiate the effect of carfentanil so that the opioid part of the combination can be reduced. however, moose immobilized with carfentanil-xylazine are usually not able to support sternal recumbency and may be more susceptible to aspiration pneumonia (kreeger 2000). unless there are overriding considerations, the addition of xylazine to opioids in moose is not recommended (kreeger et al. 2002). if carfentanil is combined with xylazine, the effects of xylazine should be antagonized by either atipamezole (antisedan® 5 mg/ml, orion pharma animal health, turku, finland) at 1 m g p e r 1 0 m g x y l a z i n e , y o h i m b i n e (antagonil® 5 mg/ml, wildlife pharmaceuticals inc.) at 1 mg per mg xylazine, or tolazoline (tolazoline® 100 mg/ml, lloyd laboratories, shenandoa, iowa, usa) at 15 mg per mg xylazine (roffe et al. 2001, kreeger et al. 2002, plumb 2002). immobilization of moose – arnemo et al. alces vol. 39, 2003 246 etorphine etorphine, alone or in combination with xylazine, has been the drug of choice for moose capture in scandinavia (sandegren et al. 1987, arnemo et al. 2001). standard doses are 7.5 mg etorphine/adult (etorphine hcl® 9.8 mg/ml, vericore veterinary products, novartis animal health uk ltd., litlington, uk) or 2.25 mg etorphine + 10 mg acepromazine/adult (large animal immobilon® 2.25 mg/ml, vericore veterinary products, novartis animal health uk ltd.) combined with 100 mg xylazine/adult (rompun® dry substance, bayer ag, leverkusen, germany). these doses fit into a standard dart of most remote drug delivery systems. data from 1,464 immobilizations carried out over a 19-year period in norway (arnemo et al. 2001; j. m. arnemo, unpublished data) show that ethorphine alone is an extremely safe and effective drug in moose and there is no indication for combining etorphine with an alpha-2 agonist. due to the potentiating effect and muscle relaxing properties of alpha-2 agonists, moose immobilized with e t o r p h i n e x y l a z i n e o r e t o r p h i n e medetomidine are usually not able to maintain sternal recumency and regurgitation of rumen contents are frequently seen (j. m. arnemo, unpublished data). diprenorphine is a specific antagonist for etorphine and is marketed in the same package as etorphine at a concentration of 1.2 times the concentration of etorphine (diprenorphine hcl® 12 mg/ml and large animal revivon® 3 mg/ml, vericore veterinary products, novartis animal health uk ltd.). for reversal of etorphine effects in moose, the volume of diprenorphine should be equivalent to the total volume of etorphine administered. if etorphine is combined with xylazine or medetomidine, the effects of the alpha-2 agonist should be reversed by atipamezole (antisedan® 5 mg/ml, orion pharma animal health, turku, finland) at 1 mg per 10 mg xylazine or 5 mg per mg medetomidine (kreeger et al. 2002). thiafentanil we are aware of only two reports on the use of thiafentanil for immobilization of free-ranging moose. in one study average down time in moose (n = 18) darted with a standard dose of 10 mg thiafentanil was 1.5 min compared to 4.5 min in moose (n > 100) injected with a standard dose of 4.5 mg carfentanil (stanley et al. 1989). reversals of immobilization were achieved with either nalmefene or diprenorphine (no data on a n t a g o n i s t d o s e s w a s p r o v i d e d ) . renarcotization in animals immobilized with thiafentanil was not observed and the authors state that the elimination half-life of thiafentanil is only half as long as the elimination time of carfentanil. later, mcjames et al. (1994) reported that a standard dose of 10 mg thiafentanil was used to immobilize moose from a helicopter in winter. the mean induction time in 59 moose immobilized after one injection was 3.6 min. the 10 mg dose was effective for large bulls and safe for calves. three animals required a second dart to become immobilized and received a total dose of 20 mg thiafentanil. reversals after different doses of nalmefene (50 and 300 mg) and naltrexone (50 and 100 mg) were rapid and complete with no residual ataxia. mean standing times ranged from 1.9 to 2.4 min after intramuscular administration of the antagonist in all groups. renarcotization was not seen and no deaths occurred. although more studies on its efficacy and safety are required, there are strong indications that thiafentanil may be a very useful drug for immobilization of moose in the future: small volume (1 ml), induction time is rapid, duration of action is short, no major clinical side effects have been reported, and renarcotization has not been observed. this view is supported by several studies on thiafentanil in other alces vol. 39, 2003 arnemo et al. – immobilization of moose 247 artiodactylid species (stanley et al. 1988, janssen et al. 1993, mcjames et al. 1993, citino et al. 2001, grobler et al. 2001, kreeger et al. 2001, citino et al. 2002). currently, thiafentanil is only available for investigational purposes (a3080® 10 mg/ ml, wildlife pharmaceuticals inc.). medetomidine-ketamine studies on medetomidine (med), ketamine (ket), and atipamezole (ati) in free-ranging moose were performed in norway and finland from 1992 to1997. although some of the data from these studies have been printed in non-indexed sources (arnemo et al. 1994, arnemo 1995, arnemo et al. 1996), they are not easily available to the scientific community. in addition, a lot of useful information is not yet published (j. m. arnemo and t. soveri, personal observations). a summary of the results is therefore included here. in summer, 30 mg med + 400 mg ket (n = 15), 30 mg med + 500 mg ket (n = 3), and 40 mg med + 500 mg ket (n = 4) were used to immobilize adults from ground (on foot, stalking, and from a motor vehicle). for reversal, all animals received 5 mg ati per mg med, half intravenously or intramuscularly and half subcutaneously. only a few of the animals were actually seen going down and to avoid stress and excitement during induction, the standard procedure was to wait for 10 min after darting before tracking with a dog was initiated. mean time (range) from darting until the animal was found was 18 (1-35) min for animals completely immobilized after one dart injection. mean estimated distance (range) covered after darting was 300 (10-750) m. two animals darted with 30 med + 500 ket and one animal darted with 40 mg med + 500 mg ket required reiteration with a full initial dose to become completely immobilized. one cow in poor body condition darted with 30 mg med + 400 mg ket developed periodic apnea after 45 min and was treated with ati (half intravenously and half subcutaneously) to reverse immobilization. one cow injected twice with 40 mg med + 500 mg ket apparently stopped breathing 40 min after the initial darting and was treated with doxapram (dopram® 20 mg/ml, wyeth lederle, wyeth-ayerst international inc., philadelphia, pennsylvania, usa) at 1 mg/ k g i n t r a v e n o u s l y a n d a t i ( h a l f intramuscularly and half subcutaneously). one cow (400 kg) was found drowned in a small creek 13 min after darting with 40 mg med + 500 ket, 200 m from where she was shot. necropsy (national veterinary institute, oslo, norway) revealed no underlying pathological conditions. no other immediate mortalities occurred. recoveries were uneventful and all animals were standing in less than 11 min after administration of ati. animals monitored by radiotracking (n = 17) survived at least 2 months post capture. data on physiologic parameters (rectal temperature, heart rate, respiratory rate, and relative arterial oxygen saturation) are found in arnemo et al. (1994) and arnemo (1995). in winter, 30-40 mg med + 500 mg ket induced complete immobilization in 6 out of 8 adult cows darted from a motor vehicle at a bait. four animals were observed going down after a mean induction time (range) of 7 (4-11) min while 2 individuals were found after 22 and 43 min, respectively. in two animals the initial dose did not induce recumbency, one of them was manually restrained while the other was captured after an additional dose of 6 mg etorphine. all animals received 5 mg ati per mg med for reversal. recoveries were uneventful and all animals were on their feet in less than 13 min after administration of ati. all animals were monitored by radiotracking and survived for at least 9 months post capture. physiologic data are immobilization of moose – arnemo et al. alces vol. 39, 2003 248 found in arnemo et al. (1994). in winter, 8 adult cows and 5 bulls were darted from a helicopter with 30 mg med + 400 mg ket (n = 2) or 40 mg med + 500 mg ket (n = 11). two animals receiving the highest dose required reiteration with a full initial dose. mean time (range) from darting to recumbency in animals completely immobilized after one injection was 8 (4-15) min. two animals injected with the highest dose showed signs of respiratory depression with shallow breathing and periodic apnea. reversals were achieved with ati at 5 mg per mg med injected half intramuscularly and half subcutaneously. one cow that apparently stopped breathing 40 min after darting was treated with doxapram at 1 mg/kg intravenously in addition to ati, while inspirations were induced by manual chest compressions. this individual recovered completely. twelve of the animals were on their feet in less than 10 min after administration of ati. one bull immobilized with the lowest dose became fully alert after injection of ati but was apparently unable to get up. the bull was net-lifted with a helicopter to a safe area and was left in sternal recumbency 2.5 hrs post darting. next morning the bull was tracked for > 1 km with the helicopter but was not observed. all animals were monitored by radiotracking and survived for at least 6 months post capture. based on clinical examination of each individual and the actual body mass of the bull that remained recumbent (240 kg), all animals in this part of the study were in very poor body condition. physiologic data are found in arnemo et al. (1994). a major part of the med-ket and ati evaluation was carried out during 5 winters in finland from 1993 to 1997. a total of 92 moose were darted from a helicopter: 26 calves (10 females, 16 males), 20 yearlings (17 were of known age) (7 females, 13 males), 26 adult cows, and 20 adult bulls. standard initial doses were 30 mg med + 400 mg ket in calves, 40 mg med + 400 ket in yearlings, and 40 or 50 mg med + 600 ket in adults. mean times (range) from darting to recumbency in animals that became completely immobilized after one dart injection were 4.4 (2-7) min in calves (n = 20), 7.6 (5-11) min in yearlings (n = 14), 6.0 (3-10) min (n = 22) in cows, and 5.9 (1-12) min in bulls (n = 14). animals that required additional dosing to induce complete immobilization, animals that were darted more than once due to malfunctioning darts or bounce offs, and animals that were not observed going down, were not included in the analyses. no animals died during immobilization. however, a total of 4 animals (4.3%) died or were euthanized within 24 hrs post capture. one bull was unable to get up after administration of ati and was found dead next day. necropsy (national veterinary institute, oulu, finland) showed very poor body condition, massive lungworm infestation, and signs of circulatory failure. one small calf which was unable to get up after injection of ati was euthanized next day. necropsy (national veterinary institute) showed very poor body condition, osteoporosis, and fractures in scapula and metatarsus. one small calf and one yearling, both in very poor body condition, recovered after injection of ati but were found dead next day 100 and 300 m, respectively, from the marking place. necropsies were not carried out on these two individuals. three of the deaths (both calves and the yearling) occurred in 1996, a year with extremely harsh winter conditions which caused poor body conditions in most of the captured animals. all animals were monitored by radiotracking and no other mortalities occurred within 2 months post capture. the complete set of data from this trial, including serum biochemistry, is currently being analyzed for publication elsewhere. alces vol. 39, 2003 arnemo et al. – immobilization of moose 249 other drugs drug combinations like xylazineketamine, xylazine-tiletamine/zolazepam, or medetomidine-tiletamine/zolazepam are not recommended for capture of free-ranging moose (franzmann 1982, kreeger et al. 2002; j. m. arnemo and t. j. kreeger, unpublished data). rapid induction is of paramount importance in wildlife capture operations and the enzyme hyaluronidase has been used to increase drug absorption rate (haigh 1979, kreeger et al. 2002). however, induction time is more dependent on the injection site and drug dose, and hyaluronidase is probably of benefit only for sub-optimal hits and doses. in addition, there are concerns regarding the stability of the drug mixture and also the epizootiological aspects of the enzyme that is a biological product extracted from bovine testes. moose are often captured during low ambient temperatures in winter and propylene glycol has been added to immobilizing mixtures (xylazine-tiletamine/zolazepam) to avoid freezing (kreeger et al. 1995). however, in moose darted with 7.5 mg etorphine (1 ml) from a helicopter in winter using standard remote drug delivery equipment (dan-inject®, børkop, denmark), addition of propylene glycol (0.5 ml) caused delayed inductions (j. m. arnemo, unpublished data). ten adult cows were immobilized on 10-11 december 1999, half of them received etorphine and propylene glycol (group 1) and the other half received etorphine only (group 2). mean times (range) from initial darting to recumbency were 18 (11-32) min in group 1 and 5 (2-7) min in group 2. two animals in group 1 required a second dart for immobilization. mean estimated distances (range) covered after darting were approximately 3 (2-5) km in group 1 and 0.4 (0.1-0.7) km in group 2. the use of propylene glycol as an antifreeze in etorphine mixtures cannot be recommended for moose immobilization. monitoring and risks anesthetic monitoring after capture, immobilized moose should be examined and monitored by a wildlife veterinarian. clinical problems or injuries should be treated according to established standards in veterinary medicine (kreeger et al. 2002). dart wounds are extremely rare in moose if lightweight darts with low impact energy and modern remote drug delivery equipment are used. to avoid bloat and to reduce the risk of regurgitation and aspiration of rumen contents, captured moose should be kept in sternal recumbency with the head higher than the body and the nose lower than the neck. the use of head covers/blinds and ear plugs will reduce stress in animals during handling. for safety reasons, the feet of immobilized animals should be hobbled. franzmann et al. (1984) established baseline values for rectal temperature (rt), heart rate (hr), and respiratotory rate (rr) in chemically immobilized moose and safe expected ranges were 38.4-38.9 ºc, 70-91 beats/min, and 13-40 respirations/min, respectively. critical values for corrective actions were rt 40.2 ºc, hr 102 beats/ min, and 40 respirations/min. based on personal experience with moose immobilization, we consider these values to be conservative. assessment of respiration in immobilized animals can be done by monitoring of the relative arterial oxygen saturat i o n ( s p o 2 ) w i t h a p u l s e o x i m e t e r . hypoxemia is defined as spo 2 < 90%. in field situations, however, spo 2 values markedly below 90% are often recorded, apparently with no harm to the animal. a critical spo 2 value has not been defined but one of the authors (j. m. arnemo) usually institutes corrective actions (administration of supplemental oxygen, respiratory stimulants, or specific antagonists) when the spo 2 falls immobilization of moose – arnemo et al. alces vol. 39, 2003 250 below 70%. the trend of spo 2 values is probably more important than the absolute values and if the spo 2 steadily decreases, it can be presumed that the animal is in some sort of respiratory crisis (kreeger et al. 2002). exertional myopathy exertional myopathy (commonly referred to as capture myopathy) is a wellknown, usually fatal syndrome in free-ranging artiodactylids (spraker 1993, williams and thorne 1996). exertional myopathy may be caused by several factors, such as stress, chasing, restraint, and transportation. clinical signs of exertional myopathy may become apparent during the capture p r o c e s s o r m a y o c c u r w i t h i n h o u r s postcapture. it is, however, important to note that the pathologic manifestations of exertional myopathy can be delayed for up to a month following capture before the animal eventually dies (spraker 1993, williams and thorne 1996). any evaluation of capture methods and drugs in free-ranging moose should therefore include a minimum of 4 weeks follow-up by radiotelemetry to detect delayed mortalities caused by exertional myopathy. risk of chemical capture in moose, chemical immobilization is an invaluable tool both for management and research. since the pioneer days of the 1950s and 1960s, a large number of freeranging moose have been chemically immobilized for various purposes. during the initial phase of moose chemical capture, mortality rates were often very high. in some instances as much as 26% of the animals died, either during the capture process, during transport, or shortly after release (franzmann 1982). main causes of mortality were respiratory depression, cardiovascular collapse, hyperthermia, trauma, stress, and exertional myopathy. efficient drugs and antagonists have been available for reversible immobilization of moose for at least 2 decades. in addition, remote delivery systems and lightweight darts were developed for non-traumatic administration of drugs. access to portable and easy to use monitoring devices like pulse oximeters further improved animal safety during field anesthesia. in spite of this progress, reported mortality rates often range from 6 to 19% in moose captured with carfentanil combinations (roffe et al. 2001). in contrast, only 7 animals (0.5%) died during 1,464 immobilizations carried out with etorphine from helicopter over a 19-year period in norway (arnemo et al. 2001; j. m. arnemo, unpublished data). more than 97% of the animals in this study were monitored by radiotracking and no mortalities due to resedation, predation, or exertional myopathy occurred. in a review of stress and exertional myopathy in artiodactylids, spraker (1993) stated that a mortality rate greater than 2% during trapping is not acceptable. we believe that this rule should be applied also to chemical capture situations: a capture related mortality rate greater than 2% during chemical immobilization and a 1-month follow-up period is not acceptable for routine captures of moose and requires that the capture protocol be re-evaluated. at least this should be the rule of thumb when a large number (n > 100) of free-ranging moose are being chemically immobilized. human safety although animal welfare is important, the first concern when dealing with wild animals should be the safety of humans (fowler 1995). carfentanil, etorphine, and thiafentanil are up to 10,000 times more potent than morphine and minuscule amounts of drug are theoretically lethal to people (kreeger et al. 2002). extreme care should therefore be taken when working with poalces vol. 39, 2003 arnemo et al. – immobilization of moose 251 tent opioids and lost darts should be of major concern. other drugs and drug combinations at doses prepared for moose are also potentially dangerous and all personnel involved in moose captures should therefore be qualified to perform first aid on humans. a brief update on human medical treatment following accidental exposure to immobilising drugs is found in kreeger et al. (2002). most drugs used for moose capture are colourless. as a safety precaution, drugs may be coloured to make it easier to detect leakage from vials, needles, darts, and injection sites. congo red and cobalt blue are commonly used for this purpose (nielsen 1999). the use of dart guns requires an understanding of ballistics and gun safety and readers are referred to recent publications on wildlife chemical immobilization (nielsen 1999, kreeger et al. 2002). overviews of safety aspects regarding helicopter operations were provided by nielsen (1999). recommendations for routine immobilization of free-ranging moose, we recommend carfentanil at 0.01 mg/kg or etorphine at 7.5 mg total dose per adult. at these doses most animals are able to maintain sternal recumbency. we do not advocate combining opioids with xylazine or other sedative drugs because this will often induce lateral recumbency and thereby increase the risk for tympany, regurgitation, and aspiration of rumen contents. carfentanil and etorphine have a wide safety margin in moose and the risk of severe anesthetic side effects during immobilization is minimal. medetomidineketamine is a useful non-opioid alternative. neuromuscular blocking agents should never be used in moose. in our opinion, a mortality rate greater than 2% is not acceptable for routine captures of moose. references arnemo, j. m. 1995. immobilization of free-ranging-moose (alces alces) with medetomidine-ketamine and remobilization with atipamezole. rangifer 15:1925. , a. s. blix, ø. os, and t. soveri. 1996. reversible immobilization of arctic ungulates using medetomidineketamine. page 59 in proceedings of the 45th annual wildlife disease association conference, fairbanks, alaska, usa. 22-25 july 1996. , e. o. oen, and m. heim. 2001. risk assessment of etorphine immobilization in moose: a review of 1,347 captures. page 179 in proceedings of the society for tropical veterinary medicine and wildlife disease association i n t e r n a t i o n a l j o i n t c o n f e r e n c e , pilanesberg national park, south africa. 22-27 july 2001. , t. soveri, and n. e. søli. 1994. immobilization of free-ranging moose (alces alces) with medetomidineketamine and reversal with atipamezole. pages 197-199 in r. e. junge, editor. proceedings of the american association of zoo veterinarians and association of reptilian and amphibian veterinarians annual conference, pittsburgh, pennsylvania, usa. 23-27 october 1994. carpenter, l. h., and j. i. innes. 1995. helicopter netgunning: a successful moose capture technique. alces 31:181184. citino, s. b, m. bush, d. grobler, and w. lance. 2001. anaesthesia of roan antelope (hippotragus equinus) with a combination of a3080, medetomidine and ketamine. journal of the south african veterinary association 72:2932. , , , and . 2002. anaesthesia of boma-captured immobilization of moose – arnemo et al. alces vol. 39, 2003 252 lichtenstein’s hartebeest (sigmoceros lichtensteinii) with a combination of t h i a f e n t a n i l , m e d e t o m i d i n e , a n d ketamine. journal of wildlife diseases 38:467-462. delvaux, h., r. courtois, l. breton, and r. patenaude. 1999. relative efficiency of succinylcholine, xylazine, and carfentanil/xylazine mixtures to immobilize free-ranging moose. journal of wildlife diseases 35:38-48. fowler, m. e. 1995. restraint and handling of wild and domestic animals. second edition. iowa state university press, ames, iowa, usa. franzmann, a. w. 1982. an assessment of chemical immobilizatiuon of north american moose. pages 393-407 in l. m. nielsen, j. c. haigh, and m. e. fowler, editors. chemical immobilization of north american wildlife. the wisconsin humane society, milwaukee, wisconsin, usa. . 1998. restraint, translocation and husbandry. pages 519-537 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , c. c. sc h w a r t z, and d. c. johnson. 1984. baseline body temperatures, heart rates, and respiratory rates of moose in alaska. journal of wildlife diseases 20:333-337. grobler, d., m. bush, d. jessup, and w. lance. 2001. anaesthesia of gemsbok (oryx gazella) with a combination of a3080, medetomidine and ketamine. journal of the south african veterinary association 72:81-83. haigh, j. c. 1979. hyaluronidase as an adjunct in an immobilizing mixture for moose. journal of the american veterinary medical association 175:916917. jalanka, h. h., and b. o. roken. 1990. t h e u s e o f m e d e t o m i d i n e , medetomidine-ketamine combinations, and atipamezole in nondomestic mammals: a review. journal of zoo and wildlife medicine 21:259-282. janssen, d. l., g. e. swan, j. p. raath, s. w. mcjames, j. l. allen, v. devos, k. e. williams, j. m. anderson, and t. h. stanley. 1993. immobilization and physiological-effects of the narcotic a 3 0 8 0 i n i m p a l a ( a e p y c e r o s melampus). journal of zoo and wildlife medicine 24: 11-18. kock, m. d., r. k. clark, c. e. franti, d. a. jessup, and j. d. wehausen. 1987a. effects of capture on biological parameters in free-ranging bighorn sheep (ovis canadensis): evaluation of normal, stressed and mortality outcomes and documentation of postcapture survival. journal of wildlife diseases 23:652662. , d. a. j essup, r. k. clark, and c. e. franti. 1987b. effects of capture on biological parameters in free-ranging bighorn sheep (ovis canadensis): evaluation of drop-net, drive-net, chemical immobilization and the net-gun. journal of wildlife diseases 23:641-651. , , r. a. kock, c. e. franti, and r. a. weaver. 1987c. capture methods in five subspecies of freeranging sheep: an evaluation of dropnet, drive-net, chemical immobilization and the net-gun. journal of wildlife diseases 23:634-640. kreeger, t. j. 2000. xylazine-induced aspiration pneumonia in shira’s moose. wildlife society bulletin 28:751-753. , j. m. arnemo, and j. p. raath. 2002. handbook of wildlife chemical immobilization. international edition. wildlife pharmaceuticals inc., fort collins, colorado, usa. , w. e. cook, c. a. piché, and t. alces vol. 39, 2003 arnemo et al. – immobilization of moose 253 smith. 2001. annesthesia of pronghorns using thiafentanil or thiafentanil plus xylazine. journal of wildlife management 65:25-28. , d. l. hunter, and m. r. johnson. 1995. immobilization protocol for freeranging gray wolves (canis lupus) translocated to yellowstone national park and central idaho. pages 529530 in r. e. junge, editor. proceedings of the american association of zoo veterinarians, wildlife disease association, and american association of wildlife veterinarians joint conference, east lansing, michigan, usa. 12-17 august 1995. mcjames, s. w., j. f. kimball, and t. h. stanley. 1994. immobilization of moose with a-3080 and reversal with nalmefen hcl or naltrexone hcl. alces 30:2124. , i. l. smith, t. h. stanley, and g. painter. 1993. elk immobilization with potent opioids: a-3080 vs. carfentanil. pages 418-419 in r. e. junge, editor. proceedings of the american association of zoo veterinarians annual conference, saint louis, missouri, usa. 10-15 october 1993. nielsen, l. 1999. chemical immobilization of wild and exotic animals. iowa state university press, ames, iowa, usa. olterman, j. h., d. w. kenvin and r. c. kufeld. 1994. moose transplant to southwestern colorado. alces 30:1-8. plumb, d. c. 2002 veterinary drug handbook. third edition. iowa state university press, ames, iowa. usa. rausch, r. a., and r. w. ritcey. 1961. narcosis of moose with nicotine. journal of wildlife management 25:326328. roffe, t. j., k. coffin, and j. berger. 2001. survival and immobilizing moose with carfentanil and xylazine. wildlife society bulletin 29:1140-1146. sandegren, f., l. pettersson, p. ahlqvist, and b. o. röken. 1987. immobilization of moose in sweden. swedish wildlife research supplement 1:785-791. spraker, t. r. 1993. stress and capture myopathy in artiodactylids. pages 481488 in m. e. fowler, editor. zoo and wild animal medicine: current therapy 3. w. b. saunders, philadelphia, pennsylvania, usa. stanley, t. h., s. mcjames, and j. kimball. 1989. chemical immobilization for the capture and transportation of big game. pages 13-14 in j. h. olsen and m. eisenacher, editors. proceedings of the a m e r i c a n a s s o c i a t i o n o f z o o v e t e r i n a r i a n s a n n u a l m e e t i n g , greensboro, north carolina, usa. 1419 october 1989. , , , j. d. port, and n. l. pace. 1988. immobilization of elk with a-3080. journal of wildlife management 52:577-581. williams, e. s., and e. t. thorne. 1996. exertional myopathy. pages 181-193 in a. fairbrother, l. l. locke, and g. l. hoff, editors. noninfectious diseases of wildlife. second edition. manson publishing, london, uk. immobilization of moose – arnemo et al. alces vol. 39, 2003 254 alces37(2)_475.pdf alces vol. 47, 2010 kantar controlled moose hunt in maine 83 broccoli and moose, not always best served together: implementing a controlled moose hunt in maine lee e. kantar maine department of inland fisheries and wildlife, 650 state st., bangor, maine 04401, usa abstract: in eastern aroostook county, maine abundant populations of moose (alces alces) within an agricultural-woodland setting negatively impact cole crops and incur a high rate of moose-vehicle collisions. despite increases in antlerless hunting permits and relatively high hunter success rates, the recreational hunting framework was not effective in reducing these negative impacts, and hunter behavior had strained landowner relations and reduced access. continuing landowner relation problems and loss of access were counterproductive to the effective distribution of hunters and reducing moose abundance. in 2009 a controlled moose hunt was implemented to reduce immediate impacts on cole crops by moose, affect short-term population reduction, and facilitate cooperation and communication among stakeholders. this paper describes the rationale and framework for implementation of the controlled moose hunt, use of a co-managerial approach, and how the hunt addressed moose management goals and objectives. development and application of this controlled moose hunt in maine provides managers with another critical tool to affect population trajectory and address tangible social issues associated with moose populations above social carrying capacity. alces vol. 47: 83-90 (2011) key words: agricultural damage, conflict, harvest strategies, hunting, maine, moose, population management, stakeholder participation. agricultural damage by a wide range of wildlife species has been a chronic problem across north america (conover and decker 1991). damage and economic losses from deer (odocoileus spp.), elk (cervus elaphus), and moose (alces alces) can be both extensive and intensive depending on local populations and specific agricultural crops. as a consequence, controlling damage by deer has become a critical element of state and federal agency duties (smith and coggin 1984). however, moose damage to agricultural crops is not well documented in north america which is likely due to few moose occupying agriculturally dominated landscapes, and the relative scarcity of commercial farmlands in typical moose habitat. research and management of moose related to “crop” damage is usually associated with commercial forestlands (e.g., andren and anglestam 1993, gunderson et al. 2004, bergeron et al. 2011) as opposed to orchard damage by deer (mower et al. 1997) or elk damage to haystacks (kantar 2002). moose are managed by the maine department of inland fisheries and wildlife (mdifw) under a moose management system (morris 2002) that describes both the decision-making process and management actions that develop population goals and objectives set by a public working group. the 3 primary management approaches include recreational hunting, public safety, and compromise areas that seek to balance the positive social aspects of moose hunting and viewing with the negative impacts of road collisions and crop damage. cole crops (i.e., broccoli and cauliflower) are important commercially in eastern aroostook county in the northeastern portion of maine, and are highly palatable to moose. moose cause extensive damage by feeding directly on plants and as they move through croplands. spruce (picea spp.)-fir (abies balsamea) woodlands and wetlands provide controlled moose hunt in maine kantar alces vol. 47, 2011 84 ample cover, forage, and resting sites for moose in close proximity to these croplands. farmers and mdifw personnel have documented >40 moose in a single field (r. hoppe, mdifw, pers. comm.), and moose may intensively and continuously use these areas for >4 months. in particular, broccoli is very frost tolerant and becomes more palatable after a heavy frost increases its sugar content (d. hentosh, local farm manager, pers. comm.). its use and associated damage increase throughout fall as woody browse senesces. techniques to minimize and prevent wildlife damage typically follow a step-down approach (mdifw administrative nuisance policy j1.6) incorporating deterrents, repellents, hazing (i.e., cracker shells, trained domestic dogs), and fencing. however, several years of local hazing proved ineffective especially during the september-october breeding season. when non-lethal approaches fail to prevent damage, provide necessary relief, or cannot be applied practically, lethal removal may ensue. using hunters to affect population change can help reduce crop damage (conover 2001), increase cooperation with landowners, and improve agency credibility in resolving conflict (chase et al. 2000). in 1999 a big game public working group (wg) was formed to “guide and develop” moose management goals and objectives over the next 10 years. the wg defined a compromise management area as a wmd where current (i.e., 2000) population levels were too high and would be reduced to minimize moosevehicle collisions (table 1). the population reduction would be balanced with the ability to provide hunting opportunity and maintain a population comprised of 17% bulls >4 years old. herd composition would be determined annually from moose surveys by deer hunters and the age of harvested moose. moose hunting had occurred in and around eastern aroostook county croplands since 1980, and wildlife management districts (wmd) 3 and 6 were managed as a compromise management area. permits in 1999 were either any-moose (amp) or antlerless permits (aop). in 2003 to provide better control over harvest composition, these districts received both bull-only (bop) and aop allocations; by 2003 wmds 3 and 6 were allocated 670 permits, 290 aop and 380 bop. the following year with the moose population still above the desired level, an additional 195 aop and 65 bop were added; in 2007 an additional 25 aop were added for a total allocation of 510 aop and 445 bop (table 2). in 2001 moose populations in wmds 3 and 6 were estimated at 4.9 and 1.2 moose/km2 based on deer hunter survey data (bontaites et al. 2000). from 2001-2009 moose populations remained above goal and damage to crops and annual moose-vehicle collisions (234) were town 2001 2002 2003 2004 2005 2006 2007 2008 2009 caribou 24 15 21 19 19 19 33 15 21 connor twp 6 12 10 8 11 8 5 5 4 easton 2 6 8 2 9 5 5 1 3 fort fairfield 9 6 6 7 13 6 9 7 6 limestone 4 4 4 3 4 1 2 2 4 presque isle 12 9 16 22 15 19 27 12 10 washburn 3 3 1 5 5 4 5 3 3 westfield 6 8 3 6 6 1 3 1 3 woodland 8 2 6 6 6 7 10 2 6 total 74 65 75 78 88 70 99 48 60 table 1. annual moose-vehicle collisions within townships in the controlled hunt area, 2001-2009, maine, usa. alces vol. 47, 2010 kantar controlled moose hunt in maine 85 at high levels relative to the rest of maine; in response, aop continued to increase (table 2) to reach the population objective. hunter success was consistently high for both permit types (82 and 81%); however, despite an apparent downward trend of moose in wmd 6 (fig. 1), local moose-vehicle collisions and crop depredation warranted further remediation. under maine statute, authority is given under the nuisance animal law (chapter 921, sec. 12402-1 and 2) to address specific crop or orchard damage. except for grasses, clovers, and grain fields, farmers “may take or kill wild animals night or day, when wild animals are located within the orchard or crop, and where substantial damage to the orchard or crop is occurring.” section 12402-2 specifies that a game warden may issue depredation permits authorizing farmers to employ agents to kill wildlife observed damaging qualifying crops or nursery and orchard stock. depredation permits typically identify a specific individual(s) as the shooter, a specific location and crop, and a specific number of offending animals to be killed over a specified time frame. farmers with depredation permits had removed 2-10 moose annually from croplands, and hunting pressure was high around croplands during the recreational hunt; approximately 60-70 moose were removed from a single farm during 10 years of recreational hunting (d. hentosh, pers. comm.). while this removal likely alleviated some crop depredation, it also resulted in trespass issues, damage to agricultural fields, and associated problems within farmlands. one alternative to depredation permits is a controlled hunt, and under maine statute (chapter 903, sec. 10105-1), the mdifw commissioner has the authority to issue permits for the taking of wildlife, including controlled hunts. the purpose of a controlled hunt is to reduce negative impacts caused by wildlife. although a controlled hunt can occur within a recreational hunt, hunts outside this timeframe are permissible. the mdifw may limit the number of participants during controlled hunts, and biologists authorize hunting methods, weapons, bag limits, and other provisions to ensure the harvest. importantly, moose killed during controlled hunts would not count against bag limits specified for the recreational hunting season. to initiate a controlled hunt the mdifw proposes rule making that is reviewed by the mdifw advisory council (mdifw commissioner and 10 county representatives) after receiving public comments for 3 months within a 3-step process; at the third step the rule is voted on by the advisory council. to address negative landowner-hunter interactions, reduce the number of moose dam2001 2002 2003 2004 2005 2006 2007 2008 2009 wmd 3 amp* 175 175 0 0 0 0 0 0 0 aop* 100 100 150 220 220 220 230 230 230 bop* 0 0 160 225 225 225 225 225 225 total 275 275 310 445 445 445 455 455 455 wmd 6 amp 220 220 0 0 0 0 0 0 0 aop 100 100 140 265 265 265 280 280 280 bop 0 0 220 220 220 220 220 220 220 total 320 320 360 485 485 485 500 500 500 table 2. annual moose permit allocations for wildlife management districts 3 and 6 from 2001-2009 in maine, usa. *amp = any moose permit, aop = antlerless only permit, bop = bull only permit. controlled moose hunt in maine kantar alces vol. 47, 2011 86 aging cole crops, and reduce moose numbers within the surrounding cropland areas, the mdifw designed and implemented a controlled moose hunt in 2009. the controlled hunt targeted moose prior to the recreational hunt when crops were most vulnerable, and exerted additional pressure on localized moose populations that were causing damage without putting undue burden on landowners to remove additional moose. thus, implementation of a controlled hunt provided a 3-tiered approach to managing moose numbers: recreational hunting within a traditional framework to achieve publicly derived population goals, a controlled hunt to alleviate crop depredation and reduce moose-vehicle collisions, and depredation permits to provide immediate relief from crop damage. this approach serves to manage moose abundance at both the landowner and wmd scales, while providing flexibility and responsiveness to moose-human conflicts. this paper describes the implementation of the controlled hunt as a novel management tool to help reduce crop damage by a locally overabundant moose population in maine. study area aroostook county, maine is large (17,687 km2) with its eastern portion comprised mostly of farmland, >131,118 ha with about 76,000 ha as cropland; currently <1% of croplands contain cole crops (maine department of agriculture, food and rural resources). however, cole crops are distributed across numerous townships and active fields are rotated annually; the size of fields range from about 5->200 ha with the majority 16-40 ha; fields are typically on a 4-year rotation (d. hentosh, pers. comm.). wmds 3 and 6 overlap these lands and comprise about 5,970 km2 (fig. 2); forested areas are dominated by spruce, balsam fir, northern white cedar (thuja occidentalis), and white pine (pinus strobus) with mixed hardwoods of aspen (populus spp.), birch (betula spp.), beech (fagus grandifolia), and maple (acer spp.). other species highly palatable to moose include red-osier dogwood (cornus stolonifera) and willow (salix spp.). 0 5 10 15 20 25 30 35 40 45 50 2001 2002 2003 2004 2006 2007 y ear m o o s e s e e n /1 0 0 h o u rs wmd 3 wmd 6 fig. 1. moose seen per 100 hours of deer hunting in wildlife management districts 3 and 6, 2001-2007, maine, usa. due to administrative error, the 2005 deer hunter survey was invalid. alces vol. 47, 2010 kantar controlled moose hunt in maine 87 methods the controlled moose hunt was modeled after similar hunts for white-tailed deer (o. virgininaus) in maine where traditional hunting seasons aimed at providing recreational opportunity failed to reduce negative impacts from deer and high human density restricted hunter access. over the course of a year mdifw biologists met in both informal meetings with local landowners, and formal meetings with invited stakeholders including local farmers, sportsmen, landowners, the farm bureau, and warden service to discuss moose numbers, crop depredation problems, recreational hunting, landowner access, and hunter behavior. regional mdifw biologists and district wardens had addressed moose crop depredation complaints over time and had in-depth knowledge of the issues, layout of croplands, and the dynamics of recreational moose hunting in the region. discussion focused on exerting additional hunting pressure around the cropland area and designing a season structure outside of the normal recreational hunt. different approaches were presented informally to stakeholders to identify issues and potential problems. once local stakeholders accepted the preliminary framework, mdifw staff refined and formalized the proposed controlled hunt. since damage to broccoli crops can start as early as july and extend into october and november after the initial frost, the controlled hunt needed to include a longer time period than the traditional 6-day recreational season. therefore, the controlled hunt was scheduled from 17 august-19 september (5 weeks excluding sundays), leaving a 1-week interlude prior to the recreational season. since any moose regardless of sex or age can damage crops, it was determined that the controlled hunt would target a total of 100 moose based on historical recreational permit levels in the area (~10% of recreational permit levels, table 2). this harvest level had the objective of reducing impacts in a localized area (approximately half the wmd) without (presumably) affecting the benefit for those in the recreational hunt. of the 100 permits, 55 were assigned to qualifying landowners to hunt on their property if it was >80 acres (32 ha) in size and within the controlled hunt area (fig. 2); landowner permits were designated as amp to facilitate harvest and increase success rates. the remaining 45 permits were issued (3 each, 1 amp and 2 aop) to 15 registered maine guides fig. 2. location of controlled moose hunt and associated wildlife management districts 3 and 6 in aroostook county, maine, usa. controlled moose hunt in maine kantar alces vol. 47, 2011 88 selected by lottery; guides were used to ensure positive landowner relations, and to facilitate harvest and care of moose. prior to issuance of permits, each guide was required to attend a training session conducted by mdifw; failure to attend resulted in permit forfeiture and designation to runner-ups in the lottery. all hunters were required to register their moose and provide for collection of biological data including a canine for aging, sex, weight, antler measurements, hunter information, town of kill, date of kill, and caliber of firearm. all hunters were required to fill out a survey to document number and type (adult/calf and sex) of moose seen, hours hunted, date of hunt, wmd, and number of deer and grouse (bonasa umbellus) seen. guides and hunters had to personally contact farmers to identify cropland areas open to hunting and specific landowner rules. guides were instructed to harvest moose directly in the headlands or adjacent woodlands outside cropland to avoid direct damage to cole crops. at the conclusion of the controlled hunt, agricultural interests, mdifw biologists and wardens, and the moose registration station owner/operator gathered for a debriefing of the controlled hunt. comments requesting input on the development and implementation of the controlled hunt were solicited from other stakeholders. notes from the debriefing and letters from agricultural interests were reviewed by mdifw staff. results mdifw biologists conducted a mandatory evening seminar about the controlled moose hunt that outlined hunting rules and regulations, moose biology and behavior, and ethical hunting conduct. all permitted maine guides were in attendance as well as representatives from the 2 farms where the majority of hunting would occurr. a total of 81 moose were harvested: 37 adult bulls (yearling and older), 41 adult cows (yearling and older), and 3 female calves. landowners harvested 45 moose (82% success) and clients of guides harvested 36 (80% success). hunters included 67 maine residents and 24 non-residents from 15 states, the territory of guam, and quebec, canada. both written responses (formal letter and email) and phone calls from the primary agricultural interests provided consistently positive responses about the management of the hunt and its outcome. the maine warden service reported full compliance with the hunt rules with a single exception, a bull taken on an aop; adherence to landowner rules and respect for landowner property met expectations. discussion harvest rates in the recreational hunt in eastern aroostook county were considered moderate, but were meeting district-wide objectives to reduce moose abundance in wmd 6. however, the palatability and distribution of cole crops provides a highly valuable and concentrated food source that likely reduces foraging and handling time for moose. this effect created persistent, locally overabundant populations despite reduction at the districtwide scale. while hunters are required to hunt in a specified wmd, within a district they are only restricted by landowner permission and firearm ordinances where applicable. while the distribution of hunters is acutely linked to where moose occur during the hunting season, landownership patterns and hunter density also influence hunter distribution. within the context of the agricultural-woodland dynamic, the controlled hunt area is not characteristic of what hunters think of as moose habitat, access is limited, and within the traditional 6-day season hunters can quickly overwhelm croplands and dramatically elevate hunter density. thus the benefits of hunting the croplands (increased visibility and moose density) are confounded by a reduced quality of the hunt. the relative small geographical size of the croplands relative to the larger wmd also confounds the ability to control alces vol. 47, 2010 kantar controlled moose hunt in maine 89 moose populations and damage through the traditional hunting framework. several key elements were critical to the implementation of the controlled hunt. the hunt needed to be biologically sound and provide relief to cole crop damage during the growing season with high compliance with landowner requests, and have potential to reduce moose-vehicle collisions. based on known success rates of resident and nonresident hunters for amp and aop permits in the northeast zones, mdifw predicted a harvest success rate of 88%, with a slight skew towards adult bulls. the actual harvest resulted in 20% fewer adult bulls, 11% more adult cows, and 40% fewer calves than predicted; relative to population control, this harvest was beneficial biologically and increased stakeholder satisfaction. the intent of the controlled hunt was to strategically remove moose associated with localized croplands with emphasis on hunting as a management tool and not as another “opportunity.” although a focal point of the mdifw throughout the process was to identify and describe the hunt in terms of a targeted and focused effort, certain stakeholders recommended that permits be allocated to other hunting interests with specific needs (e.g., veterans and disabled hunters). despite indepth explanation of the rationale and purpose of the hunt and mdifw interest in keeping the framework straightforward, stakeholders continued to press for modification. while it remained critical to address stakeholder concerns, alleviating crop damage and addressing agricultural concerns were paramount to implementing a hunt that met the needs of all stakeholders. while the bulk of the permits went to 2 farms, local landowners and the general hunting public maintained a high level of satisfaction with this hunt believing it provided relief from crop damage. farmers were included during the entire process, and participated in the guide training session and hunt debriefing. thus, the controlled hunt represents another step along the stakeholder continuum (decker and chase 2001) by incorporating elements of co-management rather than a transactional approach as exemplified by maine’s strategic planning process. the co-managerial approach provides flexibility that parallels the current legal framework of mdifw and their responsibility to manage wildlife populations and respond to negative impacts. current authority provides flexibility to design and implement management activities that help resolve both biological and social problems due to moose. when provided with these tools, biologists can formulate management actions necessary to meet publicly derived goals and objectives, and better address social issues that can either increase or detract from agency accountability and credibility. farmers had experienced substantial crop damage resulting in both financial loss and loss of resource investment in the controlled hunt area. while a mechanism was in place to alleviate immediate problems (depredation permits), most farmers do not have the time or are unwilling to remove moose throughout the growing season, and continual removal by farmers is impractical. importantly, farmers preferred providing public opportunity in removing and utilizing the moose resource, although moose hunting in and around croplands during the recreational hunt incurred a cost (i.e., property abuse and damage). in its initial year the controlled hunt was considered successful because it provided relief from crop depredation and property abuse from hunters accessing croplands. thus the “burden” of managing nuisance moose was born by multiple stakeholders that benefited from each other. farmers realized lower depredation and property abuse, hunters provided a service to the mdifw and landowners, and mdifw was able to facilitate their moose management program and improve communication and credibility among stakeholders. controlled moose hunt in maine kantar alces vol. 47, 2011 90 acknowledgements i thank r. hoppe, m. stadler, s. ritchie, and lt. t. ward, mdifw, and d. hentosh. all were integral to the design and planning of the controlled hunt. i greatly appreciate p. pekins editorial and procedural advice. references andren, h., and p. angelstam. 1993. moose browsing on scots pine in relation to stand size and distance to forest edge. journal of applied ecology 30: 133-142. bergeron, d. h., p. j. pekins, h. f. jones., and w. b. leak. 2011. moose browsing and forest regeneration: a case study in new hampshire. alces 47: xxx-xxx. bontaites, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69-76 chase, l. c., t. m. schusler, and d. j. decker. 2000. innovations in stakeholder involvement: what’s the next step? wildlife society bulletin 28: 208-217. conover, m. r. 2001. effect of hunting and trapping on wildlife damage. wildlife society bulletin 29: 521-532. _____, and d. j. decker. 1991. wildlife damage to crops: perceptions of agricultural and wildlife professionals in 1957 and 1987. wildlife society bulletin 19: 46-52. decker, d. j., and l. c. chase. 2001. stakeholder involvement: seeking solutions in changing times. pages 133-152 in d. j. decker, t. l. brown., and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. gundersen, h., h. p. andreassen, and t. storaas. 2004. supplemental feeding of migratory moose alces alces: forest damage at two spatial scales. wildlife biology 10: 213-223. kantar, l. e. 2002. evaluating perceived resource conflicts in context with spatial dynamics of an interstate wintering elk herd. m. s. thesis. new mexico state university, las cruces, new mexico, usa. mower, k. j., t. w. townsend, and w. j. tyznik. 1997. white-tailed deer damage to experimental apple orchards in ohio. wildlife society bulletin 25: 337-343. morris, k. i. 2002. moose management system. maine department of inland fisheries and wildlife, augusta, maine, usa. smith r. l., and j. l. coggin. 1984. basis and role of management. pages 571-600 in l. l. halls, editor. white-tailed deer: ecology and management. stackpole books, harrisburg, pennsylvania, usa. 4303.pdf alces vol. 43, 2007 scarpitti et al. neonatal habitat characteristics 29 characteristics of neonatal moose habitat in northern new hampshire david l. scarpitti1, peter j. pekins, and anthony r. musante department of natural resources, university of new hampshire, durham, nh 03824, usa abstract: habitat use by parturient moose (alces alces) may have important implications for calf or specialized and little descriptive information exists in the northeastern united states, this study was 30 maternal moose. there was no difference (p > 0.10 for each parameter) in 22 of 23 physical and n = 30) and random sites (n = 30). however, neonatal sites were about 2x farther (p occurred. most neonatal sites (> 63%) were located in pole or saw timber stands comprised of mixed or coniferous habitat (> 75%); conifers were the dominant canopy species at 67% of neonatal sites. in location of neonatal habitat. mature, mixed, and coniferous habitats may provide microhabitat that helps conceal neonates from potential predators such as black bears (ursus americana), particularly alces vol. 43: 29-38 (2007) key words: alces alces, calves, habitat, moose, neonatal, predation, survival calf survival is an important factor affectalces alces) population dynamics (gasaway et al. 1977, franzmann et al. 1980), and predation by black bears (ursus americana), brown bears (ursus arctos), and wolves (canis lupus few weeks postpartum (schwartz and franzmann 1989, ballard et al. 1991, osborne et al. 1991, testa et al. 2000, bertram and vivion is limited and movements are restricted for 1-2 weeks postpartum (addison et al. 1990, testa et al. 2000). consequently, the cow-calf pair frequently remains within 20-50 m of the 1974). habitat use by parturient moose (i.e., neonatal habitat) may have important implications for survival of newborn calves and ultimately affect population dynamics. tich and gilbert 1986), islands and peninsulas (clarke 1936, peterson 1955, stephens and peterson 1984, addison et al. 1993, testa et al. 2000), and elevated and open sites (wilton and garner 1991, bowyer et al. 1999) provide may conceal calves and reduce predation risk hudson 1986, schwartz and renecker 1998), 1 ma 01581, usa neonatal habitat characteristics – scarpitti et al. alces vol. 43, 2007 30 relate indirectly to predation rates. new hampshire’s northern moose population may have approached stability, despite moderate harvest and presumably favorable habitat. characteristics of neonatal habitat have not close to water. this study was performed to describe habitat used by parturient moose and determine whether these habitats have specialthis study was performed in tandem with seasonal habitat and reproductive measurements as part of an extensive 4-year research project. information from this study will help land use activity on parturient moose and neonatal habitat. methods study area the study area encompassed approximately 1,000 km2 of primarily commercial forest land within eastern coos county, n milan stark berlin success dummer shelburne cambridge gorham kilkenny 0 5 10 kilometers alces vol. 43, 2007 scarpitti et al. neonatal habitat characteristics 31 watershed where numerous intermittent and the primary land use. small areas of cultivated land and pasture occurred primarily adjacent were common. predators in the study area included black bear, coyote (canis latrans), and bobcat (lynx rufus). the estimated moose density was 0.7 moose/km2; white-tailed deer (odocoileus virginianus) were sympatric with moose are hunted annually by a permit-lottery system; hunter success rates typically exceed 85% within the study area (nhfgd 2003). dominant forest types were northern hardwoods (36%) as a mix of yellow birch (betula alleghaniensis), american beech (fagus grandifolia acer saccharum spruce (picea rubens abies balsamea) on poorly-drained or nutrient-poor tive communities (16%) were clearcuts and populus tremuloides), paper birch (betula papyrifera), and pin cherry (prunus serotina). numerous (castor canadensis small developed areas of residential and inwas > 100 cm and occurs mostly as snowfall from november-march; seasonal temperatures field sampling direct observations of radio-collared cow complete leaf-out in early-mid may provided optimal conditions to observe maternal moose. an observed calf was estimated as 0 – 3 days 1974, larsen et al. 1989). no births were turbed at the presumed birth site of several and the limited mobility of calves, it was assumed that sampled habitat was representative of neonatal habitat associated with the birth a random sample of 10 maternal cows natal sites. of 50 random utm coordinates coordinate was chosen randomly and sampled in an identical manner as the neonatal site. because moose are not territorial, random points could have occurred within neonatal habitat of another maternal cow. measured at neonatal and random sites to evaluate the presence and preference of varimeasurements taken in each cardinal direction from plot center. the percent shrub-level density was estimated as the proportion of a 2 m shrub density was estimated at low (0 – 1 m) percent abundance of bare soil, rock, dead a 5 m radius of plot center was estimated neonatal habitat characteristics – scarpitti et al. alces vol. 43, 2007 32 (dbh) > 5 cm were counted, the percent shrub was estimated visually within a radius of 10 m from plot center. the habitat type was recorded as northern or other. stand structure was recorded as saw or recently disturbed. the dominant canopy species was recorded and the dbh of all trees within a 10 m radius of plot center was meafrom plot center. the elevation, slope, and aspect were the spatial analyst extension within arcview gis 3.3 (esri 2002). distance (m) to the nearest road (either class iii paved road or nearest island within a lake, pond, or river, was measured with the animal movement extension version 2.0 and arcview gis 3.3 data analysis reported means are absolute values. differences between continuous parameters at neonatal and random sites were evaluated with two-sample t-tests. fisher’s exact test was used to detect differences between the dominant canopy species, aspect, and presence software and fisher’s exact test was assessed with sas version 6. results were measured each summer, 2003 – 2005. all neonatal sites were associated with cows in 2005. neonatal and random sites were located predominantly (95%) in northern forest cover type did not differ at neonatal and random sites (p = 0.154), however, > 75% of neonatal sites were located within mixed and 50% of random sites. no neonatal sites and 10% of random sites were located within cut/ different at neonatal and random sites (p = natal sites were located within pole and saw timber stands as random sites; neonatal sites much as random sites. the dominant canopy species was not different between neonatal and random sites (p = 0.144); red spruce and sites (table 1). ent at neonatal and random sites (p = 0.596). aspect at neonatal and random sites was not different (p 50% of both site types were located on south and random sites was approximately 450 m and not different (p = 0.797); percent slope random sites (p = 0.355) (table 1). p = 0.311), neonatal sites (mean = 487.2 m) were 100 m farther from roads than random sites. the distances of neonatal and random sites to wetlands, perennial or intermittent streams, and open water were not different; distance to islands was similar and > 3,000 m from both alces vol. 43, 2007 scarpitti et al. neonatal habitat characteristics 33 neonatal and random sites (p > 0.05 for each parameter) (table 2). however, random sites than neonatal sites (p = 0.032). trees within 15 m of plot center at neonatal and random sites (approximately 15.5 cm) was not different (p = 0.783), nor was the p = 0.711). mean percent canopy cover at neonatal sites (78.6%) was was not different (p = 0.228). mean shrub from 40 to 60% at neonatal sites and was not different at random sites (p > 0.05 for each different between neonatal and random sites (p > 0.05) (table 3). discussion moose could have important implications for in northern new hampshire where commercial forestry continually alters habitat composition gilbert 1986, bowyer et al. 1999), as well as and gilbert 1986; addison et al. 1990, 1993; neonatal random p-value habitat type (number of sites) 0.15 northern hardwood forest 7 11 coniferous forest 11 9 mixed forest 12 7 0 3 dominant canopy species (number of sites) 0.14 northern hardwoods species 8 10 21 15 aspen, paper birch, or cherry 1 5 stand size class (number of sites) 0.14 4 5 uneven 7 14 pole 12 5 7 6 (number of sites) 0.6 present 10 13 absent 20 17 aspect (number of sites) 0.27 northerly (n, ne, nw) 6 3 southerly (s, se, sw) 17 14 7 13 slope (%) 2 1.3 0.36 elevation (m) 464.4 457.5 0.8 table 1. absolute counts and mean distances of physical parameters measured at 30 neonatal moose sites and 30 random sites in northern new hampshire, 2003-2005. neonatal random p-value distance to road (m) 487.2 384.9 0.31 distance to wetland (m) 395.1 401.1 0.95 distance to stream (m) 612.1 655.7 0.72 distance to open water (m) 1,059.9 1,252.1 0.24 distance to island (m) 3,156.5 3,384.7 0.55 distance to cut/ 136.5 69.8 0.03 table 2. mean distances of physical parameters measured at 30 neonatal moose sites and 30 random sites in northern new hampshire, 2003-2005. neonatal habitat characteristics – scarpitti et al. alces vol. 43, 2007 34 1999). site location could enhance either or relative importance of either is probably a function of local conditions. mid-late may (schwartz 1998, scarpitti et al. preferred species (i.e., aspen, cherry, maple) at most neonatal sites was probably lower than at random sites because > 75% of neonatal sites were located in pole and saw timber stands in tal sites were not closer than random sites to open water, rivers and streams, or wetlands patches than random sites. conversely, the majority of both neonatal and random sites were located on southerly exposures where relative to other aspects (table 1). at neonatal and random sites (table 3), and associated with the diverse forest types and was within 140 m of all sites sampled) in the study area. other forestry practices such as neonatal random p-value tree diameter (cm) 15.8 15.3 0.78 4.2 3.8 0.71 percent overstory canopy cover (%) 78.6 69.8 0.23 0–1m at 15m 48.8 51.5 0.72 1–2m at 15m 44.8 47.4 0.72 0–1m at 30m 61.4 60.3 0.91 1–2m at 30m 57.3 54.7 0.75 53.1 53.5 0.95 percent cover within 10m of plot center (%) 35 46.7 0.13 forbs/ferns 38.3 34 0.52 32.7 35.4 0.62 leaf litter 28.6 33.9 0.44 moss 28.2 20.7 0.28 dead wood 14.7 13 0.64 soil 4.5 4.9 0.76 rock 2.5 2.1 0.73 northern new hampshire, 2003-2005. alces vol. 43, 2007 scarpitti et al. neonatal habitat characteristics 35 resources likely become more important when peak lactation and widely available weeks postpartum when calves are rapidly to predation (robbins 1993, schwartz and renecker 1998). site characteristics that provide security ence on selection of neonatal habitat. neonates are susceptible to predation and experience was not documented, anecdotal accounts of local bear predation were reported, and approximately 25% of neonates did not survive 2 months post-partum (unpublished data). the majority of mortality occurred within 3 weeks of birth (scarpitti et al. 2005) and some predation by black bears was suspected. use of islands, peninsulas, and sites near open water by parturient cow moose is believed to improve their ability to detect and/or escape predators (peterson 1955; altmann 1958; bai1984; leptich and gilbert 1986; addison et al. 1990, 1993). however, neither lake nor water was sparsely distributed in the study area; both neonatal and random sites were > 3 km from islands and > 1 km from open other water features were common, mostly small perennial or intermittent streams and was similar for neonatal and random sites (table 1). such features probably do not improve a cow’s ability to detect or escape 1994). tops, upper slopes) to increase visibility and help detect potential predators, as reported in ontario (wilton and garner 1991), québec (chekchak et al. 1998), and alaska (bowyer et al. 1999). however, no difference in overall elevation was measured at neonatal and landscape position was not determined, the use improve visibility because of the well stocked, dense nature of forests within the study area, particularly in mixed and coniferous habitat used by most moose (table 1). habitats. most cows (> 75%) used mixed and coniferous neonatal habitat that may conceal calves from potential predators more effectively than deciduous habitat, particularly in measured in summer after leaf-out, when were delayed to minimize disturbance of parconiferous habitats by parturient moose may occur in response to the concurrent low use many moose populations located in more nutritional condition and food resources prior neonates in this population had relatively low predation rates (20 – 25% maximum). calf from 30 to 85%, of which black bears may account for 30 – 50% (ballard 1992, ballard the potential production of the study population neonatal habitat characteristics – scarpitti et al. alces vol. 43, 2007 36 predation and restrictive food resources that cally, recent population estimates in the study area indicate stability. this study indicates factor. sure in the study area appeared minimal on a relative scale and use of neonatal habitat was selective factor on neonatal habitat use when wolves, black bears, and moose existed in islands and water bodies, or other “secluded” mixed and coniferous habitats likely offers the best conditions to conceal calves and improve when neonates are most susceptible. forest successional habitat and mature mixed and coniferous forest stands should provide optimal habitat for parturient moose. acknowledgements by the new hampshire fish and game department. the professional efforts of hawkins and powers, inc., greybull, wyoin often adverse weather and environmental conditions. this study was possible because of the cooperation and access provided by ltd., international paper, inc., plum creek timber company, inc., meade corporation, and hancock timber resource group. local dr. christopher neefus and kent gustafson provided statistical consultation. kristine references addison, e. m., j. d. smith, r. f. mclaughlin, d. j. h. fraser, and m. e. buss. 1993. observations of preand post-partum behavior of moose in central ontario. alces 29:27-33. _____, _____, _____, _____, and d. g. joachim in central ontario. alces 26:142-153. altmann the moose calf. animal behavior 6:155159. _____. 1963. naturalistic studies of maternal in h. in mammals. j. wiley and sons, incorporated, new york, new york, usa. bailey, t. n., and e. e. bangs. 1980. moose of the north american moose conference and workshop 16:289-313. ballard, w. b. 1992. bear predation on moose: a review of recent north american tions. alces supplement 1:1-15. _____, and v. van ballenberghe. 1998. predin a. w. franzmann and c. c. schwartz, north american moose. smithsonian in_____, j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south114. bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior 66:747-756. bowyer, r. t., j. g. kie, and v. van balalces vol. 43, 2007 scarpitti et al. neonatal habitat characteristics 37 lenberghe. 1999. birth-site selection by bubenik 173-222 in a. w. franzmann and c. c. ton, d.c., usa. chekchak, t., r. courtois, j-p. ouellet, and l. b. s. st-onge. 1998. caracteristiques alces alces). 76:1663-1670. clarke, c. h. d. 1936. moose seeks shelter 50:67-68. degraaf, r. m., m. yamasaki, w. b. leak, and j. w. lanier technical report ne-144. forest service, northeast forest experiment station, randor, pennsylvania, usa. (esri) environmental systems research institute san francisco, california, usa. franzmann, a. w., c. c. schwartz, and r. o. peterson. 1980. moose calf mortality in summer on the kenai peninsula, 44:764-768. gasaway, w. c., d. haggstrom, and o. e. burris. 1977. preliminary observations in an interior alaskan moose population. conference and workshop 13:54-70. hooge, p. n., and b. eichenlaub. 1997. animal movement extension to arcview, langley, m. a., and d. h. pletscher. 1994. montana and southeastern british columbia. alces 30:127-135. larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rate of moose mortality in the southwest yukon. journal leptich, d. j., and j. r. gilbert. 1986. northern maine as determined by multition. alces 22:69-81. (nhfgd) new hampshire fish and game department. 2003. annual harvest summary. concord, new hampshire, usa. osborne, t. o., t. f. paragi, j. l. bodkin, a. j. loranger, and w. n. johnson. 1991. mortality in western interior alaska. alces 27:24-30. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. renecker, l. a., and r. j. hudson. 1986. robbins nutrition. second edition. academic scarpitti, d. l., c. h. habeck, a. r. musante, and p. j. pekins of moose in northern new hampshire. alces 41:25-35. schwartz, c. c. 1998. reproduction, natalin a. w. franzmann and c. c. schwartz, editors. american moose. smithsonian institution _____, and a. w. franzmann. 1989. bears, wolves, moose, and forest succession, kenai peninsula, alaska. alces 25:110. _____, and l. a. renecker. 1998. nutrition neonatal habitat characteristics – scarpitti et al. alces vol. 43, 2007 38 in a. w. franzmann and c. c. schwartz, editors. american moose. smithsonian institution silver, h. 1957. a history of new hampshire game and furbearers. new hampshire fish and game department. report no. 6. concord, new hampshire, usa. stephens, p. w., and r. o. peterson. 1984. stringham, s. f. 1974. mother-infant relations in moose. naturaliste canadien 101:325-369. testa, j. w., e. f. becker, and g. r. lee. 2000. movements of female moose in relation to birth and death of calves. alces 36:155-163. wilton, m., and d. garner. 1991. prelimiin south central ontario, canada. alces 27:111-117. 138 previous meeting sites of the north american moose conference and workshop 1963 st. paul, minnesota 1964 st. paul, minnesota 1966 winnipeg, manitoba 1967 edmonton, alberta 1968 kenai, alaska 1970 kamloops, british columbia 1971 saskatoon, saskatchewan 8th 1972 thunder bay, ontario 9th 1973 québec city, québec 10th 1974 duluth, minnesota 11th 1975 winnipeg, manitoba 12th 1976 st. john’s, newfoundland 13th 1977 jasper, alberta 14th 1978 halifax, nova scotia 15th 1979 soldotna kenai, alaska 16th 1980 prince albert, saskatchewan 17th 1981 thunder bay, ontario 18th 1982 whitehorse, yukon territory 19th 1983 prince george, british columbia 20th 1984 québec city, québec 21st 1985 jackson hole, wyoming 22nd 1986 fredericton, new brunswick 23rd 1987 duluth, minnesota 24th 1988 winnipeg, manitoba 25th 1989 st. john’s, newfoundland 26th 1990 regina and ft. qu’apelle, saskatchewan 27th 1991 anchorage and denali national park, alaska 28th 1992 algonquin park, ontario 29th 1993 bretton woods, new hampshire 30th 1994 idaho falls, idaho 31st 1995 fundy national park, new brunswick 32nd 1996 banff national park, alberta 33rd 1997 fairbanks, alaska in conjunction with the 4th international moose symposium 34th 1998 québec city, québec 35th 1999 grand portage, minnesota 36th 2000 whitehorse, yukon territory 37th 2001 carrabassett valley, maine 38th 2002 hafjell, norway in conjunction with the 5th international moose symposium 39th 2003 jackson hole, wyoming 40th 2004 corner brook, newfoundland and labrador 41st 2005 whitefish, montana 42nd 2006 baddeck, nova scotia 43rd 2007 prince george, british columbia future meetings 44th 2008 6th international moose symposium, yakutsk, russia 45th 2009 pocatello, idaho alces36_111.pdf f:\alces\vol_38\pagemaker\3811. alces vol. 38, 2002 snaith and beazley – population viability 193 application of population viability theory to moose in mainland nova scotia tamaini v. snaith and karen f. beazley school for resource and environmental studies, dalhousie university, 1312 robie st., halifax, ns, canada b3h 3e2 abstract: populations of moose (alces alces americana) in mainland nova scotia, canada, have been reduced to approximately 1,000 individuals fragmented into a number of isolated populations. although the data required for a comprehensive population viability assessment (pva) are not currently available, there are some general rules concerning minimum viable population (mvp) size that may be applied for a preliminary assessment. genetic evidence suggests that, in general, a genetically effective population (ne) of 50 individuals is required for short-term persistence and 500 to 5,000 individuals are required for long-term survival. census population size (n) is generally larger than ne, and a 10:1 relationship between n and ne has been roughly established in moose populations elsewhere. given this relationship, n = 5,000 individuals may be required for long-term viability. based on current home range size (30-55 km2) and population density (0.05/km2), the minimum critical area required by a population of this size is estimated to be approximately 100,000200,000 km2. strategies for moose conservation and forest management should concentrate on (1) conducting genetic, population, and habitat analyses to increase understanding of population viability and limiting factors; (2) reestablishing connectedness among discrete populations to form a viable metapopulation; (3) protecting/enhancing habitat to meet the critical requirements of a viable population; and (4) increasing carrying capacity of available habitat to support a greater population density. alces vol. 38: 193-204 (2002) key words: conservation, minimum critical area, minimum viable population, moose, nova scotia, population viability prior to european colonization, moose were widely distributed and abundant throughout mainland nova scotia (pulsifer and nette 1995). however, only a few small and isolated populations currently remain (fig. 1) and little is known about their status. there are approximately 500 individuals in the cobequid highlands, 300 in the southwestern portion of the province, and scattered pockets elsewhere (a.l. nette, nova scotia department of natural resources, personal communication). because the total population is < 1,000 individuals, moose are considered to be at risk of extirpation in mainland nova scotia (cescc 2001). because small and isolated populations are more likely to become extinct than large populations (diamond 1976, terborgh and winter 1980, shaffer 1981, henriksen 1997), it is important to address the viability of these moose populations. population viability a viable population is one that will continue to exist and to function naturally so that, over the long term, reproductive rates remain higher than or equal to rates of loss (salwasser et al. 1984, newmark 1985). the minimum viable population (mvp) is the population size below which the probability of extinction is unacceptably high, population viability – snaith and beazley alces vol. 38, 2002 194 but at or above which the probability of extinction is reduced to an acceptable level over a given period of time (shaffer 1981, samson 1983, lehmkhul 1984, gilpin and soulé 1986, lacy 1993/94, henriksen 1997). population viability requires maintenance of enough individuals to form an effective breeding population. extinction, demographic, environmental, and spatial factors are among the factors that influence population viability. effective population size the effective population (ne) is that portion of the actual or census population (n) that represents a genetically ideal population (brussard 1985, samson et al. 1985). in a genetically ideal population, all individuals are breeding adults, individuals mate at random, generations do not overlap, sex ratio is equal, reproductive success does not vary among individuals, there is no migration, mutation, or selection, and all individuals contribute equally to the genetic variation of the next generation. formally defined, ne is the size of a genetically ideal population that has the same rate of inbreeding or loss of genetic diversity through genetic drift as the real population being considered (franklin 1980, brussard 1985, reed et al. 1986). ne is almost always smaller than the actual population size (n) due to demographic and genetic factors that represent a departure from the genetically ideal population, such as the presence of non-breeding individuals (brussard 1985, newmark 1985, samson et al.1985, henriksen 1997). empirical determination of ne is difficult and data-intensive because sex ratio, age structure, reproductive behaviour, variability in reproductive success, dispersal patterns, and population fluctuations must be known (soulé 1980, brussard 1985, nunney and elam 1994). very few studies have attempted to quantify the relationship of n to ne for moose. using a computer simulation model to predict ne under a variety of harvest management options, ryman et al. (1981) estimated that, for moose, ne was approximately 5% to 20% of actual population size. arsenault (2000) theoretically determined that, based on local average population strucfig. 1. current distribution of moose in nova scotia [figure adapted from snaith and beazley (2002)]. alces vol. 38, 2002 snaith and beazley – population viability 195 ture, ne was 8.5% of n. taken together, these studies suggest that a 10:1 relationship between n and ne may be conservatively applied as a preliminary general rule for moose populations. extinction factors a population’s ability to survive depends on three characteristics: resilience, fitness, and adaptability (soulé 1980, salwasser et al. 1984, brussard 1985, reed et al. 1986). resilience is the short-term ability of a population to persist, despite normal reproductive fluctuations. fitness is the ability to cope with prevailing environmental conditions, and depends on the retention of sufficient genetic variability to avoid inbreeding depression and genetic drift over the shortto mid-term (decades). adaptability is necessary for the long-term persistence of a population and involves the ability to evolve. the capacity to adjust to environmental change depends on the maintenance of enough genetic variability to accommodate the evolutionary process of natural selection and to respond to a variety of demographic, environmental, genetic, and spatial extinction factors (terborgh and winter 1980, shaffer 1981, brussard 1985, newmark 1985, samson et al. 1985, gilpin and soulé 1986). demographic factors.—demographic stochasticity refers to random fluctuations in population parameters, such as birth-rate or mortality, which influence the probability of extinction over time (shaffer 1981, samson 1983, brussard 1985, theberge 1993). stochastic variations of population processes are more likely to lead to extinction in small populations because the effects of random fluctuations are amplified (shaffer 1981, samson 1983, brussard 1985, boyce 1992, theberge 1993, henriksen 1997). small populations are also prone to the allee effect (allee 1931), whereby very low populations experience decreasing reproductive rates (henriksen 1997, reed et al. 1998). although little information is available regarding the population structure among mainland nova scotia moose, demographic factors may be important considerations due to the small and fragmented nature of the populations that currently persist at very low densities [approx. 0.05/km2 (pulsifer and nette 1995)]. environmental factors.— environmental factors which affect population demographics include characteristics of the physical environment, populations of other species, and human activity. deterministic, or long-term systemic factors, such as habitat destruction, climate change, and environmental variation through time and space create variation in habitat carrying capacity and thereby influence population size, persistence, and probability of extinction (shaffer 1981, samson 1983, salwasser et al. 1984, samson et al. 1985, lacy 1993/94, theberge 1993, henriksen 1997). environmental stochasticity refers to random environmental events that affect all individuals in a population (shaffer 1981, samson 1983, brussard 1985, samson et al. 1985, gilpin and soulé 1986, mangel and tier 1993, henriksen 1997). for example, randomly fluctuating food availability, climatic conditions, competition, disease, predation, or hunting can lead to population-wide changes in mortality or reproductive success. in situations where environmental stochasticity is frequent or severe, only large populations will have reasonable probabilities of survival. the nova scotia moose herd has been reduced due to a number of environmental factors including habitat reduction and fragm e n t a t i o n ; h u n t i n g a n d p o a c h i n g ; interspecific competition with white-tailed deer (odocoileus virginianus); black bear (ursus americanus) predation; and disease caused by environmental contamination, brainworm (parelaphostrongylus population viability – snaith and beazley alces vol. 38, 2002 196 tenuis), and the winter tick (dermacentor albipictus) (dodds 1963, pulsifer and nette 1995, snaith and beazley 2004). because the remnant populations are small, isolated, and restricted to small fragments of suitable habitat, they are increasingly at risk of extirpation due to environmental fluctuations. moose are near the southern limit of their range in mainland nova scotia, and are potentially subject to further stress resulting from climate change (peters and darling 1985, snaith and beazley 2004). genetics.— genetic variation is the key to population fitness, adaptability, and survival. in small populations, genetic drift and inbreeding reduce genetic variability and increase the probability of extinction (franklin 1980, soulé 1980, shaffer 1981, lehmkhul 1984, salwasser et al. 1984, newmark 1985, samson et al. 1985, gregorius 1991, boyce 1992). inbreeding depression is caused by the expression of deleterious genes and is associated with reduced fitness and reproductive success. genetic drift refers to the random loss of heterozygosity (genetic variation) and can contribute to inbreeding depression, especially in chronically small populations. founder populations constrained to small numbers for short periods of time may not suffer the negative consequences of genetic drift and inbreeding depression provided that the population can subsequently expand in a relatively short period of time (franklin 1980, soulé 1980, lehmkhul 1984). a population bottleneck will only have negative consequences if heterozygosity is lost, deleterious genes become fixed, and the population loses its ability to expand (franklin 1980, soulé 1980, lehmkhul 1984). some species, such as the northern elephant seal (mirounga angustirostris) (lehmkhul 1984), seem well adapted to low levels of genetic variation but may be susceptible to environmental fluctuations due to low adaptive potential (soulé 1980, lehmkhul 1984). the importance of genetic variation within natural populations is supported by genetic evidence indicating a positive relationship between heterozygosity and fitness (soulé 1980). to maintain long-term viability, a population should be large enough to retain genetic variability and adaptability. the population density and distribution of moose populations in mainland nova scotia have been significantly reduced from historic levels (pulsifer and nette 1995). nonetheless, moose populations are adapted to maintaining low densities in sub-optimal habitat, and their reproductive potential may allow rapid population expansion when good habitat becomes available (geist 1974, timmermann and mcnicol 1988). in a number of cases, where suitable habitat was readily available, moose populations have grown from very small founder populations into large, widely distributed populations (kelsall 1987, pulsifer 1995, basquille and thompson 1997, wangersky 2000). genetic evidence from newfoundland and cape breton indicated that heterozygosity was reduced by 14% to 30% due to founder events (broders et al.1999). although there have been no known negative p h e n o t y p i c c o n s e q u e n c e s , a n d t h e populations evidently maintain enough genetic variability to persist for the short term, long-term viability may be compromised by limited adaptive potential due to this observed reduction in genetic variability (broders et al. 1999). similarly, genetic evidence indicated low heterozygosity among a swedish moose population that suffered a bottleneck event but was subsequently able to expand rapidly (ryman et al. 1977). evidence suggests that mainland nova scotia moose populations, although significantly reduced, possibly have the potential to expand if enough suitable habitat is restored, and other factors, such as disease or competition, are not limiting the populations. alces vol. 38, 2002 snaith and beazley – population viability 197 however, given the lengthy period of population decline and constraint (at current levels for 20 to 70 years), it is possible that genetic drift and inbreeding have led to a decrease in heterozygosity and adaptive potential. prolonging the small and isolated condition of moose populations in nova scotia is likely to further decrease their viability. spatial considerations.—spatial factors, including habitat reduction and fragmentation, influence population structure and size, and may increase vulnerability to extinction by isolating and reducing populations. if total isolation does not occur, habitat fragmentation may force a continuous population to take on the structure of a metapopulation, where several distinct local populations are loosely associated by periodic exchange of individuals (levins 1970, wilson 1975, caughley 1977, fahrig and merriam 1994, fahrig and grez 1996, beissinger and westphal 1998). in this situation, the deleterious effects of inbreeding and genetic drift can be compensated for by the addition of genetic variation from immigrants (one reproductively successful migrant per generation is required to maintain sufficient heterozygosity) while local divergence in response to environmental conditions may still occur (soulé 1980, brussard 1985, reed et al. 1986, beier 1993). when local populations become completely isolated, migration and gene flow become impossible, the metapopulation structure is lost, and overall ne is reduced to that of the local populations (brussard 1985, gilpin 1991). mainland nova scotia currently supports scattered moose populations separated by distances of 200 to 300 km, areas of unsuitable habitat, and barriers such as a major highway system (snaith 2001). as a result, it is unlikely that exchange of individuals occurs at an adequate rate for the herd to be an effective metapopulation (a.l. nette, nova scotia department of natural resources, personal communication). therefore, for the purposes of viability considerations, each mainland population should be treated as a separate and isolated local population until connectivity, and thus genetic exchange, is reestablished. population viability analysis population viability analysis (pva) is a comprehensive approach used to determine mvp or to evaluate extinction probabilities (lehmkhul 1984, salwasser et al. 1984, shaffer 1990, boyce 1992, lindenmayer et al. 1993, theberge 1993, lacy 1993/94, reed et al. 1998). pva and mvp estimates can be used to identify threatened populations and to quantitatively identify target population size for conservation efforts. ideally, pva is a speciesand areaspecific assessment that accounts for the demographic and genetic characteristics of the population in question, the quality and quantity of available habitat, and local environmental factors. empirical evidence, model results, and genetic analyses seem collectively to indicate that for many species an effective population of less than 50 individuals will not persist beyond the short term, that 500 to 5,000 breeding individuals are required to ensure long-term adaptability and persistence, and that habitat considerations are of primary importance in determining the fate of populations (franklin 1980; soulé 1980; shaffer 1981, 1983; samson 1983; brussard 1985; samson et al. 1985; lande 1987; berger 1990; thomas 1990; henriksen 1997; belovsky et al. 1999). estimating mvp for moose in mainland nova scotia the detailed demographic and genetic data required for a reliable pva are currently not available for moose in nova scotia. however, given the current risk of extirpapopulation viability – snaith and beazley alces vol. 38, 2002 198 tion, it is important to make some preliminary estimates. thus, the general findings that an effective population of at least 50 individuals is required for short-term persistence, and 500 for the long-term, was used as a preliminary estimate of mvp (franklin 1980, soulé 1980, shaffer 1981, brussard 1985, lande 1987, berger 1990, thomas 1990, henriksen 1997, beazley 1998). assuming a 10:1 relationship between n and ne (ryman et al. 1981, arsenault 2000), as previously described, ne = 500 may require n = 5,000 individuals to ensure long-term persistence, and for short-term viability, ne = 50 may require n = 500 individuals. currently, the total population of about 1,000 individuals, fragmented among isolated local populations, is likely too small to maintain long-term viability. whether the current population maintains the ability to expand to the long-term mvp size is unclear. nevertheless, 5,000 should be the minimum target population size for longterm conservation efforts. there appear to be enough individuals in nova scotia to maintain viability over the short term. the current population in the cobequid hills (n = 500) should be large enough for short-term persistence. however, because the population has already been restricted to this size for 20 to 70 years (a. l. nette, nova scotia department of natural resources, personal communication), it is unclear how much longer the population level can be maintained, and it is likely that a significant amount of heterozygosity has been lost. for these reasons, and because other local populations do not reach ne = 50 (n = 500) on their own, the reestablishment of connectivity among nova scotia moose populations is of primary importance over both the short and long term. minimum critical area minimum critical area (mca) represents the minimum amount of suitable habitat required to support the population and is calculated based on the number of individuals and their area requirements or population density, and must also take into account the spatial distribution of suitable habitat (soulé 1980, shaffer 1981, newmark 1985, metzgar and bader 1992, theberge 1993, doncaster et al. 1996, arsenault 2000). mca for moose in nova scotia might be calculated by multiplying population size and the area requirements (home range size) of each individual (shaffer 1981, newmark 1985, theberge 1993, doncaster et al. 1996, beazley 1998). however, this method does not account for variation in home range size, overlap among individual ranges, or non-adjacent home ranges. alternatively, mca can be calculated based on dispersion by dividing population size by population density (metzgar and bader 1992, theberge 1993, arsenault 2000). this method accounts for overlap, but does not satisfactorily take into account density variation through space and absolute area requirements. because home range sizes, population density, and the degree of overlap among individual home ranges are poorly understood in nova scotia, it is not possible to calculate mca reliably. however, a number of exploratory calculations can be performed using a variety of values for moose-area relationships based on currently available data (table 1). when average population density estimates for mainland nova scotia are used for the calculation, a long-term mvp of 5,000 individuals requires 100,000 km2 of suitable habitat. when the calculation is based on home range size, the same population requires 212,500 km2. while the home range calculation is probably an overestimate because overlapping individual ranges are not accounted for, the estimate based on density is likely inaccurate because it assumes moose are continuously alces vol. 38, 2002 snaith and beazley – population viability 199 and evenly distributed. actual mca may be somewhere between these estimates. by the same calculations, the short-term mvp (n = 500) requires 10,000 to 21,250 km2 of habitat under current conditions. the preceding calculations of mca rely on current home range and density estimates, which are dependent on local habitat quality, carrying capacity and demographic factors. implicit is the assumption that a larger population will require more area of the same quality than a smaller population. however, the current population density is very low due to a variety of factors including disease, overharvesting, predation, and poor habitat suitability (dodds 1963, pulsifer and nette 1995, snaith and beazley 2004, snaith et al. 2002). if carrying capacity/population density could be increased, then mca requirements would become smaller. management recommendations given current habitat conditions and population densities, these preliminary estimates indicate that a census population of 5,000 moose and 100,000 to 200,000 km2 of habitat are required for long-term viability. currently, there are no more than 1,000 moose, and the total land area of mainland nova scotia is approximately 45,000 km2, of which only a small portion is good quality moose habitat (snaith et al. 2002). based on these figures, mainland nova scotia does not currently support a moose population large enough to persist for the longterm, nor does it contain enough habitat to support such a population in isolation. however, there likely are enough moose to persist for the short term, providing appropriate protection and management actions are taken. to ensure the persistence of moose in nova scotia, short-term conservation efforts should concentrate on the maintenance of sufficient critical habitat (10,000 to 20,000 km2) of suitable quality to maintain current populations and to prevent further declines. for long-term viability, population size, and thus the extent and/or quality of habitat, must increase. habitat connectivity must be reestablished among local populations to allow migration and genetic exchange, which will boost the provincial effective population size to that of the table 1. exploratory calculations of minimum critical area (mca). mca mca population size (n) n x average hr1 n ÷ average density2 short-term mvp n = 500, ne = 50 21,250 km 2 10,000 km2 long-term mvp n = 5000, ne = 500 212,500 km 2 100,000 km2 1 hr = 45 km2 calculated as mean of hr = 55 km2 from empirical studies in southwest nova scotia (based on figures from d. brannen, nova scotia department of natural resources, personal communication) and hr = 30 km2 in habitat similar to northeast nova scotia (dunn 1976; crossley and gilbert 1983; crête 1987; leptich and gilbert 1989; mcnicol 1990). 2 density = 0.05/ km2 (mean of 0.01 to 0.09 from pulsifer and nette 1995). population viability – snaith and beazley alces vol. 38, 2002 200 metapopulation. restoration and enhancement of historical connections to populations in new brunswick may also be required for long-term viability. this strategy will be particularly important if population levels in nova scotia cannot reach the target for long-term mvp on their own. preliminary habitat suitability analyses indicate that there is little optimal moose habitat in the province, and that road density is an important factor in determining moose location (snaith et al. 2002). areas currently occupied by moose populations represent priority areas for protection, along with areas of high suitability and areas with few or no roads. empirical research is required to refine habitat suitability assessments (snaith et al. 2002), to establish the carrying capacity of existing habitat, to identify measures that can be used to increase habitat quality, and identify areas with potential for restoration. in addition to habitat protection and management, mortality factors and population processes unrelated to habitat must also be investigated. current moose population densities are at an historical low in nova scotia and are among the lowest documented worldwide (pulsifer and nette 1995, snaith and beazley 2004). a wide range of factors has been invoked as the cause for this drastic decline. research is required to conclusively identify the cause(s) of the decline, to isolate current limiting factors, and to design appropriate strategies for recovery. if factors limiting the population can be controlled, it may be possible to increase moose density within nova scotia, thereby reducing the total area required by a viable population. given the long period of population decline and restriction, local populations may be at risk of inbreeding depression, genetic drift, or extirpation. genetic research is required to determine ne and the level of heterozygosity that remains among local populations. until habitat connectivity among local populations can be achieved, direct population management such as the t r a n s l o c a t i o n o f i n d i v i d u a l s a m o n g populations (or from new brunswick, providing genetic evidence is available suggesting that the populations are of the same stock) may be required. artificial movem e n t o f a n i m a l s a t t h e r a t e o f o n e reproductively successful individual per generation should preserve sufficient genetic variability in local populations to maintain genetic fitness and expansion potential (soulé 1980, brussard 1985, reed et al. 1986, beier 1993). although this type of management is invasive and expensive, it may prove necessary if adaptability is low among nova scotia moose populations. in summary, strategies for moose conservation and landscape management should concentrate on genetic, population, and habitat analyses; the protection and enhancement of habitat to meet the critical requirements of viable moose populations; and the reestablishment of connectedness among discrete populations. given that the area required for the long-term persistence of a viable moose population may be greater than the total size of mainland nova scotia, and appreciating the variability of habitat suitability across the landscape, these figures suggest that the long-term viability of moose in nova scotia will require increased carrying capacity of available habitat, increased population density, and enhanced/ restored habitat connectivity to new brunswick. acknowledgements we thank a. l. nette, p. duinker, dalhousie university faculty of graduate studies, school for resource and environmental studies, the ejlb foundation, the nature conservancy of canada, the canadian wildlife federation, the editors of alces and three anonymous reviewers. alces vol. 38, 2002 snaith and beazley – population viability 201 references allee, w.c. 1931. animal aggregations. a study in general sociology. university of chicago press, chicago, illinois, usa. arsenault, a. a. 2000. status and management of moose (alces alces) in saskatchewan. fish and wildlife technical report. 00-1:1-84. basquille, s., and r. thompson. 1997. moose (alces alces) browse availability and utilization in cape breton highlands national park. parks canada technical report in ecosystem science 10:1-37. beazley, k. 1998. focus-species approach for trans-boundary biodiversity management in nova scotia. pages 755771 in n. w. munro and j. h. willison, editors. linking protected areas with w o r k i n g l a n d s c a p e s c o n s e r v i n g biodiversity. proceedings of the third international conference on science and management of protected areas, wolfville, nova scotia, canada, may 12-16, 1997. beier, p. 1993. determining minimum habitat areas and habitat corridors for cougars. conservation biology 7:94-108. beissinger, s. r., and m. i. westphal. 1998. on the use of demographic models of population viability in endangered species management. journal of wildlife management 62:821-841. belovsky, g. e., c. mellison, c. larson, and p. a. vanzandt. 1999. experimental studies of extinction dynamics. science 286:1175-1177. berger, j. 1990. persistence of differentsized populations: an empirical assessment of rapid extinctions in bighorn sheep. conservation biology 4:91-98. boyce, m. s. 1992. population viability analysis. annual review of ecology and systematics 23:481-506. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8:1309-1315. brussard, p. f. 1985. minimum viable populations: how many are too few? restoration management notes 3:2125. caughley, g. 1977. analysis of vertebrate populations. john wiley and sons, toronto, ontario, canada. (cesc) canadian endangered species council. 2001. wild species 2000: the general status of species in canada. ministry of publicworks and governm e n t s e r v i c e s , o t t a w a , o n t a r i o , canada. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1:553-563. crossley, a., and j. r. gilbert. 1983. home range and habitat use of female moose in northern maine a preliminary look. transactions of the northeast section of the wildlife society 40:67-75. diamond, j. m. 1976. island biogeography and conservation: strategy and limitations. science 193:1027-1029. dodds, d. g. 1963. the present status of moose (alces alces americana) in nova scotia. proceedings of the northeast wildlife conference 2:1-40. doncaster, c. p., t. micol, and s. p. jenson. 1996. determining minimum habitat requirements in theory and practice. oikos 75:335-339. dunn, f. 1976. behavioural study of moose. maine department of inland fish and wildlife project w-66-r-6, job 2-1. augusta, maine, usa. fahrig, l., and a. a. grez. 1996. populapopulation viability – snaith and beazley alces vol. 38, 2002 202 tion spatial structure, human-caused landscape changes and species survival. revista chilena de historia natural 69:5-13. , and g. merriam. 1994. conservation of fragmented populations. conservation biology 8:50-59. franklin, i. r. 1980. evolutionary change in small populations. pages 135-150 in m. e. soulé and m. e. wilcox, editors. conservation biology: an evolutionaryecological perspective. sinauer associates, sunderland, massachusetts, usa. geist, v. 1974. on the evolution of reproductive potential in moose. naturaliste canadien 101:527-537. gilpin, m. 1991. the genetic effective size of a metapopulation. biological journal of the linnean society 42:165-175. , and m. e. soulé. 1986. minimum viable populations: processes of species extinction. pages 19-34 in m. e. soulé, editor. conservation biology: the science of scarcity and diversity. sinauer associates, sunderland, massachusetts, usa. gregorius, h. r. 1991. gene conservation and the preservation of adaptability. pages 31-48 i n a . s e i t z a n d v . loeschcke, editors. species conservation: a population-biological approach. birkhauser verlag, boston, massachusetts, usa. henriksen, g. 1997. a scientific examination and critique of minimum viable population size. fauna norvegica 18:33-41. kelsall, j. p. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research supplement 1:1-10. lacy, r. c. 1993-1994. what is population (and habitat) viability analysis? primate conservation 14-15:27-33. lande, r. 1987. extinction thresholds in demographic models of territorial populations. the american naturalist 130:624-635. lehmkhul, j. f. 1984. determining size and dispersion of minimum viable populations for land management planning and species conservation. environmental management 8:167-176. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53:880-885. levins, r. 1970. extinction. pages 77-107 in m. gerstenhaber, editor. some mathematical questions in biology. american mathematical society, providence, rhode island, usa. lindenmayer, d. b., t. w. clark, r. c. lacy, and v. c. thomas. 1993. population viability analysis as a tool in wildlife conservation policy: with reference to australia. environmental management 17:745-758. mangel, m., and c. tier. 1993. a simple and direct method for finding persistence times of populations and application to conservation problems. proceedings of the national academy of science of the usa 90:1083-1086. mcnicol, j. 1990. moose and their environment. pages 11-18 in ontario ministry of natural resources, editor. the moose in ontario, book 1 moose biology, ecology and management. queens printer for ontario, toronto, ontario, canada. metzgar, l. h., and m. bader. 1992. large mammal predators in the northern rockies: grizzly bears and their habitat. northwest environment journal 8:231-233. newmark, w. d. 1985. legal and biotic boundaries of western north american national parks: a problem of congruence. biological conservation 33:197208. nunney, l., and d. r. elam. 1994. estialces vol. 38, 2002 snaith and beazley – population viability 203 mating the effective population size of conserved populations. conservation biology 8:175-184. peters, r. l., and j. d. s. darling. 1985. the greenhouse effect and nature reserves. bioscience 35:707-717. pulsifer, m. d. 1995. moose herd perseveres. nova scotia conservation 19:67. , and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31:209-219. reed, j. m., p. d. doerr, and j. r. walters. 1986. determining minimum population sizes for birds and mammals. wildlife society bulletin 14:255-261. , d. d. murphy, and p. f. brussard. 1998. efficacy of population viability analysis. wildlife society bulletin 26:244-251. ryman, n., r. baccus, c. reuterwall, and m. h. smith. 1981. effective population size, generation interval, and potential loss of genetic variability in game species under different hunting regimes. oikos 36:257-266. , g. beckman, g. bruun-petersen, and c. reuterwall. 1977. variability of red cell enzymes and genetic implicat i o n s o f m a n a g e m e n t p o l i c i e s i n scandinavian moose (alces alces). hereditas 85:157-165. salwasser, h., s. p. mealey, and k. johnson. 1984. wildlife population viability: a question of risk. transactions of the north american wildlife and natural resources conference 49:421-439. samson, f. b. 1983. minimum viable populations a review. natural areas journal 3:15-23. , f. perez-trejo, h. salwasser, l. f. ruggiero, and m. l. shaffer. 1985. on determining and managing minimum population size. wildlife society bulletin 13:425-433. shaffer, m. l. 1981. minimum population s i z e s f o r s p e c i e s c o n s e r v a t i o n . bioscience 31:131-134. . 1983. determining minimum viable population sizes for the grizzly bear. proceedings of the international conference of bear resource management 5:133-139. . 1990. population viability analysis. conservation biology 4:39-40. snaith, t. v. 2001. the status of moose in mainland nova scotia: population viability and habitat suitability. mes thesis, dalhousie university, halifax, nova scotia, canada. , and k. f. beazley. 2004. the distribution, status and habitat associations of moose in mainland nova scotia. proceedings of the nova scotia institute of science 42:76-134. , k. f. beazley, f. mackinnon, and p. duinker. 2002. preliminary habitat suitability analysis for moose in mainland nova scotia, canada. alces 38: 73-88. soulé, m. e. 1980. thresholds for survival: maintaining fitness and evolutionary potential. pages 151-169 in m. e. soulé and b. a. wilcox, editors. conservation biology: an evolutionary-ecological perspective. sinauer associates, sunderland, massachusetts, usa. terborgh, j., and b. winter. 1980. some causes of extinction. pages 119-134 in m. e. soulé and m. e. wilcox, editors. conservation biology: an evolutionaryecological perspective. sinauer associates, sunderland, massachusetts, usa. theberge, j. b. 1993. ecology, conservation and protected areas in canada. pages 137-153 in p. dearden and r. rollins, editors. parks and protected areas in canada. oxford university press, don mills, ontario, canada. thomas, c. d. 1990. what do real populapopulation viability – snaith and beazley alces vol. 38, 2002 204 tion dynamics tell us about minimum viable population sizes. conservation biology 4:324-327. timmermann, h. r., and j. g. mcnicol. 1988. moose habitat needs. the forestry chronicle 64:238-245. wangersky, r. 2000. too many moose? canadian geographic, nov/dec:44-56. wilson, e. o. 1975. sociobiology. belknap press, cambridge, massachusetts, usa. alces37(1)_13.pdf alces39_193.pdf alces vol. 39, 2003 bowyer and neville browsing history and forage quality 193 effects of browsing history by alaskan moose on regrowth and quality of feltleaf willow r. terry bowyer1 and juliette a. neville institute of arctic biology, and department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775-7000, usa abstract: we studied effects of browsing history by alaskan moose (alces alces gigas) on regrowth and quality of feltleaf willow (salix alaxensis) during late winter 2002 in interior alaska, usa. we recorded extensive browsing on willows, with 55.6% of leaders on 43 plants browsed by moose and 3.9% browsed by snowshoe hares (lepus americanus). foraging moose removed, on average, 15.1 mm of current annual growth from willow twigs, which averaged 24.1 mm in length (62.3% removed). twigs re-growing from 2-year-old stems that were browsed previously had larger diameters at their bud scale scar than those re-growing from stems that were not browsed in the previous year. browsing history by moose, however, had no effect on nitrogen content, in vitro dry matter digestibility, or tannin content of willow twigs. willows did not respond to browsing on individual twigs with an inducible defense system that involved tannins. diameter at point of browsing (bite size) was larger on twigs that had been browsed previously than for twigs re-growing from second-year growth that had not been browsed. moose did not exhibit an optimal bite size, but took larger-diameter bites from larger compared with smaller leaders of current annual growth. forage selection by moose for previously browsed twigs likely relates to greater forage biomass on those twigs rather than to forage quality. we caution, however, that foraging behavior by moose cannot be understood fully without considering additional factors, including predation risk in relation to forage availability. alces vol. 39: 193-202 (2003) key words: alaskan moose, alces alces gigas, browsing history, digestibility, feltleaf willow, foraging behavior, nitrogen, salix alaxensis, tannins moose (alces alces) are among the largest browsing mammals, with adult males of a. a. gigas (the largest subspecies) attaining > 770 kg and adult females reaching > 570 kg (schwartz et al. 1987). moose possess a narrower incisor arcade relative to body mass than do other ruminants, especially grazers—an allometric relationship that ostensibly is an adaptation for selective browsing (spaeth et al. 2001). indeed, browse is a critical component in the diet of moose throughout their distribution in north america, especially during winter (peek 1974, ludewig and bowyer 1985). moreover, in alaska, usa, willows (salix spp.) are the mainstay in the diet of moose (molvar et al. 1993, bowyer and bowyer 1997, maccracken et al. 1997, weixelman et al. 1998, bowyer et al. 2001, and many others), and those shrubs are consumed year-round in some areas (van ballenberghe et al. 1989; molvar et al. 1993; bowyer and bowyer 1997; bowyer et al. 1998, 1999a). a more-complete knowledge of interactions between moose and this crucial food supply is essential for understanding their distribution (telfer 1978), population dynamics (bowyer et al. 1999b), reproductive 1present address: department of biological sciences, idaho state university, pocatello, id 83209, usa browsing history and forage quality bowyer and neville alces vol. 39, 2003 194 performance (schwartz and hundertmark 1993), life-history characteristics (keech et al. 2000), and effects on ecosystem structure and function (pastor and naiman 1992, molvar et al. 1993, bowyer et al. 1997, berger et al. 2001). large herbivores tend to congregate in areas where they have foraged previously (fryxell 1991) and use of traditional areas is common among some ungulates (hjeljord 2001). for instance, moose sometimes use the same migratory routes (andersen 1991) as well as locations for mating (van ballenberghe and miquelle 1993) and giving birth (bowyer et al. 1999a). likewise, cervids (duncan et al. 1998, moore et al. 2000), including moose (molvar et al. 1993, bowyer and bowyer 1997, bergquist et al. 2001), preferentially forage on leaders of new growth that have re-grown from twigs that were browsed previously. such regrowth on previously browsed twigs is characterized by larger twigs and leaves than on unbrowsed leaders (bergstrom and danell 1987, molvar et al. 1993, bowyer and bowyer 1997). whether previous browsing of plants and their subsequent reuse by foraging herbivores influences use of traditional areas by these large mammals is unknown. we contend, however, that understanding why large herbivores re-browse particular plants, or parts thereof, is an essential step in understanding the overall process of diet selection and habitat use. the browsing history on trees and shrubs is well known to affect subsequent foraging by moose (molvar et al. 1993, bowyer and bowyer 1997, bergquist et al. 2001). consequently, we tested whether re-browsing by alaskan moose on leaders of new growth on feltleaf willow (s. alaxensis) was a result of increased size of twigs, quality of new growth, or both variables. we also tested whether moose would vary bite size in relation to the size of twigs available to browse. study area we conducted research concerning moose browsing on feltleaf willows in interior alaska, usa, about 15 km northwest of fairbanks (64° 54’ n, 147° 50’ w). the study site was near ballaine road and followed goldstream creek northeast of the road. the area is a low-elevation bog (185 m a.s.l.) underlain with intermittent permafrost, and includes a riparian zone dominated by willows and scattered alders (alnus crispa), which gradates into mixed stands of paper birch (betula papyrifera) and white spruce (picea glauca) on betterdrained soils. willows in this area were mechanically crushed with a bulldozer in march 1996 to improve habitat for moose; substantial re-sprouting of those shrubs has occurred since that treatment (bowyer et al. 2001). the crushed area, which encompasses 119 ha, extends 3 km northeast along goldstream creek, and is 100-800 m in width (bowyer et al. 2001). the climate of interior alaska is characterized by short, warm summers and long, cold, and often severe winters—temperatures range from -10 to -45° c in winter and snow depth averages 80 cm (gasaway et al. 1983), but has been about one-half that depth in recent years (keech et al. 2000, bowyer et al. 2001). moose density in nearby areas has increased recently (keech et al. 2000). although the area is near fairbanks, large mammalian carnivores, including wolves (canis lupus), are present (bowyer et al. 2001). the crushing of willows resulted in substantial stump sprouting, which created favorable foraging conditions for moose (bowyer et al. 2001). moose using the crushed area were mostly (~80%) adult males; females and young occurred infrequently on the open crushed area (bowyer et al. 2001). alces vol. 39, 2003 bowyer and neville browsing history and forage quality 195 methods we sampled feltleaf willows during late winter (march 2002), while those plants were dormant, on 43 quadrats, each 5 x 5 m, which were located randomly along 500 m of gold stream creek and extended <30 m on either side of the creek. samples were concentrated in this riparian zone to enhance the probability of locating feltleaf willows. only 1 willow, the individual plant (clump) closest to the cartesian coordinates used to select the random plot, was sampled from each quadrat. if no feltleaf willows were present on a particular quadrat, another random quadrat was chosen for sampling. all quadrats were >10 m apart. selecting only 1 plant per quadrat minimized the likelihood of obtaining multiple samples from clones of the same plant (molvar et al. 1993). we estimated length of leaders of current annual growth for feltleaf willow from their diameter at the bud scale scar using regression equations developed for this crushed area by bowyer et al. (2001). all leaders of current annual growth were counted on each willow, and the number of twigs foraged upon by moose and snowshoe hares (lepus americanus), which were distinguished easily from one another (bowyer and bowyer 1997), were recorded. we assumed all current annual growth was available as forage, but recognize that moose may not forage on some small twigs (bowyer and bowyer 1997). the crown of the willow was examined for leaders with particular patterns of browsing history by moose. those patterns of browsing included: (1) unbrowsed 2-year-old growth with unbrowsed current annual growth; (2) unbrowsed 2-year-old growth with browsed current annual growth; (3) browsed 2-yearold growth with unbrowsed current annual growth; and (4) both 2-year-old and current annual growth from that twig browsed by moose. diameter at the bud scale scar for 1-year-old twigs (i.e., current annual growth), and diameter at point of browsing for current annual growth (where appropriate) were recorded with dial calipers to the nearest 0.1 mm. more than one pattern of browsing on twigs was recorded for some individual willows, whereas other willows lacked various combinations of browsing history—not all patterns of browsing occurred on each plant. we sampled 90 twigs of current annual growth on 43 willows in evaluating effects of browsing history on diameter of twigs and re-browsing by moose. we clipped current annual growth of willows that had re-grown from 2 patterns of previous browsing by moose from 35 quadrats: second-year growth that was browsed and unbrowsed. those twigs were placed in labeled paper bags in the field, and later dried at 50°c in a forced-air oven for 4 days. samples were then ground in a wiley mill and passed through a 1-mm screen. in vitro dry matter digestibility (ivdmd) was determined with the method of tilley and terry (1963) modified to use ankom technology filter bags (fairport, ny). rumen inoculum for the digestion trial was obtained from a fistulated captive reindeer (rangifer tarandus) held at the r. g. white large animal research station at the university of alaska fairbanks, which had been accustomed to a diet of willows. we determined nitrogen content of willows using an elemental analyzer (leco # cns 2000). tannins were extracted in 50% ethanol and assayed with the folindennis method (martin and martin 1982, spaeth et al. 2002). a standard developed from salix pulchra was analyzed with samples for s. alaxensis. we tested for differences in use of leaders of current annual growth with a 2sample z test for proportions (remington and schork 1970). we used analysis of variance (anova) to test for differences in diameters of willow twigs in relation to browsing history by moose (neter et al. browsing history and forage quality bowyer and neville alces vol. 39, 2003 196 1985). in that analysis, diameter of current annual growth was the dependent variable, with location on the twig (at the bud scale scar or bite) where diameter was measured, previous browsing history (browsed or unbrowsed), and their interaction as main effects. we tested for and met assumptions of homogeneous variances prior to analysis, and performed a posteriori tests with tukey’s hsd. we arbitrarily reduced α to 0.02 for that analysis to compensate for a potential lack of statistical independence from sampling >1 combination of browsing histories from the same plant. we tested for effects of forage quality of willow twigs in relation to browsing history by moose using weighted multivariate analysis of variance (johnson and wichern 1982). nitrogen concentration (n), ivdmd, and tannin concentration were dependent variables, and whether current annual growth re-grew from browsed or unbrowsed second-year growth was the main effect. mass of current annual growth for a particular browsing history on a willow sometimes was insufficient to allow nutritional analyses. in those circumstances, we combined plants to obtain adequate sample mass. consequently, we weighted those combined samples by the number of plants included in the analysis. percentage data were square-root, arc-sine transformed prior to conducting analysis to meet assumptions of manova; we also tested for and met the assumption of multivariate homogeneity of variances. we set α = 0.05 for manova. we also examined partial correlations among dependent variables from the error sum-of-squares cross-product matrix in that analysis of forage quality conducted with manova. results mean (± sd) number of leaders of current annual growth occurring on 43 feltleaf willow was 115.3 ± 98.2 (range = 12-389 leaders). all 43 willows we selected exhibited some use by moose, but hares fed on 39.5% of those plants—typically on twigs near the base of the plant. of leaders available as forage on willows, a mean (± sd) of 55.6 ± 21.2% (range = 19.491.0%) were browsed by moose, whereas 3.9 ± 9.3% (range = 0-39.6%) of leaders were fed upon by hares; that difference was highly significant (z = 231.1, p < 0.0001). comparatively low levels of browsing by hares probably had limited effects on regrowth of twigs by willows. based on regression analysis predicting leader length from twig diameter, a mean (± sd) of 24.1 mm (± 1.2 mm) of current annual growth was available on each twig for foraging by moose. foraging moose removed, on average, 15.1 mm (± 1.0 mm) of current annual growth, or 62.3% of each twig. overall removal of current annual growth on 43 feltleaf willows by foraging moose (based on leader length) was estimated at 34.6% (55.6% of potential leaders browsed x 0.623 proportionally removed from each leader). previous foraging on feltleaf willow by moose had an effect on subsequent regrowth of twigs, and on the bite size (diameter of twigs) taken by moose (fig. 1; f = 29.83, p < 0.0001). tukey’s hsd indicated twig diameters at the bud scale scar were larger for leaders of current annual growth regrowing from second-year growth that had been browsed by moose than from secondyear growth that was not browsed (p < 0.001). likewise, moose took larger bites from twigs re-growing from previously browsed second-year growth than from that same age of twigs that had not been browsed (p < 0.001). measure of forage quality for feltleaf willow, including n content, ivdmd, and tannins, did not differ based on previous browsing history by moose (table 1). partial correlation coefficients, from that analyalces vol. 39, 2003 bowyer and neville browsing history and forage quality 197 sis of forage quality with manova (table 1), indicated a weak positive relationship between n and ivdmd (r = 0.30, p = 0.21), and a weak negative correlation between n and tannins (r = -0.17, p = 0.51). there was a significant positive relation between ivdmd and tannins (r = 0.52, p = 0.03) for current annual growth of feltleaf willow. discussion our findings concerning re-growth of twigs on feltleaf willow following browsing by moose confirm other results documenting that willows respond in the subsequent year with increased current annual growth (molvar et al. 1993, bowyer and bowyer 1997). increased re-growth of twigs from browsing by moose may relate to 3 factors: release of lateral twigs from apical dominance; greater plant resources to invest in fewer growing points; and fertilization of plants from inputs of urine and feces from browsing herbivores (molvar et al. 1993, bowyer and bowyer 1997 for reviews). the latter two factors, however, cannot explain variation in size of individual leaders from differences in browsing history, which should be similar for all current annual growth on the same plant. whatever the cause, moose foraging selectively would table 1. effects of previous browsing by moose on measures of forage quality, including percent dry mass of nitrogen (n), in vitro dry matter digestibility (ivdmd), and tannins, for current annual growth of feltleaf willow, interior alaska, usa, winter 2002. individual willows (n = 35) were combined from random quadrats into 19 samples to provide sufficient material for nutritional analyses; consequently, statistical analysis (multivariate analysis of variance; manova) is weighted by the number of plants in each sample. browsing history unbrowsed browsed (n = 10) (n = 9) 1measures of forage quality (%) x se x se n 1.18 0.053 1.17 0.056 ivdmd 37.9 1.14 37.2 0.78 tannins 11.4 0.79 12.7 0.69 1 weighted manova indicated no effect of browsing history on measures of forage quality (wilks’ lambda, f1, 17 = 0.79, p = 0.39). fig. 1. mean (+ se) diameter of current annual growth measured at the bud scale scar and at the bite on 90 twigs from 43 feltleaf willows in relation to browsing history by alaskan moose in late winter 2002, interior alaska, usa. browsing history and forage quality bowyer and neville alces vol. 39, 2003 198 benefit from an increase in available size of bites on willows. large mammalian herbivores, including moose, forage preferentially on leaders of new growth that have re-grown from twigs that were browsed previously (molvar et al. 1993, bowyer and bowyer 1997, moore et al. 2000, bergquist et al. 2001). moose obtained more forage per bite, and perhaps decreased handling time of forage, from browsing re-growth on willows that had been foraged upon formerly. moose varied the size of their bite with respect to the size of the available leader; a pattern also noted by shipley et al. (1999). consequently, moose do not take an optimum-sized bite, but may still be attempting to optimize diet quality. whether patterns of browsing history and subsequent re-growth of twigs help explain use of traditional areas, including travel paths, by moose is unknown. clearly, forage plays an important role in selection of areas by moose during critical periods of the year (bowyer et al. 1997, bowyer et al. 1999a), and has far-reaching implications for nutritional condition, reproduction, and s u r v i v a l o f m o o s e ( s c h w a r t z a n d hundertmark 1993, keech et al. 2000). variation in abundance of forage may help explain the distribution of sexes outside the mating season (bowyer et al. 2001). spatial separation of sexes in ruminants (sensu bowyer 1984, bowyer et al. 1996, kie and bowyer 1999, bowyer et al. 2002) likely has a dietary component related to forage abundance, with males using areas with more but not necessarily higher-quality food than localities inhabited principally by females (barboza and bowyer 2000, 2001). indeed, males predominated in the crushed areas where we studied browsing by moose, which had more available forage than an adjacent area where females and young were more common (bowyer et al. 2001). whether browsing history is related to resource partitioning by the sexes of moose (miquelle et al. 1992, bowyer et al. 2001), however, requires additional study. we rejected the hypothesis that browsing history of willows was related to the quality of subsequent re-growth of twigs (table 1). that conclusion is in keeping with results from molvar et al. (1993), who also reported little variation in quality of twigs re-sprouting from browsed or unbrowsed leaders of salix pulchra, or for re-growth of twigs on plants experiencing differing levels of browsing intensity. likewise, mechanical crushing of s. alaxensis had limited effects on quality of current annual growth in our study area (bowyer et al. 2001). we caution, however, that our data and those of others on browsing history and forage quality of willows come mostly from autumn and winter—we are uncertain whether browsing might enhance quality of current annual growth in spring and summer. browsing of twigs during winter results in re-growth of large leaves the following spring (molvar et al. 1993), and moose often leaf-strip that productive growth (miquelle 1983). in addition, we could only measure quality on leaders of current annual growth that moose had not browsed. if moose selectively foraged on higher-quality twigs from those re-growing from a previously browsed twig (thereby removing our opportunity to sample those twigs), our analysis would underestimate quality (bowyer et al. 1999a). we suspect such a bias is slight because of the overall low quality of willows during winter (bowyer et al. 2001; table 1). nonetheless, slight differences in forage quality can be magnified over time as herbivores accumulate resources via foraging (white 1983). understanding differences in quality of forage for moose and other large herbivores is complicated by variation in growing conditions for plants (chapin 1983, molvar et al. 1993, lenart et al. 2002), which may alces vol. 39, 2003 bowyer and neville browsing history and forage quality 199 result in fine-scale divergence in quality of individual willows (spaeth et al. 2002). such spatial variation in quality of trees and shrubs throughout the year in relation to browsing history by large herbivores is a topic in need of additional research. salix alaxensis is an important component in the winter diet of moose in interior alaska (van ballenberghe et al. 1989, miquelle et al. 1992, bowyer et al. 2001). yet, overall quality of those willows on our study site was low (table 1; bowyer et al. 2001). that outcome is not likely biased by our relatively small sample sizes, because of the near-identical values in measure of forage quality between current annual growth re-growing from browsed and unbrowsed twigs (table 1). moreover, levels of tannins, which are thought to be an important deterrent to browsing mammals (bryant and kuropat 1980, robbins et al. 1987), were comparatively high (table 1) in salix alaxensis on our study area, yet moose browsed those willows extensively. surprisingly, ivdmd was positively correlated with tannin concentrations in willows, an outcome antithetical to the hypothesis that those secondary compounds interfere with digestion of structural carbohydrates. how widespread that positive relationship is between tannins and ivdmd among willows or other species of browse is uncertain, but warrants further investigation. moreover, moose possess tannin-binding proteins in their saliva, which would further ameliorate affects of tannins on digestion (hagerman and robbins 1993). we caution that measures of forage quality have been difficult to link with diet selection in freeranging moose (weixelman et al. 1998), and that a more complete understanding of that process will require data on other aspects of the ecology and behavior of moose. population density relative to environmental carrying capacity (k) and risk of predation unquestionably affect foraging behavior by large herbivores, including m o o s e ( m o l v a r a n d b o w y e r 1 9 9 4 , weixelman et al. 1998, bowyer et al. 1999b). tradeoffs between avoiding predation and acquiring essential resources have been well-documented for large herbivores (molvar and bowyer 1994, rachlow and bowyer 1998, bowyer et al. 1999a, kie 1999, barten et al. 2001, and numerous references therein). we propose that a full understanding of foraging behavior, including hypotheses explaining bite size, browsing history, and diet selection by moose will require a synthesis of these two important fields. we hope our data on forage abundance and quality in relation to browsing history by moose will help lay the groundwork for such research. acknowledgements this research was funded, in part, by the institute of arctic biology, and department of biology and wildlife at the university of alaska fairbanks (uaf). we are indebted to c. adler, c. brockman, a. e. collins, l. j. huges, s. j. isham, y. kawaguchi, k. kellie, n. i. kohlenberg, r. a. long, t. m. massie, t. a. mclaughlin, h. moncrief, b. soiseth, and d. stewart from uaf for their assistance in the field and laboratory. we thank r. kedrowski for conducting nutritional analysis on our willow samples at the forage quality analysis laboratory at uaf. we acknowledge j. g. kie, d. f. spaeth, p. s. barboza, and k. m. stewart for their advice and assistance in the preparation of this manuscript. references andersen, r. 1991. habitat deterioration and the migratory behavior of moose (alces alces) in norway. journal of applied ecology 28:102-108. barboza, p. s., and r. t. bowyer. 2000. sexual segregation in dimorphic deer: a new gastrocentric hypothesis. journal browsing history and forage quality bowyer and neville alces vol. 39, 2003 200 of mammalogy 81:473-489. , and . 2001. seasonality of sexual segregation in dimorphic deer: extending the gastrocentric model. alces 37:275-292. barten, n. l., r. t. bowyer, and k. j. jenkins. 2001. habitat use by female caribou: tradeoffs associated with parturition. journal of wildlife management 65:77-92. berger, j., p. b. stacey, l. belis, and m. p. johnson. 2001. a mammalian predator-prey imbalance: grizzly bear and wolf extinction affect avian neotropical migrants. ecological applications 11:229-240. bergquist, g., r. bergstrom, and l. edenius. 2001. patterns of stem damage by moose (alces alces) in young pinus sylvestris stands in sweden. scandinavian journal of forest research 16:363-370. bergstrom, r., and k. danell. 1987. effects of simulated winter browsing by moose on morphology and biomass of two birch species. journal of ecology 75:533-544. bowyer, j. w., and r. t. bowyer. 1997. effects of previous browsing on the selection of willow stems by alaskan moose. alces 33:11-18. bowyer, r. t. 1984. sexual segregation in s o u t h e r n m u l e d e e r . j o u r n a l o f mammalogy 65:410-417. , j . g . k i e , a n d v . v a n ballenberghe. 1996. sexual segregation in black-tailed deer: effects of scale. journal of wildlife management 60:1017. , m. c. nicholson, e. m. molvar, and j. b. faro. 1999b. moose on kalgin island: are density-dependent processes related to harvest? alces 35:73-89. , b. m. pierce, l. k. duffy, and d. a. haggstrom. 2001. sexual segregation in alaskan moose: effects of habitat manipulation. alces 37:109-122. , k. m. stewart, s. a. wolfe, g. m. blundell, k. l. lehmkuhl, p. j. joy, t. j. mcdonough, and j. g. kie. 2002. assessing sexual segregation in deer. journal of wildlife management 66:536-544. , v. van ballenberghe, and j. g. kie. 1997. the role of moose in landscape processes: effects of biogeography, population dynamics, and predation. pages 265-287 in j. a. bissonette, editor. wildlife and landscape ecology: effects of pattern and scale. springer-verlag, new york, new york, usa. , , and . 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79:1332-1344. , , , and j. a. k. maier. 1999a. birth-site selection in alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80:1070-1083. bryant, j. p., and p. j. kuropat. 1980. selection of winter forage by subarctic browsing vertebrates: the role of plant chemistry. annual review of ecology and systematics 11:261-285. chapin, f. s., iii. 1983. direct and indirect effects of temperature on arctic plants. polar biology 2:47-52. duncan, a. j., s. e. hartley, and g. r. iason. 1998. the effect of previous browsing damage on the morphology and chemical composition of sitka spruce (picea sitchensis) saplings and on their subsequent susceptibility to browsing by red deer (cervus elaphus). forest ecology and management 103:57-67. fryxell, j. m. 1991. forage quality and aggregation by large herbivores. amerialces vol. 39, 2003 bowyer and neville browsing history and forage quality 201 can naturalist 138:478-498. gasaway, w. c., r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. hagerman, a. e., and c. t. robbins. 1993. specificity of tannin-binding salivary proteins relative to diet selection by mammals. canadian journal of zoology 71:628-633. hjeljord, o. 2001. dispersal and migration in northern forest deer—are there unifying concepts? alces 37:353-370. johnson, r. a., and d. w. wichern. 1982. applied multivariate statistical analysis. prentice hall, englewood cliffs, new jersey, usa. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64:450-462. kie, j. g. 1999. optimal foraging in a risky environment: life-history strategies for ungulates. journal of mammalogy 80:1114-1129. , and r. t. bowyer. 1999. sexual segregation in white-tailed deer: density dependent changes in use of space, habitat selection, and dietary niche. journal of mammalogy 80:1004-1020. lenart, e. a., r. t. bowyer, j. ver hoef, and r.w. ruess. 2002. climate change and caribou: effects of summer weather on forage. canadian journal of zoology 80:664-678. ludewig, h. a., and r. t. bowyer. 1985. overlap in winter diets of sympatric moose and white-tailed deer in maine. journal of mammalogy 66:390-392. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1997. habitat relationships of moose on the copper river delta in coastal and south-central alaska. wildlife monographs 136. martin, j. s., and m. m. martin. 1982. tannin assays in ecological studies: lack of correlation between phenolics, proanthocyanidins and protein-precipitating constituents in mature foliage of 6 oak species. oecologia 54:205-211. miquelle, d. g. 1983. browse regrowth and consumption following summer defoliation by moose. journal of wildlife management 47:17-24. , j . m . p e e k , a n d v . v a n ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monographs 122. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75:621630. , , a n d v . v a n ballenberghe. 1993. moose herbivory, browse quality, and nutrient cycling in a n a l a s k a n t r e e l i n e c o m m u n i t y . oecologia 94:472-479. moore, n. p., j. d. hart, p. f. kelly, and s. d. langton. 2000. browsing by fallow deer (dama dama) in young broadleaved plantations: seasonality, and the effects of previous browsing and bud eruption. forestry 73:437-445. neter, j., w. wasserman, and m. h. kutner. 1985. applied linear statistical models: regression, analysis of variance, and experimental designs. second edition. irwin, homewood, illinois, usa. pastor, j., and r. j. naiman. 1992. selective foraging and ecosystem processes in the boreal forests. american naturalist 134:690-705. peek, j. m. 1974. a review of moose food h a b i t s t u d i e s i n n o r t h a m e r i c a . naturaliste canadien 101:131-141. rachlow, j. l., and r. t. bowyer. 1998. habitat selection by dall’s sheep (ovis browsing history and forage quality bowyer and neville alces vol. 39, 2003 202 dalli): maternal trade-offs. journal of zoology (london) 345:457-465. remington, r. d., and m. a. schork. 1970. statistics with applications to the biological and health sciences. prenticehall, englewood cliffs, new jersey, usa. robbins, c. t., t. a. hanley, a. e. hagerman, o. hjeljord, d. l. baker, c. c. schwartz, and w. w. mautz. 1987. role of tannins in defending plants against ruminants: reduction in protein availability. ecology 68:98-107. schwartz, c. c., and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. journal of wildlife management 57:454-468. , w . l . r e g e l i n , a n d a . w . franzmann. 1987. seasonal weight dynamics of moose. swedish wildlife research supplement 1:301-310. shipley, l. a., a. w. illius, k. danell, n. t. hobbs, and d. e. spalinger. 1999. predicting bite size selection of mammalian herbivores: a test of a general model of diet optimization. oikos 84:5568. sp a e t h, d. f., r. t. bo w y e r, t. r. stephenson, p. s. barboza, and v. van ballenberghe. 2002. nutritional quality of willows for moose: effects of twig age and diameter. alces 38:143-154. , k. j. hundertmark, r. t. bowyer, p. s. barboza, t. r. stephenson, and r. o. peterson. 2001. incisor arcades of alaskan moose: is dimorphism related to sexual segregation? alces 37:217-226. telfer, e. s. 1978. cervid distribution, browse and snow cover in alberta. journal of wildlife management 42:352361. tilley, j. m. a., and r. a. terry. 1963. a two-stage technique for the in vitro digestion of forage crops. journal of the british grassland society 18:104111. van ballenberghe, v., and d. g. miquelle. 1993. mating in moose: timing, behavior and male access patterns. canadian journal of zoology 71:1687-1690. , , and j. g. maccracken. 1989. heavy utilization of woody plants by moose during summer in denali national park, alaska. alces 25:31-35. weixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces 34:213-238. white, r. g. 1983. foraging patterns and their multiplier effects on productivity of northern ungulates. oikos 40:377384. 1 an assessment of moose and elk train collisions in ontario, canada joe hamr1, mike hall2, and jesse n. popp3 1department of biology, laurentian university, sudbury, on p3c 2c6; 2sudbury district office, ministry of natural resources and forestry, 5-3767 hwy 69 s, greater sudbury, on p3g 1e7; 3department of geography and environment, mount allison university, 144 main st, sackville nb, e4l 1a7 abstract: to better understand train collision mortality of moose (alces alces) and elk (cervus elaphus) in ontario, we measured collisions along a 20 km segment of railway using post-winter railbed surveys (11 consecutive years), remote cameras, and radio-telemetered elk. we used these data to estimate provincial moose-train collision rates by extrapolating collision rates, moose density, and amount of high use railway per wildlife management unit (wmu). the annual collision rate varied from 0 to 7 moose and 2 to 22 elk on the 20 km section of railway; the combined collision rate of moose and elk was highest in winters with above average snowfall. the extrapolated collision rates of moose indicated that ~1/3 of wmus had a rate > 0.08 moose/km high use railway/yr; ~2/3 had a rate > 0.04. a conservative estimate of annual mortality was ~265 moose province-wide. given that railway expansion is predicted globally, and specifically in ontario, planning should include potential mitigation strategies that minimize ungulate-train collisions. alces vol. 55: 1–12 (2019) key words: elk, moose, railroad, railway mortality, wildlife-train collisions moose (alces alces), elk (cervus elaphus), boreal caribou (rangifer tarandus caribou), and white-tailed deer (odocoileus virginianus) are economically important in ontario as harvestable game and in ecotourism. ungulates are critical prey species for large carnivores, common food sources for scavengers, and as browsers and grazers, influence and maintain forest openings and grasslands (frank et al. 1998). they also provide subsistence and culturally important items such as meat, hides, teeth, and antlers for indigenous peoples. concern about decline in numerous north american moose populations (timmerman and rodgers 2017), the recent listing of boreal caribou as threatened by the committee on the status of endangered wildlife in canada (cosewic) (thomas and gray 2002), and the current stagnation in two of four restored elk populations in central ontario (popp et al. 2014) together warrant study of potential mortality sources that influence population dynamics of these species. although both natural and anthropogenic causes of ungulate declines are recognized and reasonably well understood, including overharvest, habitat degradation, parasites, and diseases (e.g., see toweil and thomas 2002, franzmann and schwartz 2007), the potential population impact of train collisions has received less attention and study. although minimal research has addressed impacts of train traffic on wildlife (popp and boyle 2017), collisions obviously cause direct mortality of multiple wildlife species (van der grift 1999, dorsey 2011, heske 2015), and indirect effects include habitat ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 2 fragmentation, noise, light, chemical pollution, and general stress (waller and servheen 2005, bartoszek and greenwald 2009, dorsey 2011). more relevant to our research, train collisions have been directly implicated in declines of local moose populations in alaska (becker and grauvogel 1991) and norway (gunderson and andreassen 1998). several factors make the collection of reliable, wildlife-train collision data difficult, including inaccessibility of remote railways, lack of experienced observers to accurately identify collisions and mortality, and the inherent difficulty of identifying and investigating collisions from moving trains. thus, wildlife mortality estimates along railways typically lack sufficient resolution to identify specific issues and mitigation strategies (wells et al. 1999). however, increased local and regional impacts on ungulate populations is likely, given that the global railway network of 1.4 million km is predicted to increase 45% by 2050 (dulac 2013, dorsey et al. 2015) and this expansion will bring more train traffic at higher speeds. in this study we measured and assessed railway use and train collision rates by moose and elk along a 20 km section of a busy railway in central ontario, canada. with these data, we estimated local and regional collision rates with the goal of providing natural resource managers a preliminary assessment and methodological approach to address moose-train collisions at these scales. study area the study area was located south of the city of greater sudbury in central ontario (46° 20’ 30”, 80° 50’ 30” 46° 11’ 30”, 80° 50’ 00”) and focused on a 20 km railway section which is part of the transcontinental canadian national railway (cnr) network (fig. 1). the railway was situated in the great lakes-st. lawrence forest that is a mixture of northern coniferous and deciduous trees. the canadian shield topography consists of numerous rocky ridges that promote growth of red oak (quercus rubra) in soils mostly composed of shallow surface deposits of silt and sand (rowe 1972). mean daily temperature ranges from -13.6 °c in january to 19 °c in july, average annual rainfall is 656.5 mm, and average snowfall is 274.4 cm with measurable snowfall occurring 78 days annually (sudbury weather station data 2006–2016, environment canada). the railway ran northwest to southeast at elevations between 200 and 230 m asl. the wanapitei river ran parallel to the railway and included ~ 0.2 km² flooded area created by a small hydroelectric dam, and sled lake (~1 km²) lay within 1.5 km. the 20 km study section transected 8 open marsh areas of various size. in all, ~30% of the habitat along the railway consisted of wetlands other than the river, and the tracks bisected or ran adjacent to ~ 3 km of open grasslands. there were 5 unprotected road crossings at which approaching trains were obliged to sound the whistle several times within 0.5 km. train speed varied from 60–80 km/h depending on the railway topography. although some long straightaways were present, the track had moderately to strongly winding sections, with at least 4 curves approaching 90–100 degrees (fig. 1). the track was regularly cleared of snow with a specialized rail-plough. ontario’s recreational hunting regulations are based on 95 wildlife management units (wmu) of various size and shape, as set by the ontario ministry of natural resources and forestry (omnrf). moose occur in at least 65 wmus and this study was conducted in wmu 42 which is located south-centrally within the provincial moose range. in 2015, the estimated moose density in wmu 42 was 36.7 animals/100 km² (omnrf 2016). the elk population in the alces vol. 55, 2019 ungulates and railways in ontario – hamr et al. 3 study area was estimated at 95–200 animals, and a core subset (~ 60–80 females, yearlings, and calves) regularly travelled along and seasonally traversed the railway (martin 2011, mcgeachy 2014). white-tailed deer were present in the area, but primarily in summer and not during the winter study period (popp 2017). ungulate predators in the study area included american black bear (ursus americanus), gray wolf (canis lupus), eastern (algonquin) wolf (c. lycaon), coyote (c. latrans), and their hybrids. methods to monitor train traffic during winter when most ungulate-train collisions occur, a reconyx© motion-triggered trail camera was placed on a tree directly adjacent to the railway within the 20 km study section in december 2012 and january 2013. radiocollared elk (20–30 adult females) in the area were located 1–2 times weekly via vhf ground telemetry to provide a temporal assessment of movement, railway use, and collisions. as all collisions with radio collared elk occurred in december-april, mortality surveys were conducted after the spring thaw in late march-early april 2006 2016. the 20 km section of railway was surveyed entirely during a single day each year. surveys were completed by 3 crews that walked separate 6–7 km sections. each crew of 2–4 observers walked on both sides of the railway scanning the immediate rail-bed and a 20–30 m margin for animal remains. mortalities were identified to species by size, hair, antlers, and skull shape. the location coordinates were recorded with gps units (5–10 m accuracy). partial skeletons were classified as either recent or old by the degree of bone bleaching, presence of dried muscle on bones, moisture content of bone marrow, and by comparing the current year’s locations with those of the previous year. these surveys were supplemented with opportunistic reports from people in passing trains or vehicles who observed fig. 1. location of the 20-km study section of cnr railway (bordered by dashed line) located 20–40 km south of sudbury in central ontario, canada. ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 4 collisions/carcasses. on several occasions cnr workers reported elk collisions, and if accessible, these animals were checked for pregnancy and physical condition. it was assumed that the number of carcasses was a minimal estimate of train collisions because a struck animal could move and die outside the survey boundary, a dismembered carcass could have been removed by a large scavenger (e.g., bears and wolves), and carcasses salvaged for meat by cnr employees at the time of, or shortly after a collision, would be unaccounted for in the survey. it is possible that an animal could have died within the boundary of another cause; however, the majority of collisions were confirmed by either a severely mangled carcass, or a broken leg(s) and/or spine. the estimated 2015 moose density (animals/100 km²) in each wmu that had railways was obtained from a public information website (omnrf 2016), and the extent of the provincial railway network in moose range was calculated from transportation corridor layers in gis (arcmap 10.3.1, fig. 2 and 3). to avoid overestimation of moose mortality on railway segments with low train volume, estimates were calculated only for those cnr and canadian pacific fig. 2. location of the ontario provincial railway network depicting the main cnr and cpr lines traversing the province from the southern canadian shield to the manitoba border (portions used for moose mortality estimates are bolded); moose distribution range in ontario roughly coincides with the canadian shield (shaded). the railroad section between cartier and white river was used in an earlier unpublished survey of train-induced moose mortality (heerschap 1982). the present study location is shown by a dot. alces vol. 55, 2019 ungulates and railways in ontario – hamr et al. 5 railway (cpr) mainlines transecting the province, and only 38 wmus in largely unpopulated regions of the precambrian shield (fig. 3). marginal areas of moose range with substantial human populations and agricultural activity (southern canadian shield), and sections running through larger urban centers (e.g., sudbury and thunder bay) were omitted from the analysis. hereafter, the wmu railway segment used in the analysis is referred to as “high-use railway”. as the baseline, the 20 km study section was part of the cnr transcontinental line and considered representative of train traffic associated with the main railway system. the 2015 population density estimate (36.7 moose/100 km²) in wmu 42, and the 11-year (2006–2016) mortality estimate on the 20 km study section within, were used to estimate mortality rates in all wmus transected by cnr and cpr mainlines within moose range. the 2015 moose density estimate in each wmu was divided by the density estimate in wmu 42 (36.7 moose/100 km²). this fraction was multiplied by the annual mortality rate on the 20 km study section during the 11 years of study (1.3 moose/km). this estimate was multiplied by the length (km) of high-use railway within a wmu to calculate an equivalent estimate of the 11-year mortality rate; mortality estimates were expressed to the nearest whole number. as easily recognized topographical features such as rivers, highways, and railways often delineate wmu boundaries, if a railway segment formed the boundary between adjacent wmus, 1/2 of the shared segment length was attributed to each adjacent wmu. results railway traffic volume most train traffic moved in a southerly direction in the 20 km study section during the 11-year study period; however, the schedule varied with seasonal customer requirements. trains passed as frequently as every 20–30 min, but more commonly once every 2–3 h. the remote camera recorded 304 train passes in 26 days of monitoring in the winter 2012–2013. the maximum number of passes was 16 and the minimum 5 during a 24-h period; the average was fig. 3. estimated annual rate of moose-train collisions in wildlife management units containing high use railway lines (bolded black) within the ontario moose distribution range. ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 6 ~12 trains/d (0.5 train/h). the majority (70.4%) of traffic was in darkness with up to 13 trains certain nights, or nearly 1 train/h; daylight was defined as 0800–1700 hr. train collision rates a total of 26 moose collisions were identified during the annual surveys (2006–2016) in the 20 km study section (annual count = 0–7); 0–2 collisions occurred in 7 of 11 years (table 1). the annual collision rate was 0.13 moose/km on high-use railway. the highest collision rates occurred in 2013 (n = 7) and 2006 (n = 4) when snowfall was above average (1–30 cm; environment canada, sudbury weather station, 2006–2016). conversely, only a single moose collision occurred in 2009 when snowfall was above average, and 4 collisions occurred in 2011 when snowfall was below average (table 1); small sample size precluded statistical analysis. collision rate (# moose/km/yr) ranged from 0.02–0.15 in the 38 wmus that were extrapolated with data from wmu 42 (fig. 3). nearly 1/3 of these wmus had collision rates >0.08 moose/km/yr (table 2, fig. 3). the highest annual mortality (n = 27) was in wmu 12b in northwestern ontario with 226 km of major railways and a table 1. the number of moose, elk, and deer killed in train collisions as identified during 11 annual surveys of a 20 km section of cnr railway located in wmu 42 in central ontario. * identifies winters with above average snow depth. year moose elk white-tailed deer 2006* 4 8 0 2007 0 2 0 2008 0 1 0 2009* 1 22 0 2010 3 4 1 2011 4 3 0 2012 1 2 0 2013* 7 14 0 2014 2 9 1 2015 2 2 0 2016 2 5 0 total 26 72 2 table 2. ontario wildlife management units (wmu) transected by high use railways with the highest estimated moose mortality rates due to trains (fig. 3, red color). moose-train collision rates were obtained by multiplying the ratio of the 2015 moose density estimate (wmux/wmu 42) by the rate of moosetrain collisions measured in wmu 42. the length of high use railway within each wmu was then factored into the collision rate calculation (see methods). wmu 2015 moose density (#/100 km²) high use railway (km) collision rate (moose/km/year) mortality (# moose/year) 5 31.2 182.0 0.12 22 11b 39.4 16.3 0.12 2 12a 35.2 34.9 0.14 5 12b 33.8 225.9 0.12 27 13 25.3 220.4 0.09 20 14 40.6 40.1 0.15 6 15a 32.8 115.0 0.12 14 15b 25.1 98.9 0.09 9 21b 26.2 170.1 0.09 15 22 28.6 101.0 0.10 10 29 24.8 30.5 0.10 3 39 24.6 147.5 0.09 13 alces vol. 55, 2019 ungulates and railways in ontario – hamr et al. 7 population density of 34 moose/100 km² (fig. 3, table 2). the lowest annual mortality (n = 0.1) was in wmu 53a at the southern fringe of the canadian shield with only 1.9 km of high use railway and a moose population density of 17 moose/100 km² (fig. 3). three northwestern ontario wmus had an estimated annual mortality of ≥20 moose, and 7 central ontario wmus (including wmu 42) had an estimated annual mortality of 10–19 moose (table 2). total moose mortality estimated in 38 wmus transected by major railways during the 11-year period was 2,642 animals, or an annual minimum of ~265 moose. elk collisions within the 20 km study section varied from 1–22 annually, averaging 6.5 collisions/year (table 1). the highest collision rates occurred in 2006 (8), 2009 (22), and 2013 (14) during winters with above average snowfall (environment canada, sudbury weather station, 2006–2016). many collisions occurred where the railway curved (popp et al. 2018), passed through rock-cuts, or was bordered by steep embankments. in such cases, the visibility of an approaching train was obscured and escape from the railbed was hindered by the topography (fig. 4). due to the herding behavior of elk during winter, collisions often resulted in multiple simultaneous casualties. only 2 white-tailed deer were located over the 11 years of surveys (table 1), largely confirming that most migrate from the study area prior to winter; neither mortality occurred in a winter with above average snowfall. given the smaller body size of deer, it is possible that carcasses were scavenged prior to the surveys. total ungulate collisions within the 20 km study section varied from 1–23 animals annually; 7 of 11 years had <10 collisions, whereas annual fig. 4. adult female elk killed by a cnr train in winter 2013. the curving railbed conceals the approaching train and rock cuts on both sides prevent lateral escape. browse trees on railbed margins provide easily accessible forage for ungulates (photo j. hamr). ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 8 collisions were >10 in the 3 years (2006, 2009, 2013) with above average snow. discussion moose-vehicular collisions have presumably received more research attention because of the associated human mortality and direct economic losses; e.g., in newfoundland (oosenbrug et al. 1986), quebec (grenier 1973, jolicoeur and crete 1994), and maine (danks and porter 2010). more recently, increasing attention towards moose-train collisions is occurring globally. although belant (1995) reported moosetrain collisions were infrequent (3–5 annually) on 1,200 km of railways in northeastern minnesota nearly 3 decades ago, other studies in western canada, alaska, and europe indicated that railway mortality can be locally significant and possibly influence population dynamics of moose. for example, child et al. (1991) reported an annual average of 200 moose-train collisions in british columbia, annual collisions ranged from 9 to 725 along a 756 km stretch of railway in alaska (0.01–0.96 moose/km; modafferi 1991), and the collision rate on a 240 km stretch of railway in norway was 0.4 moose/km, exceeding 80 fatalities annually (gunderson and andreassen 1998). our data and estimates (0.02< and < 0.15 moose/km) are similar in both magnitude and variation with these north american and scandinavian examples. the aforementioned, our, and other studies (child 1983, muzzi and bisset 1990, anderson et al. 1991, gundersen and andreassen 1998, bertwhistle 1999, danks and porter 2010, dorsey 2011) identify that moose-train collisions are influenced by several factors including the railway network, volume and frequency of train traffic, animal density, seasonal range use, migration patterns, snow conditions, and the local topography and habitat associated with the railway. although moose are mostly solitary and adapted to moving in deep snow and wetlands, elevated moose-train collisions were documented in winters with high snowfall in northwestern ontario (muzzi and bisset 1990). similarly, we found higher collision rates with moose and elk in years with above average snow when these animals used the railway as a snow-free travel corridor (table 1). even in 2015–2016 when snow depth was less than average, our camera captured multiple images of moose (n = 15) and elk (n = 13) on the 20 km study section (popp and hamr 2018, fig. 5). analysis of videos taken from moving trains indicates that moose are often reluctant to leave a railway when trains approach and may attempt to outrun them (rea et al. 2010). a joint omnrf-cpr survey of moose mortality was conducted 3 decades earlier on the railway (417.6 km) from cartier to white river between sudbury and thunder bay in central ontario, and only 75 km from the northern limit of our study area (fig. 2; heershap 1982). cpr engineers on the cartier-white river railway section recorded all wildlife-train collisions and documented 31 moose collisions between june 1981 and may 1982. the extrapolated annual mortality was 0.07 moose/km, about half of our annual estimate of 0.13 moose/ km. the cartier-white river section traverses 3 wmus with estimated moose densities of 13–22 moose/100 km² in 2015, densities ~40–70% lower than that in our study area (36.7 moose/100 km²), indicating proportional similarity relative to population density, despite the 30-year period between the studies. moose-train collisions were sporadically reported throughout ontario for decades (e.g., forbes and theberge 1992), but no systematic surveys were conducted previously. our annual estimates of provincial mortality are moderate (minimally 250–300 animals), alces vol. 55, 2019 ungulates and railways in ontario – hamr et al. 9 yet should be considered during harvest planning and allocation of moose hunting permits (tags), particularly in wmus with higher and local impacts (fig. 3). wildlife managers are encouraged to incorporate railway mortality into population modelling where major railroads transect moose and elk habitat (e.g., stocker 1983, messier 1994, eberhardt 2010). although arguably conservative and based on a single site, our model provides for easy estimation of annual mortality based on moose density and length of railway per wmu. the elk-train collision rate was ~3x higher than that of moose during the 11-year survey (table 1). telemetry locations of reintroduced elk (1998–2001) indicated that this local population incorporated the railway within their seasonal range. it was used as a travel and forage corridor (mcgeachy 2014), and importantly, elk were closest to and used the railway most during winter (popp et al. 2018). multiple mortalities were documented in winters 2009 and 2013 when trains ploughed through groups composed mainly of pregnant cows and calves (table 1). combined, railway mortality and wolf predation have prevented the growth of this un-hunted elk population through constant removal of adult females with live and unborn calves (popp et al. 2014, 2018). further, the humane dispatch of injured animals which is rarely addressed, introduces another management concern of train-collisions for both moose and elk. because concern about the impact of railways on ungulates and other wildlife was non-existent when most north american railways were constructed over a century ago, it is unsurprising that many railways transect critical wildlife habitats and traditional migratory corridors. as railway networks expand in the near future, it is imperative that ungulate ecology (e.g., movements and seasonal habitat use) be considered in the planning process to minimize what are often local impacts. in ontario, there is impending potential for railway network expansion associated with the planned ring of fire chromite mining operation in the mineral-rich james bay lowlands. the project would involve about 400 km of new railway northward through wmus 18a, 17, and 1d (fig. 3), currently with low or no fig. 5. adult female moose using the railway as a travel corridor during winter 2016 (photo j. popp). ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 10 wildlife-train collision impact, but located within moose and threatened boreal caribou range. our study used three methods to obtain valuable information – radio-telemetry, cameras, and field surveys – to monitor and predict seasonal use and mortality of moose and elk on railways. as in this study with elk (popp et al. 2018), local hot-spots of moose-train collisions (anderson et al. 1991, andreassen et al. 2005, gundersen et al. 1998) are typically associated with specific habitat features that function as ecological traps along railways. where possible, mitigation could include fencing, eliminating the attractant, wildlife crossing structures, and reduced speed (jaren et al. 1991, wells et al. 1999). a management strategy that incorporates research, monitoring, and specific mitigation strategies aimed at reducing train collisions would proactively address the projected increase in railways and train traffic in the face of provincial and regional moose decline. as railway companies periodically invest in their infrastructure (e.g. https://www. c n . c a / e n / n e w s / 2 0 1 8 / 0 7 / c n i n v e s t i n g approximately315-millionto-expandandstrengthen/), mitigation efforts addressing wildlife-train collisions should be considered and incorporated where feasible into improvement and expansion projects. acknowledgements we acknowledge cambrian college nature-based adventure tourism and environmental monitoring & impact assessment students, laurentian university graduate and undergraduate biology students, and sudbury elk restoration committee (serc) volunteers for their participation in railroad surveys. in-kind and financial support was provided by cambrian college, laurentian university, and serc. references anderson, r., b. wiseth, p. h. pedersen, and v. jaren. 1991. moose-train collisions: effects of environmental conditions. alces 27: 79–84. andreassen, h. p., h. gundersen, and t. storraas. 2005. the effect of scent marking, forest clearing, and supplemental feeding on moose–train collisions. journal of wildlife management 69: 1125–1132. bartoszek, j., and k. r. greenwald. 2009. a population divided: railroad tracks as barriers to gene flow in an isolated population of marbled salamanders (ambystoma opacum). herpetological conservation and biology 4: 191–197. becker, e. f., and c. a. grauvogel. 1991. relationship of reduced train speed on moose-train collisions in alaska. alces 27: 161–168. belant, j. l. 1995. moose collisions with vehicles and trains in northeastern minnesota. alces 31: 1–8. bertwhistle, j. 1999. description and analysis of vehicle and train collisions with wildlife in jasper national park, alberta. < h t t p : / / e s c h o l a r s h i p . o r g / u c / item/4w71z50t> (accessed march 2017). child, k. n. 1983. railways and moose in the central interior of british columbia: a recurrent management problem. alces 19: 118–135. _____, s. p. barry, and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27: 41–49. danks z. d., and w. f. porter. 2010. temporal, spatial, and landscape habitat characteristics of moose–vehicle collisions in western maine. journal of wildlife management 74: 1229–1241. dorsey, b. p. 2011. factors affecting bear and ungulate mortalities along the canadian pacific railroad through banff and yoho national parks. m. s. thesis, montana state university, bozeman, montana, usa. https://www.cn.ca/en/news/2018/07/cn-investing-approximately-315-million-to-expand-and-strengthen/� https://www.cn.ca/en/news/2018/07/cn-investing-approximately-315-million-to-expand-and-strengthen/� https://www.cn.ca/en/news/2018/07/cn-investing-approximately-315-million-to-expand-and-strengthen/� https://www.cn.ca/en/news/2018/07/cn-investing-approximately-315-million-to-expand-and-strengthen/� http://escholarship.org/uc/item/4w71z50t http://escholarship.org/uc/item/4w71z50t alces vol. 55, 2019 ungulates and railways in ontario – hamr et al. 11 _____, m. olsson, and l. j. rew. 2015. ecological effects of railways on wildlife. pages 219–227 in r. van der ree, d. j. smith, and c. grilo, editors. handbook of road ecology. wiley blackwell, west sussex, united kingdom. dulac, j. 2013. global land transport infrastructure requirements: estimating road and railway infrastructure capacity and costs to 2050. international energy agency, paris, france. eberhardt, l. l. 2010. models of ungulate population dynamics. rangifer special issue no. 7: 24–29. forbes, g. j., and j. b. theberge. 1992. importance of scavenging on moose by wolves in algonquin park, ontario. alces 28: 235–241. frank, d. a., s. j. mcnaughton, and b. tracy. 1998. the ecology of the earth’s grazing ecosystems. bioscience 48: 513–521. franzmann, a. w., and c. c. schwartz. 2007. ecology and management of the north american moose. university press of colorado, louisville, colorado, usa. grenier, p. 1973. moose killed on the highway in the laurentides park quebec, 1962 to 1972. proceedings of the north american moose conference and workshop 9: 155–194. gundersen, h., and h. p. andreassen. 1998. the risk of moose (alces alces) collision: a predictive logistic model for moose-train accidents. wildlife biology 4: 103–110. _____, _____, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34: 385–394. heerschap, a. 1982. big game mortality by trains: cartier to white river, june 1981-june 1982. unpublished report of the chapleau district, ontario conservation officer service, chapleau, ontario, canada. heske, e. j. 2015. blood on the tracks: track mortality and scavenging rate in urban nature preserves. urban naturalist 4: 1–13. jaren, v., r. andersen, m. ulleberg, p. h. pedersen, and b. wiseth. 1991. moose-train collisions: the effects of vegetation removal with a cost-benefit analysis. alces 27: 93–99. jolicoeur, h., and m. crete. 1994. failure to reduce moose-vehicle accidents after a partial drainage of roadside salt pools in quebec. alces 30: 81–89. martin, m. m. 2011. spatial behaviour and habitat use by elk (cervus elaphus) in response to highway construction and interprovincial relocation. m. s. thesis, laurentian university, sudbury, ontario. mcgeachy, d. n. c. 2014. population distribution and seasonal resource selection by elk (cervus elaphus) in central ontario. m. s. thesis, laurentian university, sudbury, ontario. messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478–488. modafferi, r. d. 1991. train-moose kill in alaska: characteristics and relationships with snowpack depth and moose distribution in lower susitna valley. alces 27: 193–207. muzzi, p.d., and a.r. bisset. 1990. effectiveness of ultrasonic wildlife warning devices to reduce moose fatalities along railway corridors. alces 26: 37–43. ontario ministry of natural resources and forestry (omnrf). 2016. moose population management. (accessed july 2017). oosenbrug, s. m., r. w. mcneily, e. w. mercer, and j. f. folinsbee. 1986. some aspects of moose-vehicle collisions in eastern newfoundland 1973–85. alces 22: 377–394. popp, j. n. 2017. population dynamics of reintroduced elk (cervus elaphus) in eastern north america. ph. d. thesis, laurentian university, sudbury, ontario. https://www.ontario.ca/page/moose-population-management>� https://www.ontario.ca/page/moose-population-management>� https://www.ontario.ca/page/moose-population-management>� ungulates and railways in ontario – hamr et al. alces vol. 55, 2019 12 _____, and s. p. boyle. 2017. railway ecology: underrepresented in science? basic and applied ecology 19: 84–93. _____, and j. hamr. 2018. seasonal use of railways by wildlife. diversity 10, 104:doi10.3390/d10040104. _____, _____, c. chan, and f. f. mallory. 2018. elk (cervus elaphus) railway mortality in ontario. canadian journal of zoology 96: 1066–1070 http://dx.doi. org/10.1139/cjz-2017-0255. _____, t. toman, f. f. mallory, and j. hamr. 2014. a century of elk restoration in eastern north america. restoration ecology 22: 723–730. rea, r. v., k. n. child, and d. a. aitken. 2010. youtube insights into moose-train interactions. alces 46: 183–187. rowe, j. s. 1972. forest regions of canada. publication no. 1300. canadian forest service, ottawa, canada. stocker, m. 1983. ungulate population dynamics and optimization models. ecological modeling 18: 121–139. thomas, d. c., and d. r. gray. 2002. cosewic assessment and update status. report on the woodland caribou (rangifer tarandus caribou) in canada. committee on the status of endangered wildlife in canada, ottawa, canada. timmermann, h. r., and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. toweill, d. e., and j. w. thomas. 2002. north american elk, ecology and management. smithsonian institution press, washington, d. c., usa. van der grift, e. 1999. mammals and railroads: impacts and management implications. lutra 42: 77–91. waller, j. s., and c. servheen. 2005. effects of transportation on grizzly bears in northwestern montana. journal of wildlife management 69: 985–1000. wells, p., j. g. woods, g. bridgewater, and h. morrison. 1999. wildlife mortalities on railways: monitoring methods and mitigation strategies. pages 85–88 in g. l. evink, p. garrett, and d. zeigler, editors. proceedings of the third international conference on wildlife ecology and transportation (icowet), 13–16 september, missoula, montana, usa. http://dx.doi.org/10.1139/cjz-2017-0255 http://dx.doi.org/10.1139/cjz-2017-0255 _gjdgxs _hlk536800738 _30j0zll _1fob9te _3znysh7 _2et92p0 alces37(1)_61.pdf 4110.pdf alces vol. 43, 2007 stumph and wright moose browse quality 129 effects of willow quality on moose distribution in a montane environment bradley p. stumph1 and r. gerald wright2 1department of fish and wildlife resources, university of idaho, p.o. box 441136, moscow, id 83844-1136, usa; 2usgs idaho cooperative fish and wildlife research unit, department of fish and wildlife resources, university of idaho, p.o. box 441136, moscow, id 83844-1136, usa salix alces alces shirasi nutrients. key words: salix alces alces shirasi salix moose browse qualitystumph and wright alces vol. 43, 2007 130 communities. cervus elaphus 2 study area pinus contorta picea engelmannii abies lasiocarpa alces vol. 43, 2007 stumph and wright moose browse quality 131 methods forage sample collection s. geyeriana, s. monticola, s. planifolia planifolia s. planifolia ing collection. forage quality analyses crude protein moose browse qualitystumph and wright alces vol. 43, 2007 132 fiber analysis moose distribution statistical analyses tion. results forage sample collection alces vol. 43, 2007 stumph and wright moose browse quality 133 s. planifolia s. planifolia forage quality analyses crude protein . s. geyeriana s. planifolia senecio triangularis a b moose browse qualitystumph and wright alces vol. 43, 2007 134 fiber analysis s. planifolia p s. planifolia s. planifolia moose distribution n n n n mountain national park, co, 2003-2004. crude proteincontentof low-elevationwillow species 10 15 20 25 15-jun 30-jun 15-jul 30-jul 15-aug % c ru d e p ro te in s. geyeriana s. monticola s. planifolia 2003 10 15 20 25 15-jun 30-jun 15-jul 30-jul 15-aug date sample collected % c ru d e p ro te in 2004 a b salix planifolia rocky mountain national park, co, 2003-2004. statistical analyses. 5 10 15 20 25 15-jun 30-jun 15-jul 30-jul 15-aug % c ru d e p ro te in low (2560 2803m) mid (2804 3047m) high (3048 3291m) 2003 crude proteincontentof s. planifolia byelevation 5 10 15 20 25 15-jun 30-jun 15-jul 30-jul 15-aug date sample collected % c ru d e p ro te in 2004 alces vol. 43, 2007 stumph and wright moose browse quality f p r p f p x1 x2 x3 x4 x 2,808m, x 2 p 2 p f p discussion forage sample collection forage quality salix planifolia co, 2003-2004. ndfcontentof s. planifolia byelevation 15 25 35 45 55 15-jun 30-jun 15-jul 30-jul 15-aug n d f a s a % o f d ry w e ig h t low (2560 2803m) mid (2804 3047m) high (3048 3291m) 2003 15 25 35 45 55 15-jun 30-jun 15-jul 30-jul 15-aug date of collection n d f a s a % o f d ry w e ig h t low mid high 2004 moose browse qualitystumph and wright alces vol. 43, 2007 30 communities. alces vol. 43, 2007 stumph and wright moose browse quality 137 moose distribution moose browse qualitystumph and wright alces vol. 43, 2007 138 alces vol. 43, 2007 stumph and wright moose browse quality 139 moose browse qualitystumph and wright alces vol. 43, 2007 140 acknowledgements references agresti, a. 1990. categorical data analysis. york, usa. albon, angvatn. 1992. plant (a.o.a.c)associationof official analytical chemists ton, d.c., usa. baker bergström anell 22:91-112. bø jeljord. 1991. do continencochran 2 test. biodorn rocky mountain region. r2-rr-97-01. dungan right edwards itcey ericsson all anell. 2002. ranzmann chwartz goering an soest. 1970. usa. hjeljord istol. 1999. rangehouston association, moose, wyoming, usa. lein omarek irios ence 71:284. ubota ieger azar. 1970. ufeld owden. alces vol. 43, 2007 stumph and wright moose browse quality 141 alces alces shirasi laca, e. a., l. a. shipley eid. larter agy. 2001. seasonal leresche, r. e. 1974. moose migrations mcmillan ecology 34:102-110. nowlin eitz enney. oldemeyer ranzmann, a. l. brundage, p. d. arneson lynn peek reese obbins regelin, w. l., c. c. schwartz ranzmann sand, h., g. cederlund anell alces alces 102:433-442. schwartz, c. c., m. e. hubbert ranzmann 33. shipley, l. a., s. blomquist anell. palinger short, h. l., d. r. dietz emmenga. singer eigenfuss, r. g. cates, arnett spalinger, d. e., t. a. hanley obbins stevens, d. r. 1988. moose in rocky mountain national park. rocky mountain stewart, r. r., r. r. maclennan innear moose browse qualitystumph and wright alces vol. 43, 2007 142 van ballenberghe, v., d. g. miguelle accracken summer at denali national park, alaska. van soest usa. alces36_197.pdf 37 internal gross pathology of moose experimentally infested with winter ticks edward m. addison1,2, and robert f. mclaughlin3 1wildlife research and development section, ontario ministry of natural resources and forests, 2140 east bank drive, peterborough, ontario, canada k9j 7b8; 2present address: 26 moorecraig road, peterborough, ontario, canada k9j 6v7; 3r. r. #3, penetanguishene, ontario, canada l0k 1p0 abstract: captive moose (alces alces) infested with 21,000 and 42,000 larval winter ticks (dermacentor albipictus) in september-october, and unifested moose were studied to assess impact of winter ticks on moose. study animals were euthanized the following april near the end of the parasitic phase of winter ticks. major organs and selected superficial lymph nodes were examined and compared among treatment groups. no visible lesions were evident in spleen, lung, liver, thyroid, heart, adrenal, and kidney of most moose. several foci of necrosis in the liver of 1 moose were considered minor and unrelated to tick infestation. prescapular and prefemoral lymph nodes, but not popliteal nodes, were significantly heavier and reddened in infested than uninfested moose. hyperactive, hypertrophied lymph nodes may compromise the immune defense of moose and may predispose infested moose to increased risk of bacterial infection. while not a proximate cause of death in heavily infested moose, bacterial infections may contribute as a secondary cause of death. alces vol. 55: 37–41 (2019) key words: alces alces, dermacentor albipictus, lymph nodes, moose, pathology, winter tick since the early 1900s winter ticks (dermacentor albipictus) have been associated with numerous die-offs of moose (alces alces) in the canadian provinces (samuel and barker 1979, samuel 2004), and most recently, annual epizootics (>50% calf mortality) are occurring with unprecedented frequency in the northeastern united states (jones et al. 2017, 2019). experimental studies with captive moose infested with winter ticks have demonstrated that the amount of grooming, rubbing, and hair loss are related directly with the level of infestation (mclaughlin and addison 1986, addison et al. 2019). captive studies further revealed that shivering by infested calves in winter is seldom observed in uninfested animals (addison and mclaughlin 2014), infested moose have less pericardial fat and abdominal visceral fat than uninfested animals (mclaughlin and addison 1986), and infested calves grow more slowly than uninfested calves (addison et al. 1994). in addition, glines and samuel (1989) reported transitory anemia in a captive calf, and the concentrated blood loss associated with feeding by adult female winter ticks is directly related to mortality of wild calves (samuel 2004, musante et al. 2007). the first record of the bacterium erysipelothrix rhusiopathiae in lymph nodes and other tissues collected from dead wild moose with high infestation of winter ticks was by campbell et al. (1994). a common route of infection for e. rhusiopathiae is from contamination of wounds (leighton 2001). because extensive grooming and rubbing induced by winter ticks (addison gross pathology – addison and mclaughlin alces vol. 55, 2019 38 et al. 2019) can cause extensive dermal wounds on moose (authors’ personal observation), high infestations of winter ticks may predispose moose to bacterial and fungal infections. here we present data and observations from internal gross pathology of uninfested captive moose and those experimentally infested with winter ticks to identify any differences in physiological response possibly associated with tick infestations. methods twelve moose captive-reared in 1982 in algonquin provincial park, ontario (45° 33’n, 78° 35’w) were used in this experiment which was part of a larger study (see addison et al. 1983). these animals were divided into 3 treatment groups: 4 uninfested (control) moose that were administered no winter ticks, 4 infested with 21,000 larvae, and 4 infested with 42,000 larvae; infestations occurred from 17 september to 12 october 1982. control moose were sprayed with an acaricide (dursban m., dow chemical of canada ltd., sarnia, ontario, canada) twice in november, and powdered with rotenone in december, january, and february in an attempt to prevent accidental infestation. moose were euthanized by initially immobilizing them with 300 mg of xylazine hydrochloride (rompun, haver-lockhart laboratories, mississauga, ontario, canada) followed with a lethal dose of t-61 (n-[2( m e t h o x y p h e n y l ) 2 e t h y l b u t y l ( 1 ) ] ghydroxy-butyramide and 4,4’ –methylene bis(cyclohexyl-triemthylammonium iodide)) (hoechst canada inc., montreal, quebec, canada). necropsies were performed on 18–29 april 1983. heart, lungs, liver, kidney, spleen, and thyroid and adrenal glands were extracted and examined for visible lesions. in addition, the prefemoral, prescapular, and popliteal superficial lymph nodes were extracted, trimmed, weighed (0.01 g), and photographed. the prescapular and prefemoral lymph nodes are bifurcated into distinct upper and lower portions in moose. an analysis of variance (anova) and tukey’s test were used to examine for difference in weight of the prescapular, prefemoral, and popliteal lymph nodes among the 3 treatment groups; significance was set at p = 0.05. all studies were approved under an animal care protocol and with close scrutiny by a provincial veterinarian who set the april termination date of the experiment. results and discussion the numbers of ticks in the treatment groups should be considered in a relative rather than absolute sense. for example, although ticks were not applied to the control moose, and despite our preventative exercises, there were 0, 4, 21, and 85 ticks recovered from their hides at the termination of the experiment. for all practical purposes, however, they served as control animals given the infestation levels of the other groups. further, the infested moose successfully removed a measurable number of ticks by the end of the experiment (addison et al. 2019). the treatments are best considered as uninfested, low infestation, and moderate infestation because infestation is typically >35,000 ticks at death (jones et al. 2019). in 11 moose (all groups) there were no gross lesions in the heart, lungs, liver, kidney, spleen, and thyroid and adrenal glands. the other (12th) animal had several foci of necrosis in its liver that were considered minor and unrelated to tick infestation. the popliteal lymph nodes were of similar size in all 3 groups (p = 0.79) (table 1), and the prescapular (p = 0.17) and prefemoral (p = 0.45) lymph nodes were of similar size in the 2 infested groups (fig. 1). conversely, the prescapular and prefemoral nodes in infested moose were hyperplastic, and 3–4 × heavier alces vol. 55, 2019 addison and mclaughlin – gross pathology 39 table 1. mean weight (g) and range of superficial lymph nodes collected from 3 groups of euthanized moose that received different infestation treatments of winter ticks (4 animals per treatment) the previous autumn, ontario, canada. samples were collected and weighed (0.01 g) on 18–29 april 1983 when adult ticks typically drop from moose. lymph node body side treatment level (# ticks) 0 21,000 42,000 prescapular right 10.25 (7–14) 40.75 (34–50) 38.50 (21–57) left 11.25 (7–16) 35.75 (32–46) 34.00 (27–40) prefemoral right 9.00 (6–12) 30.75 (23–46) 29.50 (24–34) left 8.75 (3–15) 28.50 (23–37) 29.00 (19–34) popliteal right 5.25 (4–6) 5.25 (4–6) 6.25 (4–11) left 5.25 (4–7) 6.00 (5–7) 5.25 (4–7) fig. 1. superficial lymph nodes from moose not infested with winter ticks (a), infested with 21,000 ticks (b), and infested with 42,000 ticks (c). lymph nodes in (a) to (c) from top to bottom are upper prescapular (left and right), lower prescapular, upper prefemoral, lower prefemoral and popliteal nodes; (d) is a right lower prescapular lymph node of a year-old wild moose found dead. gross pathology – addison and mclaughlin alces vol. 55, 2019 40 (p < 0.001) than those in control moose (table 1, fig. 1). among the control moose, the heaviest prescapular and prefemoral lymph nodes were in the animal that harboured the most ticks (85; fig. 1a) and groomed and rubbed most, suggesting that even a light infestation may initiate a physiological response in the host. the apparent response of the prescapular and prefemoral nodes in infested captive moose has also been noted in heavily infested wild moose with hyperplastic and completely red lymph nodes (fig. 1d). an emaciated year-old wild moose also had hyperplastic and completely red prefemoral and prescapular lymph nodes with erysipelothrix rhusiopathiae also recovered from this animal (campbell et al. 1994). it follows that infested wild moose are immunocompromised and more susceptible to infection, and that the relative degree of such is probably related directly to infestation level. that grooming behaviour to reduce infestation may simultaneously increase the probability of secondary infection is somewhat ironic. although high infestations are clearly linked to mortal blood loss and epizootics (musante et al. 2007, jones et al. 2019), further pathology of infested moose might identify the relative influence of secondary bacterial infections. acknowledgements we thank d. fraser, s. fraser, s. gadawaski, a. jones, s. mcdowell, l. berejikian, k. long, k. paterson, l. smith, d. bouchard, v. ewing, j. jefferson, m. van schie, a. macmillan, a. rynard, n. wilson, c. pirie, and m. mclaughlin for their strong commitment to some or all of capturing, raising, and maintaining of moose calves and collection of winter tick larvae. we appreciate the guidance of i. k. barker in recommending the superficial lymph nodes to examine and in joining in examination for gross lesions in organs and glands collected. field work was conducted at the wildlife research station in algonquin park, ontario, canada. references addison, e. m., d. j. h. fraser, and r. f. mclaughlin. 2019. grooming and rubbing by moose with dermacentor albipictus and their relationship with hair loss and removal of ticks. alces 55: 23–35. _____, and r. f. mclaughlin. 2014. shivering by captive moose infested with winter ticks. alces 50: 87–92. _____, _____, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and not infested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469–1476. doi:10.1139/z94-194 _____, _____, and d. j. h. fraser. 1983. raising moose calves in ontario. alces 19: 246–270. campbell, g. d., e. m. addison, i. k. barker, and s. rosendal. 1994. erysipelothrix rhusiopathiae, serotype 17, septicemia in moose (alces alces) from algonquin park, ontario. journal of wildlife diseases 30: 436–438. doi:10.7589/ 0090-3558-30.3.436 glines, m. v., and w. m. samuel. 1989. effect of dermacentor albipictus (acari:ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. experimental and applied acarology 6: 197–213. doi:10.1007/bf01193980 jones, h., p. j. pekins, l. e. kantar, m. o’neil and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. _____, _____, _____, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’ neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi:10.1139/cjz-2018-0140 alces vol. 55, 2019 addison and mclaughlin – gross pathology 41 leighton, f. a. 2001. erysipelothrix infection. pages 491–493 in e. s. williams and i. k. barker, editors. infectious diseases of wild mammals. iowa state university press, ames, iowa, usa. mclaughlin, r. f., and e. m. addison. 1986. tick (dermacentor albipictus)induced winter hair-loss in captive moose (alces alces). journal of wildlife diseases 22: 502–510. doi:10.7589/ 0090-3558-22.4.502 musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101–110. samuel, b. 2004. white as a ghost: winter ticks and moose. natural history series, volume 1. federation of alberta naturalists, edmonton, alberta, canada. samuel, w. m., and m. j. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869), on moose alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15: 303–348. alces vol. 44, 2008 heard et al a gis modified survey design. 111 using gis to modify a stratified random block survey design for moose douglas c. heard1, andrew b. d. walker2, jeremy b. ayotte1, and glen s. watts1 1british columbia ministry of environment, 4051 – 18th ave., prince george, british columbia, canada, v2n 1b3; 25657 simon fraser ave., prince george, b.c., canada, v2n 2c4 abstract: we modified the standard, stratified random block design used typically in aerial surveys of moose (alces alces). we laid a grid of approximately 9 km2 cells over our study area, and gis was then used to allocate polygons into one of 2 strata within each grid cell. the 2 strata were based upon vegetation attributes that were predicted to have either high or low moose density from previous research. we assumed that polygons of early seral forest stands (<40 yr), shrubs, and meadows would have high moose density relative to other vegetation attributes. vegetation polygons were often <1 km2, consequently, single grid cells usually included >1 high and low density polygons. adjacent cells were amalgamated to produce sample units with >4 km2 of high density stratum area. real-time navigation was used and the flight track was recorded over a map of sample units, strata boundaries, and topographic features to accurately identify polygon boundaries and assign each sighted moose to the appropriate strata. we concluded that our approach was efficient and effective in fine-grained environments where the relative selection by moose for vegetation patches is well understood, and those patches are mapped in digital databases. alces vol. 44: 111-116 (2008) key words: alces alces, british columbia, gis, moose, stratified random block survey, vegetation attributes. moose population parameters such as density, and age and sex composition are typically gathered from aerial surveys that incorporate a stratified random block design (boertje et al. 1996, timmerman and buss 1998). the often used stratified random block design of gasaway et al. (1986) was developed to survey moose in the northern boreal forest, the subalpine zone, and the northern coastal shrub zone where vegetation patches are large at the northern range of moose distribution. in this paper we describe a modification of the stratified random block design (gasaway et al. 1986) for use in central british columbia where distribution of moose in early winter is predictable, and population density varies substantially among small, discrete, and mapable patches of vegetation (nielsen et al. 2005). methods study area beginning in the late 1960s, clearcut logging superseded fire as the primary landscape disturbance in central british columbia (heard et al. 1999, nielsen et al. 2005). the early seral vegetation communities created by clearcutting were largely responsible for the relatively high abundance of moose throughout much of interior british columbia (spalding 1990, thompson and stewart 1998, heard et al. 1999, shackleton 1999). the resulting landscape contained a mosaic of multi-aged stands (i.e., regenerating cutblocks) ranging from 10-10,000 ha within a mature forest matrix; for a landscape view go to google earth maps () at 53º 55’ n x 122º 45’ w at an eye altitude of 200 km. a gis modified survey design – heard et al. alces vol. 44, 2008 112 stratification previous research revealed that moose selected low-elevation (<1200 m) cutblocks with a few, specific vegetation attributes during early winter (thompson and stewart 1998, nielsen et al. 2005). therefore, we developed a stratification design that used 2 strata (s1 and s2) based upon use of different vegetation types by moose. although potential high population density patches were small compared to the sample unit (su) size of 30 km2 suggested by gasaway et al. (1986), they were usually discernible in the field, and their boundaries were available in digital databases. stratum 1 (s1) was the high population density stratum and included young forest (≤40 years), areas with shrub crown closure ≥60%, and vegetation resource inventory descriptors for natural shrubby and open areas including meadow, open range, non-productive brush, non-commercial brush, and not sufficiently restocked. stratum 2 (s2), the low population density stratum, was classified as forest >40 years old and the remainder of study area polygons not classified as s1 including gravel bars, riparian areas, and cutblocks or burns <5 years old. we obtained forest, cutblock, and other vegetation patch attributes from 3 provincial digital databases (i.e., vegetation resources inventory, forest inventory polygon, and results) stored in british columbia’s land and resource data warehouse. unlike the stratified random block design of gasaway (1986), we assumed that our stratification process did not require a pre-census stratification flight. sample unit definition to determine stratum-specific su's, a grid of approximately 9 km2 (3.2 × 2.8 km) cells was laid out over the study area, specifically the grid layer from the land and resource data warehouse named “a5k sampling tiles.” all polygons within each grid cell were classified as s1, s2, or outside of the survey zone (land >1200 m elevation and large lakes). to improve the likelihood of observing at least 1 moose in each su (bergerud and manual 1969, heard et al. 1999), adjacent cells were arbitrarily amalgamated until the sum of all the s1 polygon areas added up to >4 km2 (fig. 1). the high population density su was the set of all s1 polygons within that group of cells, and the low population density su was the set of all s2 polygons within that group of cells. to estimate moose numbers in the high population density stratum, a random sample of sus was chosen from the entire study area and all moose were counted within all the s1 polygons in each selected su. to estimate moose numbers in the low popualtion density stratum, a random sub-sample of sus from the first sample was selected, and moose were counted within all the s2 polygons in those sus. field techniques to be certain of locating polygon boundaries within an su, real-time navigation was used where our flight track was recorded on a map of su and strata boundaries and topographic features using arcpadtm 7.0 (environmental systems research institute 2006) on a hewlittpackard ipaqtm handheld computer (hewlittpackard development company 2006) connected to a garmin mobile 10tm wireless gps unit (garmin international, inc. 2006). a comparable system includes the dnr garmin extension (t. loesch, minnesota department of natural resources; ) used in conjunction with arcviewtm (environmental systems research institute) and a garmin gps receiver (garmin international, inc. 2006). real-time navigation allows the navigator to instantly determine the location of observers and moose relative to strata and su boundaries and ensure complete sampling of a su. poole et al. (1999) suggested that incorporating realtime navigation into moose inventories can reduce flying time by 10-20%. alces vol. 44, 2008 heard et al a gis modified survey design. 113 moose were counted from a bell 206b helicopter using a search pattern over the entire su that consisted of transects spaced 200-400 m apart depending on vegetation cover. each sighted moose was circled and age (adult or calf) and sex were recorded: calves were distinguished by size, bulls by antlers, and cows by the presence or absence of a white vulva patch, bell length and shape, and facial colouration (timmermann and buss 1998). if additional moose were sighted during circling, they were categorized likewise and added to the data. sightability correction corrections for sightability bias were made according to anderson and lindzey (1996). we estimated vegetation cover vis-we estimated vegetation cover visually to the nearest 5% within a 9 m radius of where a moose was first observed according to the standards of unsworth et al. (1998). vegetation cover estimates were grouped into 5 classes and we applied the class-specific detection probability and corresponding sightability correction factor (scf) of quayle et al. (2001) which includes data from sightability tests carried out in central british columbia (d. heard, unpublished data). the recommendations of gasaway et al. (1986) were followed with respect to conducting surveys during early winter and after a fresh snowfall when there was complete snow cover in the study area. fig. 1. an example of a high moose population density stratum sample unit (s1; all dark gray polygons combined) and a low moose population density stratum sample unit (s2; all stippled area) where 7 adjacent grid cells were amalgamated to form an area within which the s1 polygons would add to >4 km2. both the williston reservoir and elevations >1200 m (black area) were excluded from the survey zone. a gis modified survey design – heard et al. alces vol. 44, 2008 114 data analysis for each stratum we calculated a naïve population and sampling variance estimate for unequal sized sus as in jolly (1969). we then multiplied the naïve population estimate by the mean stratum-specific scf (sum of the corrected number of moose/number of moose observed) to obtain the corrected population estimate. the variance of the corrected population estimate was the sum of 1) the naïve sampling variance multiplied by the squared mean scf (goodman 1960, heard 1987), 2) the sightability variance, and 3) the model variance. the sightability and model variance were calculated with the program aerial survey (unsworth et al. 1998) and the detection probabilities from quayle et al. (2001). we used jolly (1969) rather than aerial survey (unsworth et al. 1998) to calculate the sampling variance because aerial survey calculates a population estimate using a sampling fraction based on the number of censused sus divided by the total number of sus in the study area. our analysis used a sampling fraction equal to the censused area divided by the total stratum area; we were not limited to sus of equal size with this approach. aerial survey calculates variances assuming that all unseen moose are in the same su as the observed moose. where there are few moose in each su, aerial survey will overestimate the variance among sus (i.e., the sampling variance). our approach assumed that unseen moose were divided among sus in proportion to the number of moose observed in each su. the total population estimate was then calculated as the sum of the corrected stratumspecific population estimates and its variance was the sum of the 2 stratum-specific variances. overall population density was obtained by dividing the total population estimate by the area of both strata combined. results and discussion the methods described in this paper have been used 6 times to estimate moose population density in central british columbia (e.g., heard et al. 1999, walker et al. 2006, 2007). the s1:s2 population density ratios in the 6 estimates were 8.7, 3.1, 2.8, 2.4, 2.3 and 1.7. overall, the moose population density estimates in s1 sample units averaged 3.5 times higher than those in s2 sample units. the survey coefficients of variation were always <20% of the total population estimate. thus, we concluded that our stratification design using 2 strata (s1 and s2) based on use of different vegetation characteristics by moose in early winter was effective. we also improved our cost and time efficiency because pre-survey stratification flights were not necessary. our approach resulted in relatively high sampling variance in the low population density stratum because many s2 sample units had no moose. the s2 variance would likely be reduced if we constructed larger s2 sample units or counted more s2 sample units; our sus were only 20-25% as large as the su size recommended by gasaway et al. (1986). our higher variances in the low population density stratum were contrary to the findings of gasaway et al. (1986) who concluded that high population density strata generally have the greatest variance and require the most sampling effort. an additional source of variation in the population estimates was related to inaccurate stratification resulting from discrepancies between the land cover database and the actual forest attributes. we suspected that both database input errors and out-of-date map attributes contributed to that discrepancy. time since logging and silvicultural treatments affect the availability and composition of moose forage (eschholz et al. 1996, thompson and stewart 1998, rea and gillingham 2001) hence, the distribution and abundance of moose (nielsen et al. 2005). the precision of the population estimate might have been improved with a more precise habitat use model. we believe that the design described alces vol. 44, 2008 heard et al a gis modified survey design. 115 in this paper should be useful and effective in fine-grained environments where the relative selection for vegetation patches by moose is well understood, and those patches are mapped in digital databases. acknowledgements we appreciated the gis and remote sensing expertise provided by k. bush and v. michelfelder, and the comments provided by 2 anonymous reviewers. funding for this project was provided by the british columbia ministry of environment. references anderson, c. r., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24:247-259. bergerud, a. t., and f. manuel. 1969. aerial census of moose in central newfoundland. journal of wildlife management 32:722-728. boertje, r. d, p. valkalkenburg, and m. e. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal wildlife management 60:474-489. eschholz, w. e., f. a. servello, b. griffith, k. s. raymond, and w. b. krohn. 1996. winter use of glyphosate-treated clearcuts by moose in maine. journal wildlife management 60:764-769. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers university of alaska, fairbanks. no. 22. goodman, l. a. 1960. on the exact variance of products. journal american statistical association 55:708-713. heard, d. c. 1987. a simple formula for calculating the variance of products and dividends. government of the northwest territories. manuscript report. _____, k. l. zimmerman, g. s. watts, and s. p. barry. 1999. moose density and composition around prince george, british columbia, december 1998. final report for common land information base. project no. 99004. jolly, g. m. 1969. sampling methods for aerial censuses of wildlife populations. east african agricultural and forest journal 34:46-49. nielsen, s. e., c. j. johnson, d. c. heard, and m. s. boyce. 2005. can models of presence-absence be used to scale abundance? two case studies considering extremes in life history. ecography 28:197-208. poole, k. g., g. mowat, and d. pritchard. 1999. using gps and gis for navigation and mark-recapture for sightability correction in moose inventories. alces 35:1-10. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37:43-54. rea, r. v., and m. p. gillingham. 2001. the impact of the timing of brush management on the nutritional value of woody browse for moose alces alces. journal of applied ecology 38:710-719. shackleton, d. 1999. hoofed mammals of british columbia. royal british columbia museum and university of british columbia press, vancouver, british columbia, canada. spalding, d. j. 1990. the early history of moose (alces alces): distribution and relative abundance in british columbia. contributions to natural science #11. royal british columbia museum, victoria, british columbia, canada. thompson, i. d., and r. w. stewart. 1998. management of moose habitats. pages 377-401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washinga gis modified survey design – heard et al. alces vol. 44, 2008 116 ton, d. c., usa. timmermann, h. r., and m. e. buss. 1998. population and harvest management. pages 559-615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. walker, a. b. d., d. c. heard, v. michelfelder, and g. s. watts. 2006. moose density and composition in the parsnip river watershed, british columbia, december 2005. final report for the ministry of environment. project no. 2914568. _____, _____, j. b. ayotte, and g. s. watts. 2007. moose density and composition in the northern williston watershed, british columbia, january 2007. final report for the ministry of environment. project no. 2914568. unsworth, j. w., f. a. leban, e. o. garton, d. j. leptich, and p. zager. 1998. aerial survey: user’s manual. electronic edition. idaho department fish and game, boise, idaho, usa. 4306.pdf alces vol. 43, 2007 hurley et al. spatial analysis of mvc 79 a spatial analysis of moose-vehicle collisions in mount revelstoke and glacier national parks, canada michael v. hurley, eric k. rapaport, and chris j. johnson university of northern british columbia, natural resources and environmental studies graduate program, 3333 university way, prince george, bc, canada, v2n 4z9 abstract: moose (alces alces)-vehicle collisions (mvc) can be costly ecologically by affecting population numbers, economically by vehicle damage, and socially through human injury or mortality. the purpose of this paper is to identify factors related to moose ecology, driver behaviour, and road design that are useful for predicting the spatial location of mvc on the trans canada highway dismodels and used akaike’s information criteria (aic) to determine the most parsimonious model within geographic information system (gis). the receiver operator characteristic (roc) discriminated subsets. a mvc probability map along the highway was created using the gis model, providing a planning to reduce mvc risk within the parks should begin by assessing landscape-scale variables with emphasis on distance to wetland and landscape slope. this landscape-scale analysis should be predictors of moose tracks, game trails, and coniferous forest habitat. if highway planning cannot be effective in decreasing mvc, mitigation measures should include a public awareness program, speed reduction, and consideration of an alternative intercept foraging plan. alces vol. 43: 79-100 (2007) key words: driver visibility, evidence, gis, habitat, highway design, moose, mvc, roadside vegetation wildlife-vehicle collisions (wvc) are a serious problem in north america (bashore et al. 1985, child et al. 1991, del frate and spraker 1991, oosenbrug et al. 1991, romin nearly 3,000 moose-vehicle collisions occur annually in north america (child 1998) and 200 – 300 moose are killed on major british columbia highways each year (child et conservative and do not take underreporting into consideration or the unknown number of mortalities on mining, logging, and rural roads. if the impacts of trains are included, this numcolumbia (child et al. 1991). collisions can be costly ecologically by affecting population numbers, economically by vehicle damage and lost hunting opportunities, as well as socially through human injury and mortality. for predicting areas of high mvc. seiler (2005) stated that more detailed knowledge of occurrence of preferred moose forage (ball and dahlgren 2002, seiler 2005), embankment of the road (clevenger et al. 2003), and spatial analysis of mvc hurley et al. alces vol. 43, 2007 80 driver visibility (bashore et al. 1985) would increase the predictive power of past modelling attempts. seiler (2005) noted how new edge of the spatial distribution of collisions. malo et al. (2004) suggest that wvc models should be used at both the landscape and local scales during the process of road design and implementation of mitigation measures. the purpose of this paper is to predict the spatial occurrence of moose-vehicle collisions (mvc) along the trans canada highway through mount revelstoke and glacier national parks along with the associated correlated factors. mvc rates along this stretch of 0.045 per kilometre per year (sielecki 2004) for a total of 0.5 – 3 mvc per year within the parks. this mvc rate is relatively similar to outside of the park boundary; however, the reporting procedure within the park is more accurate for modelling purposes. the area is of high concern due to both the trans canada highway and wildlife having limited movement options through narrow and high mountain passes. in addition, parks canada has a management objective to reduce the environmental impact of the transportation corridor, particularly on wildlife, vegetation, tional parks. to predict mvc and determine the related process, models were developed using or using a geographic information system (gis) (finder et al. 1999, malo et al. 2004, were included based on their contribution to ecological processes, moose biology, and driver attributes. the predictive capability of subsets. the model with the best predictive representative model to be compared among local-scale model subsets included highway design, moose evidence, roadside vegetation management, moose habitat, and driver visibility. by predicting mvc locations, their reduction could be looked upon from a proactive perspective. by focusing on preventative measures as opposed to relying on mitigation measures, the implementation is not as costly, ecologically, economically, or socially. study area the study site was restricted to the trans canada highway dissecting glacier and mount revelstoke national parks within the rocky mountain highway district in south-eastern british columbia, canada (fig. 1). rugged, have resulted in limited transportation corridor operation of this segment of the trans canada highway therefore faces numerous challenges weather, slope, rock instability, and collisions parks canada is responsible for the planning, fig. 1. regional setting of glacier and mount revelstoke national parks alces vol. 43, 2007 hurley et al. spatial analysis of mvc 81 construction, and operation of the highway within the national park boundaries. glacier and mount revelstoke national parks encompass 3 biogeoclimatic zones, the interior cedar hemlock (ich), englemann spruce-subalpine fir (essf), and the alpine tundra zone (at). the ich is primarily comprised of old-growth cedar (thuja plicata) and mountain hemlock (tsuga mertensiana). in the essf, the lower subalpine forests are dominated by englemann spruce (picea engelmannii abies lasiocarpa), and mountain hemlock. mean annual precipitation is 700-3,000 mm, most of which (70 – 80%) falls as snow (meidinger and polar 1991). methods data collection mvc data were contributed by mount revelstoke and glacier national parks (john flaa, personal communication, parks canada). of each mvc was recorded by park wardens by either marking the collision on a map or by recording the collision location using a global positioning system (gps) unit. an assumption was made that the reporting system has the primary reporting method transformation was from map marking to gps use in the year 2000, representing 80% and 20% of mvc locations using each respective method. the utm co-ordinates were recorded in a database along with date of kill, hour of kill, and information regarding the number and species of wildlife. the utm coordinates for each mvc were plotted onto the highway layer within the study area using arcgis (esri 2005). the study encompassed a spatial analysis generated reference points so that logistic regression could be used to contrast highway points with and without mvc (fig. 2). reference points were created by randomly generating numbers that represented distances along the highway. road distances started at 0 km from the southern entrance of mount the northern entrance of glacier national park. random reference points that shared the coordinates with a snow shed were not included. changes in land cover due to natural or human disturbance over time were assessed using parks canada stand origin data. this lations could be studied between independent variable data collected in one season with mvc data spanning nearly 4 decades. both coniferous and deciduous cover has regenerated since the right of way was cleared for on the assumption does not warrant concern after highway construction. since highway natural disturbances have occurred within the 500 m highway buffer area since highway construction. fig. 2. topographical slope classes assessed at each point. the thick lines represent the highway and the thin lines represent the adjacent spatial analysis of mvc hurley et al. alces vol. 43, 2007 82 depending on accuracy and availability of spatial data, each variable was either mea(landscape scale). we chose variables based we used past studies and our knowledge of the study area to select potentially relevant landscape-scale variable analysis (gis) we used a gis to measure 15 landscapescale variables (500 m radius) (table 1). all continuous variables were averaged within the 500 m radius buffer centered on each collision and random point. a 500 m buffer around each location represented the road-effect zone effect zone is the area that encompasses the majority of ecological effects resulting from road construction and use and is typically the focus of planning and mitigation (forman 1999). a minimum of 500 m was kept between random reference points upon creation in order to ensure independence. the 500 m radii represented the area over which collision attributes were sampled using a gis at the landscape-scale. we used british columbia provincial government terrain resource information management (trim) spatial data in gis to represent highway segments, elevation, slope, and aspect. all trim data had a scale of 1:20,000 with a resolution of 25 m by 25 m cell size. topographical criteria were included due to the inherent nature of moose migration from hills to valleys during the winter (gundersen et al. 1998, hundertmark 1998). thus, measures of slope and aspect were included in an effort to gain insight into the effects of moose movement on mvc. the distance to water bodies and wetland were measured due to the fact that moose seek variable unit aspect (gis) mean aspect within 500 m buffer degrees built (gis) distance to the nearest human development m crossroad m elevation (gis) elevation above sea level generated using a digital elevation model m forest edge m hiking (gis) distance to the nearest hiking trail m high use habitat (gis) area of high moose habitat within 500 m buffer as per parks canada data m2 land cover (gis) dominant land cover type within 500 m buffer shrub/ coniferous/ lines (gis) distance to the nearest communication line m rail (gis) distance to the nearest railway line m risk sign (gis) distance to nearest wildlife-risk sign m slope (gis) mean slope within 500 m buffer degrees water (gis) distance to the nearest water body boundary m water int m wetland (gis) distance to the nearest wetland boundary m table 1. landscape-scale variables measured at each mvc site and reference point to model the factors that determine moose-vehicle collision locations within mount revelstoke and glacier national alces vol. 43, 2007 hurley et al. spatial analysis of mvc 83 (peek 1998). the distance to water and wetland trim data. we measured the presence/absence of high use habitat at each collision and reference point to determine the relationship of mvc with critical habitat range. the dominant land cover type was determined within a 500 m buffer to further assess habitat-related attributes and also potential effects on driver data within mount revelstoke and glacier national parks were based on parks canada scale of 1:50,000 (achuff et al. 1984). we used gis to record the distance from each mvc to rail lines, power lines, hiking trails, and built areas. the distance to rail and power lines was based on parks canada spatial data while the distance to built areas was rail lines are plowed in the winter, providing a potential movement corridor. in addition, the vegetation clearance within rail line and power line corridors creates the potential for the presence of early seral forage. the human development affects the occurrence of mvc by means of habitat alteration, human activity, and potential predator avoidance (malo et al. 2004, seiler 2005). hiking trails potential for increased movement, predation, and effect of human use on moose distribution. found moose to be more vulnerable to wolves at sites closer to trails and streams. distance of each mvc location from the nearest wildlife risk sign and highway curvature. we used the distance to wildlife risk sign criteria to assess the role of driver awareness on mvc. the distance to highway curvature was analyzed to assess driver visibility at a landscape-scale. spatial representation of gis model — using landscape-scale gis data, we developed a model with the structure: 0 1 1 k k) ————————————— 0 1 1 k k) where y is the predicted probability of a mvc k k (manly et al. 1993). the predictive mvc probability surface was created using local-scale variable analysis from june to august 2005, we collected data for local-scale analyses. we used a gps to locate each mvc and random reference sites scale variables (table 2). variables ranged from habitat related to driver and highway attributes, each contributing to one of the 5 local-scale model subsets. habitat — at each site, habitat characteristics were measured using a variety of methfig. 3. probability surface showing the likelihood of mvc for mount revelstoke and glacier national parks using the gis model. spatial analysis of mvc hurley et al. alces vol. 43, 2007 84 variable unit ang 5 m mean distance at which an observer standing 5 m from the pavement edge could no longer see passing vehicles taken from each direction on both sides of the highway m ang 10 m mean distance at which an observer standing 10 m from the pavement edge could no longer see passing vehicles taken from each direction on both sides of the highway m browse presence of browse within 100 m transect p/a browse (roadside) presence of browse within 25 m transect p/a corridor width width of highway corridor clearance including pavement m dist cover mean distance to vegetative cover (trees and shrubs >1 m high) taken from both sides of the road m ditch presence of ditch adjacent highway p/a ecotone presence of an ecotone p/a game trail absent/low/high a/l/h habitat class (mf)/coniferous forest(cf)/wetland(w)/shrub(s) ofm/cf/w/s inline mean distance at which an observer standing at the pavement edge could no longer see passing vehicles taken from each direction on both sides of the highway m jersey barrier presence of jersey barrier p/a median presence of median p/a passing lane presence of a passing lane p/a pellets presence of pellets within 100 m transect p/a pellets (roadside) presence of pellets within 25 m transect p/a roadside age class highest age of shrub within 25 m transect (1-3 yrs) (7-10 yrs) roadside vegetation type of vegetation species within 25 m transect p/a slope (0-5 m) mean slope of the land 0-5 m perpendicular to the pavement edge taken from both sides of the road degrees slope (5-10 m) mean slope of the land 5-10 m perpendicular to the pavement edge taken from both sides of the road degrees slope (10-30 m) mean slope of the land 10-30 m perpendicular to the pavement edge taken from both sides of the road degrees speed mean recorded speed of passing vehicle km/h topo terrain slope category tracks presence of tracks within 100 m transect p/a tracks (roadside) presence of moose tracks within 25 m transect p/a table 2. local-scale variables measured at each site and reference point to model the factors that determine moose-vehicle collision locations. alces vol. 43, 2007 hurley et al. spatial analysis of mvc 85 attractants of moose to highway corridors. we placed 25 m transects perpendicular to the highway and measured plant species presence and age at 5 m intervals within 4 m2 determine the most recent year of roadside clearing. the highest age of a shrub within the 25 m transect was used as an indicator of time since the roadside was cleared. we also recorded evidence of browsing, moose tracks, some roadside vegetation species were grouped into families due to their low occurrence. western mountain ash (sorbus scopulina) and saskatoon berry (amelanchier alnifolia) were grouped into the rose family. narrow-leaved hawkweed (hieracium umbellatum), common dandelion (taraxacum ), pearly everlasting (anaphalis margaritacea), yarrow (achillea millefolium), leucanthemum vulgare) were that were rarely present (1 – 3 occurrences) and could not be grouped into a family were roadside vegetation was modelled at the species level (table 3). we placed a 100 m transect perpendicular and assess the roadway for presence of moose. coniferous forest (cf), wetland (w), or shrub (s). we recorded the dominant land cover class at 10 m intervals on the transect. evidence of moose included wildlife trails, pellets, tracks, or browse. if the highway bisected two habitat types, this ecotone was noted. ecotone was used as a variable to investigate any habitat edge effect that could potentially be correlated with mvc. the distance to the nearest forest edge perpendicular to the road was measured the distance to crossroads and water bodies intersecting the road were also measured in the same manner. the distance to crossroads was tested to determine whether intersections opportunity of a collision. human and wildlife movement — we recorded a number of highway attributes that the ability of drivers to avoid a mvc. we used an inclinometer to measure the slope immediate to the roadbed (0 – 5 m), the verge (5 – 10 m), and the adjacent land (10 – 30 m). we and topographic measurements tested whether embankments had positive or negative relationships with moose-vehicle collisions. driver visibility — driver visibility was measured as the shortest distance to the point at which a car becomes out of sight of an observer from 3 different locations adjacent the highway. field visibility variables meamoose on the right-of-way. since it could species modelling name common horsetail (equisetum arvense) horsetail grass grass willow (salix sp.) willow red-osier dogwood (cornus stolonifera) dogwood sitka alder (alnus crispa) alder western red cedar (thuja plicata) cedar spruce (picea sp.) spruce thimbleberry thimbleberry common red paintbrush (castilleja miniata) paintbrush black twinberry (lonicera involucrata) twinberry spreading dogbane (apocynum androsaemifolium) dogbane lupine (lupinus sp.) lupine aspen (populus tremuloides) aspen table 3. roadside vegetation species present within spatial analysis of mvc hurley et al. alces vol. 43, 2007 not be determined from what side or which direction a vehicle struck an animal, 4 visibility measurements were taken at each site, 2 facing each direction, on each side of the highway. one in-line (from road edge) and 2 angular measurements were measured (5 m and 10 m from the road edge). recognising that trucks were more visible at greater distances than cars or motorcycles, visibility distances were always measured using trucks. the mean distance to vegetative cover (trees and shrubs > 1 m high) was measured on both sides of the road to determine driver visibility. the corridor width was the total area cleared for the highway including a combination of roadside clearance on both sides of the highway and the highway pavement width. the presence/absence of roadside ditches was reand animal movement. the presence/absence of jersey barriers, passing lanes, and medians resulting from highway design and construction. the average speed limit was read by means of a bushnell radar gun. highway speed was recorded as the mean of 20 vehicles (10 vehicles going in each direction). actual vehicle speed was recorded as opposed to speed limit due to the inherent nature of vehicles not included in model development due to the absence of variability within the study area. all distances were measured using a range ence/absence and continuous/discontinuous variables were estimated visually. data analysis due to the binary nature of the dependent variable (0 = reference, 1 = collision), and the inclusion of categorical independent variables, the data were analyzed using bivariate logistic regression. the variables were grouped into and akaike’s information criteria (aic) was used to determine the most parsimonious model within each subset. the use of model selection criteria enabled inference to be drawn from several models simultaneously, so that a ‘best set’ of similarly supported models could be chosen (johnson and omland 2004). we isolated, understood, and adapted to mitigation strategies. five subsets modelled local-scale/ ined gis landscape-scale hypotheses. the that affected the driver visibility of moose. the second subset included the variables that indicated the evidence of moose in the terrain perpendicular to the highway. highway design was assessed in the third subset. the fourth and age in order to relate mvc to roadside local-scale subset tested moose habitat features completed among the best aic local-scale in order to identify the most parsimonious model overall. this round of aic did not include the landscape-scale gis models in its comparison due to the difference in scale relative to the 5 local-scale models. variables grouped into common hypothesized subsets, 2 combination models were developed to help further reveal the mvc phenomena. we recognise that these interaction models were not initial hypotheses, but arose as model subsets. variables chosen for interactions included those that previously showed and lemeshow 2000). to reduce multicollinearity among the modelled variables (zar 1998), correlation screening was completed prior to model dealces vol. 43, 2007 hurley et al. spatial analysis of mvc 87 which compared each variable combination, and removed those that were highly correlated (r > 0.75) (seiler 2005). in the gis model subset, the distance to communication lines was omitted from further analysis as it was highly correlated (pearson correlation coefshowed a lower correlation with mvc points posed to 0.49). in the driver visibility model subset, angular visibility 5 m was eliminated as it was highly correlated with inline visibility. angular visibility 5 m was chosen to be eliminated as opposed to inline as angular visibility 5 m is measured in between inline and angular visibility 10 m thus providing a larger range of measurements. in addition, inline is also taken from the road edge closer to where a collision occurs. also in the driver visibility model subset and the highway design model subset, slope (5 – 10 m) was highly correlated with slope (0 – 5 m). to provide a greater range of slope measurements, slope (5 – 10 m) was eliminated as it is intermediate to the other two slope measurements (0 – 5 m and 10 – 30 m). contribution that a unit increase in the independent variable made to the outcome probof the individual independent variables. we variables on the collision probability. each topographic and distance variable was modelled as a simple linear and then a for further model comparisons if the more relative to the simple linear form. for the gis further modelling for the 3 topographic variables of elevation, slope, and aspect. for the for inline visibility and angular visibility at 10 m were included for further modelling. we used the change in deviance to assess amine high-leverage points which may have the 3 points with the highest leverage were investigated to determine the location in the parks, and the corresponding change in coefboth statistical and biological consideration, the points remained in the model as 95% of the cases were within +/2 (menard 2001). autocorrelation had to be corrected, as (neilsen et al. 2002). autocorrelation was assessed using passage by calculating the moran’s i using the unstandardized model residuals and distance between points. robust standard errors were estimated using the huber/white sandwich estimator in the program stata (2002) to correct for autohuber/white sandwich estimator is robust decreased the potential for type i errors by levels (lennon 2000). model validation the receiver operating characteristic (roc) was used to determine the degree of that is independent of probability cut-off levels (boyce et al. 2002). roc validation was developed using independent data not included during model creation. twenty percent of validation. to represent the variance associated with the process of choosing validation data, we repeated the roc procedure 5 times. each iteration used a different set of randomly selected collision and reference points. this validation procedure was followed for each spatial analysis of mvc hurley et al. alces vol. 43, 2007 88 results aic model comparison the use of aic in model comparison showed selection uncertainty being within models in certain subsets meaning a small difference in performance. this development be interchanged as the model of choice were eses often only differing in one variable to assess whether that certain factor is of critical importance to the susceptibility of mvc. driver visibility model subset — of the 10 driver visibility candidate models, the vided support as the most parsimonious with an aicw included the variables of vehicle speed, corridor width, and presence/absence of passing lanes. adding variables of roadside slope or visibility distance to this model did not contribute to the aicw (aicw = 0.283 and 0.224, respectively). the aicw for the additional the odds of mvc (table 5). corridor width visibility models; mvc were more likely with increasing corridor widths. gis model subset — the topographic models within the gis subset (aicw = 0.537) variables included slope, aspect, and elevation while water bodies included lakes, rivers, and wetland model hypothesis resulted in w (aicw = 0.299) while w w such as the human built model using variables of hiking trails, distance to rail, and distance to built area. ence was found with mvc being correlated to to mvc in the gis/driver visibility model but not the gis model alone included elevation and aspect. the distance to wetland had closer to wetland. the gis model produced roadside vegetation model subset — of the roadside vegetation models, the forage species hypothesis had the greatest aicw, although the weight was only 0.504, suggesting hypothesis/model variables -2ll aic aicw speed + passing lane + corridorwidth 4 132.7 140.71 0.44 adjacent roadside slope and slope(0-5m) + slope(10-30 m) + passing lane + speed + corridorwidth 129.53 141.53 0.28 and visibility inline + inline2 + ang10 + ang102 + passing lane + speed + corridorwidth 8 125.93 141.93 0.22 table 4. results of driver visibility aic candidate model selection within mount revelstoke and glacier national parks. alces vol. 43, 2007 hurley et al. spatial analysis of mvc 89 considerable uncertainty in model selection (table 8). variables included in this model as reported in the literature. shrub age alone or when combined with forage species did model hypotheses were not included in table w. one of based on non-forage species with an aicw of 0.041. within the roadside vegetation model, the presence of grasses was positively correlated to mvc sites, while the presence of a kill (table 9). moose habitat model subset — the land cover type hypothesized model was the most parsimonious of the moose habitat candidate models (aicw = 0.479) (table 10). the addition of the distance to water intervariable s.e. (robust) w p (robust) speed* 0.05 10.05 0.00 corridor width* 0.05 0.02 0.03 passing -0.13 0.49 0.07 0.80 constant 4.7 12.09 0 table 5. logistic regression analysis results for the best driver visibility aic model. *p hypothesis/model variables -2ll aic aicw water bodies elevation + elevation2 + slope + slope2 + aspect aspect2 + wetland + water 9 44.08 0.54 wetland wetland + elevation + elevation2 + slope + slope2 51.35 0.3 moose movement elevation + elevation2 + slope +slope2 + aspect + aspect2 + rail 9 parks. variable s.e. (robust) w p (robust) wetland* -0.00 0.00 9.80 slope* -1.05 0.30 7.44 slope2* 0.02 0.01 0.00 aspect2* 4 4 4.31 0.04 elev2 5 3.309 0.058 aspect -0.085 0.0479 elev 0.055 0.034 2.475 0.112 water 0.001 0.003 0.135 0.707 constant 0.958 0.004 0.953 table 7. logistic regression analysis results for the best gis aic model. *p spatial analysis of mvc hurley et al. alces vol. 43, 2007 90 sections to this land cover model decreased the aicw (aicw = 0.441). the remainder of the candidate hypotheses all had aicw under alone resulted in an aicw of 0.004. coniference on the odds of a mvc within the moose habitat model (table 11). moose evidence model subset — the aicw was 0.529 for the trails and transect evidence hypothesized model, providing support as the most parsimonious of the moose evidence candidate models (table 12). this model included moose evidence within the 100 m transect as well as the presence/absence of game trails. the candidate models with only trails (aicw = 0) or only transect evidence (aicw = 0.048) performed poorly on their own and were not included in table 12. the inclusion of roadside tracks, browse, of the best model (aicw = 0.315) nor were the roadside variables effective predictors on their own (aicw = 0). evidence of moose was positively correlated with mvc sites with the presence of tracks being the most important, followed by the presence of game trails (table 13). this best aic moose evidence model of trails and transect evidence correctly clashighway design model subset — the comparison of the 9 highway design candidate models resulted in the highway corridor (aicw = 0.553) (table 14). the full model, which included the additional variable of distance to crossroad, was no more parsimonious hypothesis/model variables -2ll aic aicw forage species willow + dogwood + alder + cedar + aspen + horsetail + grass + spruce + rose 10 139.09 159.09 0.5 forage species and shrub age roadside ageclass + willow + dogwood + alder + cedar + aspen + horsetail + grass + spruce + rose 11 138.48 0.25 shrub age roadside ageclass 2 157.4 0.17 table 8. results of roadside vegetation aic candidate model selection within mount revelstoke and glacier national parks. variable s.e. (robust) w p (robust) grass* 1.08 0.52 5.19 0.04 alder -0.98 0.52 4.05 spruce 1.15 3.70 0.07 horsetail -0.88 0.47 dogwood 0.22 willow 0.73 1.41 0.23 rose -0.31 0.54 0.41 0.57 cedar 1.03 0.37 0.59 aspen -0.14 0.48 0.08 0.78 constant -0.91 0.14 table 9. logistic regression analysis results for the best roadside vegetation aic model. *p alces vol. 43, 2007 hurley et al. spatial analysis of mvc 91 hypothesis/model variables -2ll aic aicw land cover type cf + ofm + wetland + shrub 5 141.53 151.53 0.48 and land cover type ofm + cf + shrub + wetland + waterint 0.44 full model ecotone + forestedge + waterint + cf + ofm + wetland + shrub 8 table 10. results of moose habitat aic candidate model selection within mount revelstoke and glacier national parks. variable s.e. (robust) w p (robust) coniferous forest* 0.04 0.01 0.00 shrub -0.04 0.02 3.42 0.05 wetland 0.04 0.03 2.71 0.18 0.00 0.01 0.05 0.80 constant -0.97 0.81 1.22 0.23 table 11. logistic regression analysis results for the best moose habitat aic model. *p hypothesis/model variables -2ll aic aicw trails and transect evidence pellets 5 94.43 104.43 0.53 full model browseroad 7 91.4 105.42 0.32 roadside evidence and transect evidence 95.73 107.73 0.1 table 12. results of moose evidence aic candidate model selection within mount revelstoke and glacier national parks. variable s.e. (robust) w p (robust) tracks* 1.89 0.00 pellets 2.47 5.53 0.13 trail trail(high)* 1.33 2.04 0.02 trail(low) 0.21 0.04 0.88 browse 0.59 1.95 0.11 constant -2.93 0.50 4.91 table 13. logistic regression analysis results for the best moose evidence aic model. *p spatial analysis of mvc hurley et al. alces vol. 43, 2007 92 (aicw = 0.215). the hypothesis that variables associated with moose movement resulted in a model with a lower aicw (aicw = 0.144). the additional highway design hypotheses modelling smaller variable groupings were all under aicw of 0.1 and not included in the model under an aicw of 0.1 was using the variables of topographic class, slope, and presence of ditches. corridor width displayed a and the highway design models (tables 5 and 15, respectively). in each model, mvc were more likely with increasing corridor widths. the highway design model showed the poorest performance among the model subsets with interaction models — combined gis and driver visibility models to table 14. results of highway design aic candidate model selection within mount revelstoke and glacier national parks. hypothesis/model variables -2ll aic aicw highway corridor engineering topo + slope(0-5 m) + slope(10-30m) + median + jersey + passing lane + corridorwidth + ditch 9 130.89 148.89 0.58 full model topo + ditch + slope(0-5 m) + slope(1030m) + median + jersey + passing lane + crossroad + corridorwidth 10 130.8 150.84 0.22 moose movement topo + slope(0-5 m) + slope(1030 m) + crossroad + jersey + corridorwidth + ditch 8 135.71 151.71 0.14 table 15. logistic regression analysis results for the best highway design aic model. *p variable s.e. (robust) w p (robust) corridor width* 0.03 10.05 0.03 passing lane 0.75 0.49 2.07 0.13 slope (0-5 m) -0.03 0.02 1.85 0.13 median 1.02 1.50 ditch -0.29 0.53 0.34 0.58 jersey barrier 0.20 0.47 0.14 slope (10-30m) 0.00 0.01 0.01 0.93 topo 8.87 topo(2b) 0.57 1.24 0.51 topo(3a) -1.35 0.87 -1.59 0.12 topo(3b) -1.57 0.90 -1.7 0.08 topo(3c) 0.91 1.31 0.81 0.49 topo(4) 0.35 0.91 0.41 0.70 topo(5a) -0.42 0.89 -0.44 topo(5b) -0.41 0.98 -0.44 -1.21 1.18 -1.03 0.30 constant -2.24 1.37 -1.94 0.10 alces vol. 43, 2007 hurley et al. spatial analysis of mvc 93 corridor width, wetland-speed, and wetlandcorridor width were included as interactions. in the gis/driver visibility interaction model, with greater speeds. when gis was combined with driver visibility, the interaction model was lower than gis alone, yet still impressive, correctly classifying 92.4% of points. the second combination model included variables from the moose habitat and driver visibility models. interaction terms consisted of coniferous forest with both speed and highway corridor width (table 17). no factors were habitat/driver visibility interaction model. when driver visibility was combined with moose habitat, the interaction model had a higher roc score than the driver visibility model alone, yet was still poor; only correctly roc validation nation ability as reasonable and rates higher than 90% as very good discrimination because the sensitivity rate is high relative to the false positive rate. using this 70% as a minimum threshold, the acceptable models after roc validation in descending order include gis, gis + driver visibility, moose evidence, and moose habitat. highway design, roadside variable s.e. (robust) w p (robust) -0.02 0.01 0.02 elev2 -5 -5 5.513 0.050 aspect2 -4 -4 5.343 0.008 elev 0.138 0.077 5.073 0.072 slope2 0.023 0.012 4.098 0.053 aspect -0.129 0.053 3.920 0.015 0.008 0.004 3.537 0.053 -5 2.825 -5 0.381 water -0.002 0.003 0.127 passing -0.047 0.002 constant -27.371 0.345 *p variable s.e. (robust) w p (robust) wetland 0.044 0.028 3.395 0.124 shrub 0.022 2.333 0.102 0.001 2.215 0.117 passing -0.511 0.438 1.327 0.243 0.879 0.307 0.009 0.014 0.345 0.520 constant -1.254 0.900 1.573 0.050 table 17. logistic regression analysis results for the best moose habitat/driver visibility interaction aic model. *p spatial analysis of mvc hurley et al. alces vol. 43, 2007 94 vegetation, driver visibility and the moose habitat/driver visibility models were below test among the best local-scale model from each of the 5 subsets strongly supported the moose evidence model as the most parsimonious (aicw = 1.0), adding further support to its mvc predictive model. discussion model performance although roc scores for the gis, gis and driver visibility interaction, moose evireasonably high discrimination, results should be interpreted with caution. as the study area is within a national park, the land processes outside park boundaries, such as forestry, close to the park entrance over those near the centre of the park. as the study area is situated within both a high mountain pass and a protected area, the transportation challenges model results should therefore not be directly to be used elsewhere, the structure could be priately adapted to the location and species. we assumed that land-use remained constant over the reporting period, although minor changes were most likely inevitable despite the national parks having been managed in a relatively constant ecological state. additional caution should be used when interpreting these models as not all of the collisions that have occurred in the past were reported. the total number of collisions involving motor vehicles and large animals in canada has generally been underestimated by of these reporting discrepancies include the unknown taking of carcasses before highway contractors are alerted, carcasses falling out of sight or animals moving away to die at unknown locations. in addition, drivers may report the collision to another jurisdiction or fail to report a minor collision, instead paying for the damages privately (sielecki 2004). variation present among the models can be factors. the models were developed using the sures previously shown to have successfully et al. 1999, clevenger et al. 2003, malo et al. 2004). notwithstanding the inclusion of adbe due to one simple factor such as weather, driver alertness, or moose behaviour. there is a possibility that the inclusion of mount revelstoke mvc in the overall model afdistance gap between parks. the two parks model roc validation s.e. aicw gis 0.035 n/a gis + driver visibility 92.4% n/a moose evidence 1.0 moose habitat 70.2% 0.115 0 moose habitat + driver visibility 0.117 0 driver visibility 0.12 0 roadside vegetation 59.2% 0.123 0 highway design 0 table 18. model roc validation results on the best aic model from each subset. alces vol. 43, 2007 hurley et al. spatial analysis of mvc 95 do, however, share ecosystem characteristics and are managed under one division of parks canada. in addition, the spatial error in reporting mvc locations may have affected measurements based on the assumption that locations were accurate. interpretation of contributing factors relationship with mvc in the driver visibility model. higher speeds leading to a greater vides support to the literature, although seiler (2005) and malo et al. (2004) modelled speed limit as opposed to actual radar speed. the to mvc in both the driver visibility model and the highway design model, although these 2 models were poor predictors overall. mvc sites were found at highway locations with greater corridor width than reference sites. clevenger and waltho (2000) found that wildlife use of highway passages was positively correlated with road width. improved visibility due to greater vegetation clearance may not have displayed importance as the bulk of accidents in the 2 parks occurred at correlation of mvc with distance to road curve, inline visibility, and angular visibility. to a decrease in vehicle speed while joyce and mahoney (2001) found more mvc at night due to increased moose activity. furthermore, roadside brushing likely augments the risk of collision by maintaining early seral vegetation, which attracts wildlife to the highway (child et al. 1991, rea 2003). other studies have provided support for animals preferring to cross highways that are closer to vegetation cover (jaren et al. 1991, clevenger et al. 2003, malo et al. 2004, seiler 2005). these contradicting theories of increased visibility and increased moose attraction may have led to the poor predictive abilities of the driver visibility and highway design models. the positive correlation between a wider highway corridor width and mvc may simply be a function of the highway being reduced to narrow widths along steeper sections. both corridor width and speed no longer showed seiler (2005) where the distance between correlated to mvc; however, if vehicle speed was weakened. coniferous forest as a single variable in the habitat model was, however, to be an important habitat type, with moose use ranging from 31 – 49% use per season in cant contributor to the habitat model. perry forest to be of slightly less important moose addition, moose avoid wolves by spacing out and pletscher 2000). model was observed in the slope variable which to mvc in previous studies (gunson et al. et al. (2003) found that mammals were more likely to cross when the highway was level with the adjacent terrain. where the two national parks are within the selkirk mountain range, valley corridors and limited gentle sloping landscapes. snow accumulation is less in the valley bottoms, providing important ungulate habitat in the late autumn, winter, and early ged mountain terrain forces both wildlife and human movement through the valley passes spatial analysis of mvc hurley et al. alces vol. 43, 2007 mvc within mount revelstoke and glacier national parks have occurred in winter months providing support for this theory. correlation to mvc within the gis model whereas the distance to water did not. moose (peek 1998). the distance from water to mvc locations may not show a correlation simply due to the general fact that there are lakes and rivers dispersed throughout the parks and not in one particular area. the poor prediction ability of the roadside vegetation model may be attributed to a relatively homogeneous highway corridor throughout the 2 parks. moose are browsing specialists with 90% of average diets being shrubs and trees (perry 1999). many of the preferred shrub species for moose were relatively common at both mvc and reference locations. the presence of grass was the vegetation model and this may have been due to the overall scarcity of grasses in the steeper, higher elevation reference point locations, instead being more prevalent within most likely did not contribute to the aicw in roadside vegetation candidate models due to the majority of roadside shrubs being the entire park. moose tracks and high-use game trails 100 m transect perpendicular to the highway on can be a simple indicator of mvc locations. roadside moose evidence was not included in w was not improved after inclusion in the full model or on its own. roadside evidence may not have improved the model due to the presence of roadside browsing at the majority of both mvc locations (89%) and reference points (75%). scale-dependent factors step in assessing contributing variables within this landscape-scale/gis approach shows may have shown less predictive ability than the landscape-scale model, but were nevertheless and revealing factors important at both scales of analysis. for this reason, we created the gis and driver visibility interaction model; although, the roc score for this interaction model was no higher than that of the gis model on its own. although the moose habitat and moose evidence models suggested that habitat was a strong predictor of mvc, the distance to high use habitat and land cover variables in the gis model subset were not present in the difference in predictability between the different models seems to be a scale-dependant issue where local effects within 100 m such as forest type and moose evidence are more prodirect habitat variables. often, availability of the actual use of the habitat is restricted to one scale (johnson et al. 2002). in addition, use habitat or land cover type might not have been selected for by moose and if so it may be so only at certain times of the year, thus introducing a temporal aspect to the model. joyce and mahoney (2001) suggest that mvc occur in areas of low and high moose density. 1937). predictions from an anthropogenic concept states that animals have programmed alces vol. 43, 2007 hurley et al. spatial analysis of mvc 97 neurohormonal cues in how the environment is interpreted which can be species, gender, social, or season dependant (bubenik 1998). the models were created using variables stemming from an anthropogenic perspective, however, human impressions on where moose should live do not ultimately determine where a moose will be. an opposite scale-related phenomenon may have occurred within the highway design model where the poor predictive ability may be attributed to the local-scale variables being overshadowed by landscape-scale factors. the class variable may not have been large enough topographic factors as seen in the gis model. linear landscape elements such as riparian corridors, ditches, steep slopes, and ridges may funnel animals alongside or across the roadway and thereby increase the risk of collisions (malo et al. 2004, seiler 2005). the importance of highway corridor width decreased when combined with landscapescale factors of slope and wetland in the gis/ driver visibility interaction model. the speed and slope interaction variable did, however, different scales were combined, suggesting mvc are correlated to locations with higher vehicle speeds and lower slope values. management implications gis is a powerful tool in the initial where local-scale mitigation measures are needed. if the need for local-scale analysis should be modelled due to their reasonably high predictive abilities. attention should be focused on highway segments close to wetcorridors, presence of coniferous forest, moose evidence, and at higher vehicle speeds. improved road planning is the primary practice that should be regarded as the means to reduce the ecological effects that transport infrastructure impose. this study has helped observe some of the underlying processes that contribute to mvc within the parks. the trans canada highway in mount revelstoke and glacier national parks, is a well established transportation route and mitigation measures will be the only option unless road alteration or new construction occurs. although the processes within the predictive models are best suited for highway planning, the knowledge can be used as a basis for mitigation decisions. an effective and acceptable countermeasure should reduce animal-vehicle interactions while still allowing for necessary animal behaviour and movements (bashore et al. 1985). suggested measures include reductions in vehicle speed and intercept foraging. additional enforcement which can be costly. intercept foraging involves the development of alternative feeding sites away from the transportation corridor (schwartz and bartley 1991). wood and wolfe (1988) determined that intercept foraging was an effective shorthowever, they cautioned that wildlife may become dependant on the supplemental food resulting in the attraction of additional wildlife. a fencing and wildlife underpass combination could be effective along the highway adjacent the beaver river. whenever possible, these a public awareness program such as the wildlife collision prevention program in british columbia. complete reliance should not be put into educational programs to enhance public awareness about wvc as their success has not yet proven effective (romin and bisbe a starting point. the models presented here may provide useful tools for road planners, but effective spatial analysis of mvc hurley et al. alces vol. 43, 2007 98 concrete approach that includes consideration of the landscape outside of park boundaries and more in-depth knowledge of the local work would be to investigate actual moose movement in the study area using telemetry data to map key crossing points. these data in combination with the collision points and modelling could provide invaluable informain the national parks. acknowledgements this study could not have been conducted moral support of maya dougherty. mount revelstoke and glacier senior park warden john flaa went above and beyond his call of duty in providing spatial and collision data, not to mention the generous in kind support theoretical support by our companions tony improved the manuscript. gis support from scott emmons and ping bai facilitated the analysis process. references achuff, p. l., w. d. holland, g. m. coen, and k. van tighem. 1984. ecological land classification of mount revelstoke and glacier national parks, british columbia. volume 1 integrated resource description. alberta institute of pedology. publication number m-84-11. edmonton, alberta, canada. ball, j. p., and j. dahlgren. 2002. browsing damage on pine (pinus sylvestris and p. contorta) by a migrating moose (alces alces) population in winter: relation to habitat composition and road barriers. scandinavian journal of forest research 17:427-435. bashore, t. l., w. m. tzilkowski, and e. d. bellis. 1985. analysis of deer-vehicle collision sites in pennsylvania. journal of bifulco, r., and h. f. ladd choice, racial segregation and test-score gaps. proceedings from the annual meeting of allied social science associations. boston, massachusetts, usa. boyce, m. s., p. r. vernier, s. e. nielsen, and f. k. a. schmiegelow. 2002. evaluating resource selection functions. ecological modelling 157:281-300. bubenik, a. b. 1998. behavior. pages 173-221 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. child, k. n. 1998. incidental mortality. pages 275–285 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, s. p. barry, and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27:41-49. clevenger, a. p., b. chruszcz, and k. e. gunson. 2003. spatial patterns and factors influencing small vertebrate fauna road-kill aggregations. biological conservation 109: _____, and n. waltho. 2000. factors influencing the effectiveness of wildlife underpasses in banff national park, alberta, canada. conservation biology damas, and smith. 1982. wildlife mortality in transportation corridors in canada’s national parks, volume i and ii. parks canada, ottawa, ontario, canada. del frate, g. g., and t. h. spraker. 1991. moose vehicle interactions and an associated public awareness program on the alces vol. 43, 2007 hurley et al. spatial analysis of mvc 99 (esri) environmental systems research institute. 2005. arcmap gis version 9.1. esri inc., redlands, california, usa. finder, r. a., j. l. roseberry, and a. woolf. 1999. site and landscape conditions at white-tailed deer/vehicle collision locations in illinois. landscape and urban planning 44:77-85. forman, r. t. t. 1999. horizontal processes, roads, suburbs, societal objectives, and landscape ecology. pages 35-53 in j.m. scape ecological analysis: issues and applications. springer-verlag incorporated, new york, new york, usa. _____, and l. e. alexander. 1998. roads and their major ecological effects. annual review of ecology and systematics 29:207-231. gunderson, h., h. p. andreassen, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34:385-394. gunson, k. e., b. chruszcz, and a. p. clevenger scape and highway influence ungulate vehicle collisions in the watersheds of the central canadian rocky mountains: a fine-scale perspective? proceedings from the international conference on ecology and transportation 2005. san diego, california, usa. hosmer, d. w., and s. lemeshow. 2000. applied logistic regression. second edition. john wiley and sons incorporated, new york, new york, usa. huber, p. j likelihood estimates under nonstandard conditions. proceedings of the fifth berkeley symposium on mathematical statistics and probability 1:221–223. university of california press, berkeley, california, usa. hundertmark, k. j. 1998. home range, dispersal and migration. pages 275-285 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c.,usa. jaren, v., r. andersen, m. ulleberg, p. h. pedersen, and b. wiseth. 1991. moosetrain collisions: the effects of vegetation removal with a cost-benefit analysis. alces 27:93-99. johnson, c. j., k. l. parker, d. c. heard, and m. p. gillingham. 2002. movement parameters of ungulates and scale-specific responses to the environment. journal of animal ecology 71:225–235. johnson, j. b., and k. s. omland. 2004. model selection in ecology and evolution. trends in ecology and evolution 19:101-108. joyce, t. l., and s. p. mahoney. 2001. spatial and temporal distribution of moosevehicle collisions in newfoundland. wildlife society bulletin 29:281-291. kunkel, k. e., and d. h. pletscher. 2000. habitat factors affecting vulnerability of moose to predation by wolves in southeastern british columbia. canadian journal of zoology 78:150–157. leblanc, y., f. bolduc, and d. martel upgrading a 144 km section of highway in prime moose habitat: where, why, and how to reduce moose-vehicle collisions. proceedings from the international conference on ecology and transportation 2005. san diego, california, usa. lennon, j. j. 1999. resource selection functions: taking space seriously. trends in ecology and evolution 14:399–400. malo, j. e., f. suarez, and a. diez. 2004. can we mitigate animal-vehicle accidents using predictive models? journal of applied ecology 41:701-710. manly, b. f. j., l. l. mcdonald, and d. l. thomas. 1993. resource selection by animals: statistical design and analysis for field studies. chapman & hall, meidinger, d., and j. polar. 1991. special spatial analysis of mvc hurley et al. alces vol. 43, 2007 100 report february 1991. research branch, victoria, british columbia, canada. menard, s. 2001. applied logistic regresapplied logistic regression analysis. sage publishing, thousand oaks, california, usa. neilsen, s. e., m. s. boyce, g. b. stenhouse, and r. h. m. munro. 2002. modeling grizzly bear habitats in the yellowhead ecosystem of alberta: taking autocorrelaoosenbrug, s. m., e. w. mercer, and s. h. ferguson. 1991. moose-vehicle collisions in newfoundland management considerations for the 1990’s. alces 27:220-225. peek, j. m. 1998. habitat relationships. pages 275-285 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. perry, j., editor. 1999. moose, mule deer research partnership. file report 99-5. rea, r. v. 2003. modifying roadside vegetation management practices to reduce vehicular collisions with moose alces alces. wildlife biology 9:81-91. romin, l. a., and j. a. bissonette deer-vehicle collisions: status of state monitoring activities and mitigation measures. wildlife society bulletin schwartz, c. c., and b. bartley. 1991. reducing incidental moose mortality: considerations for management. alces 27:227-231. seiler, a. 2005. predicting locations of moose–vehicle collisions in sweden. journal of applied ecology 42:371–382. sielecki, l. 2000. wildlife accident reporting system annual report. british columbia ministry of transportation, victoria, british columbia, canada. swets, j. a. 1988. measuring the accuracy of diagnostic systems. science 240:1285–1293. tabachnick, s. c., and l. s. fidell. using multivariate statistics, third edition. harper collins college publishers, new york, new york, usa. von uexkull, j. 1921. umwelt und innenwelt der tiere. second edition. springer verlag, berlin. _____. 1937. umweltforschubg. zeitschrift für tierpsychologie. 1:33-34. white, h. 1980. a heteroskedasticityand a direct test for heteroskedasticity. econometrica 48:817-830. wood, p., and m. l. wolfe. 1988. intercept feeding as a means of reducing deervehicle collisions. wildlife society bulwoods, j. g., and r. h. munroe rails and the environment: wildlife at the intersection in canada’s western mountains. proceedings from the transportation related wildlife mortality seminar, orlando, florida. usa. zar, j. h. 1998. biostatistical analysis. international editions. prentice-hall, new jersey, new york, usa. alces37(1)_109.pdf alces37(2)_403.pdf alces vol. 45, 2009 volokh – history and status of moose in ukraine 5 history and status of the population dynamics of moose in the steppe zone of ukraine anatolii m. volokh taurian state agrotechnological university, department of ecology and environmental protection. b. khmelnitsky street 18, melitopol, ukraine 72312. abstract: the moose (alces alces) population in the steppe zone of ukraine developed initially in 1955-1965. early annual population growth rates were high ranging from 13-49% partly due to immigration of moose from russia and byelorussia. however, after fully occupying forest habitat and expanding to treeless biotopes, reproductive efficiency declined. this decline was influenced by large spatial isolation of suitable habitat; 52% of solitary males and 48% of solitary females were in isolated biotopes during breeding. further, 8% of adult bulls were in herds without cows, and 21% of cows were in herds without bulls. although individual productivity was good, 1.3±0.1/pregnant cow and 0.4/ adult cow, isolation caused low participation in breeding (38.5% of adult cows), low number of calves (17.1% of population), and low annual population growth rate (≈ 6%/yr). the steppe moose population reached its maximum (n = 2776) in 1974 followed by steep decline; the decline was associated with harvests of 16.1% in 1973 and 12.5% in 1974, of which about 50% were adult animals. the population reached a second peak (n = 2147) in 1982 and declined gradually until 1992. a steep, annual population decline of 25.3±5.8% occurred after 1992; this decline was associated with excessive harvest beyond the annual population growth rate. moose were extirpated from most regions of the steppe zone by the late 20th century. the current southern range of moose is limited to forest habitat, and except for a remnant population of about 80, the unique steppe population has disappeared from ukraine. alces vol. 45: 5-12 (2009) key words: alces alces, biotopes, dynamics, hunting, moose, population, steppe zone, structure ukraine. moose (alces alces) dispersing into southern ukraine from adjacent forestland began to develop a marginal population in the steppe region in the mid-20th century. by 1974 this population numbered about 2,800 and primarily occupied planted forests, shelter belts, and agrocoenoses. however, most were killed by the end of the 20th century. thus, our research was focused upon the distribution and reproductive characteristics of the remaining population to assess population dynamics under new ecological conditions. this paper presents a summary of >25 years of data collected from 1976-2003. because field population surveys were initiated in 1961, surveys of game-keepers and hunters were used to reconstruct earlier population estimates. data were collected with a variety of methods including observations of moose on permanent plots, during field expeditions, and with observers in helicopter and airplane. as a result, data were collected about locations and movements at 91 points. a vast area of the steppe zone was covered where habitat use and distribution of animals (n = 639), herd composition (n = 334), and dynamics were measured. twenty-two females were dissected to assess their reproductive state, and necropsies and field observations of mortality were conducted when possible (n = 27). we assessed spatial structure with the nearest-neighbor method (odum 1975), and used the ministry of statistics of ukraine for statistical testing and data analysis. history and status of moose in ukraine – volokh alces vol. 45, 2009 6 dynamics of the area moose were widely distributed throughout ukraine during the paleolithic and late stone age. even then moose were highly used because they were relatively easy to hunt and had high economic value. therefore, even ancient hunters had substantial influence on moose populations. as civilization developed, hunting became the principal factor influencing the population size and distribution of moose. the first world war (1914-1917) and socialist revolution of 1917 changed the lives of millions of people and added to uncontrolled moose hunting. in many places moose were eliminated, and by the 1890s only about 10 remained in large forests in northern ukraine (migulin 1938) and a few in neighboring byelorussia (fedjushin 1929). gradual stabilization of life for the local people, renewal of work in governmental institutions, and implementation of nature conservation measures created favorable circumstances for population growth of moose in areas of the former ussr. range expansion occurred and the southern limit extended 200-400 km in the 1940s (geptner et al. 1961). however, this process was interrupted by world war ii (1941-1945) when most moose were killed during military operations. due to a hunting ban after the war, moose density in bordering regions of byelorussia and russia increased markedly and dispersal into ukraine was common. a large inflow from the northern forests was recorded in 1947-1948 and again in 1960 (boldenkov 1975). by 1961 moose occupied nearly the entire eastern forest-steppe zone, and some penetrated into the steppe zone and appeared in the foothills of the carpathians mountains (fig. 1). some moose dispersed very long distances and were observed north and south of the steppe zone in the dnieper delta and along the coast of the azov sea in 1957-1962 (filonov 1983), and even in the danube delta in romania in 1964 (аlmeshan 1966). by 1962 their range boundary shifted to the south along the dnieper floodplain for 350-500 km, and along the seversky donets floodplain for 200-250 km (galаkа 1964). unfortunately, initial migrants were often tolerant of people and the majority often succumbed to poaching. dispersal slowed afterward and reduced numbers were found in the large river valleys where floodplains were considerably transformed in the steppe zone. dispersal was also slowed by the small area of forest habitat intermingled with vast areas of non-habitat (i.e., arable land;table 1). by 1966 the southern range limit moved to the forest-steppe zone in western ukraine, however many animals were previously south of this limit (fig. 1). in the eastern steppe zone moose completely occupied forests in the seversky donets river basin and along the dnieper tributaries. later in 1967-1980, moose occupied shelter belts, planted forests, and fields in areas adjacent to the sea of azov and the black sea; some even penetrated to the crimean peninsula in 1971 and 1976 (dulitsky 2001). by 1980 the southern range boundary had reached its contemporary limit in ukraine. forest habitats were already occupied by moose, and further expansion to the south was not recorded, although occasional observations were noted in the late 20th century. during development of the steppe population, the distance between core areas quickly contracted and reached a minimum by administrative region total area arable land (%) forest (%) lugansk 2670.0 1458.0 (54.6) 155.8 (5.8) dniepropetrovsk 3190.0 2155.0 (67.6) 78.9 (3.7) donetsk 2650.0 1686.0 (63.6) 94.2 (3.6) odessa 3330.0 2081.0 (62.5) 72.6 (2.2) kherson 2856.0 1712.2 (60.0) 45.0 (1.6) zaporozhye 2730.0 1944.0 (71.2) 42.8 (1.6) nikolayev 2450.0 1716.0 (70.1) 23.0 (0.9) total: 19876.0 12752.2 (64.2) 512.3 (2.6) таble 1. characteristics of land occupied by moose in the steppe zone of ukraine (1000s ha). alces vol. 45, 2009 volokh – history and status of moose in ukraine 7 the1980-90s. the extreme limit was defined by the distance between forest biotopes, and by territorial and trophic competition. in some areas of the steppe zone moose inhabited agrocoenoses or forest belts 3-5 km apart. by the end of the 20th century, moose were exterminated everywhere due to excessive hunting and poaching, and their southern range retracted northward to the forest-steppe and steppe zones. during that process the distance between separate micro-populations and single moose increased greatly to 100s of km (table 2). in summary, the following dynamics of moose range were observed in ukraine during the 20th century: 1) reduction due to poaching and habitat loss occurred from periphery to refuge areas; restoration occurred vice versa, 2) all new pockets of moose developed only in forestland, 3) some initial dispersers were atypically tolerant of people and suffered high mortality, 4) dispersal occurred along a broad front, but major routes were along river valleys and forestland, and 5) expansion of moose range occurred when the population density was both high and low. fig. 1. the chronology and location of the southerly dispersal and expansion of the moose population in ukraine during the 20th century. phase years n mean ± se range occupation 1955-1971 47 107.6 ± 9.1 13.1 – 300.2 peak population 1980-1990 30 24.4 ± 5.7 3.1 – 135.0 depressed population 1993-2003 14 125.7 ± 7.0 90.1 – 183.2 таble 2. spatial structure of the steppe moose population during 3 phases of population dynamics in southern ukraine. data are distances (km) between pocket areas or areas of prolonged occupation. history and status of moose in ukraine – volokh alces vol. 45, 2009 8 reproduction moose reproduction was maximal in the steppe zone of ukraine during the period of its highest population (1981-1990, таble 3); however, during expansion and dispersion the population essentially declined. possibly, this was related to the dominance of young females in the population that are less fertile than animals of middle age (filonov 1983). the mean calving rate was 1.3 ± 0.10/pregnant or 0.4/ adult female (volokh 2002). on average there were 78% singletons, 25% twins, and 3% triplets; however, the adult reproductive rate was only 39%. in comparison, this rate was 46% north of and 69% south of the forest zone in russia, with1.2-1.4 embryos/pregnant female and 0.6-1.1 embryos/adult female (filonov 1983). an analysis of 27,300 licenses from harvested females indicated embryo/adult female ratios of 1.2-1.4 in european russia and 1.3 in the forest-steppe and steppe zones. in the latter case, the proportion of pregnant females to adults was 0.3-0.8 (rozhkov et al. 2001), which was similar to the fertility rate of moose at their southern range in the steppe region of ukraine. it is interesting that reproductive rates were 80-100% along the northern border of moose range in finland (rajakoski and koivisto 1970), and fertility fluctuated from 0.3 for young animals to 1.1 for animals of middle age (nygren et al. 1999). the majority (69%) of the marginal population in southern ukraine was adult animals in very small groups (herd) (таble 4); in comparison, 74% were adults in northwest russia (vereschagin and rusakov 1979) and 76% in byelorussia (kozlo 1983). few were aged 2 years (9.7%) which presumably was a result of high mortality of young animals and low productivity. maximum birth rate (25-26%) is achieved at an adult sex ratio of 1 ♂:0.91.1 ♀ (filonov 1983). the birth rate was 24.2% in a population with sex ratio of 1:1 in the forest-steppe zone of russia (pobedynskij 1990), 18.0% in a population with a sex ratio of 1:1.4 in the forest zone (filonov 1983), and 14.8% in a population with sex ratio of 1:0.9 in byelorussia (kozlo 1983). there was considerable dominance of females (1 ♂:1.4 ♀) in the northern steppe zone of ukraine with small isolated forests. the proportion of calves was 21.8%, similar to the rate reported in canada (21%; peterson 1955). in southern ukraine the sex ratio was nearly optimal (1 ♂:1.2 ♀) only in forest biotopes, whereas in open landscapes males were dominant (1.6 ♂:1 ♀). further, many solitary adult moose were observed (n = 205; table 5); because many (107 males and 98 females) were in isolated biotopes considerable distances apart, reproduction was affected negatively. there were 22 adult males (8.4%) in biotopes without adult females, and 53 females (20.6%) were likewise isolated. this isolation contributed to both low annual population growth (about 6%) and calf production (17.1%). low reproductive rates were caused by isolation of forest habitats preventing animal exchange, as well as different reproductive strategies of males and females. at the southern border of their range, adult females are usually long-term occupants of isolated forest habitats, island-like among agrocoenoses, and move little during the breeding season. conversely, males often disperse widely searching for females and forest habitats, sometimes reaching the seashore; many perish, mostly from poaching. we identified traits of gunshot wounds in 47% of mortalities (n = 62). unfortunately, this information was not considered when harvest regulations were developed for exploiting the year number of adult females number of calves/ pregnant female total with calf (%) mean ± se range 1971-1980 76 28 (37) 1.25 ± 0.08 1 – 2 1981-1990 83 37 (45) 1.40 ± 0.10 1 – 3 1991-2002 75 25 (33) 1.28 ± 0.11 1 – 2 total: 234 90 (39) 1.33 ± 0.10 1 – 3 таble 3. reproductive indices of the steppe moose population in southern ukraine. alces vol. 45, 2009 volokh – history and status of moose in ukraine 9 southern moose population, and contributed to its extirpation (volokh 2002). population dynamics humans invariably have strong influence on population size of hoofed animals, including moose. the developing moose population in the steppe zone of ukraine was related to an initial increase of resources in neighboring byelorussia (kozlo 1983) and russia (filonov 1983), and later in the forest zone of ukraine. moose hunting began in 1946 in the european part of russia, and the population and harvest increased concurrently; in 1960 there were 9,200 moose harvested, and by 1962 the harvest was 21,300 (filonov 1983). moose were first hunted in byelorussia in 1965, and about 30,000 were shot in 30 years (kozlo 1983). the associated reduction in moose density caused less emigration into the forest zone of ukraine where an initial harvest of 60 moose occurred in 1965. population growth rate was high along the southern range border reaching 28% (13-49%), similar to that in the forest zone (fig.2), as well as throughout ukraine in the 1960s (25-26%; boldenkov 1975). this high growth rate resulted from high reproduction of local animals and immigration. it should be noted that the dynamics of the moose population throughout ukraine was notable for its synchronism (fig. 3). the time period between the initial hunt (1965) and when the maximum population was reached in each area of the steppe zone was nearly identical, 15.3+0.6 years. this was irrespective of area or location, relative isolation, and relative distance to refuges occupied with moose. it was evidence of the strong influence of the age and sex total herd characteristics # % # moose/herd ± se range adult males 208 32.6 190 1.09 ± 0.03 1-5 adult females 234 36.7 197 1.19 ± 0.05 1-4 yearlings 62 9.7 54 1.15 ± 0.09 1-3 juveniles 134 21.0 90 1.49 ± 0.14 1-6 total: 638 334 таble 4. age structure and group (herd) size of the steppe population of moose at the southern range limit in ukraine. biotope single adult moose herd composition total % male (n=107) % female (n=98) аdult:total % male only % female only (n = 13) (n = 18) deciduous wood 74 36.5 63.5 51:89 5.9 13.7 coniferous wood 26 53.9 46.1 18:37 27.8 11.1 forest belt 29 72.4 27.6 19:36 5.3 10.5 plavni 22 63.6 36.4 20:31 10.0 15.0 field 21 71.4 28.6 14:34 7.1 7.1 garden 12 41.7 58.3 6:15 16.7 16.7 gully 11 45.5 54.5 3:6 33.3 steppe 6 83.3 16.7 2:3 reed beds 4 25.0 75.0 3:5 33.3 total or average: 205 52.2 47.8 136:256 8.4 20.6 таble 5. distribution and herd composition of the moose population in biotopes at the southern range of moose in ukraine. history and status of moose in ukraine – volokh alces vol. 45, 2009 10 migration rate and effective reproduction in an anthropogenic landscape. the moose population reached maximum (n = 2776) in the steppe zone of ukraine in 1974 (fig. 3). most (>57%) occupied the lugansk region, almost 23.5% the donetsk region, >9.4 % in dniepropetrovsk, and the remaining few were in other regions (table year p op ul at io n fig. 3. comparison of the moose population in the steppe zone (1) with other areas of ukraine (2), 1965-2002. p op ul at io n % p op ul at io n gr ow th year fig. 2. the estimated size (1) and % growth (2) of the moose population in the steppe zone of ukraine, 1970-2000. alces vol. 45, 2009 volokh – history and status of moose in ukraine 11 6). in that period the population density was so high in most forestland of the steppe zone that individual home ranges often overlapped, and damage was high on forest plants difficult and expensive to grow in dry southern regions. consequently, moose were hunted even in areas with low population density. however, instead of reasonable management of this valuable resource, hunting was based on maximum harvest for monetary profit from domestic and export meat markets. with harvest rates equal to 4.3-8.1% of the annual population, hunting had minimal influence on population density. however, a measurable population decline occurred after harvest rate increased to 16% in 1973 and 12.5% in 1974. this negative situation was apparently managed for and in 1982 the steppe population reached its second peak (2,147); however, in 1981-1984 the population fell >50%. in 1985-1991 the population stabilized at 1,000-1,200, with small increases observed in 1986, 1988, and 1991. since 1992 a steady and rather quick population decline (25.3+5.80% annually) occurred throughout the steppe zone (fig. 2). by the end of the 20th century moose had disappeared in the zaporozhye, odessa, nikolayev, and kherson regions, and only about 80 animals remained in the dniepropetrovsk, donetsk and lugansk regions (table 6). presently, primary moose habitat is floodplain forests of the samara and the seversky donetsk rivers in the steppe zone of ukraine. excessive hunting also caused considerable reduction (69-98%) of the population in the forest-steppe zone. further, many local populations were lost in areas (mostly forest habitats) with the highest numbers and density. although moose hunting is prohibited except for a few selected individuals (3 in 2001 and 8 in 2003), population growth has not been realized; the estimated population was 4950 in 2000 and 4490 in 2003. the primary reason for the moose population decline in the steppe zone was that the exploitation rate greatly exceeded the average annual growth rate. although only about 8% of the population was harvested annually in ukraine, more than half the harvest was adult animals (krizhanovskij et al. 1988) that ultimately caused lower productivity. importantly, this harvest estimate does not account for wounding and poaching mortality. in total, human-associated mortality was undeniably the most important factor influencing the population dynamics of moose and all hoofed animals in ukraine in the late 20th century. references аlmeshan, k. h. a. 1966. process of acclimatization and range development of some commercial animal species in socialist republic of romania. iv inter-university zoogeographic conference, odessa, russia. abstract only. (in russian). boldenkov, s. v. 1975. a modern condition of population in ukrainian ssr. pages 324-325 in proceedings of all-union conference on mammals. moscow, russia. (in russian). dulitsky, a. i. 2001. biodiversity of crimea. mammals: natural history, status, conservation, perspective. s�nаt press, sim-�nаt press, sim-nаt press, sim-аt press, sim-t press, simperfol, ukraine. (in russian). fedjushin, a. v. 1929. dynamics and geographical distribution of game fauna of belarus ssr. hunters of the belarus. location maximum population (year) population jan. 2003 (% decline) odessa 65 (1979) 0 (100) nikolaev 26 (1974) 0 (100) kherson 64 (1976) 0 (100) zaporozhye 120 (1974) 0 (100) dniepropetrovsk 260 (1973) 5 (98) donetsk 670 (1973) 31 (95) lugansk 1595 (1973) 46 (97) table 6. chronology of the population estimates of moose in 7 regions of the steppe zone of ukraine. the initial population estimate in each region occurred in the time period 1955-1965. history and status of moose in ukraine – volokh alces vol. 45, 2009 12 minsk, russia. (in russian). filonov, k. p. 1983. moose. lesnaya promyshlennost, moscow, russia. (in russian). galаkа, b. a. 1964. moose range expansion in ukraine. pages 35-43 in moose biology and hunting. moscow, russia. (in russian). geptner, v. g., a. a. nasimovich, and a. g. bannikov. 1961. the mammals of the ussr. volume 1. artiodactyles and perissodactyles. vyshaya shkola, moscow, russia. (in russian). kozlo, p. g. 1983. ecological-morphological analysis of the minsk moose population. nauka press, minsk, russia. (in russian). krizhanovskij, v. i., s. v. boldenkov, and a. a. gubkin. 1988. biological bases and priorities of hunting economy of ussr. pages 3-19 in studies of theriofauna of ukraine: its rational use and protection. kiev, russia. (in russian). migulin, a. a. 1938. the mammals of the ukraine ssr. kharkіv, ukraine ssr. (in ukrainian). nygren, t., m. pesonen, r. tykkylainen, and m. l. wallen. 1999. hirvikannan ikajakautumassa nakyvat verotuksen jaljet (population age distribution of elk). riistantutkimusken tiedote (game research bulletin). 158: 1-16. (in finnish). odum, e. p. 1975. ecology: the link between the natural and the social sciences. holt, rinehart and winston, new york, new york, usa. peterson, r. l. 1955. north american moose. university of toronto press, toronto, canada. pobedynskij, y. d. 1990. elk population structure in the forest-steppe central black soll region. 3rd international moose symposium, syktyvkar, russia, 27 august-5 september. abstract only. (in russian). rajakoski, e., and i. koivisto. 1970. possible reasons for the variations in the moose population in finland. тransactions of the ix international congress of game biologists. moscow, russia. 7: 799-801. (in finnish with english summary). rozhkov, y. i., a. v. pronaev, o. d. piskunov, n. e. ovcucova, a. v. daviodov, and l. v. rozhkova. 2001. moose populationbiological analysis of hunting license information. tsentrokhotkontrol, moscow, russia. (in russian). vereschagin, n. k., and o. s. rusakov. 1979. ungulates from the north-western part of the ussr. nauka, leningrad, russia. (in russian). volokh, a. м. 2002. ecological regulation of moose numbers in the southern part of ukraine. pages 49-54 in visnyk lvivskoho nathsionalnoho universytetu. biological series no. 30. (in ukrainian). 23 evidence of summer nutritional limitations in a northeastern washington moose population rachel c. cook1, jared oyster2, kristin mansfield2, and richard b. harris3 1national council for air and stream improvement, 1401 gekeler lane, la grande, oregon 97850, usa; 2washington department of fish and wildlife, 2315 north discovery place, spokane, washington 99216, usa; 3washington department of fish and wildlife, 600 capital way north, olympia, washington 98501, usa abstract: understanding the role of summer-autumn nutrition is critically important as moose (alces alces) populations decline along their southern range in north america because it influences dynamics through performance and susceptibility to predation, disease, and parasitism. to assess nutritional limitations during summer-autumn, we estimated body fat and protein reserves (n = 61), pregnancy rate (n = 71), and lactation status (n = 59) of adult female moose in northeastern washington state in december 2013, 2014, and 2016. adult pregnancy rate was depressed (79%) and correlated with loin muscle thickness, and 14% of adult moose had evidence of delayed conception. adult moose, particularly those that had successfully raised a calf, entered winter with low energy stores. lactating moose were thinner than non-lactating moose and overall, 79% of moose sampled had < 9% body fat, indicating at least moderate nutritional limitations linked to performance and survival. body fat was positively related to subsequent survival, and marrow fat levels indicative of starvation or severe nutritional stress were found in 56% of femurs (10 of 18) collected. combined, these data highlight the importance of accounting for reproductive history when interpreting nutritional condition data and the importance of sampling moose populations in autumn when interpreting the influence of seasonal habitats on subsequent productivity and mortality. alces vol. 57: 23–46 (2021) key words: alces alces; body condition; body fat; moose; nutrition; pregnancy; ultrasonography; washington state. although huntable populations remain in most jurisdictions, moose (alces alces) are currently declining across much of their range within the united states south of the 49th parallel (timmerman and rodgers 2017, jensen et al. 2018), including most populations in rocky mountain states (nadeau et al. 2017). many factors are hypothesized to influence southern populations, including parasites and disease (lankester and samuel 2007), predation (ross and jalkotzy 1997, patterson et al. 2013, mech and fieberg 2014), plant phenology, composition, and nutrition (monteith et al. 2015, ruprecht et al. 2016), and climate change directly through heat stress (mccann et al. 2013, lenarz et al. 2019), or more likely, indirectly through changes in host-parasite relationships (rempel 2011, jones et al. 2017, 2019, pekins 2020). nutrition rarely impacts populations in a catastrophic, obvious manner but rather through subtle, cumulative effects across many metrics of performance (cook et al. 2004, hurley et al. 2014) complicating explicit identification of its effects. inadequate nutrition impacts reproduction, sub-adult growth, and survival and also interacts with other limiting factors making animals more susceptible to predation, autumn condition of moose – cook et al. alces vol. 57, 2021 24 disease, parasites, or heat stress and winter weather (huggard 1993, bender et al. 2008, metz et al. 2012, mattisson et al. 2014, johnson et al. 2019), with these interactions challenging to quantify in wild populations. in addition, researchers commonly measure pregnancy rates or survival of adults; however, these performance metrics in adults are the least sensitive to nutritional limitations (gaillard et al. 2000, bonenfant et al. 2002, cook et al. 2004, 2018). adding to the complexity, measuring nutritional resources is difficult for ruminants that make decisions at a variety of scales to maximize energy or protein intake while minimizing ingestion of secondary metabolites (hobbs and swift 1985, hobbs and hanley 1990, cook et al. 2016, forbey et al. 2018, shipley et al. 2020). intake and diet quality are reflective not only of the abundance of high-quality foods (renecker and schwartz 1998), but also of variation in bite mass among plant species (shipley 2007, cook et al. 2016, denryter 2017, hull et al. 2020), distribution of patches, and forage density (wickstrom et al. 1984, spalinger and hobbs 1992), attributes difficult to measure with standard vegetation sampling methods (cook et al. 2016). alternatively, nutritional condition – formally defined as the state of body components (i.e., fat and protein) that influence an animal’s future fitness (harder and kirkpatrick 1994) – reflects the cumulative balance of nutritional resources, energetic expenditures, and requirements. strategic, temporal measurements of nutritional condition of moose can provide insight into the occurrence and severity of resource limitations across seasons. critical to field measurements of nutritional condition has been the development of techniques to estimate the ingesta-free body fat percentage ( hereafter referred to as “body fat”) of live ungulates. in particular, rump fat thickness measured via ultrasonography is consistently one of the most accurate predictors of body fat in moose (stephenson et al. 1998), mule deer (odocoileus hemionus; stephenson et al. 2002, cook et al. 2007), elk (cervus canadensis; cook et al. 2001a), bighorn sheep (ovis canadensis; stephenson et al. 2020), and caribou (rangifer tarandus; cook et al. in press). unfortunately, measurements of body fat in moose near their southern distribution has been limited and restricted to mid-winter captures (delgiudice et al. 2011, oates 2016, ruprecht et al. 2016, newby and decesare 2020) constraining evaluations of summer-autumn nutritional resources shown to influence performance in moose (hjeljord and histol 1999, ericsson et al. 2002, herfindal et al. 2006a, 2006b, mcart et al. 2009, rolandsen et al. 2017). by the 1970s, moose numbers were sufficient to support a recreational hunt in northeastern washington (base et al. 2006, harris et al. 2015) that continued as wolves repopulated parts of the region in the last decade (harris et al. 2021). in 2016, this moose population was estimated at 5169 animals (3510–7034 [95% ci]; oyster et al. 2018) but was declining annually in 2014– 2018 based on dynamics of radio-collared adult females (harris et al. 2021). declines in other western states including wyoming, montana, and idaho, mid-western states (minnesota and north dakota), and northeastern states (new hampshire and vermont) have been associated with multiple, often interrelated factors including disease, predation, parasitism, habitat composition, and climate change (timmerman and rodgers 2017). understanding the relative influence of these factors on performance and survival of adult females and calves is a management challenge and a priority across the southern range of moose where habitat quality is of paramount importance. our objective was to use body condition metrics to assess evidence of summer-autumn alces vol. 57, 2021 autumn condition of moose – cook et al. 25 nutritional limitations of adult moose in northeastern washington and to interpret whether limitations were severe enough to influence reproduction and survival. we considered a moose population to be nutritionally limited, defined herein as nutritional inadequacies sufficient to reduce reproduction, sub-adult growth and development, or survival regardless if these inadequacies directly influence population growth trends, if 1) lactating moose had less than 12% body fat in autumn, 2) pregnancy rate of females ≥ 2-years old was less than 95%, or 3) in addition to other indicators of nutritional limitations, starvation was evident in adult female moose that died over winter. these criteria were based on studies of captive and wild elk and mule deer (cook et al. 2004, tollefson et al. 2010, cook et al. 2013) and of alaskan moose linking nutritional condition to performance (testa and adams 1998). study area our 1262 km2 study area was in northeastern washington, usa within the north-central rocky mountain forest terrestrial ecoregion on the eastern side and the okanagan dry forest ecoregion on the western side (olson et al. 2001) (fig. 1); elevations ranged from 500 to 2200 m. climate was characteristic of both the continental and marine types with low relative humidity and moderate temperatures during summer (mean = 16.6 °c in june – august) and cool, foggy weather during winter (mean = −2.7 °c in november – february) (usgs north america climate; www.sciencebase.gov). most precipitation occurred in winter and spring as measured in colville, washington: mean precipitation = 210 mm in november–february, 135 mm in march–may, 102 mm in june-august, and 57 mm in september–october (https://www. usclimatedata.com/climate/colville/washington/united-states/uswa0606). see harris et al. (2021) for weather information during the study period and an assessment relative to conditions monteith et al. (2015) found were correlated with recruitment in a multipopulation study. land ownership was a matrix of private timber inholdings (21%), private landowners (45%), and public (34% federal and state). the study area was dominated by forests (86%) in subalpine parkland, subalpine fir (abies lasiocarpa), grand fir (abies grandis), western red cedar/western hemlock (thuja plicata/ tsuga heterophylla), douglas fir (pseudotsuga menziesii), and ponderosa pine (pinus ponderosa) potential vegetation zones (appendix a; franklin and dyrness 1988, cooper et al. 1991). other communities included grasslands (subalpine meadows, idaho fescue [festuca idahoensis] and bluebunch wheatgrass [pseudoroegneria spicatum]) and sagebrush (artemisia spp.) and bitterbrush (purshia tridentata) shrublands (integrated landscape assessment project; https://inr.oregonstate.edu/ilap). the most common deciduous shrubs included rocky mountain maple (acer glabrum), serviceberry (amelanchier alnifolia), oceanspray (holodiscus discolor), ninebark (physocarpus malvaceus), scouler’s willow (salix scouleriana), fool’s huckleberry (menziesia ferruginea), huckleberry (vaccinium spp.), and snowberry (symphoricarpos spp.), and the most common evergreen shrub was snowbrush ceanothus (ceanothus velutinus) (johnson and o’neil 2001). potential moose predators included mountain lions (puma concolor), black bears (ursus americanus), and wolves (canis lupus) that arrived in northeastern washington by 2008 with 10 packs documented in 2013 (wdfw et al. 2019, harris et al. 2021). white-tailed deer (odocoileus virginianus) and a few elk (cervus canadensis) and mule deer (o. hemionus) also occupied the study area. moose hunting was permitted in the study area during all years of the study; http://www.sciencebase.gov https://www.usclimatedata.com/climate/colville/washington/united-states/uswa0606 https://www.usclimatedata.com/climate/colville/washington/united-states/uswa0606 https://www.usclimatedata.com/climate/colville/washington/united-states/uswa0606 https://inr.oregonstate.edu/ilap autumn condition of moose – cook et al. alces vol. 57, 2021 26 hunter pressure and success varied by game management unit (harris et al. 2021). methods animal capture and handling we captured adult female moose via aerial darting (pneudart inc., williamsport, pennsylvania, usa) from a bell jet-ranger helicopter (northwest helicopters, olympia, washington, usa) on 16–20 december 2013 (n = 28), 2–6 december 2014 (n = 25), and 1–6 december 2016 (n = 28); capture crews documented accompanying calves when possible. we immobilized animals with carfentanil (3–4.5 mg) and xylazine (50 mg) in 2013 and 2014, and etorphine (7.5–15 mg) and xylazine (50 mg) in 2016. after sedation, we blindfolded moose and injected 100 mg xylazine iv to deepen anesthesia and improve muscle relaxation, administered long-acting penicillin, flunixin meglumine, and a clostridium vaccine, and fig. 1. study area, defined by 95% kernel density estimates of locations of captured moose (left panel) and in context of their location in northeastern washington (right panel). major u.s. highways 2 and 395 are illustrated along with the pend oreille river. adult female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. alces vol. 57, 2021 autumn condition of moose – cook et al. 27 collected a blood sample. after injecting 10 mg of bupivacaine hydrochloride into the mental foramen for pain relief (mansfield et al. 2006), we extracted an incisiform lower canine to determine age by examination of annuli (matson’s laboratory, milltown, montana, usa). we fitted moose with gps (geographic positioning systems) collars transmitting a location every 23 h (vectronic/globalstar survey; vectronic aerospace gmbh, berlin, germany), and subsequently reversed the immobilizing agents with 450 mg naltrexone and 700 mg tolazoline. all captures were under supervision of the washington department of fish and wildlife (wdfw) veterinarian. on each animal, one trained observer (r. cook) measured maximum subcutaneous rump fat thickness (maxfat) and longissimus dorsi (hereafter “loin”) muscle thickness (to the nearest mm between the 12th and 13th ribs adjacent to the backbone) using an ibex® pro ultrasound with a 5.0 mhz, 7.0 cm linear probe (e. i. medical imaging, loveland, colorado, usa) (stephenson et al. 1998, cook et al. 2001a, 2010), and collected a rump body condition score (rumpbcs) using the elk score developed by cook et al. (2001). we measured chest-girth circumference to the nearest cm when positioning allowed (cook et al. 2003); otherwise, we measured from the mid-line of the sternum to the backbone and multiplied by 2. we classified pregnancy status for most moose using a combination of serology, palpation, and ultrasonography. we collected serum samples and quantified pregnancy status using enzyme-linked immunosorbent assay for pregnancy-specific protein b (pspb), with an optical density reading > 0.21 considered positive for pregnancy (biotracking llc, moscow, idaho, usa; haigh et al. 1993, noyes et al. 1997, huang et al. 2000). because pspb is considered 96% accurate 40 d post-conception in elk (noyes et al. 1997), an experienced observer also used transrectal ultrasound to visually examine the uterine contents (i.e., fluid and placentomes), a method that detects pregnancy in ruminants at ~15 d post-conception and can detect the fetal heartbeat at 21 d (pohler et al. 2016). therefore, we considered moose with non-pregnant pspb levels but classified as pregnant by ultrasound, and subsequently seen with a calf the following spring, to have conceived in early to midnovember. in addition, we considered moose with no evidence of pregnancy (either method), but observed with calf the following spring, to have conceived not earlier than mid-november (i.e., a few weeks prior to capture). because peak rut for moose occurs in late september-early october, we classified these moose as having delayed conception. we documented whether a calf was observed at heel immediately prior to capture, and categorized lactation status by visually examining udder size and any extractable fluid. we grouped females into 3 categories: 1) currently lactating had a swollen udder and either thick milk (signifying a calf had nursed within the past 11 days; flook 1970, fleet and peaker 1978, noble and hurley 1999) or dilute, milk-like fluid (suggesting that weaning occurred within days or weeks; unpublished data, r. cook) was extracted from the udder; 2) non lactating were without calf and had no evidence of a swollen udder with a small amount or no extractable clear fluid; 3) calfat-heel females had an accompanying calf at capture with udder characteristics of a non-lactating female. we relied on mortality signals from collars (9 h without movement) to assess the proximate cause of mortality, typically within 24 h, but up to 3 d later; if a legal harvest, we communicated with the hunter. autumn condition of moose – cook et al. alces vol. 57, 2021 28 we surveyed mortality sites for signs of a struggle and evidence of predators (scat or tracks), performed necropsies, noted tick loads and hair loss, collected tissue samples for histopathological analysis, and collected a femur when possible; see harris et al. (2021) for additional details on assigning proximate cause of death. we disarticulated whole femur bones from the carcass, removed most of the flesh, wrapped them twice in heavy plastic and stored them frozen for up to 8 months before extracting marrow for analysis. we split the femur on the longitudinal axis, extracted 5.6 to 50.1 g of marrow fat from the entirety of the cavity, and oven-dried samples at 75 °c to constant weight (3–5 d; neiland 1970). we note that desiccation of marrow soon after death in warmer climates can result in overestimates of fat content; however, kie (1978) recommended corrections only if collection was >10 d post-death, a period beyond our recovery window. see harris et al. (2021) for a comprehensive survival analysis. statistical analysis we used the global equation in cook et al. (in press) to estimate body mass from the chest-girth circumference measurement. when we could not measure circumference due to positioning, we used the average estimated body mass (350 kg) from all other adult moose. to account for the smaller body size of moose in the northern rockies, we allometrically scaled maxfat estimates to surface area with a large-animal scaling unit [maxfat/(0.15 × body mass0.56)]; the scaled maxfat values were used to predict body fat (cook et al. 2010). for any animal without measurable rump fat, we estimated body fat using the rumpbcs equation from cook et al. (2001). although not validated with moose, we believe that using rumpbcs and associated equations validated on north american elk provided biologically relevant and useful estimates of body fat. for animals with measurable rump fat, we used a 2sample t-test (proc ttest; sas institute 1988) to compare the estimates of body fat from the scaled maxfat equations with those derived from the equations of stephenson et al. (1998) which integrated body size within the predictive equation for body fat. we evaluated the effects of age at capture and of lactation status (i.e., lactating, non-lactating, and calf-at-heel) on our 3 metrics of nutritional condition (body fat, body mass, loin muscle thickness) with an analysis of covariance (sas institute 1988). if age was insignificant, we removed it from the model and conducted a one-way fixed effects analysis of variance (anova; sas institute 1988) with duncan’s multiple range test. we assembled frequency histograms to illustrate differences in distribution of body fat among moose for each lactation status. we compared means of body fat and loin muscle thickness using one-way, fixed effects anova for moose classified as non-pregnant, pregnant, or pregnant but with delayed conception, and separately for moose classified simply as non-pregnant and pregnant. we used logistic regression to evaluate probability of pregnancy as a function of autumn body fat and loin muscle thickness and age. we did not evaluate body mass for these analyses because of insufficient sample size. we estimated femur marrow fat (%) of collared moose that died, graphed results over time, and interpreted these results relative to reported values for other ungulate species. finally, we used cox proportional hazards models to evaluate adult survival as a function of both autumn fat and loin muscle thickness. for these analyses, we omitted 2 records in which sample animals died within 1 week after handling (no other moose died within 3 weeks of capture). we quantified alces vol. 57, 2021 autumn condition of moose – cook et al. 29 exposure as the number of days between capture and the terminal event (death or censored). because we were interested in assessing potential effects of body condition on non-hunting mortalities, we censored hunter-harvested moose on the date of harvest. age-at-death was not included in the survival analyses. we used the r package coxph, with ties among sampled animals handled using the efron method. for all analyses of statistical hypotheses, we used a significance level of α = 0.10 because sample sizes were small (reporting p values for results we consider significant), although we acknowledge this increased the probability of a type i error. results we captured 69 unique female moose over 78 capture events during early december in 2013, 2014, and 2016 (table 1). one emaciated female with severe mastitis died shortly after recapture (data from that second capture were excluded from all summaries and analyses). mean age at first capture was 6.7 years (range = 1–14); by year, mean age at capture was 5.1 (se = 0.8, n = 12), 6.0 (se = 0.7, n = 23), and 8.2 years (se = 0.7, n = 24) in 2013, 2014, and 2016, respectively. we estimated nutritional condition for 61 (58 unique) adult females that were ≥ 2 years old (table 1); annual measurements were considered independent samples for the 3 animals measured 2 (n = 1) or 3 (n = 2) years apart. only 6 moose (8%) were observed with calf and classified as currently lactating, whereas 53 (77%) were classified as non-lactating of which 16 (30%) were observed with calf. estimated body mass of 45 moose (29 full chest-girth, 16 half chest-girth) averaged 350 ± 49 kg and did not differ by age or by lactation status although lactating females were ~10% lighter than other females (table 2). average loin muscle thickness was 4.7 cm (range = 3.3–5.4; n = 60) and did not differ by lactation status or its interaction with age (table 2), although age was related to loin muscle thickness (f1,52 = 3.61, p = 0.063); for each year an animal aged, loin muscle decreased 0.036 cm (fig. 2b). average body fat was 7.7% (range = 0–12.2%; n = 61) across all animals. lactating moose had less body fat than non-lactating moose (f2,58 = 3.72, p = 0.030); body fat increased from 6.5 ± 0.5% in lactating females to 8.3 ± 0.3% in non-lactating females (table 2, fig. 3). eight moose (13%) had no measurable rump fat; rumpbcs was 1.25 for 2 females, 2.0 for 1 female, 2.5 for 1 female, and 2.75 for 3 females. most moose had body fat levels indicating moderate nutritional limitations. lactating females (n = 6) had a higher proportion of animals with body fat levels indicative of severe nutritional limitation than other female categories; no lactating female had body fat levels ≥ 8% (fig. 3). no significant difference was found between body fat estimates calculated with the stephenson et al. (1998) table 1. sample sizes for each measure of nutritional condition, pregnancy status, age, and lactation status for 78 moose capture events during december 2013, 2014, and 2016 in northeastern washington, usa. measurement na pregnancy status 73 age (cementum analysis) 71 lactation status 59 body fat (%) 61 loin muscle thickness (cm) 60 estimated body mass (kg) 45 full chest-girth circumference (cm) 29 half chest-girth circumference (cm) 16 a, three animals handled originally in either 2013 or 2014 were recaptured in 2016 and are included in these sample sizes. one additional recaptured female with severe mastitis was emaciated and died shortly after capture; we excluded this capture event from all analyses and summaries. autumn condition of moose – cook et al. alces vol. 57, 2021 30 equation and scaled maxfat (cook et al. 2010). two trends were that estimates diverged as fat increased and body fat calculated with scaled maxfat was always higher (appendix b), but this divergence was minimal at moderate to severe nutritional restriction (≤ 9% body fat). pregnancy rate of moose > 1 year of age (n = 71) was 79%; 8 of 56 pregnant moose (14%; 2–11 years old) were classified as delayed conception (fig. 4a). for 2-year-olds specifically (n = 8), 87% were pregnant with 2 classified as delayed conception (fig. 4a) and 1 of 2 yearlings was pregnant. we found no difference in body fat among animals classified as non-pregnant, pregnant, or pregnant with delayed conception. when combined, pregnant moose were fatter than non pregnant moose (7.9% versus 6.8% body fat; f1, 57 = 4.98, p = 0.090; fig. 4a). loin muscle thickness was smaller in non-pregnant than in all pregnant moose (4.5 cm versus 4.8 cm; f1, 57 = 4.98, p = 0.030) and in pregnant moose with delayed conception (4.5 cm versus 4.9 cm; f2, 56 = 3.03, p = 0.056). we found no effect of age on probability of pregnancy (fig. 4c, 5a), although sample size was limited. because age was not significant as an interactive covariate in models predicting pregnancy with either body fat or loin muscle thickness, we dropped age from those models. we found no effect of body fat on probability of pregnancy although the associated coefficient suggested increased probability of pregnancy as fat increased (β = 0.2654, fig. 5b). loin muscle thickness predicted pregnancy both in a stand-alone model 0 1 to 3 4 to 6 7 to 9 >10 2.5 3 3.5 4 4.5 5 5.5 body fat (%) lo in th ic kn es s (c m ) 1 2720 102 0 2 4 6 8 10 12 14 16 3.5 4 4.5 5 5.5 age lo in th ic kn es s (c m ) a b y = 4.95 – 0.036x se = 0.38; r = 0.09; p = 0.0222y x. fig. 2. average loin muscle thickness (cm) across 5 body fat categories (graph a) and the relationship between loin muscle thickness (cm) and moose age (years; graph b). female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. table 2. mean, standard error (se), and sample size (n) of body fat, body mass, and loin muscle thickness across 3 lactation categories of moose (lactating = had swollen udder and evidence of calf at capture; calf-at-heel = no evidence in udder of lactation but was seen with a calf before the capture approach; non-lactating = no evidence in udder of lactation and was not seen with a calf before the capture approach) captured during december 2013, 2014, 2016 in northeastern washington, usa. means with the same letters following them are not significantly different. nutritional condition lactating calf-at-heel non-lactating mean se n mean se n mean se n body fat (%) 6.4a 0.5 6 7.1a, b 0.5 17 8.3b 0.3 38 body mass (kg) 314.7a 23.9 3 357.4a 21.6 11 351.4a 7.5 30 loin muscle thickness (cm) 4.5a 0.2 6 4.7a 0.1 17 4.8a 0.1 37 alces vol. 57, 2021 autumn condition of moose – cook et al. 31 (β = 1.9229, p = 0.038; fig. 5c) and when included as a covariate with body fat (body fat: β = 0.2300, p = 0.023; loin muscle: β = 1.8038, p = 0.056). among the 18 moose for which we measured femur marrow fat, the proximate cause of death was predation (39%), harvest (22%), winter ticks (17%), unknown health-related (11%), and accidental (11%). three moose (16.7%) had marrow fat levels indicative of starvation (< 12% marrow fat); these moose died in april or may (fig. 6) and death was proximately associated with tick infestation (high tick loads, substantial hair loss, and evidence of emaciation). an additional 7 moose (38.9%) had femur fat between 12 and 80%, levels indicative of nutritional limitations that could predispose animals to predation, disease, or parasites; 4 were predated, 2 died from an accident, and 1 death was unspecified health-related. six of these 7 moose died between february and june (the 7th died in november). the remaining 8 (44%) moose had femur marrow fat > 80% and died primarily between july and november (fig. 6); 3 were predated, 4 were harvested, and 1 was killed by vehicle. of 10 moose with marrow fat < 80%, 4 successfully raised a calf the previous growing season (i.e., observed calf or lactation status); no evidence was found for the other 6 animals. the average age and range of moose with < 80% (9.4 years, range = 4–14 years) and > 80% marrow fat (8.1 years, range = 5–12 years) were generally similar. increased body fat reduced mortality hazard regardless of lactation or calf status at time of capture (β = −0.2799, se = 0.1273, z = −2.1990, p = 0.028); neither calving status nor its interaction with body fat significantly predicted mortality hazard. neither loin thickness nor body mass were significant predictors of mortality hazard. 1 2 3 4 5 6 7 8 9 10 11 12 0 0.1 0.2 0.3 0.4 0.5 0.6 body fat (%) p ro p o rt io n lactating calf-at-heel non-lactating severe nutritional limitations moderate nutritional limitations mild nutritional limitations fig. 3. distribution of body fat (%) of adult female moose by lactation status: 1) lactating females with milk in udder, 2) non-lactating females lacking any udder characteristics of lactation and no evidence of calf at capture, and 3) calf-at-heel females with udder characteristics of non-lactating females, but were seen with a calf at the time of capture. vertical dotted lines indicate thresholds of nutritional limitations based on elk (cook et al. 2004). adult female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. autumn condition of moose – cook et al. alces vol. 57, 2021 32 1 2 3 4 5 6 7 8 9 10 11 12 0 0.1 0.2 0.3 0.4 0.5 0.6 body fat (%) p ro p o rt io n nonpregnant late pregnant severe nutritional limitations moderate nutritional limitations mild nutritional limitations a 1 2 3 4 5 6 7 8 9 10 11 12 13 14 0 20 40 60 80 100 age p re g n a n t (% ) 8 5 8 4 7 7 5 3 16 5 7 12 c 2 3 4 5 6 7 8 9 10 11 12 13 14 4 5 6 7 8 9 10 age b o d y fa t (% ) 7 5 7 4 5 7 5 3 16 2 5 1 bb fig. 4. distribution of body fat (%) by pregnancy status (non-pregnant, late = pregnant but evidence suggests delayed conception, pregnant) (graph a) where vertical dotted lines categorize expected severity of nutritional limitations associated with body fat levels based on captive elk trials (cook et al. 2004); average body fat (%) for each age with sample sizes imbedded in each bar (graph b); percent pregnant by age with sample sizes imbedded in each bar and moose with delayed conception distinguished by the checkered portion of each bar (graph c). female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. alces vol. 57, 2021 autumn condition of moose – cook et al. 33 discussion population trends of ungulates reflect mortality and productivity, both of which are strongly influenced by habitat and summer-autumn nutrition (cook et al. 2004, peek 2007, dale et al. 2008, hurley et al. 2014, rolandsen et al. 2017). our hypothesis that inadequate nutrition in summer and autumn limited adult moose performance in the study population was supported by several lines of evidence. in early december, 79% of moose sampled had body fat levels indicative of at least moderate nutritional limitations expected to impact performance (< 9% body fat, cook et al. 2004; fig. 3). non-lactating moose were significantly fatter than lactating moose, but even non lactating moose failed to attain body fat levels > 12%, a level indicative of little to no nutritional limitations on performance (cook et al. 2004). pregnancy rates were depressed (79%), correlated with protein stores, and evidence of delayed conception (i.e., breeding in early to midnovember) was found in 14% of adult moose. we documented a positive relationship between body fat and fig. 5. probability of pregnancy, as determined by logistic regression, based on (graph a) age of animal (years;), (graph b) ingesta-free body fat (%), and (graph c) loin muscle thickness (cm). female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. 0 2 4 6 8 10 12 14 16 0 0.2 0.4 0.6 0.8 1 age p ro b a b il it y o f p re g n a n c y a coefficient: -0.1554 se (slope): 0.10 p-value: 0.12 0 5 10 15 20 0 0.2 0.4 0.6 0.8 1 body fat (%) p ro b a b il it y o f p re g n a n c y b coefficient: 0.2654 se (slope): 0.17 p-value: 0.11 2.5 3 3.5 4 4.5 5 5.5 6 6.5 0 0.2 0.4 0.6 0.8 1 loin thickness (cm) p ro b a b il it y o f p re g n a n c y c coefficient: 1.9229 se (slope): 0.93 p-value: 0.038 2 4 6 8 10 12 0 20 40 60 80 100 sample month f e m u r m a rr o w f a t (% ) fig. 6. femur marrow fat (%) for collared moose that died, distributed by sample month (sample month = month + [month day/31.1]). femur marrow fat < 12% (lower dotted horizontal line) is indicative of starvation; marrow fat between 12 and 80% is indicative of nutritional limitations that can predispose animals to predation, disease, or parasites; marrow fat > 80% (upper dotted horizontal line) indicates only that animals have > 4.6% body fat as per equations presented for elk (cook et al. 2001a). female moose were captured and collared during early december 2013, 2014, and 2016 in northeastern washington, usa. autumn condition of moose – cook et al. alces vol. 57, 2021 34 subsequent survival, and 56% of femurs collected had marrow fat levels indicative of starvation or nutritional limitations that could predispose moose to predation, disease, or parasites (ratcliffe 1980, peterson et al. 1984, mech and delgiudice 1985, depperschmidt et al. 1987, sand et al. 2012). thus, we conclude that nutritional limitations primarily operating on the summer-autumn range are of a magnitude to affect moose populations in our study area. our inferences depend partly on the applicability of nutritional condition indices developed in part for other ungulates, particularly elk. for example, we used a rump bcs that is unvalidated for moose. however, a version of this score showed strong relationships with body fat in 3 ungulate species (cook et al. 2001a, 2007, in press) suggesting use of the elk-based bcs in combination with ultrasound measures of rump fat thickness was a biologically reasonable approach to estimate body fat in these moose. alternatively, we could have presented rump fat data without estimating body fat (e.g., ruprecht et al. 2016, newby and decesare 2020). however, a rump fat depth of zero indicates that subcutaneous rump fat has been depleted, and thus that the animal falls outside the range of condition for which this nutritional condition index can be used (cook et al. 2001a, 2007, 2010). body fat can vary from 0 to 5.7% in moose with no rump fat, a range in values equivalent to rump fat measurements of 0–2.75 cm (appendix c); hence, bias in the data would increase with an increasing proportion of “zero” animals. the proportion of moose in our data (13% of our sample) and those from other regions with no measurable rump fat demonstrate the need for additional validated indices of nutritional condition that can predict body fat in live moose with high levels of accuracy across the entire range found in wild populations. thresholds created from captive elk studies (cook et al. 2004) indicate that autumn body fat levels of 6–9% in lactating adults reflect populations experiencing moderate nutritional limitations resulting in depressed pregnancy rate, delayed conception, slower juvenile growth, and increased probability of winter mortality. at body fat levels < 6%, these impacts to performance would increase in severity. based on these criteria, our estimates of body fat indicate that most female moose in our sample were experiencing severe or moderate nutritional limitations regardless of whether they were lactating (33% severe, 67% moderate), not lactating but had a calf (6% severe, 82% moderate), or had no evidence of lactation or a calf (3% severe, 73% moderate). only 1 moose had december body fat > 12% (a non-lactating, 6-year-old pregnant animal), a level indicative of almost no nutritional limitations if the animal had raised a calf. adult moose in this population, particularly those that had successfully raised a calf, entered winter in a susceptible state (i.e., to predation, winter weather, parasites, or disease). in addition to low pregnancy rates, ungulates in poor body condition or negative energy balance may delay ovulation (cook et al. 2001b, 2013, johnson et al. 2019, this study) which can lead to delayed parturition dates thereby increasing the probability of neonatal mortality due to predation or other causes (keech et al. 2000, testa 2002, johnson et al. 2019); testa et al. (2000) found a 6.3% increase in neonatal mortality with each day’s delay of parturition in moose. body fat of our moose was positively related to subsequent survival, and of 8 moose (ages 4–12 years old) with autumn condition data that died within 6 months of capture, 63% entered winter with either < 7% body fat or loin thickness ≤ 4.4 cm. marrow fat levels indicative of starvation in 3 of 18 femurs further confirm alces vol. 57, 2021 autumn condition of moose – cook et al. 35 these findings. however, despite entering winter with low fat reserves, femur marrow fat patterns across the annual cycle indicate the peak of mortality from starvation in our sample did not occur until april, suggesting moose were employing strong compensatory mechanisms to survive most of winter. for example, as plants senesce and available energy declines in autumn, northern ungulates reduce organ size as a compensatory strategy. because organs account for 70% of resting energy use but comprise only about 10% of body mass, relatively small reductions in body mass due to reductions in organ size result in relatively large declines in metabolic energy requirements for winter maintenance (ramsey and hagopian 2006). may and june femur marrow fat suggested that moose were recovering at this time, but many did not exceed 5% body fat until mid-to late june. body fat is a critical reserve for temperate ungulates because it is a more concentrated and efficient source of energy than protein (robbins 1993) and, unlike protein reserves, nearly all fat reserves can be utilized (cahill 1970, watkins et al. 1992). however, understanding how both fat and protein influence productivity and survival may be useful (torbit et al. 1985, hilderbrand et al. 1999, parker et al. 2005), particularly at times of the year when body fat levels are low. loin muscle thickness has been related non-linearly to body fat in elk (i.e., declining slowly at high levels of body fat and more rapidly at low levels of body fat; cook 2000), leading researchers to use this measurement to identify survival thresholds (i.e., relative to probability of starvation). we found a similar relation between loin muscle thickness and body fat in sampled moose (fig. 2a) and that loin muscle thickness was correlated to probability of pregnancy. non-pregnant moose had smaller loin muscles (fig. 5c) suggesting moose in our study area were utilizing protein reserves at a level that affected performance even before winter. our data provide justification for additional research into the adequacy of summer ranges, but we also urge caution when interpreting nutritional condition data without concurrent information on other factors that influence energy and protein reserves in ungulates. nutritional condition in one season may influence nutritional condition in subsequent seasons, a “carry-over” effect documented in elk (cook et al. 2013), mule deer (monteith et al. 2013, 2014), moose (white et al. 2014), and caribou (dale et al. 2008). thus, nutritional condition in late winter or spring may be more closely related to levels the previous autumn (and reproductive status) than to winter weather or nutritional resources of the winter ranges. in addition, energy costs of lactation are 2–3 × greater than for maintenance metabolism (oftedal 1985, robbins 1993, cook 2002, national research council 2007), thus lactating ungulates are usually thinner than their non-lactating counterparts (testa and adams 1998, keech et al. 2000, crouse 2003, cook et al. 2013, white et al. 2014) in nutritionally inadequate environments. in contrast, non-lactating females are a heterogeneous mix of individuals that are not pregnant (having the lowest requirements during summer and thus highest nutritional condition), those that lose a calf early (having the second lowest requirements during summer; often in high nutritional condition), and those that lose a calf late in the growing season (nutritional requirements and thus nutritional condition often equivalent to a lactating female). as such, inferences of the nutritional value of habitats based on nutritional condition of non-lactating females can be ambiguous and misleading (cook et al. 2013). these patterns emerged in our sample; moose with no evidence of lactation autumn condition of moose – cook et al. alces vol. 57, 2021 36 or a calf were not only fatter on average but had greater variation: their body fat range was 11.2 % (1.0–12.2%), whereas the range of body fat for thinner lactating females was only 3.4% (4.3–7.7%). because moose are relatively solitary, observations of calf presence are routinely used to indicate reproductive status of females although a variety of issues, including sightability bias, could influence the reliability of those observations. although we expected a wide range of body fat in non-lactating moose, we also found high variation in animals not lactating but with a calf (1.0–10.2% body fat), suggesting certain moose may have been classified incorrectly. for example, one moose was seen with 2 juveniles and yet she had one of the highest estimates of body fat (18 mm maxfat = 10% body fat) suggesting the observed offspring may have been yearlings (testa 2004). this uncertainty in determining reproductive history, even in early december, increases as winter progresses, complicating interpretation of nutritional condition data (see bergman et al. 2020). measuring nutritional condition of moose can provide insights into seasonal and geographical patterns of nutritional limitations and the severity of these limitations but cannot identify the ecological states and processes that explain why populations may be nutritionally stressed (shipley et al. 2020). although moose did expand into our study area and ultimately occupy it at moderately high density by the early 2010s, recent studies of nutritional condition and forage resources from other ungulate species have indicated that nutritional limitations severe enough to substantially reduce performance in ungulates prevail across the northwestern united states. in the inland northwest, lactating elk had low levels of autumn body fat similar to our moose (e.g., ~5–7% body fat on average), with some populations displaying depressed pregnancy rates and delayed conception (cook et al. 2013), and at least 2 showing evidence of adult and juvenile starvation mortality in winter (cook et al. 2013, johnson et al. 2019). horne et al. (2019) reported summer nutritional influences on juvenile elk survival across much of idaho, and proffitt et al. (2016) reported significant influences of inadequate summer nutrition on elk populations in western montana. several studies that evaluated forage resources using tame ungulate foraging studies in the region found that dietary quality, particularly digestible energy content of ingested forage and their nutrient intake rates, were generally inadequate to fully support requirements of lactating females for deer and elk (cook et al. 2014, 2016, 2018, hull et al. 2020, ulappa et al. 2020). additional studies reported levels of forage quality and quantity in summer that were likely to negatively impact performance, including in western montana (proffitt et al. 2016), north idaho (monzingo 2020), and southwestern washington (geary et al. 2017). inadequate nutrition during the growing season impacted population growth in mule deer (hurley et al. 2014) and moose (schrempp et al. 2019) in idaho and mule deer in central oregon (peek et al. 2002). we found no data on body fat in moose during autumn in jurisdictions along the southern extent of their range in north american to compare with our study. but, our body fat estimates in early december (ave. = 7.7%) were lower than estimates from other regions obtained later in winter. for example, body fat averaged ~10–11% in february-march depending on winter severity in minnesota (delgiudice et al. 2011), ~8% in mid-february for females with calves and ~9.3% for those without in northwestern ontario (crouse 2003), and ranged from ~1 to 10.5% for individual moose in mid-february in wyoming (oates 2016). alces vol. 57, 2021 autumn condition of moose – cook et al. 37 clarifying the pathways through which forage quality and quantity influence seasonal nutrition, nutritional condition, and productivity may be essential for developing complete and holistic landscape assessments and resource planning on behalf of moose (rowland et al. 2018). in the northwestern united states, for example, forage quality, particularly digestible energy content, of ungulate diets and their forage intake rates are significantly higher in early seral communities (geary et al. 2017, barker et al. 2019) than in midand late seral forest communities (cook et al. 2014, 2016, 2018, hull et al. 2020, ulappa et al. 2020). in addition, digestible energy content of ungulate diets tends to be too low to satisfy requirements of lactating females and their calves, with a nutritional “bottleneck” occurring in late summer, an effect that is significantly exacerbated in vegetation communities with low forage biomass levels (cook et al. 2014, 2016, ulappa et al. 2020) and in populations existing at high densities. in northern idaho and the blue mountains ecoregion in northeastern oregon, estimated levels of biomass considered acceptable for large ungulates varied widely during summer and early autumn depending on forest type and canopy cover. in most forest zones, biomass levels of perennial forbs plus deciduous shrubs in midand advanced seral stages were generally < 200 kg/ha (cook et al. 2014, monzingo 2020), a level too low to support instantaneous intakes rates that can satisfy daily dry matter, digestible energy, and digestible protein requirements in elk (wickstrom et al. 1984, cook et al. 2014, 2016). in early seral communities with < 40% overstory canopy cover, biomass levels of perennial forbs plus deciduous shrubs in both ecoregions were 2–5 × greater than in midand late-seral stages, and more likely to satisfy nutritional requirements of large ungulates. in our study area, 75% of all forested stands had canopy closure > 40% (appendix a). among the more productive, higher elevation, and more mesic forests, 80% had canopy closure > 40% (appendix a). combined, these studies suggest that moose in our study area existed in an environment that provided marginal to inadequate forage resources during summer. we suspect that, in part, this is due to long-term fire suppression and reduced logging on federal lands that have and continue to reduce early seral communities (hessburg and agee 2003, haugo and welch 2013, schrempp et al. 2019), and thus forage resources for moose in this region. moose populations have increased after wildfires and other stand-replacing disturbances over vast landscapes, attributable at least in part to higher pregnancy and twinning (spencer and chatelain 1953, hansen et al. 1973, irwin 1974). although moose in northeastern washington appear to be in a post-irruptive state (harris et al. 2021), broad-scale habitat manipulations that promote early seral communities, particularly in productive, moist forests, may be required to increase nutritional condition and performance of this population if that is a management goal. acknowledgements principle funding came from the washington department of fish and wildlife. additional funding came from the upper columbia united tribes (ucut), spokane, washington and the national council for air and stream improvement, inc. (ncasi). for field assistance, we thank m. atamian, d. base, s. boogard, e. cosgrove, b. hoenes, m. devivo, h. ferguson, j. goerz, s. hansen, c. lowe, s. mccorquodale, b. murphie, a. prince, l. rossier, t. seitz, and p. wik. for capture work, we particularly thank j. hagerman of northwest helicopters, olympia, wa. thanks also to l. west (colville national autumn condition of moose – cook et al. alces vol. 57, 2021 38 forest), s. fisher (washington department of natural resources), g. lech (hancock forest resources), p. buckland (inland empire paper company), t. carlson (stimson lumber company), and m. sapp (riley creek lumber company). for assistance with cartography, we thank p. whelan (washington department of fish and wildlife). we appreciated suggestions by two anonymous reviewers and the editors for ways to improve this manuscript. references barker, k. j., m. s. mitchell, k. m. proffitt, and j. d. devoe. 2018. land management alters traditional nutritional benefits of migration for elk. journal of wildlife management 83: 167–174. doi:10.1002/jwmg.21564 base, d. l., s. zender, and d. martorello. 2006. history, status, and hunter harvest of moose in washington state. alces 42: 111–114. bender, l. c., j. g. cook, r. c. cook, and p. b. hall. 2008. relations between nutritional condition and survival of north american elk (cervus elaphus). wildlife biology 14: 70–80. doi:10.2981/ 0909-6396(2008)14 [70:rbncas]2.0. co;2 bergman, e. j., f. p. hayes, p. m. lukacs, and c. j. bishop. 2020. moose calf detection probabilities: quantification and evaluation of a ground-based survey technique. wildlife biology 2020(2). doi:10.2981/wlb.00599 bonenfant, c., j. m. gaillard, f. klein, and a. loison. 2002. sexand age-dependent effects of population density on life history traits of red deer (cervus elaphus) in a temperate forest. ecography 25: 446–458. doi:10.1034/j. 1600-0587. 2002.250407.x cahill, g. f. jr. 1970. starvation in man. new england journal of medicine 282: 668–675. doi:10.1056/nejm197003192 821209 cook, j. g. 2002. nutrition and food. pages 259–349 in e. d. toweill and j. w. thomas, editors. north american elk: ecology and management. smithsonian institution press, washington, dc, usa. _____, b. k. johnson, r. c. cook, r. a. riggs, t. delcurto, l. d. bryant, and l. l. irwin. 2004. effects of summer– autumn nutrition and parturition date on reproduction and survival of elk. wildlife monographs 155: 1–61. _____, r. c. cook, r. w. davis, and l. l. irwin. 2016. nutritional ecology of elk during summer and autumn in the pacific northwest. wildlife monographs 195: 1–81. doi:10.1002/wmon.1020 _____, _____, r. w. davis, m. m. rowland, r. m. nielson, m. j. wisdom, j. m. hafer, and l. l. irwin. 2018. development and evaluation of a landscape nutrition model for elk in western oregon and washington. pp. 13–30 in m. w. rowland et al., modeling elk nutrition and habitat use in western oregon and washington. wildlife monographs 199: 1–102. doi:10.1002/ wmon.1033 cook, r. c. 2000. studies of body condition and reproductive physiology in rocky mountain elk. m.s. thesis, university of idaho, moscow, idaho, usa. _____, j. g. cook, d. l. murray, p. zager, b. k. johnson, and m. w. gratson. 2001a. development of predictive models of nutritional condition for rocky mountain elk. journal of wildlife management 65: 973–987. doi:10. 2307/3803046 _____, _____, and l. l. irwin. 2003. estimating elk body mass using chest girth circumference. wildlife society bulletin 31: 536–543. _____, _____, r. a. riggs, and l. l. irwin. 2014. habitat-nutrition relationships of elk during spring through autumn in the blue mountains of eastern oregon and their implications for forest landscape management. national alces vol. 57, 2021 autumn condition of moose – cook et al. 39 council of air and stream improvement, la grande, oregon, usa. _____, _____, t. r. stephenson, w. l. myers, s. m. mccorquodale, d. j. vales, l. l. irwin, p. b. hall, r. d. spencer, s. l. murphie, k. a. schoenecker, and p. j. miller. 2010. revisions of rump fat and body scoring indices for deer, elk, and moose. journal of wildlife management 74: 880–896. doi:10.2193/2009-031 _____, _____, d. j. vales, b. k. johnson, s. m. mccorquodale, l. a. shipley, r. a. riggs, l. l. irwin, s. l. murphie, b. l. murphie, k. a. schoenecker, f. geyer, p. b. hall, r. d. spencer, d. a. immell, d. h. jackson, b. l. tiller, p. j. miller, and l. schmitz. 2013. regional and seasonal patterns of nutritional condition and reproduction in elk. wildlife monographs 184: 1–45. doi:10.1002/ wmon.1008 _____, j. a. crouse, j. c. cook, and t. r. stephenson. in press. evaluating indices of nutritional condition for caribou (rangifer tarandus): which are the most valuable and why? canadian journal of zoology. _____, d. l. murray, j. g. cook, p. zager, and m. j. gratson. 2001b. nutritional influences on breeding dynamics in elk. canadian journal of zoology 79: 845– 853. doi:10.1139/z01-050 _____, t. r. stephenson, w. l. myers, j. g. cook, and l. a. shipley. 2007. validating predictive models of nutritional condition for mule deer. journal of wildlife management 71: 1934–1943. doi:10.2193/2006-262 cooper, s. v., k. e. neiman, and d. w. roberts. 1991. forest habitat types of northern idaho: a second approximation. united states department of agriculture (usda), intermountain research station, ogden, utah, usa. crouse, j. a. 2013. health, nutritional condition, and productivity of female moose (alces alces) in northeastern ontario. m.s. thesis. lakehead university, thunder bay, ontario, canada. dale, b. w., l. g. adams, w. b. collins, k. joly, p. valkenburg, and r. tobey. 2008. stochastic and compensatory effects limit persistence of variation in body mass of young caribou. journal of mammalogy 89: 1130–1135. doi:10.1644/07-mamm-a-137.1 delgiudice, g. d., b. a. sampson, m. s. lenarz, m. w. schrage, and a. j. edwards. 2011. winter body condition of moose (alces alces) in a declining population in northeastern minnesota. journal of wildlife disease 47: 30–40. doi:10.7589/0090-3558-47.1.30 _____, r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restriction, ticks, and numbers of moose on isle royale. journal of wildlife management 61: 895–903. doi:10.2307/3802198 denryter, k. a. 2017. foraging ecology of woodland caribou in boreal and montane ecosystems of northeastern british columbia. ph.d. dissertation. university of northern british columbia, prince george, british columbia, canada. depperschmidt, j. d., torbit, s. c., alldredge, a. w., and deblinger, r. d. 1987. body condition indices for starved pronghorn. journal of wildlife management 51: 675–678. doi:10.2307/ 3801287 ericsson, g., j. p. ball, and k. danell. 2002. body mass of moose calves along an altitudinal gradient. journal of wildlife management 66: 91–97. doi:10.2307/3802875 fleet, i. r., and m. peaker. 1978. mammary function and its control at the cessation of lactation in the goat. journal of physiology 279: 491–507. doi:10.1113/ jphysiol.1978.sp012358 flook, d. r. 1970. a study of sex differential in the survival of wapiti. autumn condition of moose – cook et al. alces vol. 57, 2021 40 canada wildlife service report, serial number 11. queens printer, ottawa, ontario, canada. forbey, j. s., r. liu, t. t. caughlin, m. d. matocq, j. a. vucetich, k. d. kohl, m. d. dearing, and a. m. felton. 2018. review: using physiologically based models to predict population responses to phytochemicals by wild vertebrate herbivores. animal 12: 383–398. doi:10.1017/s1751731118002264 franklin, j. f., and c. t. dyrness. 1988. natural vegetation of oregon and washington. oregon state university press, corvallis, oregon, usa. 452 pp. gaillard, j.-m., m. festa-blanchet, n. g. yoccoz, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual reviews of ecology and systematics 31: 367–393. doi:10.1146/annurev.ecolsys.31.1.367 geary, a. b., e. h. merrill, j. g. cook, r. c. cook, and l. l. irwin. 2017. elk nutritional resources: herbicides, herbivory and forest succession at mount st. helens. forest ecology and management 401: 242–254. doi:10.1016/j.foreco. 2017. 06.028 haigh, j. c., w. j. dalton, c. a. ruder, and r. g. sasser. diagnosis of pregnancy in moose using a bovine assay for pregnancy-specific protein b. theriogenology 40: 905–911. doi:10.1016/0093-691x (93)90358-c hansen, h., l. w. krefting, and v. kurmis. 1973. the forest of isle royale in relation to fire history and wildlife. university of minnesota agriculture experimental station. technical bulletin 294. 44 pp. harder, j. d., and r. l. kirkpatrick. 1994. physiological methods in wildlife research. pages 275–306 in t. a. bookhout, editor. research and management techniques for wildlife and habitats. fifth edition. the wildlife society, bethesda, maryland, usa. harris, r. b., m. atamian, h. ferguson, and i. keren. 2015. estimating moose abundance and trends in northeastern washington state: index counts, sightability models, and reducing uncertainty. alces 51: 57–69. _____, j. g. goerz, j. oyster, r. c. cook, k. mansfield, m. atamian, c. lowe, a. prince, and b. y. turnock. 2021. bottom-up and top-down factors contribute to reversing a moose population increase in northeastern washington. alces. haugo, r., and n. welch. 2013. current ecological conditions and restoration needs in forests of the clearwater basin, idaho. the nature conservancy in idaho. 40 pp. herfindal, i., b. e. sæther, e. j. solberg, r. andersen, and k. a. høgda. 2006a. population characteristics predict responses in moose body mass to temporal variation in the environment. journal of animal ecology 75: 1110–1118. doi:10.1111/j.1365-2656.2006.01138.x _____, e. j. solberg, b. e. sæther, k. a. høgda, and r. andersen. 2006b. environmental phenology and geographical gradients in moose body mass. oecologia 150: 213–224. doi:10.1007/ s00442-006-0519-8 hessburg, p. f., and j. k. agee. 2003. an environmental narrative of inland northwest united states forests, 1800– 2000. forest ecology and management 178: 23–59. doi:10.1016/ s0378-1127(03)00052-5 hilderbrand, g. v., c. c. schwartz, c. t. robbins, m. e. jacoby, t. a. hanley, s. m. arthur, and c. servheen. 1999. the importance of meat, particularly salmon, to body size, population productivity, and conservation of north american brown bears. canadian journal of zoology 77: 132–138. doi:10.1139/ z98-195 hjeljord, o., and t. histol. 1999. rangebody mass interactions of a northern alces vol. 57, 2021 autumn condition of moose – cook et al. 41 ungulate-a test of hypothesis. oecologia 119: 326–339. doi:10.1007/ s004420050793 hobbs, n. t., and t. a. hanley. 1990. habitat evaluation: do use/availability data reflect carrying capacity? journal of wildlife management 54: 515–522. doi:10.2307/3809344 _____, and d. m. swift. 1985. estimates of habitat carrying capacity incorporating explicit nutritional constraints. journal of wildlife management 49: 814–22. doi:10.2307/3801716 horne, j. s., m. a. hurley, c. g white, and j. rachael. 2019. effects of wolf pack size and winter conditions on elk mortality. journal of wildlife management 83: 1103–1116. doi:10. 1002/ jwmg.21689 huang, f., d. c. cockrell, t. r. stephenson, j. h. noyes, and r. g. sasser. 2000. a serum pregnancy test with a specific radioimmunoassay for moose and elk pregnancy-specific protein b. journal of wildlife management 64: 492–499. doi:10.2307/3803246 huggard, d. j. 1993. prey selectivity of wolves in banff national park. ii. age, sex, and condition of elk. canadian journal of zoology 71: 140–147. doi:10.1139/z93-020 hull, i. t., l. a. shipley, s. l. berry, c. loggers, and t. r. johnson. 2020. effects of fuel reduction timber harvests on forage resources for deer in northeastern washington. forest ecology and management. doi:10.1016/j.foreco. 2019. 117757 hurley, m. a., m. hebblewhite, j. m. gaillard, s. dray, k. a. taylor, w. k. smith, p. aaber, and c. bonenfant. 2014. functional analysis of normalized difference vegetation index curves reveals overwinter mule deer survival is driven by both spring and autumn phenology. philosophical transactions of the royal society b 369: 1–15. doi:10.1098/rstb.2013.0196 irwin, l. l. 1974. relationships between deer and moose on a burn in northeastern minnesota. m.s. thesis, university of idaho, moscow, idaho, usa. jensen, w. f., j. r. smith, m. carstensen, c. e. penner, b. m. hosek, and j. j. maskey, jr. 2018. expanding gis analyses to monitor and assess north american moose distribution and density. alces 54: 45–54. doi:10.1002/wmon.1039 johnson, b. k., d. h. jackson, r. c. cook, d. a clark, p. k. coe, j. g. cook, s. n. rearden, s. l. findholt, and j. h. noyes. 2019. roles of maternal condition and predation in survival of juvenile elk in oregon. wildlife monographs 201: 1–60. johnson, d. h., and t. a. o’neil. 2001. wildlife-habitat relationships in oregon and washington. oregon state university press, corvallis, oregon, usa. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. _____, _____, _____, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’ neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi:10.1139/ cjz-2018-0140 keech, m. a, r. t. bowyer, j. m. verhoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 52: 613–615. kie, j. g. 1978. femur marrow fat in whitetailed deer carcasses. journal of wildlife management 42: 661–663. doi:10.2307/3800838 lankester, m. w., and w. m. samuel. 2007. pests, parasites and diseases. pages autumn condition of moose – cook et al. alces vol. 57, 2021 42 479–517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2019. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. doi:10.2193/2009-493 mansfield, k. g., f. j. m. verstraete, and p. j. pascoe. 2006. mitigating pain during tooth extraction from conscious deer. wildlife society bulletin 34: 201– 202. doi:10.2193/0091-7648(2006)34[2 01:mpdtef]2.0.co;2 mattisson, j., g. r. rauset, j. odden, h. andrén, j. d. c. linnell, and j. persson. 2016. predation or scavenging? prey body condition influences decision-making in a facultative predator, the wolverine. ecosphere 7: 1–14. doi:10.1002/ ecs2.1407 mcart, s. h., d. e. spalinger, w. b. collins, e. r. schoen, t. stevenson, and m. bucho. 2009. summer dietary nitrogen availability as a potential bottom–up constraint on moose in south– central alaska. ecology 90: 1400–1411. doi:10.1890/08-1435.1 mccann, n. p., r. a. moen, and t. r. harris 2013. warm-season heat stress in moose (alces alces). canadian journal of zoology 91: 893–898. doi:10.1139/ cjz-2013-0175 mech, l. d., and g. d. delgiudice. 1985. limitations of the marrow-fat technique as an indicator of body condition. wildlife society bulletin 13: 204–206. _____, and j. fieberg. 2014. re-evaluating the northeastern minnesota moose decline and the role of wolves. journal of wildlife management 78: 1143–1150. doi:10.1002/jwmg.775 metz, m. c., d. w. smith,, j. a. vucetich, d. r. stahler, and r. o. peterson. 2012. seasonal patterns of predation for gray wolves in the multi-prey system of yellowstone national park. journal of animal ecology 81: 553– 563. doi:10.1111/j.1365-2656.2011. 01945.x monteith, k. l., v. c. bleich, t. r. stephenson, b. m. pierce, m. m. conner, j. g. kie, and r. t. bowyer. 2014. life-history characteristics of mule deer: effects of nutrition in a variable environment. wildlife monographs 186: 1–62. doi:10.1002/wmon.1011 _____, r. w. klaver, k. r. hersey, a. a. holland, t. p. thomas, and m. j. kauffman. 2015. effects of climate and plant phenology on recruitment of moose at the southern extent of their range. oecologia 178: 1137–1148. doi:10.1007/s00442015-3296-4 _____, t. r. stephenson, v. c. bleich, m. m. conner, b. m. pierce, and r. t. bowyer. 2013. risk-sensitive allocation in seasonal dynamics of fat and protein reserves in a long-lived mammal. journal of animal ecology 82: 377–388. doi:10.1111/ 1365-2656.12016 monzingo, d. s. 2020. influences of habitat characteristics on forage resources of rocky mountain elk (cervus canadensis) in north-central idaho. m.s. thesis. washington state university, pullman, washington, dc, usa. nadeau, m. s., n. j. decesare, d. g. brimeyer, e. j. bergman, r. b. harris, k. r. hersey, k. k. huebner, p. e. matthews, and t. p. thomas. 2017. status and trends of moose populations and hunting opportunity in the western united states. alces 53: 99–112. national research council. 2007. nutrient requirements of small ruminants: sheep, goats, cervids, and new world camelids. the national academies press,washington, dc, usa. neiland, n. t. 1970. weight of dried marrow as indicator of fat in caribou femurs. journal of wildlife management 34: 904–907. doi:10.2307/3799158 alces vol. 57, 2021 autumn condition of moose – cook et al. 43 newby, j. r., and n. j. decesare. 2020. multiple nutritional currencies shape pregnancy in a large herbivore. canadian journal of zoology 98: 307–315. doi:10.1139/cjz-2019-0241 noble, m. s., and w. l. hurley. 1999. effects of secretion removal on bovine mammary gland function following an extended mild stasis. journal of dairy science 82: 1723–1730. doi:10.3168/ jds.s0022-0302(99)75402-0 noyes, j. h., r. g. sasser, b. k. johnson, l. d. bryant, and b. alexander. 1997. accuracy of pregnancy detection by serum protein (pspb) in elk. wildlife society bulletin 25: 695–698. oates, b. a. 2016. effects of predators and resource limitation on demography and behavior of moose in the greater yellowstone ecosystem, m.s. thesis. university of wyoming, laramie, wyoming, usa. oftedal, o. t. 1985. pregnancy and lactation. pages 215–238 in r. j. hudson and r. g. white, editors. bioenergetics of wild herbivores. crc press, boca raton, florida, usa. olson, d. m., e. dinerstein, e. d. wikramanayake, n. d. burgess, g. v. n. powell, e. c. underwood, j. a. d’amico, i. itoua, h. e. strand, j. c. morrison, c. j. loucks, t. f. allnutt, t. h. ricketts, y. kura, j. f. lamoreux, w. w. wettengel, p. hedao, and k. r. kassem. 2001. terrestrial ecoregions of the world: a new map of life on earth. bioscience 51: 933–938. doi: 10.1641/0006-3568(200 1)051[0933: teotwa]2.0.co;2 oyster, j. h., i. n. keren, s. j. k. hansen, and r. b. harris. 2018. hierarchical mark-recapture distance sampling to estimate moose abundance. journal of wildlife management 82: 1668–1679. doi:10. 1002/jwmg.21541 parker, k. l., p. s. barboza, and t. r. stephenson. 2005. protein conservation in female caribou (rangifer tarandus): effects of decreasing diet quality during winter. journal of mammalogy 86: 610–622. doi:1 0.1644/1545-1542(20 05)86[610:pcifcr]2.0.co;2 patterson, b. r., j. f. benson, k. r. middel, k. j. mills, a. silver, and m. e. obbard. 2013. moose calf mortality in central ontario, canada. journal of wildlife management 77: 832–841. doi:10.1002/ jwmg.516 peek, j. m. 2007. habitat relationships. pages 351–375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. university press, boulder, colorado, usa. _____, b. dennis, and t. hershey. 2002. predicting population trend of mule deer in south–central oregon. journal of wildlife management 66: 729–736. doi:10.2307/3803138 pekins, p. j. 2020. metabolic and population effects of winter tick infestation on moose: unique evolutionary circumstances? frontiers in ecology and evolution 8: 1–13. doi: 10.3389/ fevo.2020.00176 peterson, r. o., j. d. woolington, and t. n. bailey. 1984. wolves of the kenai peninsula, alaska. wildlife monographs 88: 1–52. pohler, k. g., g. a. franco, s. t. reese, f. g., dantas, m. d. ellis, and r. r. payton. 2016. past, present and future of pregnancy detection methods. applied reproductive strategies in beef cattle symposium, des moines, iowa, usa. proffitt, k. m., m. hebblewhite, w. peters, n. hupp, and j. shamhart. 2016. linking landscape-scale differences in forage to ungulate nutritional ecology. ecological applications 26: 2156–2174. doi:10.1002/eap.1370 ramsey, j. j., and k. hagopian. 2006. energy expenditure and restriction of energy intake: could energy restriction alter energy expenditure in companion autumn condition of moose – cook et al. alces vol. 57, 2021 44 animals? journal of nutrition 136: 1958–1966. doi:10.1093/jn/136.7.1958s ratcliffe, p. r. 1980. bone marrow fat as an indicator of condition in roe deer. acta theriologica 25: 333–340. doi:10.4098/ at.arch.80-30 rempel, r. s. 2011. effects of climate change on moose populations: exploring the response horizon through biometric and systems models. ecological modelling 222: 3355–3365. doi:10. 1016/j.ecolmodel.2011.07.012 renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behavior. pages 403–440 in a. w. franzmann, and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. robbins, c. t. 1993. wildlife feeding and nutrition. second edition. academic press, san diego, california, usa. rolandsen, c. m., e. j. solberg, b.-e. sæther, b. v. moortor, i. herfindal, and k. jørnerass. 2017. on fitness and partial migration in a large herbivore – migratory moose have higher reproductive performance than residents. oikos 126: 547–555. doi:10.1111/oik.02996 ross, p. i., and m. g. jalkotzy. 1997. cougar predation on moose in southwestern alberta. alces 32: 1–8. rowland, m. m., m. j. wisdom, r. m. nielson, j. g. cook, r. c. cook, b. k. johnson, p. k. coe, j. m. hafer, b. j. naylor, d. j. vales, r. g. anthony, e. k. cole, c. d. danilson, r. w. davis, f. geyer, s. harris, l. l. irwin, r. mccoy, m. d. pope, k. sager-fradkin, and m. vavra. 2018. modeling elk nutrition and habitat use in western oregon and washington. wildlife monographs 199: 1–69. doi:10.1002/wmon.1033 ruprecht, j. s., k. r. hersey, k. hafen, k. l. monteith, n. j. decesare, m. j. kauffmann, and d. r. macnulty. 2015. reproduction in moose at their southern range limit. journal of mammalogy 97: 1355–1365. doi:10.1093/jmammal/ gyw099 sand, h., c. wikenros, p. ahlqvist, t. h. strømseth, and p. wabakken. 2012. comparing body condition of moose (alces alces) selected by wolves (canis lupus) and human hunters: consequences for the extent of compensatory mortality. canadian journal of zoology 90: 403–412. doi:10.1139/z2012-007 sas institute. 1988. sas/stat user’s guide. release 6.03. sas institute, cary, north carolina, usa. schrempp, t. v., j. l. rachlow, t. r. johnson, l. a. shipley, r. a. long, j. l. aycrigg, and m. a. hurley. 2019. linking forest management to moose population trends: the role of the nutritional landscape. plos one 14: e0219128. shipley, l. a. 2007. the influence of bite size on foraging at larger spatial and temporal scales by mammalian herbivores. oikos 116: 1964–1974. doi:10.1111/j.2007. 0030-1299.15974.x _____, r. c. cook, and d. g. hewitt. 2020. techniques for wildlife nutritional ecology. pages 439–482 in in n. j. silvy (ed.), the wildlife techniques manual. volume 1, research methods, eighth edition. john hopkins university press, baltimore, maryland, usa. spalinger, d. e., and n. t. hobbs. 1992. mechanisms of foraging in mammalian herbivores: new models of functional response. american naturalist 140: 325–348. doi:10.1086/285415 spencer, d. l., and e. f. chatelain. 1953. progress in the management of the moose of southcentral alaska. transactions of the north american wildlife conference 18: 539–552. stephenson, t. r., v. c. bleich, b. m. pierce, and g. p. mulcahy. 2002. validation of mule deer body composition using in vivo and post-mortem indices of nutritional condition. wildlife society bulletin 30: 557–564. alces vol. 57, 2021 autumn condition of moose – cook et al. 45 _____, d. w. german, m. e. blum, m. cox, k. m. stewart, e. f. cassirer, d. p. walsh, and k. l. monteith. 2020. linking population performance to nutritional condition in an alpine ungulate. journal of mammalogy. doi:10.1093/jmammal/gyaa091 _____, k. j. hundertmark, c. c. schwartz, and v. van ballenberghe. 1998. predicting body fat and body mass in moose with ultrasonography. canadian journal of zoology 76: 717–722. doi:10.1139/z97-248 testa, j. w. 2002. does predation on neonates inherently select for earlier births? journal of mammalogy 83: 699–706. d o i : 1 0 . 1 6 4 4 / 1 5 4 5 1 5 4 2 ( 2 0 0 2 ) 0 8 3 %3c0699:dponis%3e2.0.co;2 _____. 2004. population dynamics and life history trade-offs of moose (alces alces) in south-central alaska. ecology 85: 1439–1452. doi:10.1890/ 02-0671 _____, and g. p. adams. 1998. body condition and adjustments to reproductive effort in female moose (alces alces). journal of mammalogy 79: 1345–1354. doi:10.2307/1383026 _____, e. f. becker, and g. r. lee. 2000. temporal patterns in the survival of twin and single moose calves (alces alces) in southcentral alaska. journal of mammalogy 81: 162–168. doi:10.1093/ jmammal/81.1.162 timmerman, h. r., and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. tollefson, t. n., l. a. shipley, w. l. myers, d. h. keisler, and n. dasgupta. 2010. influence of summer and autumn nutrition on body condition and reproduction in lactating mule deer. journal of wildlife management 74: 974–986. doi:10.2193/2008-529 torbit, s. c., l. h. carpenter, d. m. swift, and a. w. alldredge. 1985. differential loss of fat and protein by mule deer during winter. journal of wildlife management 49: 80–85. doi:10.2307/ 3801849 ulappa, a. c., l. a. shipley, r. c. cook, j. g. cook, and m. e. swanson. 2020. silvicultural herbicides and forest succession influence understory vegetation and nutritional ecology of black-tailed deer in managed forests. forest ecology and management. doi:10.1016/j.foreco. 2020.118216 washington department of fish and wildlife (wdfw), confederated colville tribes, spokane tribe of indians, usda-aphis wildlife services, and u.s. fish and wildlife service. 2019. washington gray wolf conservation and management 2018 annual report. washington department of fish and wildlife, ellensburg, wa, usa. (accessed april 2020). watkins, b. e., j. h. witham, and d. e. ullrey. 1992. body composition changes in white–tailed deer fawns during winter. canadian journal of zoology 70: 1409–1416. doi:10.1139/z92-197 white, k. s., n. l. barten, s. crouse, and j. crouse. 2014. benefits of migration in relation to nutritional condition and predation risk in a partially migratory moose population. ecology 95: 225–237. doi:10.1890/13-0054.1 wickstrom, m. l., c. t. robbins, t. a. hanley, d. e. spalinger, and s. m. parish. 1984. food intake and foraging energetics of elk and mule deer. journal of wildlife management 48: 1285–1301. doi:10.2307/3801789. autumn condition of moose – cook et al. alces vol. 57, 2021 46 appendix a. percent of study area, defined by 95% kernel density estimates of locations of captured moose, in each of 6 forested types (sa-prk = subalpine fir-parkland, sa-fir = subalpine fir, abgr = grand fir, thpl = western red cedar, psme = douglas fir, pipo = ponderosa pine; graph a) and by canopy cover classes grouped as wet forests (sa-prk + sa-fir + abgr + thpl) or dry forests (psme + pipo) (graph b). adult female moose were captured during early december 2013, 2014, and 2016 in northeastern washington, usa. 0 to 20 21 to 40 41 to 60 >60 0 20 40 60 80 100 canopy cover (%) p e rc e n t o f a re a dry forests wet forests sa-prk sa-fir abgr thpl psme pipo 0 10 20 30 40 50 60 potential vegetation type p e rc e n t o f a re a a b appendix b. difference in predicted body fat (%) for female moose captured in northeastern wa, usa during early december (2013, 2014, and 2016) between the original alaskan moose equation which integrated body size in the development of the prediction equation (stephenson et al. 1998), and one where maxfat is scaled for surface area prior to predicting body fat (cook et al. 2010); the dashed line is a 1:1 reference line. 6 8 10 12 6 8 10 12 body fat (%) -stephenson et al. 1998 b o d y f a t (% ) - s c a le d l iv in d e x appendix c. using data from this study, we demonstrate that moose with no measurable rump fat can range from 0 to 5.7% body fat, a range in body fat no different from rump fat measurements ranging from > 0 to 2.75cm. representing this range with the single value of zero introduces a bias into the data set; the greater the proportion of animals with no measurable rump fat, the more biased the nutritional condition data. 0 2 4 6 8 10 12 0 0.5 1 1.5 2 2.5 3 3.5 body fat (%) m a x f a t ( c m ) 6 percentage point range in body fat, no change in maxfat 6 percentage point range in body fat, 2.75 cm range in maxfat _hlk57024215 _hlk56511840 _hlk57036488 _hlk65750986 _hlk66084038 _hlk66084410 _hlk57121853 _hlk57027957 alces37(2)_303.pdf f:\alces\vol_39\p65\3932.pdf alces vol. 39, 2003 jaren et al. – moose and integrated ecosystem management 1 moose in modern integrated ecosystem management – how should the malawi principles be adapted? vemund jaren1, a. r. e. sinclair2, reidar andersen3, kjell danell4, chuck schwartz5, rolf o. peterson6, r. terry bowyer7,8, and göran ericsson4 1directorate for nature management, n-7485 trondheim, norway; 2centre for biodiversity research, 6270 university boulevard, university of british columbia, vancouver, bc, canada v6t 1z4; 3department of biology, norwegian university for science and technology, n-7491 trondheim, norway; 4department of animal ecology, swedish university of agricultural sciences, se-901 83 umeå, sweden; 5interagency grizzly bear study team, u.s. geological survey, northern rocky mountain science center, montana state university, bozeman, mt 59717, usa; 6school of forestry and wood products, michigan technological university, houghton, mi 49931, usa; 7institute of arctic biology, department of biology and wildlife, university of alaska, fairbanks, ak 99775, usa abstract: under the implementation of the convention on biological diversity, a special emphasis has been put on an integrated ecosystem approach. some of the “malawi principles” state that management objectives are a matter of societal choice, and that management should be decentralized to the lowest appropriate level. a key feature of the ecosystem approach includes conservation of ecosystem structure and function on a long term basis, while seeking an appropriate balance between conservation and use of biodiversity. the role of moose in the ecosystem and how the malawi principles can be adopted in moose management were a focus of the 5th international moose symposium. all invited speakers and session chairs were asked to provide a brief summary of how they considered the malawi principles to relate to the topic of their respective papers or sessions at the symposium. those summaries are given in this paper. alces vol. 39: 1-10 (2003) key words: alces alces, carnivores, ecosystem, local involvement, malawi principles, management, society, vegetation “moose in modern integrated ecosystem management” was the main conference theme for the 5th international moose symposium in norway 2002. in the last plenary session at the symposium, two papers specifically addressing this topic were presented, followed by a general discussion. as an introduction to the discussion, all invited speakers and session chairs were asked to give a brief summary of how they considered the so-called malawi principles to relate to the topic of their respective sessions at the symposium. their valuable contributions to highlight this issue are presented below. the malawi principles in a workshop organized in the african country malawi in january, 1998, and submitted to the 4th conference of the parties of the convention on biological diversity (unep/cbd/cop/4/inf.9), the following 12 principles/characteristics of an ecosystem approach to biodiversity management were identified: 1. management objectives are a matter of societal choice. 2. management should be decentralized to the lowest appropriate level. 3. ecosystem managers should consider the effects of their activities on adja8present address: department of biological sciences, idaho state university, pocatello, id 83209, usa. moose and integrated ecosystem management – jaren et al. alces vol. 39, 2003 2 cent and other ecosystems. 4. recognizing potential gains from management there is a need to understand the ecosystem in an economic context, considering, for example, mitigating market distortions, aligning incentives to promote sustainable use, and internalizing costs and benefits. 5. a key feature of the ecosystem approach includes conservation of ecosystem structure and functioning. 6. ecosystems must be managed within the limits to their functioning. 7. the ecosystem approach should be undertaken at the appropriate scale. 8. recognizing the varying temporal scales and lag effects which characterize ecosystem processes, objectives for ecosystem management should be set for the long term. 9. management must recognize that change is inevitable. 10. the ecosystem approach should seek the appropriate balance between conservation and use of biodiversity. 11. the ecosystem approach should consider all forms of relevant information, including scientific and indigenous local knowledge, innovations, and practices. 12. the ecosystem approach should involve all relevant sectors of society and scientific disciplines. a. r. e. sinclair: commentary on the malawi principles the malawi principles embody two fundamental principles. firstly, all stakeholders should be involved in the process of developing conservation management plans. underlying this is the idea that most of the biodiversity in the world lies in tropical regions that are owned and administered by developing countries. these countries must consider the development and advancement of their peoples and unless this is taken into account conservation problems will be ignored. in particular, we must pay attention to the mismatch where those that gain benefit from conservation are not the same peoples as those that bear the costs of conservation. secondly, the malawi principles recognize that the ecosystem is the unit of management rather than individual species. traditionally, conservation has focused on single species, particularly those that are endangered. yet all these species require habitat and other resources, often the loss of such resources being reason for the conservation problems, and so it is these resources that need to be conserved within the context of the whole ecosystem. it is no coincidence that these principles have been laid out at the conference in malawi, one of those developing countries that are confronting the trade-off between development and environment. whilst recognizing the validity of these malawi principles, we should not ignore the constraints and limitations that still have to be addressed. first, we must recognize that there are problems with time scale. the principles do refer to large spatial scales and long time periods. nevertheless, they do not recognize that poor peoples do not conserve their resources because they discount the future, often the very near future of a few years or even a few months. if a peasant farmer has to cut down a tree for fuel to cook tomorrow’s meal he is in no position to consider the problem of conserving the forest for next year let alone 10 or 100 years. poverty means that people do not save for the future. in addition, the principles have not addressed the problem of benefits for future generations. indeed, future generations are unrepresented in any discussions when it comes to natural resource economics. finally, the principles stress the need to consult and incorporate the wishes of local, indigenous peoples. however, such peoples, at the scale of villages, invariably think alces vol. 39, 2003 jaren et al. – moose and integrated ecosystem management 3 at the very small scale. they consider only their own local needs. thus, the principle of large scale is in direct conflict with the principle of local involvement. secondly, the principles have not built in any mechanism for enforcement. in essence, these principles are a form of social contract. developing peoples obtain some benefit in return for conserving the resources that benefit the whole world. what happens if they renege on their commitments? they could obtain the benefit and then exhaust the biodiversity under their control. as the principles stand, there is no penalty for noncompliance in the social contract. experience has shown that without such a penalty these social contracts have not worked and are unlikely to work. we must face the hard facts, no matter how unpalatable they are, that humans act in their own short-term selfish interests so that if there is no inducement to comply with a contract they will not do so. thirdly, the principles purport to incorporate the concept of the ecosystem. however, this concept remains vague even for biologists and many components have yet to be defined. for example, what are the bounds of an ecosystem? without some recognition of this boundary it can be tailored to suit the needs of anyone that wishes to exploit a system. in british columbia this term has been used to allow mining and logging within provincial parks, areas previously protected from exploitation, under the guise of ecosystem conservation: the new rationale is that now conservation must take into account the combined area of park and external regions together. this has allowed exploiters to obtain greater resources inside the park while conveniently ignoring the costs of greater conservation outside the park. another problem lies in understanding ecosystem function. this is a term that is now frequently used in the debate on biodiversity processes and yet we do not know precisely what this term means. it can mean a number of different things and sometimes they are in conflict with each other. thus, ecosystem function can refer to productivity, nutrient cycling, or hydrology, but it can also refer to stability, resilience, and robustness. promoting high productivity could reduce resilience. we have to be precise in what we mean by ecosystem function in implementing conservation. finally, the term biodiversity itself encompasses all of living matter. as such it is not very useful. we have to address for practical purposes particular components of biodiversity that will be essential in terms of the functioning of ecosystems. however, we do not yet know what those components are. do we, for example, pay attention to the large mammals which can act as ‘umbrella species’, thus protecting all those that fall within their large scale habitats or, in contrast, do we protect the microorganisms of the soil because they determine all processes that feed up the trophic levels to the large mammals? we do not know the answers to these questions yet. reidar andersen: malawi principles and moose management – challenges in north america, eastern europe / northern asia, and fennoscandia the malawi principles focus on an ecosystem approach, where management objectives are a matter of societal choice and where management should be decentralized to the lowest possible level. the relevance of these principles varies throughout the moose distribution area. in many parts of eastern europe and northern asia, moose populations have declined over the last decade. the decline is most pronounced in areas with high human population density, reflecting the fact that poaching and overexploitation is the main cause of decline. in some cases deterioration of moose moose and integrated ecosystem management – jaren et al. alces vol. 39, 2003 4 habitat may also be important. clearly, in a situation where hunting is not a pure recreational activity, and where some hunters are close to a subsistence level, each individual hunter has difficulty in seeing the benefits of ecosystem management. thus, the malawi principles seem to be of minor relevance, yet the goal local managers should strive for. in north america and fennoscandia, moose populations are managed at the lowest appropriate level; often at the level of forestry owners and hunters (sweden) or local management within smaller districts (norway and finland). in many areas, vehicle accidents are on a level that needs consideration. if management objectives should be a matter of societal choice, frightened car drivers should also have a voice in setting the quotas. two of the malawi principles state that ecosystems must be managed within the limits of their functioning, and at the appropriate scale. first, what is appropriate, and how could managers decide what is the limit of ecosystem function? in several places, high density moose populations are thought to have led to reduced biological diversity, still we are lacking a clear understanding of how large herbivores, like the moose, are affecting their habitats and their functioning. one malawi principle states that management must recognize that change is inevitable. change not only in management objectives, but also in population densities. while most people can agree in the principles of optimal production and yield in relation to the carrying capacity of the habitats and damage to other management operations, we also need to realize that populations of large herbivores seldom are stable in numbers over long periods of time. in some areas, managers and others ask themselves; will large herbivores in areas lacking large carnivores stabilize at a certain density, or will large herbivores grow beyond their carrying capacity, overgraze their food resources, and create a highly unstable situation with large temporal variation in numbers? in that case, will reintroduction of l a r g e c a r n i v o r e s b e a b l e t o c r e a t e stabilization? or, will hunters be able to create this stability? obviously these two scenarios have profound effects on community structure. one major conclusion from analysis of several long-term individual based population studies of large herbivores is that the population dynamics of ungulates in predator-free environments are strongly influenced by a combination of stochastic variation in the environment and population density. both factors operate through changes in life history traits, correlated with variation in body weight, which generates delays in the response of the population to changes in the environment. in such cases it is claimed that in the absence of predation, a stable equilibrium between an ungulate population and its food resources is therefore unlikely to exist. consequently, no management regime should be judged in relation to the stability of moose numbers. kjell danell: “moose in modern, integrated ecosystem management”, in relation to the malawi principles. some c o n c l u s i o n s f r o m t h e s e c t i o n o n trophic interactions between moose and vegetation management objectives are a matter of societal choice. — the food of moose constitutes, to a large extent, woody plants in boreal forests. these plants are valuable to at least 3 interest groups: moose hunters, timber industry, and nature conservation interests. for moose hunters, more food plants of better quality can mean more moose. for the timber industry, damage to scots pine (pinus sylvestris) is especially negative, both in the short and long term, alces vol. 39, 2003 jaren et al. – moose and integrated ecosystem management 5 because moose can cause losses in both quantitative and qualitative terms. for nature conservation, heavy browsing pressure can depress populations of preferred woody species, such as aspen (populus spp.) and rowan (sorbus spp.). these deciduous species have a value as such, but also because they are host plants for a wide array of plant and invertebrate species. important tasks for the future are to determine how to bring these interest groups together and to find a common procedure for discussion of the appropriate moose population level. how to find a balance between different values? private interests versus interests of society. management should be decentralized to the lowest appropriate level. — which is the appropriate level for management? a small area may be good for solving conflicts between landowners, because then few landowners are involved, but a large area is often needed because moose migrate. ecosystem managers should consider the effects of their activities on adjacent and other ecosystems. — so far, moose management in fennoscandia has been very much a single species approach, especially in a situation without large predators. we need to move away from single species management plans and strategies towards ecosystem management if we are to follow the malawi principles. a key feature of the ecosystem approach includes conservation of ecosystem structure and functioning. — for moose we need more knowledge of its effect on ecosystem structure and functioning. at high densities, moose can have a much more dramatic impact on landscape features and plant succession than we generally believe. in southern sweden, high densities of ungulates cause pine forests to be replaced by spruce forests. so far, a great effort has been directed to understand the population dynamics of moose. however, we need to broaden our research approaches. chuck schwartz: application of the malawi principles to the section on trophic interactions between moose and carnivores first, science and the application of science is a major principle. all the papers presented at the session furthered our understanding of predator-moose relationships and were very much in line with the principles. second, the principles suggest that management be reduced to the lowest level when and where possible. with rare carnivores, the opposite has occurred. the endangered species act and the bern convention are national or international laws. they are basically top-down control rather than bottom-up. however, it was pointed out by one of the members of the audience that the principle states that authority should seek the lowest practical level. this, in the case of large rare carnivores, may in fact be at the national or international level. finally, i summed up with the principle of ecosystem function and management and the fact that large carnivores are a major part of such function. society has deemed it appropriate to retain or restore large carnivores and maintain healthy systems. consequently, moose managers must view their role differently than simply providing sustainable harvests to hunters. rolf o. peterson: wolves, moose, and the malawi principles the malawi principles relate, in general, to the difficult problem of conserving the diversity of life, which underlies the human experience. predation by large carnivores, in particular the gray wolf, relates to the specific principles calling for maintenance of ecosystem structure and function as well as those recognizing the enormous moose and integrated ecosystem management – jaren et al. alces vol. 39, 2003 6 importance of local human attitudes. i view the gray wolf as the most significant “stage manager” in the evolutionary “theatre” in which moose developed. from basic data on the age structure and pattern of moose vulnerability to wolf predation, we can infer that wolf predation has been the dominant agent of natural selection for moose throughout its extensive geographic range. indeed, this notion was implied by charles darwin in the origin of species. based on ongoing simulation work shortly before his death, a. b. bubenik (personal communication) concluded that optimal social balance in moose populations was attained by ageand sex-specific mortality patterns similar to those produced by wolf predation. how can we maintain natural selective forces in a human-dominated world? remarkably, gray wolves have staged a major comeback in many parts of their historic range, attributable to changing public attitudes and formal recovery programs. yet this has accentuated a mismatch between those people who bear the cost of living with wolves and those who derive the benefits. wolf recovery prompts divisiveness in human responses. the addition of wolves can greatly complicate human affairs in rural landscapes, particularly where human presence is pervasive, so the greatest hope for recovery of large carnivores lies in extensive wildland habitats. fear of the wolf remains a key influence in the public mind. information and education programs should be management priorities anywhere wolf recovery is ongoing or contemplated, as local people and their beliefs will greatly affect the future of the wolf. ecosystem processes include those that are both fast and slow, at both large and small spatial scales. the inherent nature of wolf predation, as an intensive, culling agent operating over local scales that are measured in hundreds of square kilometers, is that of a highly dynamic element operating at large scales in an ecosystem. the responses that are demanded from human societies, operating in relatively slow bureaucracies at local levels, will pose a particularly interesting management challenge. r. terry bowyer: relationship of genetics, physiology, diseases, and parasites of moose to the malawi principles presentations in our technical session involved the malawi principles primarily in relation to effects of moose (alces alces) on biodiversity, but also with respect to societal needs for consumptive uses of these large herbivores. the evolutionary history of a species holds implications for understanding its current distribution, but also its likelihood of persisting or expanding. thus, information on the phylogeography of moose (hundertmark et al. 2002, 2003) and whether speciation has occurred (boeskorov 1996, udina et al. 2002) are data essential for the wise management of these unique large mammals. for instance, will all subspecies of moose respond in a similar manner to environmental constraints such as severe weather or various harvest regimes (sæther 1997)? indeed, how these large mammals cope with climatic variation (bowyer et al. 1998, lenart et al. 2002) could have farreaching implications for their population dynamics and subsequent interactions with their environment, which holds implications for biodiversity. differences in body mass among subspecies may result in variation in life-history strategies (sensu keech et al. 2000) of moose that may necessitate different management tactics to meet societal goals. clearly, adaptations of moose to their environment can lead to morphological differences that may be useful in their management, including antler characteristics (bowyer et al. 2001, engan 2001), and alces vol. 39, 2003 jaren et al. – moose and integrated ecosystem management 7 digestive physiology (kochan 2001). knowledge concerning how these morphological and physiological characteristics differ among populations or subspecies will promote a more complete understanding of how moose are adapted to boreal environments, and thereby enhance opportunities for their sound management. society demands the ethical treatment of mammals (animal care and use committee 1998). consequently, the manner in which researchers and managers capture, restrain, and study moose and other large herbivores affects the opportunity to meet the needs of people for subsistence or recreational uses. capture methods, then, must be humane, effective, and result in limited mortality of study animals and minimal risk to humans (arnemo and søli 1994, arnemo 1995). finally, the phylogeography of moose has implications for how moose interact with and affect their forage plants (bowyer et al. 1997). clearly, moose can drive successional patterns in boreal forests (pastor and naiman 1992, pastor et al. 1993). high densities of moose can have deleterious effects on biodiversity of other vertebrates (berger et al. 2001). nonetheless, intermediate densities of those large herbivores can promote nutrient cycling in boreal forests (molvar et al. 1993) and enhance rates of decomposition in aquatic systems (irons et al. 1991). the ability of moose to alter successional patterns and affect trophic cascades makes them a keystone species (simberloff 1998). accordingly, moose offer a unique opportunity for single-species management to directly affect biodiversity to achieve results that benefit those components of society seeking consumptive uses of moose and those hoping to enhance biodiversity. göran ericsson: human dimensions of moose management and the malawi principles the human dimension of moose management (i.e., how people value moose, how people want moose to be managed, and how people are affected by or affect moose including management decisions [ericsson 2003]), is a central part of the malawi principles with respect to natural resource management in the boreal region. all wildlife management is based on human values, with “management” itself being a human construct (decker et al. 2001). thus, our societies manage moose (and other wildlife) because we implicitly view them as a resource. central to the malawi principles is that the decision-making power should be handed down to the lowest possible level. however, that poses a great challenge for traditional natural resource management. not only local groups want to have a say today in natural resource management. most importantly, several national and international stakeholder groups want to have a say, and political oversight of management has increased. moreover, several international agreements and conventions also regulate natural resource management (e.g., moose and other large mammals). consequently, natural resource management now has to pay more attention to non-consumptive use as well. moose management is no longer just about setting harvest quotas. at the same time, the consumptive use of moose is still of central importance, making moose management more complex, as non-local interest tends to be centralized around nonconsumptive issues. therefore, if we decentralize moose management to the lowest appropriate level according to the malawi principles, we will most likely see an increased tension between local people’s interest which tends to focus on the consumptive aspects of moose management moose and integrated ecosystem management – jaren et al. alces vol. 39, 2003 8 and non-locals’ interests. because local people normally represent a small minority of any western urbanized society, the democratic process easily, but often unintentionally, overruns their interest. however, western societies now pay more attention to local groups as well (ericsson and heberlein 2002). recent data from sweden supports this and show that non-locals support the right of local people to have the final say, even in controversial management issues. when asked “i think that the local people should have the final say in large carnivore management”, 55 % of the swedish public said that they supported this (ericsson and heberlein 2002). thus, it suggests that implementation of the malawi principles in moose management is not a controversial issue. instead, the vast majority most likely support local moose management. human dimensions are still a “management challenge” in moose management (crichton et al. 1998). during the sessions at the 5th international moose symposium it became evident that human dimensions so far is mostly ad-hoc to “pure” moose projects. this is most unfortunate for successful implementation of the malawi principles and a wise, sustainable use of moose as a multidimensional resource. thus, we urgently need to involve socioeconomic expertise from the beginning when we deal with a “moose problem”, not afterwards and probably most important, don’t set their agenda. human dimensions of moose management, and a successful implementation of the malawi principles, are far too complicated to apply a single discipline solution. local management is here to stay, like it or not but “john doe” and “sven svensson” demand a voice even in moose management and research today. “since the early 1970s, citizen participation has been emphasized in natural resource management decision making” (lauber and knuth 1999). now we face the challenge to make this work in moose management. remember that management is a human construct based on human values thus we need to involve people in the decision and implementation of moose management. a. r. e. sinclair: some concluding remarks in conclusion, we must be honest with ourselves in recognizing that there are some fundamental problems that will either prevent the implementation of the malawi principles or will allow them to be distorted and corrupted if they are not addressed. secondly, we must be flexible in implementation through the use of adaptive management. sustainable use of resources requires flexible harvest quotas rather than constant numbers. thirdly, we have to implement programs to monitor any conservation initiatives to assess whether they are meeting their objectives. finally, we should remember that had we known what was present on many of the continents in the 1700s or 1800s we would now be in a better position to know what to conserve and how to conserve. in a hundred years time, future generations may be wishing that we had been wiser at the current time. it is certain that what we are currently doing is imperfect at best and we must be constantly asking ourselves this question: what are we doing wrong now that will impact future generations? references animal care and use committee. 1998. guidelines for the capture, handling, and care of mammals as approved by the american society of mammalogists. journal of mammalogy 74:1416-1431. arnemo, j. m. 1995. immobilization of free-ranging moose (alces alces) with medetomidine-ketamine and remobilization with atipamezole. rangifer 15:19-25. alces vol. 39, 2003 jaren et al. – moose and integrated ecosystem management 9 , and n. e. søli. 1994. injection darts containing drugs—a potential health hazard for several years? norsk veterinaer 106:306-308. berger, j., p. b. stacey, l. bellis, and m. p. johnson. 2001. a mammalian predator-prey imbalance: grizzly bear and wolf extinction affect avian neotropical migrants. ecological applications 11:229-240. boeskorov, g. g. 1996. chromosomal differentiation of moose (alces ) . doklady akademii nauk 348:275-278. bowyer, r. t., k. m. stewart, j. g. kie, and w. c. gasaway. 2001. fluctuating asymmetry in antlers of alaskan moose: size matters. journal of mammalogy 82:814-824. , v. van ballenberghe, and j .g. kie. 1997. the role of moose in landscape processes: effects of biogeography, population dynamics, and predation. pages 265-287 in j. a. bissonette, editor. wildlife and landscape ecology: effects and patterns of scale. springer-verlag, new york, new york, usa. , , and . 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79:1332-1344. crichton, v. f. j., w. e. regelin, a. w. franzmann, and c. c. schwartz. 1998. the future of moose management and research. pages 655-663 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. decker, d. j., t. l. brown, and w. l. siemer. 2001. evolution of peoplewildlife relations. pages 3-21 in d. j. decker, t. l. brown, and w. f. siemer, editors. human dimensions of wildlife management in north america. the wildlife society, bethesda, maryland, usa. engan, j. h. 2001. changes in the relationship between palmate and corvine antlers in moose (alces alces) in southeastern norway. alces 37:79-88. ericsson, g. 2003. of moose and man: the past, the present and the future of human dimensions in moose research and management. alces 39:11-26. , and t. a. heberlein. 2002. attityder till varg och vargjakt i sverige. slu kontakt 14. isbn 91-576-61324. hundertmark, k. j., r. t. bowyer, g. f. shields, and c. c. schwartz. 2003. mitochondrial phylogeography of moose (alces alces) in north america. journal of mammalogy 84:718-728. , g. f. shields, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. sc h w a r t z. 2002. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and popul a t i o n e x p a n s i o n . m o l e c u l a r phylogenetics and evolution 22:375-387. irons, j. g., j. p. bryant, and m. w. oswood. 1991. effects of moose browsing on decomposition rates of birch leaf litter in a subarctic stream. canadian journal of fisheries and aquatic sciences 48:442-444. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64:450-462. kochan, t. i. 2001. seasonal adaptation of metabolism and energy in the pechora taiga moose alces alces. journal of evolutionary biochemistry and physiology 37:246-251. lauber, t. b., and b. a. knuth. 1999. measuring fairness in citizen participamoose and integrated ecosystem management – jaren et al. alces vol. 39, 2003 10 tion: a case study of moose management. society and natural resources 11:19-37. lenart, e. a., r. t. bowyer, j. ver hoef, and r. w. ruess. 2002. climate change and caribou: effects of summer weather on forage. canadian journal of zoology 80:664-678. molvar, e. m., r. t. bowyer, and v. van ballenberghe. 1993. moose herbivory, browse quality, and nutrient cycling in a n a l a s k a n t r e e l i n e c o m m u n i t y . oecologia 94:472-479. pastor, j., b. dewey, r. j. naiman, p. f. mcinnes, and y. cohen. 1993. moose browsing and soil fertility in the boreal forests of isle royale national park. ecology 74:467-480. , and r. j. naiman. 1992. selective foraging and ecosystem processes in the boreal forests. american naturalist 134:690-705. sæther, b.-e. 1997. environmental stochasticity and population dynamics of large herbivores: a search for mechanisms. trends in ecology and evolution 12:143-149. simberloff, d. 1998. flagships, umbrellas, and keystones: is single species management passé in the landscape era? biological conservation 83:247-257. udina, i. g., a. a. danilkin, and g. g. boeskorov. 2002. genetic diversity of moose (alces alces l.) in eurasia. russian journal of genetics 38:951957. alces34(1)_47.pdf 4204(33-39).pdf alces vol. 42, 2006 belant et al. moose distribution and human development 33 moose distribution relative to human development in a national park jerrold l. belant1,2, jonathan a. paynter1, kenneth e. stahlnecker1,3, and victor van ballenberghe4 1national park service, denali national park and preserve, p.o. box 9, denali park, ak 99755, usa; 4 alces alces) key words: alces alces alces alces), rangifer tarandus ursus arctos ovis dalli 2 3 moose distribution and human development – belant et al. alces vol. 42, 2006 34 study area picea salix carex salix betula 2 methods alces vol. 42, 2006 belant et al. moose distribution and human development t p p results univariate analyses of resource selection t p t p t p t p t p t p t p = t p = t p moose distribution and human development – belant et al. alces vol. 42, 2006 36 2 p 2 = p 2 p = 0.002). multivariate analyses of resource selection p p p discussion sd sd sd 62 62 2 3 32 0 alces vol. 42, 2006 belant et al. moose distribution and human development se 2 p se 2 p 2.3 0.32 0.22 moose distribution and human development – belant et al. alces vol. 42, 2006 acknowledgements references belant, j. l., t. w. seamans, c. p. dwyer. l. a. tyson burson, s. l., iii, j. l. belant, k. a. fortier, w. c. tomkiewicz, iii cole, e. k., m. d. pope r. g. anthony. dalle-molle, j j. van horn in hosmer, d. w. s. lemeshow johnson, d. r m. c. todd knight, r. l k. j. gutzwiller, d.c., usa. mace, r. d. j. s. waller. manly, b. f. j., l. l. mcdonald, d. l. thomas mccullough, d. r mech, l. d miquelle, d. g., j. m. peek, v. van ballenberghe morrison, j. r., w. j. de vergie, a. w. alldredge, a. e. byrne w. w. andree. risenhoover, k. r sas institute. schultz, r. d j. a. bailey alces vol. 42, 2006 belant et al. moose distribution and human development singer, f. j j. b. beattie. tracy, d. m. van ballenberghe, v. d. g. miquelle alces35_1.pdf 4215(115-131).pdf alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 115 long-distance dispersal and population trends of moose in the central united states justin d. hoffman 1, hugh h. genoways1, and jerry r. choate2 1school of natural resources and university of nebraska state museum, w436 nebraska hall, university of nebraska-lincoln, lincoln, ne 68588, usa; 2sternberg museum of natural history, fort hays state university, hays, ks 67601, usa abstract: dispersal is a basic feature of the natural history of moose. most information about moose dispersal pertains to short-distance movements because long-distance movements are uncomand north dakota. this may have contributed to several long-distance dispersal events for moose that recently were reported in the central united states. these dispersal events provide an opportunity to investigate both the causes and the biological implications of this rare phenomenon. herein, we review long-distance dispersal events based on information obtained from a variety of sources. dispersal routes that could be measured included two with minimal distances of 1,511 and 367 km, plus several others that were shorter. these dispersal events and recent evidence of moose reproducing outside the current range of the species could be the result of increasing population trends of moose in the central united states. we suggest that the dispersing moose are founder individuals that are dispersing naturally from established populations in search of suitable habitats and mates in areas to the south. we hypothesize that this type of geographic range expansion is similar to that of moose when they dispersed across north america during the early holocene. as moose continue to move south, wildlife managers should be aware of habitats within their respective states that might sustain populations of moose. alces vol. 42: 115-131 (2006) key words: alces alces, central united states, dispersal, moose, range expansion dispersal is a basic feature of the life history of most species. local dispersal occurs within established populations of a species. it is important because it regulates population is movement into suitable habitats adjacent to the currently occupied range of the species. this type of dispersal can lead to a gradual expansion of the range of a species. longdistance dispersal differs from diffusion dispersal in that it may occur across large areas of unsuitable habitat. this type of dispersal can result in the discovery and colonization of isolated unoccupied habitats, and potentially the rapid expansion of a species’ geographic range (ricklefs and miller 2000). long-disthe moose (alces alces) is the largest member of the family cervidae and occurs primarily in the boreal forests of north america overall geographic distribution of moose expanded in the late pleistocene, with individuals dispersing from eastern asia into north american via the bering land bridge approximately 14,000-11,000 years ago (hundertmark et al. 2002). at that time, glaciers in north america had begun to retreat, producing an abundance of the early successional habitats favored by moose and facilitating rapid expansion of the geographic range of the moose across north long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 116 are highly mobile with a strong propensity for ity, they can disperse hundreds of kilometers in a short period of time. dispersers are often, but not exclusively, young individuals (hunand juveniles tend to disperse only a short distance from their natal range (gasaway et al. tend to disperse farther from their natal range than juvenile females, and the percentage of overlap of home ranges between juvenile and dam is less for juvenile males than for juvenile although moose usually disperse short distances, long-distance dispersal events have been documented. for example, mytton and for 4 young moose and a 250-km dispersal distance for a young bull moose in alberta, adult cow that dispersed a distance of 177 km in southern alaska. these accounts represent dispersal events within the geographic range of moose. to our knowledge, there have been few published accounts of longdistance dispersal of moose outside their that a dead moose was found approximately 500 km north of traditional moose range in the northwest territories, canada. in the central tracked the dispersal route of a bull moose from minnesota through iowa and into missouri. they estimated that the total distance recently, there have been several instances of moose dispersing outside their known range suggested that moose are still expanding their geographic range into areas that they have not occupied since the last glaciation. these long-distance dispersal records are important because they document a phenomenon that rarely is observed and potentially can provide insight into moose movements and biogeography. as discussed above, technical reports on long-distance dispersal events by moose outside their normal geographic range are few. most such information is found in popular media and newspaper articles, which, by themselves, provide little biological information. herein, we present a summary of recent trends in moose populations in the central united states and of long-distance dispersal events by moose in this region. we discuss possible explanations for long distance dispersal by moose and potential biological implications methods information from primary literature, government documents, and minnesota and north dakota moose harvest reports were used to summarize recent population trends of moose in the central united states. to describe long-distance moose dispersal, we collected information regarding moose sightings and potential dispersal routes for north dakota, south dakota, minnesota, iowa, nebraska, missouri, kansas, oklahoma, and texas. the sources consulted included newspapers, popular journals, books, primary literature, and communication with employees of the kansas department of wildlife and parks (kdwp), nebraska game and parks commission (ngpc), north dakota game and fish department (ndgfd), and south dakota game, fish, and parks (sdgfp). we such as newspapers and popular journals, to document the movements of moose because moose are not likely to be confused with any other species. in this paper, we consider the “historical” geographic range of moose in the central united states to be northeastern minnesota alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 117 range consists of the “historical” range plus areas outside this region occupied in subsequent years as part of the recent expansion of moose in the central united states. we calculated dispersal distances from localities at the southern edge of the current geographic because it is likely that most long-distance dispersers came from populations in those regions. we obtained potential dispersal routes by connecting chronologically ordered localities and calculated dispersal distances considered to be minimal distances travelled because they were measured as straight lines between localities. results population status when the north-central united states was settled by european immigrants and their descendents, moose occurred in northern minnesota and northeastern north dakota. been extirpated from north dakota and moose continued to inhabit the boreal forests of northeastern minnesota although they undoubtedly were less abundant than they had been in pre-settlement periods. because of this decline in moose numbers, minnesota closed numbers began to rise and moose began to reclaim their former range in northwestern gan moving back into north dakota and, by from there, they spread westward along the canadian border to the turtle mountains and southward along the red river valley, where they inhabit the rugged lands of prehistoric resident moose populations again occurred because of the recent increase of moose, both north dakota and minnesota have established hunting seasons. minnesota reopened two units were available to hunters, in the northwest and the northeast, with hunting harvested in minnesota continued to increase with considerably more moose being taken in the northwestern unit (minnesota department northwestern unit and 737 moose were harvested, as compared with 523 permits sold in the northeastern unit and 442 moose harvested (minnesota department of natural resources was noticed in the northwestern unit. as a result, restrictions on moose hunting were implemented throughout the northwestern northwest was closed. since then, population numbers have remained low in the northwest stable in the northeast (minnesota department of natural resources 2005). north dakota implemented a moose huntharvest rates and hunting units have increased. today, hunting is permitted in the pembina hills, the turtle mountains, and the red river valley, which encompass the north-central, northeastern, and extreme eastern parts of the state extending as far south as the south areas where moose occur in north dakota, population estimates continue to be highest in the pembina hills area, followed by the turtle mountains, and then the red river valley long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 of moose in north dakota are more-or-less dispersal moose recently have been reported outside the current geographic range of the species in the central united states (fig. 1). some of those individuals dispersed over long distances. the most notable dispersal was undertaken in south dakota and eventually dispersed as was of 2 young bulls seen near dell rapids, separated, because there were no reports of a second moose beyond dell rapids. in sepacross southeastern south dakota, where it was reported near the towns of alexandria, crossed the missouri river near avon, south a few days later in page, nebraska (omaha southward across nebraska, passing near the towns of elba, palmer, and chapman (hornof grand island and the platte river for the remainder of the year and was spotted near saronville, approximately 45 km southeast of reported about 101 km to the southwest near agra, kansas, in late february (kleinschmidt account of the moose’s movement in kansas. the moose remained in north-central kansas, reportedly being seen near kirwin reservoir on the north fork of the solomon river (which is just south of agra), for the remainder of it was seen west of stockton, approximately 45 km southwest of agra. the moose refrom the summer heat at webster reservoir on the south fork of the solomon river (which is just west of stockton), until september, when it was seen approximately 104 km farther south near rush center. from rush center, the moose quickly moved through south-central state reference locality day month south dakota dell rapids (1) 15 sept alexandria (2) 27 sept dimock (3) sept parkston (4) sept avon (5) 1 oct nebraska page (6) 5 oct elba (7) 13 oct 16 oct oct grand island (10) oct phillips (11) 30 oct trumball (12) 12 nov harvard (13) 26 dec saronville (14) nelson (15) 15 guide rock (16) 21 kansas agra (17) 22 feb sept kinsley (20) 20 sept ashland (21) 22 sept englewood (22) sept texas perrytown (23) nov dalhart (24) nov kansas ulysses (25) dec sublette (26) 5 feb table 1. time and locality information for the dispersal of a moose through south dakota, nebraska, kansas, oklahoma, and texas. numbers in paraenthesis correspond to those shown in figure 1. alces vol. 42, 2006 hoffman et al. long-distance moose dispersal fig. 1. records of occurrences and potential dispersal routes of moose in the central united states. symbols connected by solid lines represent potential dispersal routes described in this study. the symbols connected by a dashed line represent a dispersal route described by bowles and gladfelter and diamond represent instances in which a moose remained in the area for an extended period of time. numbers correspond to locality information listed in tables 1 and 2. long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 120 kansas, passing near kinsley, ashland, and any reports of the moose having been seen in oklahoma, it presumably passed through the oklahoma panhandle because it was next seen near perrytown, texas, in november and later location, the moose reversed directions and returned to southwestern kansas, where it area, a local veterinarian obtained permission from kdwp to tranquilize the moose. she claimed that the moose was in poor health and in need of medical attention. on 5 febnear sublette, kansas, and transported to the veterinarian’s facility (associated press whether or not the moose was, in fact, sick. after treating the animal, it apparently was released somewhere in colorado (the exact served near dell rapids, south dakota, the moose travelled an estimated straight-line distance of 1,511 km. the time it took to travel area from which this moose dispersed, and thus the total distance it moved, is impossible to know. however, the distance this moose travelled as measured from the southern edge of the current geographic range of the species in iowa and missouri, similar dispersal events by moose have been documented. distance dispersal by a bull moose that began in southwestern minnesota, continued through iowa, and ended in the vicinity of bowling green, missouri, near the mississippi river (fig. 1). a similar dispersal event took place in the same area a few years later (fig. 1). in observed in south-central minnesota near st. reported in iowa near the town of fertile in of sightings, refer to table 2). from fertile, the moose travelled directly south across iowa, passing near clear lake, thornton, latimer, observed in iowa, the moose was reported from there, the moose veered southeastward and passed near pleasantville, northeast of in early december was reported for the last time in iowa in the vicinity of georgetown and table 2. time and locality information for the dispersal of a moose through minnesota, iowa, and missouri. numbers in paraentheses correspond to those shown in figure 1. state reference locality day month minnesota oct oct iowa 2 nov clear lake (30) 2 nov thornton (31) nov latimer (32) nov alden (33) 7 nov ames (34) nov nevada (35) 11 nov ne of des moines (36) nov altoona (37) nov 30 nov 30 nov 4 dec missouri omaha (41) 31 dec e of pollock (42) 20 25 river (44) 4 feb dalton (45) feb alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 121 near fertile, it took a little over a month for the bull moose to travel 335 km across iowa. the moose then moved into missouri, where it was spotted near omaha in late december again until approximately a month later, when it was seen east of pollock, which is southwest moose was observed near the junction of highthere, it moved southwestward into chariton county, where it reportedly was seen near the the last report of this moose in missouri was ing observed in south-central minnesota, the moose travelled an approximate distance of 650 km. because the exact dates when this average dispersal speed for the entire trip was not calculated. however, we were able to calculate an average dispersal speed from in missouri, near omaha. the approximate distance travelled by the moose was 365 km in 60 days, giving it an average dispersing from which this moose dispersed is unknown; however, the distance this moose travelled from the southern edge of the geographic range of the species was about 600 km. several other, shorter dispersal events south dakota, nebraska, iowa, and minnesota (fig. 1). in south dakota a young bull moose was reported near mobridge (in the north-central part of the state), in november nebraska, a bull moose was reported in the fall of 2000 near verdigre (tom welstad, ngpc, personal communication). by december the moose had moved farther south to the vicinity had settled on the elkhorn river near battle creek. for the next few months, the moose stayed south of battle creek in madison county where it was observed feeding in a however, in the fall of 2001, it appeared 40 km back to the north near osmond. reportedly, the moose was in poor condition and having trouble standing. shortly after the moose died, the ngpc transported the carcass to the university of nebraska-lincoln veterinary diagnostic lab for necropsy (associated press 2001a). the report concluded that cause of death was pneumonia and that, otherwise, the moose was in good physical condition (dave oates, ngpc, personal communication). several additional moose have been reported in iowa in the past few decades, including 2 individuals that appeared in the northwestern part of the state and dispersed moose seen in sheldon, iowa, on 23 september travelled southward near paulina (bullard a day later, the moose was reported northeast near woodbine (world-herald news service reported near underwood, where the moose another moose was seen in rock county, for about a month until it was reported again where it was hit by a car near worthington, long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 122 as moose increasingly occur in the central united states, it is of interest to ascertain if any individuals have found suitable habits in which to reside for extended periods of time. in this regard, a cow moose was seen near crawford, nebraska (indicated by the hollow star in fig. moose was seen again by hunters in the same the moose had resided in this area for more than 4 years. in another instance, a cow moose was reported inhabiting oahe wildlife management area (wma), which is located along the missouri river south of bismark, north dakota (indicated by the hollow diamond in cow moose were sighted wandering along the missouri river just south of bismark, north dakota. soon after, only 1 of the cows was ers found the moose lying down and unable to stand up, and a ndgfd biologist was called to oahe wma to euthanize the moose. from wma until it died, the moose had lived in the moose in north dakota encompasses the entire extreme eastern part of the state. however, reproductive populations currently exist only as far south as cass county, where, in 2001, munication). because no reproductive records have been reported south of this location, we conclude that populations located to the south of fargo in ransom, richland, and sargent counties consist only of vagrant individuals and no permanent populations. these vagrants appear to have spread into roberts county in northeasternmost south dakota. reports of wandering moose in roberts county have become common, with at least one moose being observed in that area each year (higgins et al. 2000; will morlock, sdgfp, personal communication). other reports indicate that moose are establishing reproductive pairs in central tion). this report represents the westernmost reproductive record of moose in the central north dakota have been reported from steele, ndgfd, personal communication). moose have been seen in the central united states (fig. 1). we report noteworthy records in nebraska, south dakota, and north dakota in table 3. these accounts are not a comprehensive list of extralimital records of moose in the central united states. in fact, moose have become so common outside their current geographic range in north dakota that biologists are only keeping reproductive records and have stopped tracking non-reproductive communication). discussion dispersal patterns our results suggest that certain dispersal patterns exist with regard to demographics and dispersal distance. for instance, the majority of long-distance dispersal events that occurred throughout the central united states appear to have been undertaken by young bulls in accordance with the process known as “jump dispersal”. however, as proximity to regions inhabited by moose increases, the demographic composition of extralimital records changes. in areas of northeastern south dakota, records of cow moose become more common and in western north dakota alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 123 were reported. female moose, especially diffusion dispersal in that they are dispersing relatively short distances from the established populations within the current geographic range of moose in north dakota. causes of dispersal the increase in moose sightings outside their geographic range could be a direct or indirect consequence of an increase in popu341,700 moose inhabited north america, but the number of moose in north america had and to 1,000,000 by 2000 (timmermann 2003). harvest population index records indicate that moose numbers in minnesota and north dakota have been at least sporadically increasing, and we suspect, based on published accounts of dispersal routes, that the populations in north dakota and minnesota are the source of most dispersing individuals in the central united states. however, it is possible that moose are dispersing into the central united states from other regions as well. the next closest populations of moose are those in the mountainous areas of montana, wyoming, in western nebraska (fig. 1) might have dispersed from those populations. for example, a yearling moose was sighted in laramie, wyoming, which is situated between the laramie and medicine bow mountain ranges, in 2001 (associated press 2001b). the straight line distance from laramie to scottsbluff, nebraska, where a bull moose had been seen moose are capable of travelling long distances state reference locality date seen citation north dakota 7-oct-71 7 mi e of bismark 10-aug-72 crown butte dam 7-nov-72 northern billings county notheastern mclean county southern mclean county southeastern morton county south dakota near flandreau sioux falls mobridge will morlock, sdgfp, pers. comm. will morlock, sdgfp, pers. comm. 0.5 mi n, and 10.5 mi e of eureka will morlock, sdgfp, pers. comm. 3 mi n of watertown will morlock, sdgfp, pers. comm. 6 mi w of conde will morlock, sdgfp, pers. comm. nebraska west of scottsbluff 3 mi ne of rose apr-06 aherns 2002 table 3. additional records of moose in north dakota, south dakota, and nebraska. dates with asobservation dates were given. long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 124 possible that individuals can disperse from these areas into adjacent states. when moose began appearing far to the south of their historical geographic range, several hypotheses were proposed by wildlife agencies and media to explain this phenomenon. the idea that a parasite, parelaphostrongylus tenuis, causes long-distance dispersal in moose was suggested, primarily by the popular media, as the reason for this unusual behavior. p. tenuis is a nematode that causes a neurological disease known as “moose sickness” (lankester and samuel p. tenuis by incidentally ingesting infected gastropods. in many accounts, witnesses claimed that dispersing moose appeared disoriented and lost, thus ultimately leading to the perception that the moose were sick. however, to our sickness in any of the dispersing individuals. on moose of the disease that p. tenuis causes. symptoms included walking in circles, holding head and ears in abnormal positions, fearlessness, stumbling, deafness, blindness, paraplegia, and, in most cases, death. none of these symptoms suggest that long distance movements are a characteristic of this disease. given the debilitating effects of this disease, especially on moose locomotory functions, it seems unlikely that infected moose would be physically able to disperse long distances. another possible explanation is that moose are leaving areas of high population density in search of other suitable habitat and mates. in presumed marginal habitats, such as along the periphery of their geographic range, densities of moose are lower than in areas located toward the center of their geographic range (telfer areas in marginal habitat cannot support large populations of moose, and that it is necessary for moose to disperse to new habitats. given their preference for early successional habitats, it seems plausible that moose would have evolved a dispersal behavior that would allow them to travel long distances in search of such with his discussion of transient and permanent habitats are those created by disturbance and are unstable and short-lived. moose invade these areas shortly after disturbance and, as climax forest reestablishes itself, moose populations decline. typically, disturbed habitats have a patchy distribution. once an area is disturbed, species turnover rate is quite rapid and decreases as the community approaches successional habitats have relatively short life spans as compared with climax communities. in order to access these areas, moose would need to be very mobile. moose were probably one of the last species to immigrate to north america from asia via the bering land bridge (reeves and mcacross northern north america (hundertmark et al. 2002). different mechanisms by which species expand their range include jump dispersal, diffusion, and secular migration (peilou the range expansion of moose by examining the genetic diversity of moose throughout their geographic range. they noted an overall lack of variation in mtdna; however, haplotype composition was different between peripheral populations and populations inhabiting the central geographic range in north america. from this, they hypothesized that range expansion of moose occurred through a few founding individuals that dispersed from a pre-expansion population. diffusion dispersal likely would not diminish genetic variation in lations in previously occupied areas. rather, diminished genetic variation suggests that range expansion of the moose was the result described as long-distance (i.e., leptokurtic) dispersal, where a few successful, long-disalces vol. 42, 2006 hoffman et al. long-distance moose dispersal 125 tance dispersers founded new populations. well, suggesting that founder effects were the cause of genetic homogeneity among the different subspecies of moose. finally, this type of dispersal process agrees with simpson’s in which an individual disperses from an established population across a major barrier to another suitable habitat patch. although, over time, numerous individuals attempt this sort of dispersal event, few are ever successful. our results favor the hypothesis that recent long-distance dispersal events by moose simply are the result of natural dispersal, rather than being induced by disease or other causes. if this is the case, the question then becomes what is the purpose for these dispersal events. we suggest that moose are dispersing from occupied habitats and searching for other suitable habitats. consequently, we believe that moose are in the process of attempting to expand their geographic range southward and that this dynamic process is similar for most mammals. for example, genoways et al. (2000) reported on extralimital records of the mexican free-tailed bat (tadarida brasiliensis) throughout the central united states. they concluded that pioneering individuals of t. brasiliensis occurring in areas outside their normal reproductive range are primarily foraging and exploring for new roost sites. they suggested that this is a natural process by which species may extend their geographic recognizable species. an example is the ninebanded armadillo (dasypus novemcinctus), which recently has dispersed northward from texas to the central great plains (choate and the likelihood that long-distance dispersal will result in an expansion of range is small. long-distance movements are rare and generally involve only one animal--often a male. the chance that both a bull and a cow (or a bull calf and heifer) will make a long move across unsuitable habitat to colonize the same new area is poor. the likelihood that diffusion dispersal will result in an expansion of range is much greater because of better reproductive opportunities. nevertheless, it may be an evolutionary strategy of moose to send out then, as the more reproductively valuable females disperse more gradually, they will tat. this would prevent loss of reproductive potential by cows wandering around looking for a mate. in conclusion, during the past 30 years there have been repeated records of moose occurring beyond their southern range boundary in the central united states. most of these individuals consisted of solitary juveniles or young adults; however, there were some exceptions. we were able to track dispersal routes for some of the moose, whereas others were indicated by single locality records. this report documents the longest known distance a moose has dispersed from an established population. we conclude that these events were the result of natural dispersal that could lead to further expansion of their geographic range to the south. further, we suggest that these dispersal events accurately represent the means by which moose expanded their range through north america during the early holocene, as described by hundertmark et al. (2003). based on the assumption that most species share similar characteristics of range expansion, we believe that this phenomenon can serve as a model to illustrate how other species expand their geographic range. management implications the southward movement of moose in the central united states has management implications. herein, we report two instances of moose inhabiting areas well outside the long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 126 current range of moose for an extended period of time—one in western nebraska and one in southern north dakota (fig 1). these records are noteworthy because they indicate that there are areas to the south of the current geographic range that may sustain permanent moose populations. the primary limiting factor for moose in the southern parts of their geographic range reportedly is climate, particularly high temperatures (kelsall and telfer moose experience heat stress at temperatures above 14-20° c. when heat-stressed, moose actively seek areas that provide them with shade and water for cooling (schwab and pitt factor, although not as important as climate. moose can adapt to a variety of forage, but in general they prefer shoots and other woody plants, such as willows (salix spp.) and other early successional vegetation resulting from factors may include the density of deer in an area and human impacts, such as urbanization, factors may work together to prevent dispersal events from resulting in an expansion of geographic distribution. moreover, dispersal can cause expansion only if dispersers eventually likelihood that a long-distance dispersal event will succeed is limited. in central north dakota, a cow moose inwma consists of approximately 6,475 ha of missouri river bottomland with good moose habitat that provides nearly continuous access to water and shade. further, oahe wma is subject to frequent disturbances that enhance wma (ndgfd 2005b). because river bottomland usually has a large fuel load, it is disturbed habitat for moose. in northwestern nebraska, a cow moose was reported living near crawford for more than 4 years. crawford is located within the pine ridge area where the habitat consists primarily of rolling prairies interspersed with pine forest. several streams and ponds are located in the area, along with the white and niobrara river valleys, which provide access to in this area has been converted to agriculture because much of it belongs to the state and that moose do not fare well in areas where human intolerance for moose is high, thus moose tend to occur more often in areas that have not been highly developed. the pine ridge area is subject to frequent disturbances resulting from timber harvesting (blyth et al. along the white and niobrara rivers. because harvesting is currently the most important factor in that it stimulates production of early and increases the amount of edge next to these food sources. these areas, that combine edge and food, are favored by moose (courtois and beaumont 2002). given that pine ridge is dominated by coniferous forest with a good supply of water and low human development, and given that a lone cow moose was able to survive in this area for more than 4 summers, it is possible that areas in the pine ridge area and the niobrara river valley may be capable of sustaining small populations of moose. amounts of shade and access to water, supplemented by frequent disturbance that promotes new plant growth, potentially could serve as suitable habitats for moose populations south of their current geographic range. areas similar to those described above that are located in adjacent states, such as the black hills of south alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 127 dakota and areas located along the missouri river and its tributaries in south dakota, iowa, missouri, nebraska, and kansas, also might serve as suitable habitat for moose. acknowledgements we would like to thank mark sexson of kansas department of wildlife and parks, dakota game, fish, and parks, bismark, north dakota, tom welstad, nebraska game and parks commission, norfolk, nebraska, dave oates, nebraska game and parks commission, lincoln, nebraska, and will morlock, south dakota game, fish, and parks, watertown, south dakota, for providing us with moose information from their respective states, as well as their personal insight. we especially appreciate the insightful review of an earlier draft of this manuscript by gordon eason of the ontario ministry of natural resources. references aherns, d. 2002. moose patrols central alberts, c. 2000. timber is a bright spot in northwestern nebraska ag economy. institute of agriculture and natural resources news service, university of nebraskaalex arrival after iowa trek. the des moines register, ister, and missouri alias. the des moines register, moose wandering through iowa. the p. 1a, 10a. anderson in moose infected experimentally with pneumostronglylus tenuis from whitetailed deer. pathologica veterinaria 1: associated press. now in ames area. the des moines re-the des moines register, the des moines register, 12 november 2a. disease. kansas city kansan. port another moose is loose in the state. the des moines register, 24 september cherokee. the des moines register, 27 northward again. the des moines regi-the des moines register, _____. 2001a. moose found dead of natural 2001, p. 2b. _____. 2001b. wandering moose takes to mini-golf course. laramie daily boomerang. ballard hitman reed moose in south-central alaska. wildlife monographs 114. beach ocheyedan, heads south. the des moines blyth ardle, and w. b. smith. agriculture, forest service, north central forest experiment station, st. paul, minnesota, usa. bowles ladfelter movement of moose south of traditional range in the upper midwestern united long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 states. proceedings of the iowa academy bry bullard through iowa. the des moines register, cederlund, g., and h. k. g. sand dispersal of subadult moose (alces alces) in a migratory population. canadian choate inkham madillo in northeastern colorado. prairie naturalist 20: 174. courtois, r., and a. beaumont. 2002. a of habitat composition and structure on moose density in clear-cuts of north-westcronin in mitochondrial dna of north ameridickson game. minnesota conservation volunteer, dockendorf visit to parkston. parkston advance, 7 franzmann alces alces. mammalian species 154: 1-7. freeman, p. w., and h. h. genoways recent northern records of the nine-banded armadillo (dasypodidae) in nebraska. gasaway, w. c., s. d. dubois eed. moose in interior alaska. federal aid and wildlife restoration final report. alaska alaska, usa. geist in behavior and evolution. the university of chicago press, chicago, illinois, usa. genoways, h. h., p. w. freeman, and c. grell. 2000. extralimital records of the mexican free-tailed bat (tadarida brasiliensis) in the central united states tions of the nebraska academy of science hall hewitt quences of ice ages, and their role in diverhiggins, k. f., e. d. stukel oulet, and d. c. backlund. 2000. wild mammals of south dakota. south dakota department of game, fish, and parks, pierre, north dakota, usa. horn persal. pages 54-62 in i. r. swingland of animal movement. clarendon press, oxford, uk. hornbeck near elba, may be one seen in page area. of nebraska’s travelling moose. omaha hundertmark dispersal, and migration. pages 303-336pages 303-336 in a. w. franzmann and c. c. schwartz, editors. ecology and management ofecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. _____, r. t. bowyer, g. f. shields, and c. c. schwartz. 2003. mitochondrial phylogeography of moose (alces alces) in _____, g. f. sheilds, i. g. udina, r. t. bowyer, a. a. danilkin, and c. c. schwartz. 2002. mitochondrial phylogeography of moose (alces alces): late pleistocene divergence and population expansion. alces vol. 42, 2006 hoffman et al. long-distance moose dispersal molecular phylogenetics and evolution idstrom in game in minnesota. minnesota department of natural resources, technical ohnson outdoors, vol. lii(6). ones rmstrong, r. s. hoffmann ones the northern great plains. university of nebraska press, lincoln, nebraska, usa. _____, and e. c. birney of mammals of the north-central states. university of minnesota press, minneapolis, minnesota, usa. karns density and trends. pages 125-140pages 125-140 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. kelsall status of moose (alces alces) in north america. swedish wildlife research supplement 1: 1-10. _____, and e. s. telfer phy of moose with particular reference to western north america. naturaliste canadien 101: 117-130. kleinschmidt p. 2a. knue a short history. north dakota game and fish department, bismark, north dakota, usa. krefting and habitat selection in north central north america. naturaliste canadien lamberto in all the wrong places. the des moines register, lankester, m. w., and w. m. samuel pests, parasites, and disease. pagespages in a. w. franzmann and c. c. schwartz, editors. ecology and man-ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. laukaitis lincoln ournal star on the loose near scottsbluff. lincoln lincoln star miller, f. l., e. broughton, and e. m. land. minnesota department of natural resources draft. minnesota department of natural resources, 17: 1-17. in northeastern minnesota; researchers (accessed 6 sept 2005). mytton, w. r., and l. b. keith dynamics of moose populations near nebraska game and parks commission moose surprises turkey hunters. nebraska (ndgfd) north dakota game and fish department ports 2004 bighorn sheep, moose, and elk harvests. north dakota game and fish _____. 2005b. open fires banned on oahe long-distance moose dispersal – hoffman et al. alces vol. 42, 2006 130 (accessed 12 oct 2005). omaha world-herald ing “aroused the whole town”. omaha pielou peterson university of toronto press, toronto, ontario, canada. reeves, h. m., and r. e. mccabe of moose and man. pages 1-76 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. renecker, l. a., and udson seasonal energy expenditures and thermoregulatory responses of moose. canadian ricklefs, r. e., and g. l. miller. 2000. ecology. fourth edition. w. h. freeman and rins well seen near melrose. the des moines register, schwab, f. e., and m. d. pitt selection of canopy cover types related to operative temperature, forage, and snow 3071-3077. seabloom, r. w., r. d. crawford, and m. g. mckenna southwestern north dakota: amphibians, reptiles, birds, mammals. institute for ecological studies, university of north nd: northern prairie wildlife research (accessed 1 oct 2005). sheilds and the evolution of philopatry. pages in greenwood, editors. the ecology of animal movement. clarendon press, oxford, uk. shugart ett cession: similarities of turnover rates. simpson emy of sciences 30: 137-163. sioux falls argus-leader on the run may be gone with a bang. sioux stone near larrabee. the des moines registar, svihovec day in area north of mobridge. mobridge taulman obbins recent range expansion and distributional limits of the nine-banded armadillo (dasypus novemcinctus) in the united states. telfer and habitat requirements of moose (alces alces in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. the university of alberta press, edmonton, alberta, canada. timmermann, h. r. 2003. the status and management of moose in north americatrego kota outdoors vol. lvii(6): 2-4 pp. (upi) united press international moose land, but… the lincoln star, 20 south. omaha world-herald, 3 march unruh ousted…again. garden city telegram. vance dateline… alces vol. 42, 2006 hoffman et al. long-distance moose dispersal 131 vosburgh, m. r., and r. peters of the wandering moose has sad ending. worthington daily globe, 1 march wagner dell rapids. sioux falls argus leader, white, t. 2001. wandering moose makes stop in madison county. nebraskaland world-herald news service bond moose draws crowd, stars in woodbine home video. omaha world-herald, woster ost heads south. the des moines register, f:\alces\vol_38\pagema~1\3805.pdf alces vol. 38, 2002 snaith et al. – habitat suitability analysis 73 preliminary habitat suitability analysis for moose in mainland nova scotia, canada tamaini v. snaith1, karen f. beazley1, frances mackinnon2, and peter duinker1 1school for resource and environmental studies, dalhousie university, 1312 robie st. halifax, ns, canada b3h 3e2; 2applied geomatics research group, centre of geographic science, 50 elliot road, rr1 lawrencetown, ns, canada b0s 1m0 abstract: ecosystem management for biological conservation should include consideration of landscape-scale processes such as the habitat requirements of focal species. moose (alces alces americana) have been identified as an appropriate target for focal attention in mainland nova scotia. currently, the population is at risk, and strategies for conservation should include the protection of sufficient habitat to meet the spatial requirements of the population. delineation of spatial habitat requirements calls for an understanding of species-habitat associations and the distribution of suitable habitat across the landscape. to this end, habitat suitability in nova scotia was assessed relative to four criteria: (1) food availability; (2) conifer cover; (3) mixed-wood cover; and, (4) aquatic resources. model predictions were tested by comparing habitat suitability values to provincial pellet inventory data. road density was found to be more important than habitat composition in determining moose pellet distribution. alces vol. 38: 73-88 (2002) key words: alces alces americana, conservation, habitat suitability index/analysis, nova scotia ecosystem management for biological conservation should incorporate coarse-filter considerations, such as adequate protection of all natural landscapes, and finefilter considerations including the habitat requirements of focal species (noss 1995, 1996; miller et al. 1998/99; noss et al. 1999). in nova scotia, american moose (alces alces americana) has been identified as an appropriate target of focal attention (beazley 1997, snaith and beazley 2002). the spatial requirements of wide-ranging species such as moose are an important consideration in the determination of the area required to maintain biodiversity at a landscape scale (noss 1995). determination of spatial requirements must incorporate consideration of ecological processes such as population viability, range use, and habitat requirements. in addition, the supply, composition, and spatial distribution of suitable habitat across the landscape must be understood to identify and delineate the spatial habitat requirements for long-term species persistence. in this paper we describe the development and application of a habitat suitability model for moose in mainland nova scotia. this analysis forms part of a larger project which describes the current status and distribution of moose populations within the context of biodiversity conservation, and develops management recommendations for protected areas system design in nova scotia (beazley et al. 2002). moose populations in nova scotia prior to european colonization, moose were widely distributed and abundant throughout nova scotia. however, there have been fluctuations and general declines in moose numbers since the early seventeenth century (dodds 1963, pulsifer and habitat suitability analysis – snaith et al. alces vol. 38, 2002 74 nette 1995). currently, mainland nova scotia is thought to contain approximately 1,000 moose, fragmented among a number of isolated smaller populations (a.l. nette, m. pulsifer, and r. hall, nova scotia department of natural resources, personal communication). moose are at risk of extirpation in mainland nova scotia, and will require special management attention if they are to persist (pulsifer and nette 1995, cesc 2001). a wide range of factors have been invoked to explain the declining moose populations. over-harvesting, habitat conversion, brain worm (paralephostrongylus t e n u i s ) , w i n t e r t i c k s ( d e r m a c e n t o r a l b i p i c t i s ) , a n d b l a c k b e a r ( u r s u s americanus) predation are among the factors affecting moose populations in nova scotia, and may regulate or limit moose density (pulsifer and nette 1995, snaith and beazley 2004). although there is currently no legal moose hunt in mainland nova scotia, hunting has been associated with major declines in the past, and may still be a marginal factor as there is evidence to indicate that some poaching occurs (snaith and beazley 2004). human land-use, including settlement and development, land clearing, cultivation, urbanization, and recreational development, restrict and eliminate moose habitat (houston 1968, dodds 1974). roads may fragment habitat, isolate populations, and affect moose density by constraining movement and habitat use, influencing habitat quality, favouring competitors or predators, causing mortality by vehicle collision, or by allowing increased human access and poaching pressure (houston 1968, prescott 1968, peek et al. 1987, hogg 1990, noss 1995, forman et al. 1997, rempel et al. 1997, beazley et al. 2004). because roads affect habitat suitability for many large mammals, it has been suggested that road density is the best indicator of ecological integrity and the intensity of human landuse (noss 1995, forman et al. 1997) moose habitat requirements in nova scotia moose need a diverse and heterogeneous habitat. optimal moose habitat contains a dynamic mosaic of forest patches with a variety of species and successional types (eastman 1974, telfer 1984, allen et al. 1987, harcombe 1988, hjeljord et al. 1990, mcnicol 1990, puttock et al. 1996). food-producing areas, water bodies, and patches of dense mature forest are critical components of moose range. small-scale patch dynamics, where open areas are scattered within dense mature forest, are most beneficial for selective feeding and will minimize travel between habitat components (timmermann and mcnicol 1988, jackson et al. 1991, heikkila et al. 1996). if present in sufficient quantity, the productive mixed forests of nova scotia can provide ideal year-round habitat for moose. forest cover is a critical habitat element which provides refuge from snow, wind, and cold temperatures, and relieves heat stress during both summer and winter months (e.g., peterson 1955; knowlton 1960; telfer 1967b, 1970; coady 1974; renecker and hudson 1986; thompson and euler 1987; schwab and pitt 1991; miquelle et al. 1992). because nova scotia is near the southern limit of moose range, thermal cover, particularly when in close proximity to forage producing areas, may be a limiting factor for moose in this area, especially during the hot summer months (telfer 1984, mitra 1999). early successional vegetation is the primary source of moose forage and an important habitat element. open areas following disturbance such as wind-throw, insect damage, wildfire, or timber harvest often contain good moose forage, as does the understory of mature forest with abundant alces vol. 38, 2002 snaith et al. – habitat suitability analysis 75 small canopy openings (e.g., wright 1956; telfer 1967a, 1967b, 1968, 1970; prescott 1968; leptich and gilbert 1989; bontaites and gustafson 1993; hjeljord and histol 1999). studies indicate that moose avoid foraging in large open areas, and generally will not move more than 80-200 m from cover, especially during snowy periods (eastman 1974, hamilton et al. 1980, tomm and beck 1981, peek et al. 1987, jackson et al. 1991, thompson et al. 1995). aquatic resources are an important component of moose habitat in many areas (e.g., wright 1956, dunn 1976, crossley and gilbert 1983, leptich and gilbert 1989, thompson et al. 1995). however, due to the paucity of wetlands in the cobequid area, and the acidity and low productivity of aquatic systems in the southwestern region where moose nevertheless persist, the importance of aquatic vegetation for moose in nova scotia is ambiguous, and it may not be a critical habitat component (telfer 1984). moose habitat selection and the quality of available habitat are biogeographically variable. there is no clear understanding of moose habitat preferences or the distribution of suitable habitat in nova scotia. we present here the results of a preliminary assessment of habitat suitability and spatial distribution based on existing data. habitat suitability analysis quantitative habitat suitability analysis can be used to determine the potential of habitat to support moose populations, to assess the relative suitability of candidate areas for protection or special management practices, and to identify measures which may enhance habitat quality (allen et al. 1987; duinker et al. 1991, 1993; puttock et al. 1996). habitat suitability is an important consideration when applying species area requirements to the landscape, because the quality and distribution of habitat will influence spatial requirements of individuals and populations (allen et al. 1987, jackson et al. 1991). an assessment of habitat suitability must consider all critical habitat components including nutritional, reproductive, and shelter requirements, and may include environmental conditions and land-use practices (allen et al. 1987, jackson et al. 1991). allen et al. (1987) constructed a habitat suitability index (hsi) model which quantitatively measures the suitability of an area to support moose. the hsi model is based on the assumption that moose require certain habitat components, and that an appropriate relative amount of each component must be present for the habitat to be considered optimal (allen et al. 1987). however, the model is unable to account for special habitat characteristics, such as mineral licks and calving sites, and nonhabitat mortality factors, such as poaching, predation, and human land-use. thus, for this preliminary study, the hsi distribution will simply be used as a relative ranking of habitat suitability and potential to support moose, rather than an absolute index of potential carrying capacity or population density. the objectives of the habitat suitability analysis were to: (1) analyze the suitability of moose habitat in nova scotia using a hsi model; (2) based on the model, produce a theoretical distribution of habitat suitability across the landscape; (3) test the model by comparing it to pellet group inventory (pgi) data as an index of moose distribution; (4) determine which habitat components may influence moose habitat selection; and (5) examine the effects of human land use on moose habitat selection by using road density as an index of human influence. methods habitat suitability index model construction allen et al. (1987) developed two hsi habitat suitability analysis – snaith et al. alces vol. 38, 2002 76 models for moose habitat evaluation in the lake superior region. model i involves a complex assessment of seasonal cover and browse quantity, quality, and interspersion. it requires extensive data including height, density, and species composition of forest cover; biomass productivity; browse diversity and quality; and interspersion of food and cover. model ii provides a less detailed examination of habitat based on easily accessible forest cover data and can be used for rapid, low-resolution evaluation of large areas. although model ii does not consider the fine scale spatial distribution of habitat components, it is useful because it is relatively simple and the necessary data are readily available (naylor et al. 1992). the models were designed for use in the lake superior region and have been applied and validated in a number of studies and modified for use in other regions (allen et al. 1991, naylor et al. 1992, puttock et al. 1996, rempel et al. 1997). for this study, hsi model ii (allen et al. 1987) was modified based on extensive literature review and local expert opinion, and applied to mainland nova scotia for preliminary assessment of moose habitat suitability. the model was used to conduct a gisbased static inventory of the forest cover, and to estimate the relative potential of the landscape to support moose. forest cover inventory. — the 1992 provincial forest cover inventory (nova scotia department of natural resources, unpublished data) was used as the input dataset for habitat coverage. these data were provided as an arcinfo® gis coverage (utm nad83) representing landscape vegetation patterns as polygons. vegetation was classified by cover type (e.g., forested, non-forested), forest type (e.g., softwood, hardwood), species composition, age, and non-forest attributes such as wetlands and agricultural areas. for this study, the forest cover dataset was analyzed according to the attributes of 4 habitat components used in the hsi model (forage, softwood cover, hard/mixed-wood cover, and wetlands) (table 1). any forest cover attributes which did not qualify as critical habitat components had no value in the hsi calculation. due to the challenges of using polygon data for this type of habitat model, the forest coverage was transformed into point data with points on a 200 m grid (duinker et al. 1991, 1993; mccallum et al. 1993). evaluation units. — habitat suitability was calculated individually for a series of systematically distributed evaluation units, which together depict the spatial patterns of hsi values across the landscape (allen et al. 1987, duinker et al. 1991). based on considerations of moose home range size, the total size of the study area, and the original design of model ii (recommended evaluation unit of 93 km2 or larger) (allen et al. 1987, duinker et al. 1991), an evaluation unit size of 100 km2 (10 x 10 km) was selected. the evaluation units were applied to the province using the moving-window technique developed by duinker et al. (1991, 1993). this technique allows each stand to contribute to the hsi calculation several times, provides a more realistic representation of habitat heterogeneity, and accounts for the possibility that moose ranges overlap the boundaries of evaluation units. critical habitat components. — the hsi was calculated based on the relative availability of critical habitat components within each evaluation unit. following allen et al.’s (1987) hsi model ii, suitable moose habitat contains 4 habitat components; open forage-producing areas, softwood cover, hardwood or mixed-wood cover, and wetlands. the composition of habitat components was modified slightly from the original model to accommodate the nova scotia forest cover classification system and local vegetation characteristics (table 1). for alces vol. 38, 2002 snaith et al. – habitat suitability analysis 77 example, unproductive acidic wetlands were considered unacceptable, and the softwood cover component was broadened to include all softwood species. the percent availability of habitat components within each evaluation unit was extracted from the converted forest cover point data using arcinfo®. suitability index (si) values were derived for each habitat component using curves which model the predicted suitability of habitat based on percent availability (fig.1) (allen et al. 1987). si values range from 0.0 to 1.0, where 0.0 represents unsuitable habitat and 1.0 represents the optimum proportion of each habitat component. according to the model, optimum moose habitat contains 40 50% preferred forage area (si 1 ); 5 15% softwood forest cover (si 2 ); 35 55% deciduous or mixed forest cover (si 3 ); and 5 10% wetlands (si 4 ). evaluation units which did not contain forage or cover retable 1. habitat component composition and associated forest cover attributes. habitat component original composition modified for forest cover (from allen et al. 1987) this study attributes forage shrub or forested cover any forest type cover type: (si 1 ) types <20 years old <20 years old softwood mixedwood hardwood age: <20 softwood cover spruce/fir forest softwood forest cover type: (winter cover) ≥20 years old ≥20 years old softwood (si 2 ) age: ≥20 hard or mixedwood upland deciduous deciduous or mixed cover type: cover forest ≥20 years old hardwood (forage/cover) mixedwood (si 3 ) age: ≥20 wetlands riverine, lacustrine, or wetlands not nonforested: (aquatic forage) palustrine wetlands not dominated by woody wetlands (si 4 ) dominated by woody vegetation, and not beaver flowage vegetation including acidic, lake/wetland unproductive wetlands marsh/swamp ceived a si of 0.0. however, evaluation units without wetlands received a si of 0.2, rather than 0.0, because the resolution of the data may have failed to represent small wetlands, and because areas with no wetlands are not totally unsuitable as moose habitat (telfer 1984, allen et al. 1987). habitat suitability calculation: — according to the original model, ideal yearround moose habitat contains all 4 habitat components (allen et al. 1987). the suitability index values were combined to calculate the overall hsi for each evaluation unit. hsi values range from 0.0 to 1.0, where, as with sis, 0.0 represents highly unsuitable habitat and 1.0 represents optimum moose habitat. following allen et al. (1987), si values were combined using the geometric mean. although a variety of mathematical functions may be used to calculate hsi, the geometric mean is a good choice because it habitat suitability analysis – snaith et al. alces vol. 38, 2002 78 assumes that components can partially compensate for one another, but that hsi is affected by the smallest value (van horne and wiens 1991). any 0.0 si value will produce an overall hsi of 0.0, which means that in the absence of food or cover, the hsi will always be 0.0 (van horne and wiens 1991). six experimental equations were used to calculate hsi values for nova scotia (table 2). allen et al.’s (1987) original equation (hsi 1 ) was modified (hsi 2 though hsi 6 ) to attempt to account for local conditions where it was hypothesized that: (1) mature forest (thermal cover) may be especially critical; (2) wetlands may be less important; and (3) forage beyond 200 m of cover may be of little value. hsi 1 is the original hsi model ii equation, which calculated the geometric mean of the 4 suitability index values (allen et al. 1987). hsi 2 and hsi 3 were modified to explore the assumption that aquatic resources may not be a critical habitat component (telfer 1984). in hsi 2 , the wetland component was removed from the equation, and the geometric mean of the 3 remaining habitat components was taken. hsi 3 is similar to hsi 2 , but the cover components were weighted more heavily than the forage component due to the possibility that for at least some portion of the winter and/or summer, moose require dense cover and will seek shelter at the expense of food. to model the known importance of the proximity of food and cover, an additional si was calculated. si 1m was a modification of si 1 , and was derived using the same forest attributes and suitability curve as si 1 , but only included forage areas located within 200 m of cover. hsi 4 , hsi 5 , and hsi 6 substitute si 1m for si 1 into hsi 1 , hsi 2 , and hsi 3 , respectively. based on the calculated hsi index values, 6 maps were produced to illustrate the spatial distribution of habitat suitability, and its relative potential to support moose, across the province. application of the hsi model the validity of the hsi was tested using moose pellet counts collected along pgi transects throughout the province (nova scotia department of natural resources, unpublished data). although pgi data cannot provide reliable estimates of population fig.1. derivation of suitability index (si) for each habitat component. 0 0.2 0.4 0.6 0.8 1 0 15 30 45 60 75 90 forage (%area) s u it ab il it y i n d ex ( s i1 ) 0 0.2 0.4 0.6 0.8 1 0 15 30 45 60 75 90 softwood cover (%area) s u it ab il it y i n d ex ( s i2 ) 0 0.2 0.4 0.6 0.8 1 0 15 30 45 60 75 90 mixedwood cover (%area) s ui ta bi li ty i nd ex ( s i3 ) 0 0.2 0.4 0.6 0.8 1 0 15 30 45 60 75 90 wetlands (%area) s ui ta bi li ty i nd ex ( s i4 ) alces vol. 38, 2002 snaith et al. – habitat suitability analysis 79 density, they are useful indicators of population trends and habitat selection (neff 1968, franzmann et al. 1976, harkonen and heikkila 1999). nova scotia department of natural resources has conducted a province-wide pgi since 1983 as a tool for estimating the size of the provincial whitetailed deer (odocoileus virginianus) population. as transects were surveyed, moose pellet data were also recorded. because transects were established randomly, the pellet counts provide a random sample of moose pellet distribution. however, the usefulness of the pgi data is limited because the exact location of pellets within transects and the characteristics of the surrounding habitat were not recorded. additionally, since only pellets deposited between leaf-fall (november) and pellet count (april/may) were recorded, the pgi data can only be used as an indication of moose habitat selection during the period from late fall to early spring. the modified hsi equations were conducive to an analysis of fall/winter habitat suitability because during winter, aquatic resources are not critical, thermal cover is required, and forage within close proximity to cover is preferred. the pgi location data were converted into utm nad83 and plotted in arcinfo®. the transects were overlain on the previously generated hsi coverages. each transect was assigned the habitat characteristics and hsi index of the corresponding hsi evaluation unit. when a transect crossed more than one hsi evaluation unit, it was assigned the value of the unit where the majority of the transect was located. a database table suitable for statistical analysis was produced which included transect identification numbers and their associated habitat values and hsi indices. preliminary descriptive statistics were generated for the pgi pellet count data using microsoft excel® and spss® 9.0 to identify the best summary statistic for the data. scatter plots were produced for each transect to examine the distribution of moose pellet counts over time. many transects (67.39%) contained zero moose pellets in all years. pellet counts were generally low on transects where moose pellets were observed (1983-2000, 524 transects: n = 7702, mean = 0.70, median = 0, sd = 3.64, range = 0-79), and there were no observable patterns consistent among transects over time. for these reasons, and because the reliability of using pellet counts as an indication of moose density is tenuous, presence/absence was selected as the best statistic to summarize pellet counts on transects over time (franzmann et al. 1976, harkonen and heikkila 1999). all pgi data collected after 1992 were excluded from analysis because the forest inventory used to generate the hsi was current only to 1992. after transects were excluded from analysis due to missing data table 2. hsi equations. hsi 1 = (si 1 si 2 si 3 si 4 )1/4 original equation following allen et al. (1987) hsi 2 = (si 1 si 2 si 3 )1/3 wetlands removed hsi 3 = [si 1 (si 2 2) (si 3 2)]1/5 wetlands removed and cover weighted more heavily hsi 4 = (si 1m si 2 si 3 si 4 )1/4 same as hsi 1 but using si 1m only includes forage near cover hsi 5 = (si 1m si 2 si 3 )1/3 same as hsi 2 but using si 1m only includes forage near cover hsi 6 = [si 1m (si 2 2) (si 3 2)]1/5 same as hsi 3 but using si 1m only includes forage near cover habitat suitability analysis – snaith et al. alces vol. 38, 2002 80 or because they were located outside the study area (on cape breton island), 370 transects remained for analysis. to test the validity of the models, we examined the ability of hsis to predict the presence/absence of moose pellets using logistic regression analysis. additional analyses were run to determine the ability of each of the 4 critical habitat components independently to predict pellet presence/ absence, and to examine the influence of individual habitat components on overall habitat suitability. the results of these analyses can be used to construct a more reliable suitability index for nova scotia. due to their geographical nature, habitat values are inherently auto-correlated, or spatially dependent. however, auto-correlation must be identified because it violates the mathematical assumptions of regression analysis (goodchild 1986). the residuals of logistic regression analysis were tested for auto-correlation using the geary and moran indices of the grid function in arcinfo®, and wherever auto-correlation was identified, the results must be interpreted with caution. effect of road density habitat suitability is likely affected by a variety of factors, such as human land-use practices, which were not accounted for in the hsi models. road density was selected as a surrogate index of human influence. using gis, road density (including all major and secondary roads, trails, railways, carttracks, and woods roads) was divided into 6 classes (density classes (km/km2): 0, 0.10.06, 0.06-0.6, 0.6-1, 1-3, >3) and mapped on a 1 x 1 km grid (fig.2 ) (beazley et al. 2004). additional regression analyses were run to determine the ability of road density to predict the presence of moose pellets, and to examine the effects of roads on habitat suitability by running multivariate logistic regression, using roads in combinafig.2. road density in mainland nova scotia. alces vol. 38, 2002 snaith et al. – habitat suitability analysis 81 tion with the habitat suitability values, to predict moose pellet presence. results habitat suitability modeling results of the hsi modeling indicate that there is little highly suitable moose habitat in nova scotia. hsi values were mapped to show the spatial distribution of habitat suitability across the landscape (fig.3). for graphic representation, hsi values were divided into 5 suitability categories: very poor (hsi = 0.00-0.19); poor (0.20-0.39); moderate (0.40-0.59); good (0.60-0.79); and very good (0.80-1.00). according to all equations, a large amount of the mainland was of very poor (27.4829.31%), poor (14.08-31.89%), or moderate (36.29-52.44%) habitat suitability. only 2.50-12.87% was of good suitability, and 0.0-0.42% was very good (fig.4). in effect, the models predicted that only 1,114-5,737 km2 of the nova scotia mainland were good, and 0-185 km2 were very good moose habitat. the spatial distribution of hsi values across the landscape differed little among hsi equations. in general, the southern coastal areas were of very low suitability, inland areas were slightly more suitable, and the few areas of high suitability were located mostly in the hilly regions of the cobequid and pictou-antigonish highlands. application of the hsi model summary statistics indicated that of the 370 transects, only 126 (34.1%) had at least one moose pellet between 1983 and 1992, while moose pellets were completely absent from 244 (65.9%) transects (fig.5). the spatial distribution of pellets seemed to correspond roughly to known moose distributions, with concentrations in the tobeatic and the cobequid to antigonish highland fig.3. an example of theoretical habitat suitability distribution for mainland nova scotia (hsi 5 ). habitat suitability analysis – snaith et al. alces vol. 38, 2002 82 fig.4. summary of habitat suitability in mainland nova scotia from 6 experimental equations. fig.5. distribution of moose pellet presence on pgi transects. 0 10 20 30 40 50 60 hsi 1 hsi 2 hsi 3 hsi 4 hsi 5 hsi 6 hsi equation % o f n o v a s c o ti a m a in la n d 0.00 – 0.19 0.20 – 0.39 0.40 – 0.59 0.60 – 0.79 0.80 – 1.00 alces vol. 38, 2002 snaith et al. – habitat suitability analysis 83 areas. however, pellet presence was scattered throughout much of the rest of the province, and may represent occasional or low-density moose occupation, or perhaps pellet presence from the early years of pgi surveys when moose populations were more w i d e l y d i s t r i b u t e d ( k e l s a l l 1 9 8 7 , timmermann and buss 1997). unfortunately, due to the nature of pgi data, and the simplification to presence/absence, it was not possible to identify trends through time or to make any inferences about moose densities across the province. the results of logistic regression analysis indicated that none of the 6 hsi models could predict the presence of moose pellets. only 2 individual habitat components (forage and forage in proximity to cover) could significantly predict the presence of moose pellets across the landscape (table 3). results of the test for spatial dependence indicate that the habitat suitability data were autocorrelated (forage: geary 0.006, moran 0.001; forage in proximity to cover: geary 0.006, moran 0.001). effect of road density regression analysis indicated that road density could significantly predict the presence of moose pellets (table 3). a significant negative correlation suggested that as road density increased, the probability of moose pellet presence decreased. when multivariate logistic regression was used to test the combined effect of road density and hsi results on pellet presence, all hsi values and individual habitat components, when combined with the effect of roads, could table 3. the ability of habitat values and road density to predict moose pellet presence on transects (chi-square values from regression analysis). habitat values habitat and road roads after habitat is habitat after or roads alone1 combined2 accounted for1 roads are accounted for1 hsi 1 1.567 *16.412 *15.954 1.459 hsi 2 1.567 *20.297 *18.085 *5.178 hsi 3 0.338 *17.515 *16.702 2.529 hsi 4 0.000 *15.830 *15.483 0.892 hsi 5 1.059 *19.069 *17.453 *4.022 hsi 6 0.254 *17.136 *16.445 2.166 comp. 13 *16.229 *38.576 *21.519 *21.248 comp. 1m *15.312 *37.388 *21.254 *20.605 comp. 2 0.464 *14.994 *14.291 0.067 comp. 3 0.209 *14.931 *14.501 0.004 comp. 4 3.162 *19.375 *15.886 3.903 roads *14.927 n/a n/a n/a * significant result p < 0.05. 1 1 degree of freedom. 2 2 degrees of freedom. 3 comp. = habitat components as described in table 1. habitat suitability analysis – snaith et al. alces vol. 38, 2002 84 significantly predict pellet distribution. after the effect of hsi was statistically accounted for, road density was able to predict moose pellet presence in all cases. conversely, once roads were accounted for, only hsi 2 and hsi 5 , forage, and forage in proximity to cover, were able to predict pellet presence. however, there was spatial dependence among the road density data (geary 0.006, moran 0.002) and the results must be interpreted with caution. discussion the results of the hsi modeling suggest that hsi 2 and hsi 5 , in combination with road density, may provide a reliable index of moose habitat quality in nova scotia. however, these results should not be accepted as conclusive, as the validation using pgi does not account for summer habitat selection, when thermal cover is likely more critical. according to these models, there is little optimal moose habitat in nova scotia. this is not surprising given the current low densities of moose populations in the province (although a host of other factors likely contribute to limit the population). the results of this analysis may help to explain the current distribution of moose in nova scotia. high suitability indices occurred in the areas known to contain the largest and most stable moose populations (cobequid and pictou-antigonish highlands), while the south-western region (also known to contain a remnant population) likely supports moose due to low road density, despite poor habitat suitability. statistical analysis using moose pellet data as an index of habitat selection was unable to validate the hsi models alone. hsi alone was unable to predict the presence of moose pellets across the landscape, and forage was the only habitat component that significantly predicted moose pellet presence. the importance of forage within close proximity to cover, and its influence on habitat suitability, remains unclear because forage was significantly related to pellet presence when all forage areas were included and when only forage areas within 200 m of cover were included in the calculation of hsi. the results suggest that human influence, as indicated by road density, had a greater effect on moose habitat selection, and presumably habitat suitability, than habitat composition alone. this hypothesis is supported by the significant relationship between road density and moose pellet presence on provincial transects, and strengthened by the ability of 2 of the hsi indices (equations 2 and 5) to predict pellet presence after the effect of road density was accounted for. this result is interesting because it indicates that, roads being equal, these 2 hsi calculations might approximate fall and winter habitat suitability for moose in mainland nova scotia. future models should include road density as an initial habitat variable. statistical validation of the hsi models was limited by the nature of the pgi data. the intent of the original hsi model was to predict the potential carrying capacity of moose habitat, and did not suggest that moose cannot survive or will not be present in sub-optimal habitat. when further data, such as aerial surveys or more comprehensive pgi, become available, the hsi model should be compared to moose density, rather than pellet presence/absence, for validation. furthermore, year-round moose distribution data are required to ensure the inclusion of summer habitat suitability when mature forest for thermal cover is likely critical. further research is required to strengthen habitat suitability analysis, to validate its ability to delineate critical habitat, and to determine the effects of human land-use practices such as roads and forest managealces vol. 38, 2002 snaith et al. – habitat suitability analysis 85 ment. a better model for habitat suitability will incorporate human-induced habitat characteristics, such as road density, into index calculation. the results of this type of analysis can be used to identify critical habitat areas as candidates for protection, and to make management recommendations which will improve habitat suitability for moose in nova scotia. acknowledgements t. snaith and k. beazley thank tony nette, vince power, dalhousie university faculty of graduate studies, school for resource and environmental studies, ejlb foundation, canadian wildlife federation, frederik doyon, wade blanchard, and tony schellinck. f. mackinnon thanks david colville, bob maher, tony nette, nova scotia department of natural resources, the sustainable forest network, bowater mersey paper co. ltd., j.d. irving, and seventeen member groups of the nova scotia federation of hunters and anglers. we thank the editor of alces and two unknown reviewers for helpful comments and suggestions. references allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. u.s. department of the interior biological report 82 (10.155). , j. w. terrell, w. l. mangus, and e. l. lindquist. 1991. application and partial validation of a habitat model for moose in the lake superior region. alces 27:50-64. beazley, k. 1997. ecological considerations for protected area system design. proceedings of the nova scotia institute of science 41:59-76. , p. austin-smith, and m. rader. 2002. toward completing a protected areas system for nova scotia. pages 516 530 in s. bondrup-nielsen and n. munro, editors. managing protected areas in a changing world. proceedings of the fourth international conference on science and the management of protected areas. wolfville, nova scotia, canada. may 14-19, 2000. , t. snaith, f. mackinnon, d. colville, and s. brown. 2004. road density and impacts on mammals in nova scotia. proceedings of the nova scotia institute of science 42:152-170. bontaites, k. m., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29:163-167. (cesc) canadian endangered species council. 2001. wild species 2000: the general status of species in canada. ministry of publicworks and governm e n t s e r v i c e s , o t t a w a , o n t a r i o , canada. coady, j. w. 1974. influence of snow on b e h a v i o r o f m o o s e . n a t u r a l i s t e canadien 101:417-436. crossley, a., and j. r. gilbert. 1983. home range and habitat use of female moose in northern maine a preliminary look. transactions of the northeast section of the wildlife society 40:67-75. dodds, d. g. 1963. the present status of moose (alces alces americana) in nova scotia. proceedings of the northeast wildlife conference 2:1-40. . 1974. distribution, habitat and status of moose in the atlantic provinces of canada and northeastern united states. naturaliste canadien 101:5165. duinker, p. n., p. e. higgelke, and n. a. bookey. 1993. future habitat for moose on the aulneau peninsula, northwest ontario. pages 551-556 in 7th annual symposium on geographic information systems in forestry, environment and habitat suitability analysis – snaith et al. alces vol. 38, 2002 86 natural resources management, vancouver, british columbia, 15-18, february 1993. , , and s. koppikar. 1991. pages 271-275 in gis-based habitat supply modelling in northwestern ontario: moose and marten. gis’91 an international symposium. geographic information systems, vancouver, british columbia, february 1991. dunn, f. 1976. behavioural study of moose. maine department of inland fish and wildlife project w-66-r-6 job 2-1. eastman, d. s. 1974. habitat use by moose of burns, cutovers and forests in north-central british columbia. proceedings of the north american moose conference and workshop 10:238-256. forman, r. t. t., d. s. friedman, d. fitzhenry, j. d. martin, a. s. chen, and l. e. alexander. 1997. ecological effects of roads: toward three summary indices and an overview for north america. pages 40-54 in k. canters, editor. habitat fragmentation and infrastructure. ministry of transportation, public works and government services, delft, netherlands. franzmann, a. w., j. l. oldemeyer, p. d. arneson, and r. k. seemel. 1976. pellet-group count evaluation for census and habitat use of alaskan moose. proceedings of the north american moose conference and workshop 12:127-142. g o o d c h i l d , m . f . 1 9 8 6 . s p a t i a l autocorrelation. geo books, catmog 47, norwich, u.k. hamilton, g. d., p. d. drysdale, and d. l. euler. 1980. moose winter browsing patterns on clear-cuttings in northern ontario. canadian journal of zoology 58:1412-1416. harcombe, a. p. 1988. wildlife habitat handbooks for the southern interior ecoprovince. volume 5: species-habitat relationship models for mammals. ministry of the environment, victoria, british columbia, canada. harkonen, s., and r. heikkila. 1999. use of pellet group counts in determining density and habitat use of moose alces alces in finland. wildlife biology 5:227-233. heikkila, r., k. nygren, s. harkonen, and a. mykkanen. 1996. characteristics of habitats used by female moose in the managed forest area. acta theriologica 41:321-326. hjeljord, o., and t. histol. 1999. rangebody mass interactions of a northern u n g u l a t e a t e s t o f h y p o t h e s i s . oecologia 119:326-339. , n. hovik, and h. pedersen. 1990. choice of feeding sites by moose during summer, the influence of forest s t r u c t u r e a n d p l a n t p h e n o l o g y . holarctic ecology 13:281-292. hogg, d. 1990 moose management: the forest habitat. pages 30-33 in m. buss and r. truman, editors. the moose in ontario, book 1. ontario ministry of natural resources, queen’s printer for ontario, toronto, ontario, canada. houston, d. b. 1968. the shiras moose in jackson hole, wyoming. u.s. department of the interior technical bulletin 1:1-62. jackson, g. l., j. g. racey, j. g. mcnicol, and l. a. godwin. 1991. moose habitat interpretation in ontario. ontario ministry of natural resources, northwestern ontario forest technology development unit, technical report 52. kelsall, j. p. 1987. the distribution and status of moose in north america. swedish wildlife research supplement 1:1-10. knowlton, f. f. 1960. food habits, movements and populations of moose in the gravelly mountains, montana. journal alces vol. 38, 2002 snaith et al. – habitat suitability analysis 87 of wildlife management 24:162-170. leptich, d. j., and j. r. gilbert. 1989. summer home range and habitat use by moose in northern maine. journal of wildlife management 53:880-885. mccallum, i., p. higgelke, and p. duinker. 1993. gis-based habitat supply analysis for moose (alces alces) and marten (martes americana) in northern ontario forests. pages 370-375 in proceedings of the international union of game biologists xxi congress: forests and wildlife towards the 21st century. chalk river, ontario, canada. mcnicol, j. 1990. moose and their environment. pages 11-18 in m. buss and r. truman, editors. the moose in ontario, book 1. ontario ministry of natural resources, queen’s printer for ontario, toronto, ontario, canada. miller, b., r. reading, j. strittholt, c. ca r r o l l, r. n o s s , m. s o u l é , o. sanchez, j. terborgh, d. brightsmith, t. cheeseman, and d. foreman. 19981999. using focal species in the design of nature reserve networks. wild earth 9:81-92. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monographs 122. mitra, p. 1999. predictive modelling of moose habitat in the tobeatic wilderness management area. college of geographical science, lawrencetown, nova scotia, canada. naylor, b., s. christilaw, and p. wielandt. 1992. validation of a habitat suitability index model for moose in the northern portion of the great lakes-st. lawrence forest region of ontario. ontario ministry of natural resources, central ontario forest technology development unit, technical report 26. neff, d. j. 1968. the pellet-group count technique for big game trend, census, and distribution: a review. journal of wildlife management 32:597-614. noss, r. f. 1995. maintaining ecological integrity in representative reserve networks. world wildlife fund canada, toronto, ontario, canada. . 1996. protected areas: how much is enough? pages 91-120 in g. r. wright, editor. national parks and protected areas: their role in environmental protection. blackwell science, cambridge, massachusetts, usa. , j. r. st r i t t h o l t , k. vanceborland, c. carroll, and p. frost. 1999. a conservation plan for the klamath-siskiyou ecoregion. natural areas journal 19:392-411. peek, j. m., d. j. pierce, d. c. graham, and d. l. davis. 1987. moose habitat use and implications for forest management in northcentral idaho. swedish wildlife research supplement 1:195-199. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. prescott, w. h. 1968. a study of winter concentration areas and food habits of moose in nova scotia. m.sc. thesis, acadia university, wolfville, nova scotia, canada. pulsifer, m. d., and t. l. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31:209219. puttock, g. d., p. shakotko, and j. g. rasaputra. 1996. an empirical habitat model for moose, alces alces, in algonquin park, ontario. forest ecology and management 81:169-178. rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber-management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61:517-524. renecker, l. a., and r. j. hudson. 1986. habitat suitability analysis – snaith et al. alces vol. 38, 2002 88 seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64:322327. schwab, f. e., and m.d. pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69:3071-3077. snaith, t., and k. beazley. 2002. moose (alces alces) as a focal species for reserve design in nova scotia, canada. natural areas journal 22:235-240. . 2004. the distribution, status, and habitat associations of moose in mainland nova scotia. proceedings of the nova scotia institute of science 42:76134. telfer, e. s. 1967a. comparison of a deer yard and a moose yard in nova scotia. canadian journal of zoology 45:485490. . 1967b. comparison of moose and deer winter range in nova scotia. journal of wildlife management 31:418-425. . 1968. distribution and association of moose and deer in central new brunswick. transactions of the northeast section of the wildlife society 25:41-70. . 1970. winter habitat selection by moose and white-tailed deer. journal of wildlife management 34:553-558. . 1984. circumpolar distribution and habitat requirements of moose (alces alces). pages 145-182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. thompson, i. d., and d. l. euler. 1987. moose habitat in ontario: a decade of change in perception. swedish wildlife research supplement 1:181-193. thompson, m. e., j. r. gilbert, g. j. matula, and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31:233-245. timmermann, h. r., and m. e. buss. 1997. the status and management of moose in northern america in the early 1990s. ontario ministry of natural resources, thunder bay, ontario, canada. , and j. g. mcnicol. 1988. moose habitat needs. forestry chronicle 1988:238-245. tomm, h. o., and j. a. beck, jr. 1981. responses of wild ungulates to logging practices in alberta. canadian journal of forest research 11:606-614. van horne, b., and j. a. wiens. 1991. forest bird habitat suitability models and the development of general habitat models. fish and wildlife research paper 8:1-31. wright, b. s. 1956. the moose of new brunswick. new brunswick department of lands and mines, fredericton, new brunswick, canada. p75-84_4106.pdf alces vol. 41, 2005 mclaren and mercer density management 75 how management unit license quotas relate to population size, density, and hunter access in newfoundland b. e. mclaren1,2 and w. e. mercer3 1government of newfoundland & labrador, department of natural resources, p.o box 2222, gander, nl, canada a1v 2n9; 38 virginia place, st. john’s, nl, canada a1a 3g6 abstract: we recommend introducing habitat-based moose density as a management tool to be used in annual quota setting. we illustrate our recommendation with the case of newfoundland, where moose densities are much higher than elsewhere in north america, and have led to local areas of habitat deterioration and subsequent population decline. we suggest more emphasis be placed on relationships between local densities of moose and reported hunter kill locations to stabilize populations. we calculated both moose density and moose-kill density using estimates of forest and “scrub” cover in management units surveyed between 1985 and 2001, comparing aerial surveys with license sales for the same year population size estimates calculated by management unit, especially in central newfoundland. in the latter part of our study period, a strong relationship between license quotas and population estimates (r2 kills and population size were less well correlated. overall during this period, kill density increased, while moose density decreased, sometimes below target. alces vol. 41: 75-84 (2005) key words: accessibility, alces alces, density, management, moose, newfoundland, population dynamics, quotas, targets effective management of moose (alces alces) populations requires on-going assessment of their size and productivity, ideally along with habitat assessments. it also requires consideration of the feasibility of implementing various hunting strategies, based on hunter demand and capability, and further on the ability of wildlife managers to understand moose population dynamics. spatial variation in moose population dynamics and in hunter success within a management unit may demand sequence of unit-by-unit comparisons of aerial survey summaries and hunting objectives. unfortunately, innovative licensing strategies within a management unit are rarely undertaken (one example is described in mclaren scale, habitat-based approach to management has long been a recommendation to managers (timmermann and buss 1998). unlike the usual approach to moose management, carried out by estimating population size and assigning a sustainable “share” or “quota” to hunters, we recommend approaching moose populations by introducing habitat-based density measures as the basic management parameter. this paper is directed to illustrating our recommendation using the case of moose management on the island of newfoundland, canada. an interesting perspective granted by newfoundland moose also comes from our 2present address: faculty of forestry and the forest environment, lakehead university, 955 oliver road, thunder bay, on, canada p7b 5e1 density management – mclaren and mercer alces vol. 41, 2005 76 experience that for game species, much focus is spent by managers on cases of declining populations, which are the short-term interest of the hunting public. much less is understood about overpopulations, even though the situation ironically also leads in the long term to declining populations (mclaren et al. 2004). overpopulation, or overabundance, usually occurs following introduction into unexploited habitat and persists in situations without natural predators (mcshea et al. 1997). this is the case for newfoundland, where wolves (canis lupus) are absent due to their extirpation in the 1920s, and as a result, moose densities are much higher than elsewhere in canada. active control strategies to optimize condition of moose and moose habitat then arguably become a primary responsibility of managers (mclaren et al. 2004). several decades ago, pimlott (1953) suggested that inadequate hunting was the most pressing issue facing moose management, particularly in newfoundland. associated with his argument were concerns for inaccessibility of large areas, refusal of logging companies to permit hunter access to their licensed lands, public pressure against liberalizing hunting seasons and against the killing of females, and the lack of inclination and/or inability of most hunters to penetrate into a hunting area from an established road. another concern recognized during the 1960s was the inaccuracy of many moose aerial surveys (bergerud and manuel 1969). mercer (1995) and mercer and mclaren (2002) reviewed some of these same concerns and determined that problems still exist regarding the access by hunters of remote areas and the ability of moose surveys to detect local overpopulations in certain management units. current stated goals for moose in newfoundland include a target density of 2 moose / km2 of forested habitat in each moose management unit. yet, achievement of this goal is assessed only as an average for an entire management unit. mercer (1995:92) presents the position that more emphasis be placed on relationships between densities of moose and hunter moose-kill to achieve this target and also to stabilize populations. this paper considers these variables explicitly and reviews moose management in newfoundland in two periods, before and after mercer’s (1995) recommendation. we will focus our discussion on local moose overabundance in newfoundland. study area the island of newfoundland, 112,000 km2, is situated off mainland canada in the north atlantic, near 50° n latitude and 55° w longitude (fig. 1). about two-thirds of the island is forested, in an area roughly bisected by the route of the trans canada highway, plus in additional areas of the northern peninsula, avalon peninsula, west coast, and in river valleys along the south coast. most forests abies balsamea) and spruce (picea spp.), with a mix of pioneer (betula spp. and populus tremuloides), and tolerant (acer spp. and sorbus spp. and other) hardwoods. moose (a. a. andersoni) were introduced to central newfoundland in 1878 with the release of a male and female from nearby nova scotia (pimlott 1953). a second release of two males and two females from new brunswick, into western newfoundland, followed in 1904. the arrival of moose to the avalon peninsula appears to have been delayed by several decades due to slower migration across a narrow isthmus (broders et al. 1999). the island-wide population increased to record high numbers, about 150,000, by 1986 (mercer 1995), after which populations decreased, to a 1999 post-hunt estimate of 125,000 animals (mercer and mclaren 2002). moose now occupy all parts of the island of newfoundland, with higher densities in forested than in non-forested areas. more than half a million newfoundland moose have been taken by licensed hunters since 1945. the majority of licenses are sold alces vol. 41, 2005 mclaren and mercer density management 77 to residents, from access linked to the trans canada highway (ferguson et al. 1989). license sales offer limited-entry opportunities to hunters, with a combination for most units of a selective hunt by sex (mostly male-only and some female-only tags) and a more restrictive draw for either-sex moose tags. young of the year (calves) can be taken on both license types. calculation of the combined license quota for each management unit begins with a population estimate from the last available aerial survey and a management objective (i.e., an increasing, decreasing, or stable population). for each year since the last aerial survey, an estimate of population recruitment is based on observed productivity (i.e., the number of calves) at the time of the survey. recruitment is adjusted by estimates of natural, latewinter mortality. mortality estimates in the population also include the number of moose killed by hunters based on hunter information (submitted mandibles and a mandatory report of success on individual licenses), and losses to poaching, crippling, and moose-vehicle collisions (mercer 1995). recruitment and mortality estimates are combined with reports of the number of moose seen by hunters with either-sex tags. the reconstructed population is used to create a “quota.” quotas are most often changed when a management unit is periodically resurveyed (about once every 5-10 years) and found to be above or below “target.” hunters provide in many cases a written description of their kill location, as well as information on maps attached to their license returns as to the 5 km × 5 km area where they took their animal. demand for resident hunting in newfoundland has historically been high, and the number of licenses sold in most management units is very near to the quota. allocations to non-resident hunting vary by management unit, are often given special consideration in more remote areas, and usually complete age for newfoundland. there is no special consideration for first nations hunters in this part of the province. resident hunter success varies considerably both among and within newfoundland management units and is correlated both to the variability in moose density (ferguson and messier 1996) and to road accessibility (ferguson et al. 1989). in 2002, 26,360 moose licenses were sold for insular newfoundland. an average resident ing a season approximately from september to december. methods twenty-three sample moose management units, ranging in size from ca 700 to ca 4,500 km2, were selected from among those in insular newfoundland considered most accessible to resident hunters. this selection, chosen to represent a standardized level of access to hunting, excluded special management units (created in a few remote locations) but included any management unit intersecting the trans canada highway, plus two management units fig. 1. location of 23 moose management units along the trans canada highway in newfoundland, for which aerial survey data are available for two periods, 1985-1991 and 1993-2001. density management – mclaren and mercer alces vol. 41, 2005 78 on the avalon peninsula less than a 30 minute drive from the highway and less than an hour from metropolitan st. john’s (fig. 1). choice of management units was also based on availability of aerial population estimates in our two periods of interest, 1985-1991 and 1993-2001, roughly before and after management recommendations made in the early 1990s (mercer 1995). these two periods also correspond to the last peak moose density recorded and to the most recent population estimates for most areas (mercer and mclaren 2002). only two management units adjacent to the highway, units 8 and 10 in western newfoundland, were excluded from our analysis due to lack of data. moose population estimates were obtained from calculations (inland fish and wildlife division, 1985-2001, unpublished) made folblock survey (gasaway et al. 1986), usually conducted from mid-january to late march. cally from cessna-185 aircraft at an altitude of 50-150 m. census blocks ranged from 2-4 km2 and moose were counted by 3 observers and a pilot in a helicopter (typically bell 206b and 206-l). search intensity was usually 4-5 minutes / km2 has been used in all calculations to account for observer bias (unseen moose), based on averages from a set of calibration efforts in the 1980s (oosenbrug and ferguson 1992). moose density was calculated from estimates of the area of merchantable forest and nonmerchantable (“scrub”) cover from forest inventory maps of each management unit. license sales and estimates of moose kill were obtained for the same years as the moose survey, within the two study periods tourism, culture and recreation. kill density was calculated using estimates of forest and “scrub” cover in each management unit as same factor as in the calculation of moose density. kills were estimated from the reports of hunters responding to a questionnaire and adjusted for all non-respondents by the reports obtained in one mailed questionnaire reminder. kill estimates, then, are calculated as a proportion of license sales. thus, sales rather than quotas are reported in our correlations, although the difference between license sales comparisons of license sales and kill density with population estimates, all correlations p > 0.05 were reported. results moose license quotas in eastern newfoundland lated by management unit, especially in central newfoundland (table 1, fig. 2). by 1993-2001 population in all sampled management units, management units we sampled, including all but unit 44 on the avalon peninsula (table 2). aerial survey data became available for western newfoundland during the second period, and in this area, all units except unit 14 were asestimate. thus, a strong relationship between license quota and population size developed over time for the management units in our the estimated moose population size (r2 = 0.81) in the 1993-2001 period (fig. 2). a consistent and strong relationship existed between the number of licensed kills and the moose population in management units on the avalon peninsula, r2 = 0.99 (n = 3) during 1985-1991, and r2 = 0.91 (n = 6) during 1993-2001 (fig. 3). however, for central newfoundland, the same relationship was stronger during 1985-1991, r2 = 0.94 (n = 4), than during 1993-2001, r2 = 0.66 (n = 8). in a comparison of the earlier and the later periods, kill density increased while moose density decreased in most management units on the avalon peninsula and in central alces vol. 41, 2005 mclaren and mercer density management 79 newfoundland (tables 1 and 2). kill density with population density in our sample from western newfoundland, r2 = 0.39, where data were available only for 2001. discussion our recommendation to set quotas using density and habitat information for effective moose management that incorporates a “biodiversity” approach, data from aerial surveys must be combined with habitat information both before periodic aerial ment unit for survey purposes) and in the intervening time between surveys (when we argue that both moose density and habitat be given consideration during a quota-setting exercise). the easiest calculation of density in our calculation, to arrive at density from a measure of available habitat from forest inventory maps. we recognize that closer attention to habitat might distinguish categories of forest types and forest age, known to have very different carrying capacities (e.g., parker and morton 1978). variation in moose density, alternatively variation in carrying capacity, is important and is regrettably too often ignored from our review that the density of moose appears to be highly variable among newfoundland management units fig. 2. moose quotas, expressed as license sales, relative to population size in management units in western newfoundland (open circles), in central newfoundland (triangles), and on the avalon peninsula (squares), in 1985-1991 and 19932001. management units were selected along the trans canada highway, as shown in fig. 1. the dashed lines indicate a range, in which lation size. the solid line indicates the correlation between license sales and population size during 1993-2001 for all units, r2 = 0.81. even in a coarse measure of habitat availability. this variation is acknowledged by managers recognized for its potential to lead to overpopulation and habitat deterioration unless it land moose management, 1985-1991. target moose density in newfoundland is 2 moose / km2 of ment unit.1 1 hand, if moose use forest preferentially, then their density in forest cover alone would have a higher expected value than in our average for forest and “scrub” cover. unit survey year population estimate density1 quota license sales percent hunter success kill density percent of population killed central newfoundland (4 management units surveyed) 22 1990 6270 3.86 180 180 71.9 0.08 2.1 23 1991 8557 3.00 340 340 61.9 0.07 2.5 24 1985 3663 6.45 350 350 53.6 0.33 5.1 27 1989 6032 3.00 180 180 58.0 0.05 1.7 avalon peninsula (3 management units surveyed) 33 1987 1680 3.48 170 170 63.3 0.22 6.4 34 1986 2100 2.56 100 100 86.1 0.10 4.1 36 1986 5738 5.34 950 950 60.5 0.54 10.0 density management – mclaren and mercer alces vol. 41, 2005 80 is given special attention during quota setting. in the newfoundland example, moose management during 1985-1991 failed to address overpopulation by even a simple approach to proportional quota setting for management units 22, 23, and 27, an apparent failure particular to central areas (fig. 2). these management units, while recognized as containing some of the more remote hunting areas, do not actually have the highest moose densities. according to our calculations, they are nevertheless above target, like all areas during 1985-1991 (table 1). areas above target density should be managed for higher hunter kill, which has generally occurred, but with more apparent success on the avalon peninsula (fig. 3). during the period after recommendations to address density in management, closer correlation occurs between license sales (quotas) and population size (fig. 2). however, during the same period, less correlation occurs between kill density and population density, particularly in western newfoundland (fig. 3). the proportion of periods and the variation in hunter success and activity that we report suggest that management has improved in terms of a simplistic and proportional approach to population size, a goal probably typical of ungulate management. at the same time, there is no apparent effort to assign quotas to manage kill density proportionally to moose density. differences in kill density in a perfect manobjectives like targeting a population for reduction. however, manipulation of license quotas to achieve an increase or decrease in hunter kill is always a “best guess,” considering the accuracy and currency of data on moose and on hunting. moose population estimates are always inaccurate, especially given observer example. population estimates also have wide within a density stratum. lag time between surveys of up to 10 years creates further difkill is imperfectly estimated, because hunters do not completely or accurately report their kill activity or location. moreover, directing fig. 2. moose quotas, expressed as license sales, relative to population size in management units in western newfoundland (open circles), in central newfoundland (triangles), and on the along the trans canada highway, as shown in fig. 1. the dashed lines indicate a range, in which tion size. the solid line indicates the correlation between license sales and population size during r2 = 0.81. alces vol. 41, 2005 mclaren and mercer density management 81 hunter kill to target geographically isolated overpopulations will create a lower overall hunter success, contrary to normal management objectives. all the same, we present the case that optimal management will certainly not be achieved if quotas are assigned by assuming that a maximum sustained yield exists proportional to all population sizes. problems are certain to arise if hunter success is not viewed and monitored as a spatial variable, especially as overall hunter success declines with demographic changes and busier lifestyles. our very feasible recommendations are to factor habitat into an explicit measure of moose density, and further to acknowledge and account for spatial variation in hunter ingly, this approach has a few precedents in newfoundland, where management objectives to target overpopulations met with success. for example, in the 1960s, a small area of central newfoundland, located within management unit 16, had experienced failure in forest regeneration due to moose overabundance. and by a moose aerial census (bergerud and table 2. data references and calculations used in this study for the second of two periods in newfoundland moose management, 1993-2001. calculations are as in table 1. unit survey year population estimate density quota license sales percent hunter success kill density percent of population killed western newfoundland (8 management units surveyed) 4 1997 4992 1.69 1200 1198 57.8 0.23 13.9 5 1993 3548 2.75 900 899 82.0 0.57 20.8 6 1994 4674 3.03 1000 999 72.2 0.47 15.4 7 1994 2478 2.48 610 610 70.6 0.43 17.4 9 1996 1217 2.07 500 500 78.2 0.66 32.1 13 1997 1870 1.55 400 399 62.4 0.21 13.3 14 1997 5117 1.89 700 700 79.5 0.21 10.9 41 1997 2039 2.24 500 500 62.8 0.34 15.4 central newfoundland (8 management units surveyed) 15 1996 7759 3.02 1700 1694 65.9 0.43 14.4 16 1994 2695 2.37 500 495 54.4 0.24 10.0 21 1997 2736 1.92 500 500 61.6 0.22 11.3 22 2000 7490 4.61 1350 1334 71.9 0.59 12.8 24 1991 1237 2.18 300 300 53.6 0.28 13.0 27 1997 1600 0.80 400 399 58.0 0.12 14.5 28 2001 3226 1.80 400 400 62.4 0.14 7.7 42 1997 5106 4.65 800 764 45.3 0.32 6.8 avalon peninsula (6 management units surveyed) 31 1996 2319 4.48 600 600 65.1 0.75 16.8 33 1995 833 1.72 250 250 63.3 0.33 19.0 34 1997 2876 3.5 650 650 86.1 0.68 19.5 35 1995 548 0.91 150 150 45.4 0.11 12.4 36 1995 3402 3.17 700 700 60.5 0.39 12.4 44 1997 1710 6.13 300 300 82.3 0.88 14.4 density management – mclaren and mercer alces vol. 41, 2005 82 targeted for a reduction of moose using special hunting licenses (bergerud et al. 1968). in a contemporary example, mclaren et al. (2000) described the creation of a management subunit, within unit 15, to direct additional hunter effort where the central newfoundland forest industry was also sustaining losses. it seems unfortunate to us that examples have not arisen outside of forest economic concerns, when, as we stated in the introduction, negative ecological effects associated with habitat and subsequent moose declines can be avoided by addressing any case of overabundance. a modern, updated forest inventory as to clasabout moose habitat, which can be combined with fairly accurate reports by hunters of locations of moose seen and killed, to create a sophisticated, annual review of predicted moose density between aerial survey years. land discrepancies between the comparisons of license sales and population size (fig. 2) and of kills and density (fig. 3) may for any moose management unit or larger area indicate an explicit management objective, a failure to manage hunter success, or a failure to address overpopulation. reductions in moose density in newfoundland seem to have been typical between 1985 and 2001, as shown by increased license quotas (tables 1 and 2) and by changes to the relationship between kill density and population sizes (fig. 3). whether these reductions were part of an explicit or implicit management strategy is unclear. from our review of moose management in newfoundland between 1985 and 2001. a difference in management of central newfoundland and the avalon peninsula is easily apparent. in both time periods, central newfoundland experienced a lower proportional number of moose taken by huntfig. 3. moose kill density calculated from license sales and adjusted success rates, relative to population density in management units in newfoundland, in 1985-1991 and 1993-2001 (management units as in fig. 1, symbols as in fig. 2). the solid lines, a, indicate the correlations between moose kill density and population density for the avalon peninsula, r2 = 0.99 in 1985-1991 and r2 = 0.91 in 1993-2001. the solid lines, b, indicate the correlations between moose kill density and population density for central newfoundland, r2 = 0.94 in 1985-1991, and r2 = 0.63 in 1993-2001. kill is only weakly correlated with population size in western newfoundland, r2 = 0.39 (data are available only for 1993-2001). alces vol. 41, 2005 mclaren and mercer density management 83 ers. moreover, this difference was larger for management units of higher moose density. a general explanation of the trend may lie in the fact that road access is much more limited in central newfoundland than on the avalon peninsula, where about half the residents of the province have their permanent homes. for example, in 1985 license sales in unit 24 were proportional to the population estimate, similar to management units on the avalon peninsula (table 1, fig. 2). yet a density calculation illustrates an apparent failure to address overpopulation in the habitat-limited unit 24, relative to the apparently more intensive management approach that was possible on the avalon peninsula (fig. 3). the overby 1991 (table 2, fig. 3), but this change is actually the result of creating a new unit, 42, by subdividing unit 24. unit 42 is not only smaller in area with more moose than the revised unit 24, but also has proportionally less forested habitat and was, up until recently, less accessible (table 2). thus, by a concerted effort to direct the quota to a less accessible area, license sales in both management units fell near the expected value proportional to the two population estimates (fig. 2), but hunter success was low and calculated kill density in unit 42 fell well below the regression line predicted by moose density (table 2, fig. 3). recent resurvey of unit 42 (inland fish and wildlife division, unpublished data, 2004) 1997 and corroborates our suggestion of an unaddressed overpopulation that led to decline. on the avalon peninsula, where hunter accessibility is relatively high, moose management strategies have generally been quicker to respond to overpopulations. examples of this principle come from reviewing both the more and less accessible management units within the region. in the more accessible category accessible category is unit 36. (unit 31, like unit 36, is not accessed directly from the trans canada highway but has a direct highway has only regional highway access and contains a large wilderness area, where motorized vehicles are prohibited.) unit 36 was among the most densely moose populated managea population estimated at about a one-third reduction in density by the second period (fig. 3). while some of this change may have been due to a density-dependent decline resulting from overpopulation, we believe much is attributable to higher kill density documented even for the earlier period (table 1). in the other management units where high moose density occurs, for example, in units 31, 34, and 44, managers responded by increasing license quotas that resulted in higher kill density (table 2). in western newfoundland, kill density and moose density are not correlated (fig. 3), and the highest proportional quotas occur (fig. 2). because management units in this area are similar in density to the target for newfoundland (table 2), we argue that they are actually managed inconsistently and in some cases hunted unsustainably toward population 6 and 41, which are among the highest in kill density despite near target moose density. we view these cases as further examples of density in assigning quotas. acknowledgements we are grateful to the inland fish and wildlife division of the government of newfoundland and labrador for providing the information required for this review. the idea for this paper arose from the management plan for moose drafted by w. e. mercer for the government of newfoundland and labrador. density management – mclaren and mercer alces vol. 41, 2005 84 references bergerud, a. t., and f. manuel. 1969. aerial census of moose in central newfoundland. journal of wildlife management 33:910–916. , , and h. whalen. 1968. the harvest reduction of a moose population in newfoundland. journal of wildlife management 32:722–728. broders, h. g., s. p. mahoney, w. a. montevecchi, and w. s. davidson. 1999. population genetic structure and the effect of founder events on the genetic variability of moose, alces alces, in canada. molecular ecology 8:1309–1315. ferguson, s. h., w. e. mercer, and s. m. oosenbrug. 1989. the relationship between hunter accessibility and moose condition in newfoundland. alces 25:36–47. ______, and f. messier. 1996. can human predation of moose cause population cycles? alces 32:149–161. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. fairbanks, alaska, usa. mclaren, b. e., s. p. mahoney, t. s. porter, and s. m. oosenbrug. 2000. spatial and temporal patterns of use by moose of pre-commercially thinned, naturallytral newfoundland. forest ecology and management 133:179–196. ______, b. a. roberts, n. djan-chékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40:45-59. mcshea, w. j., h. b. underwood, and j. h. rappole, editors. 1997. the science of overabundance, deer ecology and population management. smithsonian institution press, washington, d.c., usa. mercer, w. e. 1995. moose management with inland fish and wildlife division, newfoundland and labrador, corner brook, nl, canada. , and b. e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose. alces 38:123–141. oosenbrug, s. m., and s. h. ferguson. 1992. moose mark-recapture survey in newfoundland. alces 28:21-30. parker, g. r., and l. d. morton. 1978. the estimation of winter forage and its use by moose on clearcuts in northcentral newfoundland. journal of range management 31:300–304. pimlott, d. h. 1953. newfoundland moose. transactions of the north american wildlife conference 18:563–581. timmermann, h. r., and m. e. buss. 1998. population and harvest management. pages 559–616 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice 4010.p65 alces vol. 40, 2004 brimeyer and thomas wyoming management 133 history of moose management in wyoming and recent trends in jackson hole douglas g. brimeyer1 and timothy p. thomas2 1wyoming game and fish department, po box 67, jackson, wy 83001, usa; 2wyoming game and fish department, po box 6249, sheridan, wy 82801, usa abstract: moose are believed to have entered wyoming from montana and idaho within the past 150 years. moose did not become established in jackson hole until the early 1900s. in 1903, the wyoming state legislature closed moose hunting seasons. in 1908, agency reports indicate moose were distributed along the tetons, the upper yellowstone river, and at the head of the green river. by 1912, population estimates increased to 500 moose and the hunting season was reopened. moose began to occupy portions of the wind river range during the 1930s and became quite numerous by the 1960s. moose were first introduced in the bighorn mountains in 1948. moose moved into the medicine bow mountains from colorado in the 1980s. moose presently occupy habitats in western, north central, and southeastern wyoming. statewide, managers recognize 14 distinct herd units. each herd has a postseason population objective that is set according to biological, sociological, and political considerations. population estimates are based on aerial surveys, hunter harvests, and age structure of the harvest. check stations are utilized to monitor harvest rates and collect data from harvested animals along exit routes from popular hunting areas. wyoming currently utilizes population modeling, indices, and in some cases sample estimates such as sightability models to estimate moose populations. the 2001 statewide population was estimated at approximately 13,657 moose. the statewide population objective is 14,650 moose. hunting seasons in wyoming are traditionally conservative and hunter success has generally remained in the 80–90% range. a harvest questionnaire is sent annually to all moose permit holders. response is typically 80%. harvest data are used to calculate hunter success, days of hunter effort per moose harvested, total harvest, and harvest composition. in 2001, a total of 1,379 hunters harvested 806 bulls, 337 cows, and 72 calves. these hunters had an 89% success rate and spent an average of 6.2 days hunting / animal harvested. in 2001, a total of $360 per moose was generated from license revenue while management costs were $411 per moose. a restitution value of $7,500 for illegally taking moose has been established. statewide, a total of 2,308 moose were classified in 4 herd units and the calf:cow ratio was 35 calves per 100 cows and the bull:cow ratio was 62 bulls per 100 cows. statewide, populations have remained relatively stable, however, in the jackson area moose populations have declined in recent years. license quotas have decreased from an average of 410 licenses during 1991 –1995 to 248 during 1999 – 2003. calf:cow ratios declined from a 1963 – 1993 average of 48 calves:100 cows (se=8.9) to a 1998 – 2003 average of 34:100 (se=8.3) low pregnancy and twinning rates have been reported in this herd. habitat changes and predator population increases are thought to be contributing to this decline. alces vol. 40: 133-143 (2004) key words: classification, distribution, harvest, history, hunting seasons, population, wyoming george shiras iii described a mountain race of moose during his explorations in yellowstone national park, from 1908 – 1910. in honor of shiras, nelson (1914) named the yellowstone moose alces alces shirasi. moose in the yellowstone region continue to be a keystone species among wyoming wildlife and symbolic of wyowyoming management – brimeyer and thomas alces vol. 40, 2004 134 ming’s wildlife heritage. history in wyoming few if any moose were believed to exist in the yellowstone and jackson hole areas of wyoming prior to 1850 (houston 1968) and moose that populated wyoming were thought to have migrated from montana (koch 1941, spaulding 1956, curtright 1969, schladweiler 1974). between 1834 and 1843, osborn russell kept detailed records of his travels and observations in western wyoming where moose are now found in abundance yet he did not mention them in his journals. moose likely increased and expanded shortly after russell’s travels. in july, 1872, members of the hayden expedition killed a cow and a calf moose while camped in jackson hole (reeves and mccabe 1998). on the west side of the teton mountains, richard (beaver dick) leigh mentioned the abundant moose sign he observed and the encounters he had with moose during june 1875 (blair 1987). the first wyoming territorial legislature convened in cheyenne on october 1869 and a rudimentary act for the protection of game and fish in the territory of wyoming was passed. this act provided for the sale of elk, deer, antelope, and mountain sheep or their young between february 1 and august 1. there was no closed season for the hunting of big game, only a restriction on the sale of game. penalization however was through civil rather than criminal law and so the act was largely ignored. the first hunting season for big game was set in 1875 and the season ran from august 15 to january 15. in 1882 moose and mountain goat were added to the list of protected species. section 1 of the game code read: “it shall be unlawful to pursue, hunt or kill any deer, elk, moose, mountain sheep, goat, antelope or buffalo save only from august 1 to november 15 inclusive in each year, or kill or capture by means of any pit, pitfall or trap any of the above named animals”. following early settlement, the numbers of moose declined and in 1899 the legislature enacted laws granting moose complete protection (wgfc 1948). blunt (1950) reported that in 1903 the state legislature granted protection for 10 years. the 1908 annual report of the state game warden indicated moose were distributed along the teton range, the upper yellowstone river, and at the head of the green river. moose seasons were reopened in 1912. in the 1912 annual report it was estimated that there were 500 moose in wyoming, principally in the northwest region. in 1915 a total of 19 moose licenses were sold for $160 each and 16 moose were harvested. by 1916, it was estimated that 2,000 moose roamed western wyoming. between 1935 and 1948 the legal harvest totaled 1,515 moose in wyoming. moose were still an unusual sight, often mistaken for elk and accidentally killed. during the 1938 elk season a total of 23 moose were accidentally killed. it was also mentioned in the 1937 –1938 biennial report that there was an abundance of ticks on moose during the spring of 1938 (wgfc 1938). managers estimated the population at 3,210 moose in 1940. moose began to occupy portions of the wind river range during the 1930s (smith 1982). while natural dispersal was occurring in northwestern wyoming, efforts were underway to expand moose distribution in other parts of the state. in 1934, 17 moose were captured in jackson hole. these animals were individually crated for shipping and died in transit. in 1948, 8 moose were baited into a trap and transported by trucks to the bighorn mountains in north central wyoming. several moose calves were also captured during 1948, by roping from a horse. all of the calves captured this way died in the holding corral within a month. in 1950, an additional 8 moose were alces vol. 40, 2004 brimeyer and thomas wyoming management 135 captured in jackson and released in the bighorn mountains. an additional 13 moose were captured in the jackson area and released in the bighorns in 1974 and 1987. twelve moose were relocated from jackson to north park, colorado in 1979. another 12 moose were transplanted to the upper laramie river in colorado in 1987. moose emigrated from the colorado population into southeast wyoming and by 2000, a huntable population inhabited the snowy range. artificial feeding although the department does not support supplemental feeding of moose, feeding has occurred in the past in western wyoming (johnson et al. 1985). moose feeding areas were started at the urging of landowners in areas where extensive damage repeatedly occurred to stored hay and livestock feed lines. the wyoming game and fish department is legally responsible for paying for damage to private property caused by big and trophy game. as early as 1949 the wyoming game and fish department fed 40 moose on an emergency basis near moran junction. through the late 1960s approximately 75 moose fed with private ranch horses in grand teton national park. in the 1970s the feed ground was moved from the park to the moosehead ranch, a private in-holding within the park. moose were fed on this ranch from december through april. the wyoming game and fish department annually supplied 15 – 20 tons of hay for moose on this ranch through the early 1980s. during the 1970s the department began providing alfalfa (medicago sativa) to ranchers to minimize damage to stored hay. by the early 1980s there were 5 moose feeding areas approved by the wyoming game and fish commission along with several other unofficial moose feedgrounds. moose were fed 1 kg/day during the early winter and up to 7 kg/day by february. feeding usually began in january and ended in mid-march. it is likely that moose only supplemented their normal browse diets at these feed lines. most of these sites were phased out by the early 1990s. between 1970 and 1984, up to 586 moose were fed each year at 4 sites south of jackson near big piney. these sites were located on cottonwood creek, horse creek, beaver creek, and the green river (johnson et al. 1985). the feedgrounds in these drainages were unsuccessful in eliminating damages to private property from moose. moose generally fed at these sites and then ranged as far away as 2 miles. unofficial moose feeding sites were also started in the gros ventre drainage and in bondurant along the hoback river. in the gros ventre drainage east of jackson, moose regularly frequented the patrol cabin elk feed ground. in the mid-1970s up to 75 moose visited the elk feed line where baled alfalfa hay was scattered. in general, moose would consume hay before elk came into the feed line and then bed down in the willows adjacent to the feed ground. currently less than 12 moose visit this feed line. in bondurant along the hoback river, moose were fed in 2 locations on private land. the first was near the town of bondurant and the other was on jack creek. current status and management moose presently occupy habitats in western, north central, and southeastern wyoming (fig. 1). statewide, managers recognize 14 distinct herd units. each herd unit contains a discrete population for which emigration and immigration between adjacent herds accounts for less than 10% of the herd unit’s population. hydrologic divides, major rivers, and other natural and man-made barriers generally constitute herd unit boundaries. herd units are further wyoming management – brimeyer and thomas alces vol. 40, 2004 136 subdivided into hunt areas to provide management flexibility. there are a total of 41 hunt areas in wyoming. each herd has a postseason population objective that is set according to biological, sociological, and political considerations. statewide, managers estimated this population at 13,657 moose in 2001. the statewide postseason population objective was 14,650 moose. population census and modeling population estimates are based on population simulations along with population indices and in some cases sample estimates such as sightability models. since 1976, wgfd has been modeling big game populations using pop-ii or its precursors fig. 1. delineated moose herd units in wyoming 2003. (hnilicka and zornes 1994). models are useful tools for estimating populations and evaluating harvest strategies. modeling assumptions for pop-ii include: (1) effects of emigration and immigration are negligible; (2) natural mortality rates affect all age cohorts in a predictable linear fashion; (3) estimates of sex and age classifications, harvest, natural mortality, and wounding loss mirror reality; and (4) effects of density-dependent or other feed-back mechanisms are negligible (bartholow 1992, guenzel 1994). population modeling requires data collected by field personnel including herd classifications, harvest composition, and mortality rates, and are validated using trend count data. population alces vol. 40, 2004 brimeyer and thomas wyoming management 137 estimates prior to 1976 are based on trend counts only. accurate age and sex ratios are needed to analyze herd dynamics and reliably estimate moose populations. aerial surveys are the most practical method of classifying moose over large areas and diverse habitats during the post hunt period. aerial surveys often enable managers to meet sampling assumptions and more observations are recorded per unit effort than during ground surveys. all aerial surveys follow protocol outlined in the aircraft operation procedures and safety policy of the wyoming game and fish department policy manual (wgfc 2002). aerial surveys utilizing fixed-wing aircraft were first initiated in 1952 (wgfc 1952). previous surveys were conducted from the ground or incidentally during aerial surveys for elk. currently wgfd conducts helicopter surveys to classify herds by sex and age and to conduct trend counts. surveys are flown annually or biennially during the winter months when moose are concentrated on winter ranges. summer helicopter surveys are conducted in the bighorn herd unit when moose are in more open habitats. desired sample sizes are calculated on herds to produce an 80% confidence interval + 10 percent for sex and age classification ratios. moose are the least gregarious of all wyoming ungulates and are often observed alone or in small groups. during winter sightability flights in the jackson area during 1998 and 1999, the average group size was 2.2 moose per group (range 1-15) out of 358 groups observed. in addition to small group size, moose are often segregated by sex and age making some classes of animals more difficult to observe than others (peek et al. 1974). because of group size and their distribution in various habitats, it is uneconomical to get a total count of moose. adequate samples are usually only achieved on the larger moose herds. on smaller herds, ground or fixed-wing surveys are often conducted. preseason classifications are conducted by helicopter surveys in some herds because moose are more visible during the preseason period than during postseason when animals move into conifer habitats (wgfd 1999). preseason classifications are conducted over a week period in july or august. a shorter time frame is used to avoid duplicate counts. consistent survey techniques are utilized to make valid comparisons from year to year. flights are conducted in early morning and limited to 23 hours in duration. preseason classifications are used to estimate fall recruitment, verify population trend, and used in population simulation models. ground classifications are done in some herd units due to low moose densities and budget constraints. postseason classifications are conducted during december through february. a helicopter is generally used, however ground classifications may be warranted in herd units with low moose densities and due to budget constraints. surveys are conducted when good snow cover is present, preferably within a few days of fresh snow. surveys cover representative areas of riparian, deciduous, and conifer habitats frequented by moose. partial surveys of an area may produce biased composition estimates. in alaska, gasaway et. al. (1986) determined that classifications that were conducted during surveys designed for population estimation purposes resulted in higher, more representative calf:cow ratios than did less intensive composition surveys. all moose encountered during aerial surveys are assigned to one of the following classes: bulls, yearling bulls, cows, calves, or unclassified adults. in addition, cows with calves are tallied and those having twins are noted. body size tends to be the most useful criteria for identification of wyoming management – brimeyer and thomas alces vol. 40, 2004 138 calves. head features are used to avoid classifying large calves as yearlings. calf moose have relatively small ears and short, pointed noses compared to the larger bulbous nose of an adult moose. moose calves also tend to remain close to the cow and will often follow close behind her when disturbed. during the post hunt period observers identify 2 or more criteria when sexing adult moose. the criteria include: antler or pedicel scars, vulva patch, behavior, bell shape and size, group composition, and body conformation. not all females have the characteristic vulva patch and some males have a small light brown area that can be mistaken for a vulva patch. timmermann and buss (1998) provide a summary of the different criteria helpful in identifying the sex and age of moose. classification results are recorded by location on a hand-held gps and then d o w n l o a d e d t o a m i c r o s o f t e x c e l spreadsheet and mapped using various softw a r e p r o g r a m s s u c h a s a r c m a p , arcexplore, and all topo. location records help managers identify important habitats and their proximity to human development, recreational areas, timber sales, etc. in some herd units, a wyoming moose sightability model has been utilized to calculate a population estimate using correction factors that compensate for effects of vegetation cover, group size, etc. utilizing procedures described by unsworth et al. (1991), anderson and lindzey (1995) developed a sightablity model for moose in wyoming. sightability surveys are designed to sample a portion of a stratified survey area in order to improve accuracy and precision of population estimates while reducing survey costs. the survey is conducted in early winter (i.e., december and january) when moose are still located in open habitats. procedures outlined in the aerial survey user’s manual (unsworth et al. 1994) are followed when observing and recording moose and evaluating vegetation cover. data are then transcribed and entered into the model based upon the format described by anderson and lindzey (1995), and evaluated using the aerial survey software (unsworth et al. 1994). the wyoming sightability model was utilized in 1996 – 1997 on a portion of the jackson herd unit. estimates from 1996 – 1998 indicate that an average of 63% of moose were observed during helicopter surveys in willow, cottonwood, and coniferous habitat. in colorado, bowden and kufeld (1995) observed 58% of radiocollared moose during helicopter surveys. herd unit statistics statewide, the calf:cow ratio was 35 calves per 100 cows following the hunting season in 2001. a total of 2,308 moose were classified in 4 herd units. typical lower to upper ranges include 31 calves per 100 cows in the jackson herd unit and 45 calves per 100 cows in the lander herd unit. adequate classification sample size at the 80% confidence level is regularly obtained in only the sublette herd unit. the average calf:cow ratio was 43 calves per 100 cows in this herd during the 19972002 period. the bull:cow ratio was 63 bulls per 100 cows statewide with a high of 68 bulls per 100 cows in the lincoln herd unit. a total of 37 bulls per 100 cows were reported in the lander herd unit. the 1997 – 2001 average in the sublette herd was 63 bulls per 100 cows. the age at which cow moose first reproduce varies from 1.4 to 2.4 years (schwartz 1992) and may be delayed to 3.4 years in poor quality range (albright and keith 1987). in jackson hole, houston (1968) reported a yearling pregnancy rate of 5 – 6 % in jackson hole during the 1960s. more recently it is suspected that female moose reproduce at age 3 in jackson hole. a 2 % t w i n n i n g r a t e w a s o b s e r v e d alces vol. 40, 2004 brimeyer and thomas wyoming management 139 during the 1999 post hunt period in the jackson and targhee herd units. houston (1968) reported a 4% twinning rate in the jackson area during the 1960s. range of natural mortality for population simulation models in wyoming, we estimate calf mortality rates at 10 – 25% preseason. post hunting season calf mortality rates range from 15 – 25%. as large predator populations become established in wyoming, mortality rates used in population modeling rates are likely to change. simulation models in wyoming utilize a 98 – 100% survival rate for adult females and a 100% survival rate for adult males during the pre-hunt period. age and sex specific survival rates between 85100% are used for simulation models in wyoming during the winter period. hunting seasons hunting seasons in wyoming have been conservative and hunter success has generally remained in the 80 – 100% range. in an effort to improve calf moose survival, hunting regulations were changed in 1998 to prohibit hunters from taking a cow moose accompanied by a calf. this action was taken to improve survival of dependent calves and increase recruitment. at the beginning of the hunting season, all moose hunters are sent a packet of information that includes a hunter survey card, a tooth collection box, a n d i n s t r u c t i o n s f o r t o o t h r e m o v a l . in 2001 a total of 1,379 hunters harvested 806 bulls, 337 cows, and 72 calves. these hunters had an 89% success rate and spent an average of 6.2 days hunting / animal harvested. in 2001, a total of $360 per moose was generated from license revenue while management costs were $411 per moose. wyoming statute §23-6-204(e) grants the wyoming game and fish commission authority to recommend to courts the amount of restitution for the value of wildlife taken in violation of title 23, the game and fish act, that the court might impose as a penalty on any person. in 2003 the commission established a $7,500 restitution value to courts for individual animals. the restitution value was derived from average costs incurred and/or lost to the state. they include hunter expenditures, management costs, license revenue, and other commercial revenues. during the hunting season, hunter check stations and random field checks give managers an opportunity to collect biological data from harvested animals along exit routes from popular hunting areas. in addition, check stations enable field personnel to contact sportsmen, monitor regulation compliance, and improve public relations by informing the public of department operations (wgfc 2003). age data are collected from harvested animals by inspecting and/or collecting the front incisor teeth of harvested moose. specific ages of harvested moose are determined from sectioned teeth (sergeant and pimlott 1959). aging is based on counting cementum annuli, which can be microscopically examined in the tooth cross-sections. this information has several management uses. harvested antlerless moose are considered an un-biased sample of yearling and older age classes within the female segment of the population. accordingly, managers can estimate the age structure of the female segment using age data from harvested, antlerless moose. ages of harvested bulls do not represent a valid age structure of the male segment because hunters tend to select older age classes. however, the age distribution of harvested bulls is useful to assess current harvest trends in relation to harvest objectives. multipleyear shifts in age composition of the male harvest may indicate a need to adjust license numbers, or may provide evidence of changing moose numbers or sex ratios. agewyoming management – brimeyer and thomas alces vol. 40, 2004 140 specific harvests are incorporated into population simulation models each year. longerterm age data can be used to determine the oldest age classes and female age structure for population modeling purposes. in 2001, 578 male and 237 female tooth samples were collected during the hunting season. a total of 508 (89%) of the antlered animals were 1 – 7 years of age and 60 (11%) of the antlered samples were older than 8 years of age. the average of males > 1 year of age was 4.8 years old and the oldest male in the harvest was 12.3 years of age. of the antlerless harvest excluding calves, 185 (84%) were 1 – 7 years of age and 34 (15%) were older than 8 years of age. the average of females > 1 year of age was 4.6 years old and the oldest female in the harvest was 18.3 years of age. following the hunting season a second harvest questionnaire is sent to all moose permit holders who did not return survey cards. initial response is typically 75-80%. nonrespondents are contacted by phone. harvest data are used to calculate hunter success, days of hunter effort per moose harvested, total harvest, and harvest composition. questionnaire results are considered accurate. distribution and movement moose distribution and movement patterns have been documented and mapped in varying degrees throughout wyoming. distribution information is used to determine seasonal ranges for each herd. biologists and others use distribution maps to develop comments on proposed land use developments and other planning purposes. jackson moose trends statewide, populations have remained relatively stable, however, in the jackson area moose populations have declined in recent years. the jackson moose herd is located in northwest wyoming and encompasses 2,000 square miles. the area covers the gros ventre, buffalo fork, and upper snake river drainages in teton county. the jackson herd unit is primarily public land. the majority of moose in this herd spend spring, summer, and fall at mid-tohigh elevations and drift down drainage to winter ranges along the gros ventre, buffalo fork, or snake rivers. some moose stay at higher elevations and winter in the spruce/fir zone. upland vegetation communities in the herd unit include subalpine fir (abies lasiocarpa), englemann spruce (picea 0 20 40 60 80 100 120 0.5 2.3 4.3 6.3 8.3 10.3 12.3 14.3 16.3 18.3 age n u m b er male frequency female frequency fig. 2. male and female age frequency in the 2001 harvest of the jackson moose herd. alces vol. 40, 2004 brimeyer and thomas wyoming management 141 engelmanni ), lodgepole pine (pinus c o n t o r t a) , d o u g l a s f i r ( p s u e d o t s u g a m e n z i e s i i ) , a n d a s p e n ( p o p u l u s tremuloides). riparian areas are charact e r i z e d b y c o t t o n w o o d ( p o p u l u s angustifolia) and willow (salix spp.). extensive stands of mountain big sage (artemisia tridentata) and antelope bitterbrush (purshia tridentata) are found in a portion of jackson hole north of the town of jackson. the bitterbrush is used extensively by moose in fall and spring and when available during the winter. elevation of occupied habitat ranges from 6,000 10,000 feet. population simulations indicate that this population has declined from a high of 3,500 moose to 2,100 moose from 1992 through 2003. since 1981, a bell 47 helicopter has been used to classify moose. a total of 477 moose were classified in 2001: 161 bulls (34%), 240 cows (51%), and 74 calves (16%). herd ratios were 67 bulls:100 cows:31 calves (fig. 2). the bull:cow ratio was higher than 2000 (64 bulls:100 cows) and the 1996 2000 average (66 bulls:100 cows). the calf:cow ratio observed in 2001 was lower than 2000 (39 calves:100 cows) and the 1996 2000 average (36 calves:100 cows). lower moose numbers were most apparent on winter ranges located 30 miles north of jackson in the buffalo valley. between 1985 –1989 an average of 335 moose were observed on winter range in the buffalo valley. between 1995 – 1999 an average of 183 moose were observed on this same winter range and between 2000 and 2002 an average of 123 moose were classified. correspondingly license quotas have also decreased. between 1991-1995 an average of 410 licenses were issued in the jackson moose herd unit. from 1999 – 2003 the average number of moose licenses was 248. a total of 145 moose licenses are proposed for 2003 (table 1). calf:cow ratable 1. jackson moose license quotas, harvest, trend count, and population estimates 1991 2003. tios declined from a 1963 – 1993 average of 48 calves:100 cows (se=8.9) to a 1998 – 2003 average of 34:100 (se=8.3). while moose numbers have declined over time, it is thought that the large predator populations have increased. the cause and effect of the predator/prey relationship 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 a ntlered licenses 290 265 240 180 195 200 205 205 205 205 200 185 135 a ntlerless licenses 205 200 190 135 150 145 115 105 95 80 60 45 10 t otal p erm its 495 465 430 315 345 345 320 310 300 290 260 265 145 a ntlered harvest 213 188 181 164 163 170 159 142 153 158 159 136 a ntlerless harvest 180 176 139 130 138 117 93 92 69 62 48 30 h unter success (% ) 80 80 78 93 88 87 81 76 74 78 80 74 d ay s/anim al 6.5 6.4 7.4 4.8 6 6.2 6 7.3 11.7 8.49 7.8 10.1 t rend count 683 927 757 975 970 803 712 815 830 589 480 513 c alf:cow ratio 44 43.3 34 58.5 37.5 25.7 36.2 42.6 35.1 39.2 30.6 21.1 wyoming management – brimeyer and thomas alces vol. 40, 2004 142 between moose and grizzly bears is largely unknown in the southern portion of the yellowstone ecosystem. it is thought that moose pregnancy rates are low (~75 %) because of present habitat conditions and that moose calf survival is being influenced by the recolonization of large carnivores. an increased effort is being made to monitor pregnancy rates in other portions of the jackson herd. references albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south-coast barrens of newfoundland. canadian field naturalist 101:373-387. anderson, c. r., and f. g. lindzey. 1995. a sightability model for moose developed from helicopter surveys. wildlife society bulletin 24:247-259. bartholow, j. 1992. pop-ii system documentation. fossil creek software. fort collins, colorado, usa. blair, n. 1987. the history of wildlife management in wyoming. wyoming game and fish department, cheyenne, wyoming, usa. blunt, f.m. 1950. untitled notes on moose in wyoming. wyoming wildlife 14(1):21. bowden, d.c., and r.c. kufeld. 1995. generalized mark-sight population size estimation applied to colorado moose. journal of wildlife management 59:840851. curtright, p. r. 1969. lewis and clark: pioneering naturalists. university of illinois press, urbana, illinois, usa. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, fairbanks, no.22. gueznel, r. j. 1994. adapting new techniques to population management: wyoming’s pronghorn experience. transactions of the north american wildlife and natural resources conference 59:189-200. hnilicka, p., and m. zornes. 1994. status and management of moose in wyoming. alces 30:101-107. houston, d. b. 1968. the shiras moose in jackson hole, wyoming technical bulletin 1. grand teton natural history association. johnson, b. k., j. k. straley, and g. roby. 1985. supplemental feeding of moose in western wyoming for damage prevention. alces 21:139-148. koch, e. 1941. big game in montana from early historical records. journal of wildlife management 5:358-370. nelson, e. w. 1914. description of a new subspecies of moose from wyoming. proceedings of the biological society, washington, d.c. 27:71-74. peek, j. m., r. e. leresche, and d. r. stevens. 1974. dynamics of moose aggregations in alaska, minnesota and montana. journal of mammalogy 55:126-137. reeves, h. m., and r. e. mccabe. 1998. of moose and man. pages 1–75 in a. w. franzmann and c. c. scwhartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. schladweiler, p. 1974. ecology of shiras moose in montana. federal aid job final report. nos. w-98-r and w-120r. montana department of fish, wildlife and parks, helena, montana, usa. schwartz, c. c. 1992. reproductive biology of north american moose. alces 29:165-173. sergent, d. e., and d. h. pimlott. 1959. age determination in moose from sectioned incisor teeth. journal of wildlife management 23:315-321. alces vol. 40, 2004 brimeyer and thomas wyoming management 143 smith, b. l. 1982. the history, current status and management of moose on the wind river reservation. u.s. fish and wildlife service, lander, wyoming, usa. spaulding, k. a. (editor) 1956. the fur hunters of the far west. university of oklahoma press, norman, oklahoma, usa. timmermann, h. r., and m. e. buss. 1998. population and harvest management. pages 559 – 615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. unsworth, j.w., f.a. leban, d.j. leptich, e. o. garton, and p. zager. 1994. aerial survey: user’s manual, second edition. idaho department of fish and game, boise, idaho, usa. _____ , _____, g.a. s a r g e a n t , e.o. garton, m.a. hurley, j.r. pope, and d.j. leptich. 1991. aerial survey: user’s manual. idaho department of fish and game, boise, idaho, usa. (wgfc) wyoming game and fish commission. 1938. biennial report. cheyenne, wyoming, usa. _____. 1948. biennial report. cheyenne, wyoming, usa. _____. 1949. annual report. cheyenne, wyoming, usa. _____. 1952. annual report. cheyenne, wyoming, usa. _____. 2002. aircraft operations procedures and safety. policy vd. wyoming game and fish department, cheyenne, wyoming, usa. _____. 2003. restitution values for wildlife. policy vii m. wyoming game and fish department, cheyenne, wyoming, usa. (wgfd) wyoming game and fish department. 1999. sheridan region annual big game herd unit reports 1998. wyoming game and fish department, cheyenne, wyoming, usa. p25-35_4103.pdf alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 25 integrating habitat use and population dynamics of moose in northern new hampshire david scarpitti1, christopher habeck2, anthony r. musante, and peter j. pekins department of natural resources, university of new hampshire, durham, nh 03824, usa abstract: the new hampshire fish and game department and the university of new hampshire initiated research in northern new hampshire to better understand population dynamics and seasonal habitat use of a moose population that has apparently stabilized, despite optimal habitat and modest harvest levels. in total, 94 moose were captured by helicopter (81 net-gunned and 13 tranquilized) in december 2001-2003 and 2 were darted at salt-licks in july of 2002. capture mortality attributed to myopathy and injury was 4%. in comparison to measured reproduction during capture (63 and 100%), our ability to measure pregnancy by direct observations (69 and 100%) was validated in 2002-2003. production was 0.82 and 0.85 calves per adult cow; rate of twinning was 20 and 10%. calf mortality 2 months post-partum was similar (26 and 27%) each year. annual mortality of adult/yearling moose was 27 and 12%. hunting and vehicle collision mortality was 4 (all adult cows) and 6% (all calves but 1) each year. high annual winter calf mortality (38-43%) in late march and early april was associated with the combined effects of malnutrition and winter tick/lung nematodes. winter home range size was not restricted, and composition of available habitat was similar across seasons although overlap was minimal between seasons. consideration of habitat and population dynamics data suggests that alces vol. 41: 25-35 (2005) key words: capture, core areas, home range, mortality, population dynamics, production, survival moose (alces alces) have become an extremely valuable resource in northern new hampshire in the past 20 years. moose watching tours and casual visitation of moose viewing areas (predominantly road-side salt licks) by tourists provide measurable eco(silverberg 2000). revenue derived from hunting permits increases annually and fuels other hunting related purchases (bontaites and recreational and economic importance of the moose population, and the direct relationship of herd health to commercial forest management, it is important to manage this population moose research in the northeastern united cally, only a 2-year study (miller 1989) with marked moose has addressed habitat use in new hampshire. population indices derived from deer hunter surveys, road collisions, and infrared aerial surveys suggest that new hampshire’s northern moose population has approached stability, despite perceived high quality and quantity of suitable habitat and modest harvest levels (k. bontaites, new hampshire fish and game department, unpublished data). it is presumed that substantial non-hunting mortality of unknown sources occurs in the population. mortality and survival rates, rate 1present address: massachusetts division of marine fisheries, 50 portside drive, pocasset, ma 02559, usa 2present address: department of zoology, 1630 linden drive, university of wisconsin, madison, wi 53706, usa integrating habitat and populaton dynamics – scarpitti et al. alces vol. 41, 2005 26 of production, and habitat use are fundamental parameters in understanding moose population dynamics (van ballenberghe and ballard 1998). however, habitat use and life history of moose in new england are different from that in much moose range, and population dynamics has not been extensively researched in this region. seasonal home range and habitat use were measured in northern new hampshire (miller 1989) and maine (thompson et al. 1995) previously, however, both studies were relatively short-term. the overall objective of this study was to measure seasonal home range and habitat use, productivity, and rate, timing, and cause of mortality of cow and calf moose for 4 continuous years. this paper highlights population dynamics and habitat use data collected in december 2001-2003, and evaluates the from a helicopter. study area the study area was in eastern coos county in northern new hampshire (fig. 1) encompassing approximately 650 km2 of rolling to mountainous, forested terrain. the study area includes numerous ponds and lakes, with the androscoggin river located centrally within. the majority of the study area is working commercial forestland. dominant forest types include northern hardwoods (34%) as a mix of yellow birch (betula alleghaniensis), american beech (fagus grandifolia), and sugar maple (acer saccharum) on more well drained sites (degraaf et al. 1992). on more common (23%), consisting almost entirely of red spruce (picea rubens (abies balsamea). approximately 17% of the study area was mixed forest, typically including yellow and paper birch (betula papyrifera other important communities included clearcuts and regeneration stands (14%) of aspen (populus tremuloides) and paper birch (betula papyrifera) where pin cherry (prunus serotina) was common. tamarack (larix laricina) and northern white cedar (thuja occidentalis) were found on very poorly drained soils. wetlands, including open water, accounted for 10% of the total study area. mean weekly snow depth mea< 50 cm in 2002 and 42.6 cm (8-70 cm) in 2003. seasonal temperatures range from < -30 to black bears (ursus americana), coyote (canis latrans), and bobcat (lynx rufus); white-tailed deer (odocoileus virginianus) were sympatric with moose throughout the area. both sexes of moose are hunted annually by a permitlottery system; hunter success rates typically exceed 85% (nhfg 2002). methods moose were captured from a helicopter with net-guns (n = 81, carpenter and innes 1995) or tranquilizers (n = 13) in december vhf or gps radio collar (vhf: model 600, gps: model tgw-3700, telonics, inc., mesa, arizona, usa). moose were tranquilized with a mixture of carfentanil citrate and xylazine hydrochloride and were antagonized with a mixture of naltrexone hydrochloride and tolazoline hydrochloride given intravenously. respiration and temperature were monitored by a veterinarian. approximately 20 ml of blood, fecal pellets, and hair and parasite samples were collected from each moose. certain moose, typically adults, were tranquilized from the helicopter in 2001 because the lack of snow cover affected the ability of the pilot to identify snags/branches that could snag the net or rotors. spotter planes were used to help locate potential target moose in 2002 and 2003. the primary capture goal varied each year to ensure that 25 adult cows were available for productivity measurements each spring. the capture targets were 25 cows and 15 calves alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 27 in 2001, 4 cows and 21 calves in 2002, and 1 cow and 25 calves in 2003. all non-calves of calves was not problematic. in july 2002, 2 adult cows were tranquilized at a salt lick to redeploy gps collars. radio collared moose were located by aerial and ground-based telemetry methods (mech 1983) 1-3 times weekly year-round, or by direct observation during spring-summer as described below. ground-based telemetry was performed with tr-5, tr-3, tr-2 digital and analog receivers (telonics, inc., mesa, arizona, usa) and 3-element yagi directional antennas (af antronics, inc., urbana, illinois, usa). because moose travel extensively at night during summer, monitoring occurred evenly in 4 discrete time periods to avoid bias associated with diurnal only sampling hours), bedding (0900–1500 hrs), feeding (1500–2100 hrs), and salt lick use (2100–0300 hrs). dead moose, as indicated by mortality sensors, were located within 48 hours and a the probable cause of death. pregnancy of cows captured in 2001 was tein b in blood samples (huang et al. 2000), and with a portable ultrasound unit in 2002 (stephenson et al. 1995). beginning 1 may, fig. 1. study area located in eastern coos county, new hampshire, usa. n 6km0 androscoggin river pontook reservoir integrating habitat and populaton dynamics – scarpitti et al. alces vol. 41, 2005 28 close observation of adult cows was attempted 2-3 times weekly to document production, fecundity, and calf survival/mortality for 2 months post-partum. those cows believed barren were observed weekly through july to identify late births. technicians used homing techniques (mech 1983) to stalk adult cows within sighting distance to identify the presence and number of calves. annual mortality rates of all moose were calculated each calendar year (2002: n = 28, 2003: n = 47); moose with dropped collars were precluded from this analysis. calf winter (1 december – 1 may) mortality rates were calculated from observations of radio-collared calves (approximately 7 months old) following capture in 2001-2003. because 8 of 14 radio-collared calves dropped collars in winter 2002 following capture and the remaining 6 died, reported winter mortality in 2002 represents a minimum mortality rate. telemetry data and locations from visual observations were recorded with handheld gps units (garmin, miami, florida, usa). triangulation data were recorded in universal transverse mercator (utm) coordinates and analyzed with the locate ii program (nams 2000), which utilizes maximum likelihood estimators (lenth 1981) to estimate tion). telemetry error was estimated from a radio-collar at known locations. telemetry locations were used to construct annual and seasonal 95% minimum convex polygon (mcp, mohr 1947) home ranges using the floating amean method (rodgers and (16 december–15 april), spring (16 april–15 june), summer (16 june-15 september), and fall (16 september–15 december). locations were plotted and analyzed with arcview™ geographic information system (gis) 3.3 (esri, redlands, california, usa) and hre: the home range extension for arcview™ (rodgers and carr 1998). seasonal core areas of locations was plotted against mean home range size to evaluate if correlation existed. cows that were monitored for 2 consecutive years and were observed with a newborn calf each year (n = 9) were included in the home range analysis. this increased statistical power by accounting for variability between moose and years. percent overlap was calculated as the proportion of the overlapped area between 2 seasons divided by the total area encompassed by the 2 seasonal ranges. proportional data form a binomial distribution, thus percent overlap was arcsine transformed to allow use of parametric statistics (zar 1999), though data presented are actual proportions. analysis of variance (anova) was used to examine for differences in mean seasonal home range size, overlap, and percent overlap; moose and year were blocking factors. pairwise comparisons were made with the proportion of 6 major habitat types within annual and seasonal home ranges, and seasonal core areas of 10 reproductive cow moose were calculated in 2002. limitations and sample size prevented comparisons in multiple years. forest cover types interpreted from granit 2001 landcover assessment were grouped into 6 categories characterized by the dominant vegetation; northern hardwood (< 25% coniferous basal area per acre), coniferous (> 65% coniferous basal area), mixed (> 25% and < 65% coniferous basal area), wetland, cut-regeneration, and developed/other (includes agricultural land, residential and commercial housing, and bedrock). due to statistical issues of non-independence and 0-values, the proportion of available habitat was not statistically analyzed, but rather was used as an overall indication of home range composition and compared across seasons. alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 29 results captures capture targets were met each year in conditions ranging from minimal to >75 cm snow depth. the capture rate ranged from 6-17 moose daily with captures spread throughout the day. the highest capture rates were in 2003 when calves were targeted (24 of 25 targeted animals), snow depth was greatest, and a spotter plane helped locate moose for the helicopter crew. only 1 of 94 moose escaped without being collared during the helicopter capture procedure. mortality (all calves) during or associated (myopathy within 6 days) with capture was 4% (4 of 94). the composition of captured moose was 31 yearling/adult cows and 63 calves (31 female and 32 male). the rate of pregnancy measured in 2001 and 2002 was 63 (15 of 24) and 100% (4 of 4). eight calves dropped their collars in late winter 2002. population dynamics calves were observed with 15 of 22 (68%) cows and 20 of 26 (77%) cows monitored in spring 2002 and 2003. neonates were observed from 14 may-14 july with 69% of births occurring 14-24 may. all but 2 observed with a calf. fecundity was 0.82 and 0.85 calves per cow in 2002 and 2003, respectively; twinning rate was 3 of 15 (20%) and 2 of 20 (10%) each year. calf mortality (observations of calves with radio-collared cows) at 2 months post-partum was similar both years (26 and 27%); cause of death was unknown in all cases, but was assumed to include predation by black bears and incidental mortality (e.g., abandonment, drowning; child 1998). the annual yearling/adult mortality was 27 (6 of 22) and 12% (4 of 26) in 2002 and 2003. annual mortality of all moose due to vehicle collisions was 6% (all calves but 1) both years and occurred mostly (66%) in spring-summer. hunting accounted for 4% mortality each year (all adult cows). radio-collared calves represented 56% (14 of 25) of overall mortality. of 14 radio-collared calves in 2002, 8 dropped collars and the remaining 6 animals died over winter (1 december-1 may). overwinter mortality of collared calves in 2003 was 38% (8 of 21). of those calves that died over winter, most (66 and 63% each year) died from natural causes attributed to the combined or cumulative symptoms of malnutrition and parasites (winter tick, dermacentor albipictus, lung nematode, dictyocaulus viviparus); 93% of this type of mortality occurred during a 4 week period, 27 march-27 april. mortality was in proportion to the sex ratio of marked calves. habitat use the number of seasonal locations per cow (n = 9) ranged from 9-43; the number of locations was not correlated to home range size (p = 0.12). annual home range ranged from 15.5 to 66.4 km2. seasonal home range ranged from 2.2 to 46.5 km2 but were largely similar, with mean fall range (17.4 km2 larger than mean spring range (7.8 km2; p = 0.008). although not statistically different, mean winter (15.1 km2) and summer (14.1 km2) ranges were both nearly 2x larger than mean spring range. core areas (70% mcp) were 46% smaller, on average, than seasonal ranges (95% mcp) and varied from 1.1 to 37.0 km2; no seasonal differences were detected, although the trend in size was similar to that with seasonal home range (table 1). the percent overlap of seasonal home range and core areas ranged from 20 to 34 and 12 to 27% (table 2). the mean area and percent overlap of seasonal home ranges and core areas was greatest in summer and fall (7.7 km2 and 34%, 2.8 km2 and 27%) and smallest in winter and spring (3.6 km2 and 20%, 1.0 km2 and 12%). the area of overlap of home range and core area in summer and fall was greater than in winter and spring (p = integrating habitat and populaton dynamics – scarpitti et al. alces vol. 41, 2005 30 0.006 and 0.008). the proportion of overlap of home ranges (p = 0.016) and core areas (p = 0.019) showed similar patterns; percent summer and fall overlap was greater than that in winter and spring. the composition of habitat within annual and seasonal home ranges (fig. 2) and core areas was similar. deciduous (northern hardwood) forest was the dominant habitat in home ranges and core areas across seasons, comprising the largest percentage of home ranges and core areas in spring (38 and 37%). coniferous forest was the second most abundant habitat type in all seasonal home ranges and core areas; availability of coniferous habitat was largest in summer (24 and 28%). availability of mixed forest habitat was highest in fall home ranges (19%). clearcut/regeneration stands were most prevalent in home ranges and core areas during fall (16 and 15%). available wetland habitat within home ranges was lowest during spring (6%) and similar in other seasons. core areas contained a higher proportion of available table 1. mean size and range (km2) of seasonal home ranges (95% mcp), core areas (70% mcp), and mean number of locations per home range of 9 radio collared adult cow moose observed with a calf in consecutive years in northern new hampshire, 2002-2003. standard error for 95% mcp= 2.0 and for 70% mcp = 1.3. 95% mcp 70% mcp mean # of loc./home rangeseason mean range mean range winter 15.1 4.9-29.3 6.7 1.2-16.5 36.4 spring 7.8 2.2-17.7 3.9 1.1-10.1 16.1 summer 14.1 5.1-39.5 5.9 2.8-19.5 35.4 fall1 17.4 3.3-46.5 8.6 0.8-37.0 19.4 note: mcp = minimum convex polygon. 1greater than spring (p = 0.008) for seasonal range (95%). table 2. mean size (km2) and proportion of range overlap between consecutive seasons for seasonal home ranges (95% mcp) and core areas (70% mcp) of 9 radio collared adult cow moose observed with a calf in successive years in northern new hampshire, 2002-2003. standard error for overlap size = 0.8 for 95 and 70% mcp, for overlap proportion = 0.03 for 95 and 0.04 for 70% mcps. overlap size overlap proportion season 95% mcp 70% mcp 95% mcp 70% mcp winter-fall 6.7 1.4 0.28 0.15 winter-spring 3.6 1 0.2 0.12 spring-summer 4.8 0.9 0.29 0.12 summer-fall 7.71 2.72 0.343 0.274 winter-summer 5.7 1.8 0.27 0.22 spring-fall 4.4 1.3 0.26 0.15 note: mcp = minimum convex polygon. 1different from winter-spring (p = 0.006). 2different from winter-spring (p = 0.008). 3different from winter-spring (p = 0.016). 4different from winter-spring (p = 0.019). alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 31 wetland habitat in winter (16%). discussion the daily capture rates in 2001 and 2002, without the use of a spotter plane and in variable snow conditions, were similar to those reported by carpenter and innes (610; 1995). our capture rates were higher in 2003 (7-17) because snow (> 70 cm) impeded moose movement, increased sightability of moose, and provided optimal contrast for both pilot and net-gunner to recognize the optimal capture situation. further, the spotter plane effectively reduced the search time of the helicopter crew, thereby increasing their effort in the actual capture process. early winter, as suggested by carpenter and innes (1995), was an optimal time to capture moose was improved with adequate snow cover and a spotter plane. the capture mortality rate (4%, all calves) was between that reported by carpenter and innes (< 1%, 1995) and olterman et al. (7%, 1994). calves were apparently more susceptible to injury and mortality, as also suggested by data of carpenter and innes (1995). when net guns were not used, moose were immobilized chemically from the helicopter with no injury or mortality. the high capture rate measured in this study point to the advantage of capturing moose from helicopters when large numbers of animals are desired. the presumed accuracy of pregnancy testing in december (2 months post-breeding) was 90-95% with radio immunoassays and ultrasound (stephenson et al. 1995, huang et al. 2000). the pregnancy rates measured at capture in 2001 and 2002 (63% and 100%), and the proportion of marked cows observed with calves in spring (68% and 100%) were nearly identical, supporting the validity of our fecundity measurements derived from direct observations while stalking moose. the annual pregnancy estimates (68 and 77%) were lower than those of adult cows reported in populations below carrying capacity in newfoundland (87%, pimlott 1959), new brunswick (79%, boer 1987), alaska (84100%, schwartz 1998), and ontario (97%, bergerud and snider 1988). the observed fig. 2. mean percent composition of annual and seasonal habitat types within home ranges of 10 reproductive adult cow moose in northern new hampshire, 2002. 0% 10% 20% 30% 40% deciduous coniferous mixed cut-regeneration wetland developed/other habitat type p er ce n t co m p o si ti o n ( % ) annual winter spring summer fall integrating habitat and populaton dynamics – scarpitti et al. alces vol. 41, 2005 32 twinning rates (10-20%) were analogous to populations considered near or above carrying capacity (1-25%; edwards and ritcey 1958, blood 1974, albright and keith 1987). as expected, calf survival (74 and 73%) through summer (2 months post-partum) was much higher than in areas with several large predators. neonatal survival was only 17-27% in regions with bears and wolves in alaska (testa et al. 2000, bertram and vivion 2002). substantial neonatal mortality (> 25%) was when calves are most susceptible. however, calves radio-collared in december also experienced higher than expected over-winter mortality in 2002 (> 43% of calves died; 6 of 14 calves died, the rest dropped collars) and in 2003 (38%) presumably a result of malnutrition and parasites in march-april. the annual survival rate of calves and yearling/adult cows and fecundity of yearling/ adult cows measured in this study were used to calculate the predicted rate of population change with the leslie-lewis matrix method (goodman 1980). this analysis, although premature with 2 years data, was conducted to assess the impact of the perceived high annual mortality of calves. input data were 34% annual calf survival, 79% adult cow survival, and fecundity of 0.41 (female calves rate of population change was 0.93, a value suggesting slight decline in the annual population associated mostly with calf mortality. the current (2001-2003) estimated population density in the study area is of 0.7 moose/km2 (k. bontaites, new hampshire fish and game department, unpublished data), has not changed since 1995, and is not the highest in the state (adams et al. 1997). calf mortality was associated with the combined effects of malnutrition and high parasite loads. although a high parasitic load is often described as a density-dependent characteristic, and tick-related mortality appears to be related to nutrition and overall body condition of the host, mortality is most pronounced when tick numbers are highest (lankester and samuel 1998). presumably, calves are more susceptible than adult cows to the deleterious effect of ticks, particularly at high tick loads that produce excessive hair loss and anemia (glines and samuel 1989). however, tick abundance is not necessarily associated with moose density because environmental conditions are most important in determining tick density. snow cover and cold temperatures reduce tick transmission rates to moose in fall and survival of engorged females in spring (drew and samuel 1989, samuel and welch 1991). moderate snowfall within the project area may have contributed to increased tick density and subsequent tick loading on calf moose. if malnutrition and mortality of calves in march-april was related to habitat quality, analysis of use and availability of seasonal habitats should indicate differential use. however, winter home range was not restricted in size relative to other seasons or compared to previous studies in new hampshire and maine (table 1), and composition of available habitat was largely similar in all seasons (fig. 2), implying that moose habitat was universally distributed across the study area. further, moderate snow depth in both winters suggests that environmental conditions did not limit mobility or access to forage. habitat in northern new hampshire is generally considered high quality, with a mosaic of different age class stands distributed throughout the study area providing ample forage and cover as a result of commercial timber harvesting. consideration of the habitat use data in concert with population dynamics data suggests that both density dependent and the study population. pregnancy rates and fecundity rates were moderate, suggesting use and availability appear unconstrained. summer home range sizes were smaller than alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 33 reported in other studies in this region (table 3). adult cow mortality was low, yet calf mortality with minimal predation was high, and the calculated annual rate of change predicts population decline. further, most calf mortality was associated with a parasite that does these collective data point to the complexity of habitat and population relationships, and the need to conduct long-term population studies. we presume that collection of similar data for two additional years will better population dynamics of moose in northern new hampshire. acknowledgements funding for this research was provided by the new hampshire fish and game department. the professional efforts of hawkins and powers, inc., greybull, wyoming are recognized for capturing moose in often adverse weather and environmental conditions. this study was possible because of the cooperation and access provided by many commercial landowners including, but not limited to, wagner forest management, ltd., international paper, inc., plum creek timber company, inc., meade corporation, and hancock timber resource group. local residents granted access to other private lands. ron hamel provided aerial surveys critical to data collection. numerous unefforts. dr. christopher neefus and kent gustafson provided statistical consultation. kristine bontaites, nhfg moose biologist, was essential in creating and implementing this project. references adams, k. p., p. j. pekins, k. a. gustafson, and k. m. bontaites. 1997. evaluation of infrared technology for aerial moose surveys in new hampshire. alces 33:129-139. albright, c. a., and l. b. keith. 1987. population dynamics of moose, alces alces, on the south –coast barrens of newfoundland. canadian field-naturalist 101:373-387. ballard, w. b., j. s. miller, and d. j. reed. 1991. population dynamics of moose in south central alaska. wildlife monographs 114. table 3. comparison of seasonal home range size estimates (km2) of adult cow moose from selected studies using radio telemetry. partially reconstructed from hundertmark (1998). 1summer range estimate consists of summer and fall locations. 2winter range estimate consists of winter-spring locations and summer range consists of summer-fall locations. season location method winter spring summer fall reference new hampshire mcp 15.1 7.8 14.1 17.4 this study new hampshire hm 3.9 55.3 81.7 miller (1989) maine mcp 4.3 24.8 2.6 thompson et al. (1995) minnesota1 hrf 3.6 12.7 phillips et al. (1973) alberta mcp 47 21.6 27 15.9 lynch and morgantini (1984) colorado mcp 5 7.3 5.8 10.2 kufeld and bowden (1996) montana2 mcp 21.2 16 van dyke et al. (1995) alaska mcp 43.1 39.8 60.6 ballard et al. (1991) sweden hm 4.9 6.9 9.1 5.6 cederlund and okarma (1988) integrating habitat and populaton dynamics – scarpitti et al. alces vol. 41, 2005 34 bergerud, a. t., and j. b. snider. 1988. predation in the dynamics of moose populations: a reply. journal of wildlife management 52:559-564. bertram, m. r., and m. t. vivion. 2002. moose mortality in eastern interior alaska. journal of wildlife management 66:747-756. beyer, d. e., jr., and j. b. haufler. 1994. diurnal versus 24-hour sampling of habitat use. journal of wildlife management 58:178-180. blood, d. a. 1974. variation in reproduction and productivity of an enclosed herd of moose (alces alces). transactions of the international conference of game biology 11:59-66. boer, a. h. 1987. reproductive productivity of moose in new brunswick. alces 23:49-60. bontaites, k. m., and k. guftason. 1993. the history and status of moose management in new hampshire. alces 29:163-168. carpenter, l. h., and j. i. innes. 1995. helicopter netgunning: a successful moose capture technique. alces 31:181-184. cederlund, g., and h. okarma. 1988. home range and habitat use of adult female moose. journal of wildlife management 52: 336-343. child, k. n. 1998. incidental mortality. pages 275-301 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. degraaf, r. m., m. yaminski, w. b. leak, and j. w. lanier. 1992. new england wildlife: management of forested habitats. general technical report ne144, randor, pa: usda, forest service, northeast forest experiment station. drew, m. l., and w. m. samuel. 1989. reproduction in winter tick, dermacentor albipictus alberta, canada. canadian journal of zoology 64:714-721. edwards, r. y., and r. w. ritcey. 1958. reproduction in a moose population. journal of wildlife management 22:261-268. glines, m. v., and w. m. samuel. 1989. the effect of dermacentor albipictus (acarina: ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. experimental and applied acarology 6:197-213. goodman, d. 1980. demographic interactions for closely managed populations. pages 171-195 in m. e. soule and b. a. wilcox, editors. conservation biology: an evolutionary ecological perspective. sinaer association incorporated, sunder-sunderland, massachusetts, usa. huang, f., d. c. cockrell, t. r. stephenson, j. h. noyes, and r. g. sasser. 2000. radioimmunoassay for moose and elk wildlife management 64:492-499. hundertmark, h. j. 1998. home range, dispersal and migration. pages 303-350 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. kufeld, r. c., and d. c. bowden. 1996. movements and habitat selection of shiras moose (alces alces shirasi) in colorado. alces 32: 85-99. lankester, m. w., and w. m. samuel. 1998. pests, parasites, and diseases. pages 479-517 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. lenth, r. v. 1981. robust measures of location for directional data. technometrics 23:77-81. lynch, g. m., and l. e. morgantini. 1984. sex and age differential in seasonal home range of moose in northwestern alberta. alces vol. 41, 2005 scarpitti et al. – integrating habitat and populaton dynamics 35 alces 20: 61-78. mech, l. d. 1983. handbook of animal radio tracking. university of minnesota press, minneapolis, minnesota, usa. miller, b. k. 1989. seasonal movementseasonal movement patterns and habitat use of moose in northern new hampshire. m.sc. thesis, university of new hampshire, durham, new hampshire, usa. mohr, c. o. 1947. table of equivalent populations of north american small mammals. american midland naturalist 37:223-249. nams, v. o. 2000. locate ii user guide. truro, nova scotia, canada. (nhfg) new hampshire fish and game department. 2002. annual harvest summary report. new hampshire fish and game department, concord, new hampshire, usa. olterman, j. h., d. w. kenvin, and r. c. kufeld. 1994. moose transplant to southwestern colorado. alces 30:1-8. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37: 266-278. pimlott, d. h. 1959. reproduction and productivity of newfoundland moose. journal of wildlife management 23:381401. rodgers, a. r., and a. p. carr. 1998. hre: the home range extension for arcview™. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. samuel, w. m., and d. a. welch. 1991. winter ticks on moose and other ungulates: alces 27:169-182. schwartz, c. c. 1998. reproduction, natality and growth. pages 141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. silverberg, j. k. a. 2000. impacts of wild-impacts of wildlife viewing: a case study of the dixville notch wildlife viewing area. ph.d. thesis, university of new hampshire, durham, new hampshire, usa. stephenson, t. r., j. w. testa, g. p. adams, r. g. sasser, c. c. schwartz, and k. j. hundertmark. 1995. diagnosis of pregnancy and twinning in moose by ultrasonography and serum assay. alces 31:167-172. testa, j. w., e. f. becker, and g. r. lee. 2000. temporal patterns in the survival of twin and single moose (alces alces) calves in southcentral alaska. journal of mammalogy 81:162-168. thompson, m. e., j. r. gilbert, g. j. matula jr., and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31:223245. van ballenberghe, v., and w. ballard. 1998. population dynamics. pages 223-245 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. van dyke, f., b. l. probert, and g. m. van beek. 1995. seasonal habitat use characteristics of moose in south-central montana. alces 31: 15-26. zar, j. h. 1999. biostatistical analysis. fifth edition. prentice-hall, incorporated, englewood cliffs, new jersey, usa. << /ascii85encodepages false /allowtransparency false /autopositionepsfiles true /autorotatepages /all /binding /left /calgrayprofile (dot gain 20%) /calrgbprofile (srgb iec61966-2.1) /calcmykprofile (u.s. web coated \050swop\051 v2) /srgbprofile (srgb iec61966-2.1) /cannotembedfontpolicy /warning /compatibilitylevel 1.4 /compressobjects /tags /compresspages true /convertimagestoindexed true /passthroughjpegimages true /createjdffile false /createjobticket false /defaultrenderingintent /default /detectblends true /detectcurves 0.0000 /colorconversionstrategy /leavecolorunchanged /dothumbnails false /embedallfonts true /embedopentype false /parseiccprofilesincomments true /embedjoboptions true /dscreportinglevel 0 /emitdscwarnings false /endpage -1 /imagememory 1048576 /lockdistillerparams false /maxsubsetpct 100 /optimize true /opm 1 /parsedsccomments true /parsedsccommentsfordocinfo true /preservecopypage true /preservedicmykvalues true /preserveepsinfo true /preserveflatness true /preservehalftoneinfo false /preserveopicomments false /preserveoverprintsettings true /startpage 1 /subsetfonts true /transferfunctioninfo /apply /ucrandbginfo /preserve /useprologue false /colorsettingsfile () /alwaysembed [ true ] /neverembed [ true ] /antialiascolorimages false /cropcolorimages true /colorimageminresolution 300 /colorimageminresolutionpolicy /ok /downsamplecolorimages true /colorimagedownsampletype /bicubic /colorimageresolution 300 /colorimagedepth -1 /colorimagemindownsampledepth 1 /colorimagedownsamplethreshold 1.50000 /encodecolorimages true /colorimagefilter /dctencode /autofiltercolorimages true /colorimageautofilterstrategy /jpeg /coloracsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /colorimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000coloracsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000colorimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasgrayimages false /cropgrayimages true /grayimageminresolution 300 /grayimageminresolutionpolicy /ok /downsamplegrayimages true /grayimagedownsampletype /bicubic /grayimageresolution 300 /grayimagedepth -1 /grayimagemindownsampledepth 2 /grayimagedownsamplethreshold 1.50000 /encodegrayimages true /grayimagefilter /dctencode /autofiltergrayimages true /grayimageautofilterstrategy /jpeg /grayacsimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /grayimagedict << /qfactor 0.15 /hsamples [1 1 1 1] /vsamples [1 1 1 1] >> /jpeg2000grayacsimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /jpeg2000grayimagedict << /tilewidth 256 /tileheight 256 /quality 30 >> /antialiasmonoimages false /cropmonoimages true /monoimageminresolution 1200 /monoimageminresolutionpolicy /ok /downsamplemonoimages true /monoimagedownsampletype /bicubic /monoimageresolution 1200 /monoimagedepth -1 /monoimagedownsamplethreshold 1.50000 /encodemonoimages true /monoimagefilter /ccittfaxencode /monoimagedict << /k -1 >> /allowpsxobjects false /checkcompliance [ /none ] /pdfx1acheck false /pdfx3check false /pdfxcompliantpdfonly false /pdfxnotrimboxerror true /pdfxtrimboxtomediaboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxsetbleedboxtomediabox true /pdfxbleedboxtotrimboxoffset [ 0.00000 0.00000 0.00000 0.00000 ] /pdfxoutputintentprofile () /pdfxoutputconditionidentifier () /pdfxoutputcondition () /pdfxregistryname () /pdfxtrapped /false /description << /chs /cht /dan /deu /esp /fra /ita /jpn /kor /nld (gebruik deze instellingen om adobe pdf-documenten te maken voor kwaliteitsafdrukken op desktopprinters en proofers. de gemaakte pdf-documenten kunnen worden geopend met acrobat en adobe reader 5.0 en hoger.) /nor /ptb /suo /sve /enu (use these settings to create adobe pdf documents for quality printing on desktop printers and proofers. created pdf documents can be opened with acrobat and adobe reader 5.0 and later.) >> /namespace [ (adobe) (common) (1.0) ] /othernamespaces [ << /asreaderspreads false /cropimagestoframes true /errorcontrol /warnandcontinue /flattenerignorespreadoverrides false /includeguidesgrids false /includenonprinting false /includeslug false /namespace [ (adobe) (indesign) (4.0) ] /omitplacedbitmaps false /omitplacedeps false /omitplacedpdf false /simulateoverprint /legacy >> << /addbleedmarks false /addcolorbars false /addcropmarks false /addpageinfo false /addregmarks false /convertcolors /noconversion /destinationprofilename () /destinationprofileselector /na /downsample16bitimages true /flattenerpreset << /presetselector /mediumresolution >> /formelements false /generatestructure true /includebookmarks false /includehyperlinks false /includeinteractive false /includelayers false /includeprofiles true /multimediahandling /useobjectsettings /namespace [ (adobe) (creativesuite) (2.0) ] /pdfxoutputintentprofileselector /na /preserveediting true /untaggedcmykhandling /leaveuntagged /untaggedrgbhandling /leaveuntagged /usedocumentbleed false >> ] >> setdistillerparams << /hwresolution [2400 2400] /pagesize [612.000 792.000] >> setpagedevice alces34(1)_107.pdf alces35_159.pdf alces35_59.pdf 75 aquatic areas provide high nitrogen forage for moose (alces alces) in isle royale national park, michigan, usa keren b. tischler1, william j. severud2, rolf o. peterson1, john a. vucetich1, and joseph k. bump1,3 1school of forest resources and environmental science, michigan technological university, houghton, michigan 49931, usa; 2department of natural resource management, south dakota state university, brookings, south dakota 57007, usa; 3present address: department of fisheries, wildlife, and conservation biology, university of minnesota, saint paul, minnesota 55108, usa abstract: the distribution of ungulates reflects spatial and temporal heterogeneity in forage quality and quantity across the landscape. aquatic habitats have a patchy spatial distribution and are readily used by moose (alces alces) and other ecotone specialists. however, the importance of aquatic feeding to moose has largely been attributed to acquisition of sodium, with little consideration given to the relative and comparative quality of aquatic and terrestrial forage types. we show differences in forage quality as measured by crude protein content and carbon:nitrogen (c:n) ratios between aquatic and terrestrial summer forage in isle royale national park, michigan, usa. aquatic macrophytes had higher crude protein content and lower c:n ratio than preferred terrestrial plant species of moose. consequently, measurable consumption of aquatic forage may provide high quality forage in less than optimal habitats. because the distribution of aquatic habitats on isle royale exhibits strong spatial trends, the benefits of aquatic feeding may have spatial influence on the population dynamics of isle royale moose. alces vol. 58: 75 – 90 (2022) key words: alces alces, aquatic macrophytes, c:n ratio, crude protein, forage quality, isle royale, moose, nitrogen ecotone specialists that consume both terrestrial and aquatic resources link food webs between these two realms (bartels et al. 2012, severud et al. 2013, johnston 2017, bump 2018). the patchy spatial and temporal distribution of aquatic cover types, and the quality and quantity of forage contained therein, are important factors in predicting the landscape distribution and density of ecotone specialists (crawley 1983, mcnaughton 1988, fryxell 1991, wallis devries 1996, johnston and windels 2015). indeed, moose (alces alces) link aquatic and terrestrial biomes due to their extensive foraging activities in both habitats during summer (peterson 1955, qvarnemark and sheldon 2004, peek 2007, tischler et al. 2019). moose are restricted to northern latitudes characterized by a high degree of seasonal variability (e.g., short growing season) in forage quality and quantity (timmermann and rodgers 2017). due to the low quality of winter forage, moose are in negative energy balance during winter and rely upon energy reserves attained during late summer and autumn to maintain energy balance yearround (delgiudice et al. 1997, 2011, moen et al. 1997, schwartz and renecker 2007). to maximize these reserves, moose face a trade-off between the benefits of exploiting high quality forage patches with potential costs of predation risk (edwards 1983), high ambient air temperature (renecker and hudson 1986, 1990), and insect avoidance aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 76 (renecker and hudson 1990). as a result, moose exhibit distinctive habitat use patterns including summer foraging in aquatic habitats. aquatic habitats are typically abundant across boreal landscapes and readily used by moose in summer (peterson 1955, qvarnemark and sheldon 2004, peek 2007). moose also use deeper water to protect against predation (mech 1966, gasaway et al. 1983, jordan et al. 2010) and seek relief from high ambient temperatures (renecker and hudson 1986, 1990). the quality of aquatic forage, in terms of available energy, protein, and essential nutrients, has been the focus of many studies and some debate (botkin et al. 1973, fraser et al. 1980, 1984, jordan 1987). due to its scarcity in continental ecosystems, sodium may be a limiting nutrient for north american herbivores (hutchinson and deevey 1949), and belovsky (1981) and jordan (1987) noted the high sodium content of aquatic macrophytes. several investigators have advanced the hypothesis that moose seek aquatic habitats explicitly to satisfy sodium requirements (hutchinson and deevey 1949, jordan et al. 1973, belovsky 1981, fraser et al. 1984, jordan 1987). however, sodium was not considered the predominant factor for moose consuming emergent aquatics on the copper river delta, alaska. rather, maccracken et al. (1993) considered that the impetus for aquatic foraging in some systems was that aquatic forage was nutritious, high in digestible energy and crude protein. however, few studies have assessed the overall nutrition, protein, and energy associated with aquatic forage for moose, and importantly, as compared with terrestrial forage (but see fraser et al. 1984 and maccracken et al. 1993). our goal was to measure and compare the relative nutritional value of aquatic and terrestrial moose forages as measured by crude protein content and carbon:nitrogen (c:n) ratios at isle royale national park, michigan, usa. we predicted that aquatic macrophytes contain more protein and have lower c:n ratios than terrestrial plants. in this system, sodium is a seasonally important nutrient for moose (jordan 1987), with aquatic habitats dominated by submergent (plants either free-floating or entirely submerged beneath water surface) rather than emergent (plant parts emergent above water surface) species. apart from vegetation, sodium is also available at mineral licks and springs on isle royale which may be in sufficient abundance to meet the nutritional needs of moose given that these sources are well-used by moose (risenhoover and peterson 1986). we additionally compared forage quality among plant species composing the principal terrestrial summer diet of moose on isle royale, and evaluated spatial differences in forage quality between eastern and western sides of the island due to different glacial history (huber 1973; see study area). study area isle royale is a 544 km2 island archipelago located in the boreal forest region of northwestern lake superior, usa, 24 km from the nearest shoreline (48° n, 89 °w). the primary island consists of precambrian-aged basalt and conglomerate bedrock shaped by the last glaciation into a series of parallel ridges and valleys including numerous water bodies (huber 1973). lakes and ponds (n = 84 ≥ 1 ha) comprise 36 km2 of the surface area, with an additional 8 km2 of palustrine emergent wetlands. additional shoreline is found in numerous bays of lake superior, particularly at the east end of the island. as a result of glacial activity, soils are more developed on the west end of isle royale (huber 1973). fire has historically burned the entire east end, and relatively recent fires (1936 and 1948) have burned the alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 77 midsection of the island. moose have been on isle royale since the early 1900s, with winter density ranging from 1 to 4 moose/km2 across the island in the past ~50 years (vucetich and peterson 2004). the legacy of the island’s disturbance regime has resulted in forest succession following different trajectories on the east and west ends of the island. forests on the west end are in a late successional stage and dominated by deciduous species, while forests on the east end are younger and conifer-dominated (janke et al. 1978). mean daily high temperature is 20°c in summer and −3°c in winter (de jager et al. 2020). snow and ice cover persist from november through april, and the islands receive ~750 mm of precipitation annually (risenhoover and maass 1987). in winter, moose concentrate along shoreline areas where balsam fir (abies balsamea) is ~ 60% of the diet, with the remainder woody browse and arboreal lichens (risenhoover 1987, parikh et al. 2017, tischler et al. 2019). the summer spatial distribution and local density of moose on isle royale is unknown, but the spring diet of moose includes newly emergent leaves and the summer diet is largely composed of current leaf growth of deciduous plants and aquatic macrophytes (ackerman 1987, tischler et al. 2019). methods sample design we collected aquatic macrophytes and leaves of terrestrial plant at the east and west ends of the island (hereafter e and w, respectively). the e and w sampling sites were delineated by the boundary of the 1936 and 1948 fires, leaving the central portion of the island unsampled. we collected samples between 13 july and 3 august 2002 when plants were mature, as opposed to emergent or senescent. samples were collected at ≥ 5  e and w sites from the 6 terrestrial species composing the principal summer diet of moose on isle royale: mountain maple (acer spicatum), sugar maple (a. saccharum), mountain-ash (sorbus spp.), paper birch (betula papyrifera), yellow birch (b. alleghaniensis), and beaked hazelnut (corylus cornuta) (ackerman 1987). sites were separated by >200 m and at each we collected 5 green leaves of each species (including the petiole and excluding twigs) at browse height (0.5–3.0 m) from separate but neighboring stems; samples were pooled for analysis. moose commonly forage in aquatic habitats during summer and appear to consume aquatic species in proportion to abundance at isle royale (qvarnemark and sheldon 2004). consequently, we opportunistically collected dominant (i.e., most abundant) aquatic macrophyte species (identified to the genus) at 3 e and 3 w sites (lakes) used by moose. at each site, 5 subsamples of each species along the shoreline were collected (where available) and pooled for analysis. since moose are not known to discriminate among aquatic plant parts (i.e., rhizome, stem, flower), we attempted to collect the entire plant, excluding only large and wellrooted rhizomes. to minimize the collection of benthic sediment, we rinsed samples in lake water to remove loose debris prior to placing in plastic sample bags. due to the paucity of inland lakes in the w, all 3 w sites were in bays of lake superior, versus 1 e site. to reduce the potential effect of sampling in lake superior, in 2003 we expanded aquatic plant sampling (6–18 july) to include 5 e and 5 w inland aquatic sites. the w sites included lakes, small ponds, or wetland habitats containing open water where moose were observed feeding or evidence of use was identified (e.g., tracks, fecal pellets). to minimize degradation prior to analyses, we cooled samples until freezing them at −20°c  within 12 h of collection. aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 78 we collected winter forage samples at 7 e and 7 w sites spaced ≥ 200 m apart between  12 january and 10 february 2003. at each site, we collected 5 twigs (current annual growth) from adjacent stems of individual plants from balsam fir, white cedar (thuja occidentalis), mountain-ash, red-osier dogwood (cornus stolonifera), paper birch, and quaking aspen (populus tremuloides). we clipped twigs at the average diameter for each species eaten by moose in winter (risenhoover 1987). at each site, arboreal lichens of the genera usnea and parmelias were collected from the branches/bark of standing or newly fallen white spruce (picea glauca) and paper birch. samples were handled and frozen as with the aquatic macrophytes. metrics of quality and analysis indices of forage quality are based upon either the presence of essential plant nutrients (e.g., water, carbohydrate, fat, protein, vitamins, and minerals) or the absence of indigestible structural carbon (c) compounds and toxins (crawley 1983). nitrogen (n) availability is considered the most limiting aspect of herbivore nutrition (crawley 1983). since rumen microbes can incorporate both organic and inorganic sources of n into the synthesis of amino acids, crude protein (n × 6.25) is a sufficient metric of digestible protein in ruminants (schwartz and renecker 2007). the elemental ratio of c:n is also a useful index of gross forage quality as it provides a measure of the relative investment in c structural compounds (associated with reduced quality) per atom of n (associated with enhanced quality) (crawley 1983, sterner and elser 2002). we report crude protein content (%) and c:n mass ratios of forage types as metrics of overall forage quality. we oven-dried plant tissue at 60 °c for 48 h to constant mass (or longer as needed for aquatic macrophytes) and ground it to fine powder in a ball mill (spex certiprep inc., metuchen, new jersey, usa). all samples were re-dried overnight and stored in a dessicator until subsamples (c: 1.5 ± 0.1 mg, n: 3.0 ± 0.1 mg) were weighed into tin cups. subsamples were combusted in a costech elemental combustion system 4010 elemental analyzer (costech analytical technologies, valencia, california, usa) to measure c and n content. the instrument was calibrated with acetanilide and internal organic check standards were analyzed every 10 samples; analytical precisions were %c ± 0.20 and %n ± 0.05. duplicate samples were analyzed every 5 samples and results were accepted only if the variance between duplicates was less than that of the standards. we calculated standard errors of c:n ratios using error propagation, which derives a composite error from that of its component parts (sterner and elser 2002). differences in crude protein and c:n ratios among pooled forage types (terrestrial, aquatic), terrestrial species, sampling location (e and w), and sampling year (2002, 2003) of aquatic macrophytes were tested separately using univariate analysis of variance (anova) (sas institute inc., cary, north carolina, usa) due to the unbalanced nature of the data with respect to aquatic sampling location. we examined crude protein content and c:n ratios among all known moose forage types (i.e., summer and winter) for correlation without a priori predictions as to the nature of the relationship. we included data on winter forage types (terrestrial plant twigs and lichens) in this analysis to increase the range of values used to model this correlation. we used the best-fit model describing the correlation for only summer terrestrial leaves as a baseline for comparing the observed and predicted relationship for submergent and emergent aquatic macrophytes using univariate anova. we determined the best fit model by visual assessment and alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 79 improvements in r2. where anova results were significant (p < 0.05), we used tukey’s honestly significant difference (hsd) to determine which samples differed. all tests were considered significant at the α = 0.05  level and assumptions of normal distribution and homogeneous variance were tested. where the assumption of homogeneous variance was violated, individual comparisons were made with two-sample t-tests assuming unequal variance. we report means ± standard errors, unless otherwise noted. results the aquatic macrophyte samples (n = 26; 17 lake superior, 9 inland) were from 7 genera in 7 families in 2002, increasing to 88 samples from 27 genera in 15 families in 2003 (table 1). samples collected from lake superior in 2002 had lower c:n ratios (−x = 13.7 ± 0.7) than those collected from inland lakes (= 18.0 ± 1.2; f1,24 = 11.07, p = 0.0028); crude protein content did not differ between lake types (t22 = 0.10; p = 0.92). we detected no annual difference in crude protein content of aquatic plants by sample year (f1,86 = 1.01, p = 0.3172); however, the c:n ratio of aquatic samples was higher (t76 = −2.95, p = 0.0043) in 2003 (c:n = 18.3 ± 0.8) than in 2002 (c:n = 15.2 ± 0.7). the c:n ratio distribution was non-normal in the 2003 aquatic samples, in large part due to two outliers that were emergent taxa that contain more structural compounds and expected to have a higher c:n ratio than submerged plants. removal of the emergent species (n = 15, all collected in 2003) from the analysis resulted in a normally distributed dataset with no difference in crude protein and c:n ratios between years. therefore, the aquatic macrophyte data we present in comparisons among forage types represent only submergent aquatic data pooled across years, and includes both inland and lake superior samples. alternative analyses using unpooled aquatic data including all taxa (submergent and emergent) did not alter the statistical significance of the comparison (supplemental table). submergent aquatic macrophytes had higher crude protein (t110 = 2.9, p < 0.0001; fig. 1) and lower c:n ratios (t104 = −10.6, p = 0.004; fig. 1) than terrestrial plant leaves. since aquatic macrophyte table 1. mean crude protein content (%) and carbon:nitrogen (c:n) ratios (±sd) of aquatic macrophyte taxa sampled at isle royale national park during 2002–2003. taxon n crude protein c:n emergent juncus 1 17.7 15.4 sagittaria 1 14.8 14.6 asteraceae: unk. 1 14.0 15.1 lysimachia 2 13.7 ± 2.7 18.5 ± 2.1 eupatorium 1 13.1 17.4 eliocharus 2 13.0 ± 3.1 14.9 ± 3.4 menyanthes 1 12.0 24.3 poaceae: unk. 3 10.4 ± 1.9 25.3 ± 7.9 equisetum 1 9.6 25.3 dulichium 1 9.5 28.1 carex 1 5.8 49 submerged nuphar 5 22.9 ± 2.7 12.0 ± 1.0 potamogeton 28 15.3 ± 3.0 16.9 ± 3.4 brasenia 2 14.9 ± 7.3 17.8 ± 5.1 elodea 1 14.2 17.0 myriophyllum 5 14.0 ± 1.7 14.6 ± 2.0 lemna 1 13.8 18.4 najas 3 13.7 ± 4.3 13.7 ± 1.6 sparganium 10 13.4 ± 0.6 19.4 ± 0.9 utricularia 7 12.1 ± 1.1 15.0 ± 1.8 megalodonta 1 11.6 17.0 scirpus 2 11.4 ± 1.2 20.3 ± 1.7 isoetes 1 11.3 16.5 vallisneria 1 8.9 21.5 sclerolepis 1 7.0 11.7 chara 3 6.8 ± 0.7 17.5 ± 1.8 ranunculus 1 4.5 15.0 aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 80 taxa were collected opportunistically, sample sizes were too small to compare forage quality among taxa. nevertheless, yellow pond lily (nuphar) had the highest mean protein content and a low c:n ratio (table 1). among terrestrial forage species, leaves from yellow birch, beaked hazelnut, and paper birch had higher crude protein (f5,54 = 6.86, p < 0.0001; fig. 2a) and lower c:n ratio (f5,54 = 8.53, p < 0.0001; fig. 2b) than leaves from sugar maple. the trend in c:n ratios across these terrestrial species mirrored that of crude protein content with no exceptions (figs. 2a, b). crude protein content of aquatic macrophytes was higher (f1,86 = 6.56, p = 0.012) at w sites (−x = 15.1% ± 0.5) than e sites (−x = 12.6% ± 0.9), whereas the c:n ratio was lower at w sites (t2,24  =  −3.3,  p = 0.003). among summer terrestrial species, only sugar maple differed by location, with crude protein higher (f1,10 = 5.35, p = 0.046) at w (−x = 10.3% ± 0.5) than e sites (−x = 8.5% ± 0.5). the c:n ratio of sugar maple at w sites (−x = 29.4 ± 1.7) was correspondingly lower than at e sites (−x = 35.0 ± 2.2), but not different (f1,10 = 3.67, p = 0.088). the crude protein content and c:n ratios of terrestrial forage types (i.e., summer leaves, winter twigs, and winter lichens) were strongly correlated (r2 = 0.97; fig. 3a). this relationship was best explained by a negative exponential model: y = β0 × e − β1x, where x = percent crude protein and y = c:n ratio. aquatic macrophytes did not fit the exponential model describing summer terrestrial leaves (y = 69.006 × e −0.0843x, seb1 = 0.228, f2,145 = 23.7, p < 0.0001), showing instead a high degree of variability in c:n ratios, particularly at low levels of protein (fig. 3a). among aquatic macrophytes, the relationship between crude protein content and c:n ratio of emergent taxa followed the terrestrial curve more closely than submergent taxa (fig. 3b), but was not different (t-test, p = 0.065). sample variances among forage types were non-homogenous, although the anova and t-test results agreed. discussion because crude protein content and c:n ratios are indicative of forage quality, our results support the hypothesis that aquatic macrophytes provide high quality summer forage to moose, complementing their consumption of terrestrial plants. on isle royale, aquatic macrophytes have ~20% higher crude protein and 40% lower c:n fig. 1. boxplot of crude protein content and carbon:nitrogen (c:n) ratios of aquatic and terrestrial moose forage types from isle royale national park, michigan, usa, 2002. boxes depict interquartile range, dark lines are median values, circles are outliers, and whiskers are 1.5× interquartile range. aquatic macrophytes are submergent species pooled across sampling years. alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 81 ratios than terrestrial plant leaves collected in mid-summer. additionally, the quality of aquatic forage found in bays of lake superior appears to be higher than in inland lakes – similar crude protein content and lower c:n ratios. these data support studies on the copper river delta in alaska suggesting that submergent aquatic plants represent an important protein source for moose during summer (maccracken et al. 1993). in ontario, moose in a “cafeteria” food trial preferred aquatic species with higher sodium, phosphorus, and crude protein (fraser et al. 1984). however, crude protein content of terrestrial and aquatic plants did not differ in the alaskan study, perhaps due to highly variable data, low sample size, and the relatively high protein content (16%) of one terrestrial species, pin cherry (prunus pennsylvanica) (maccracken et al. 1993). the crude protein content of terrestrial plants in our study was generally similar to levels reported in other studies; however, the aquatic macrophytes had lower crude fig. 2. boxplot of crude protein content and carbon:nitrogen (c:n) ratios among leaves of summer terrestrial forage species from isle royale national park, michigan, usa, 2002. boxes depict interquartile range, dark lines are median values, circles are outliers, and whiskers are 1.5× interquartile range. letters indicate species that are significantly different. aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 82 protein than measured in alaska and ontario (table 2; fraser and hristienko 1983, maccracken et al. 1993). indeed, crude protein content was highly variable within and between lakes on isle royale with the minima and maxima ranging 16 and 18%, respectively. the lower crude protein content may reflect differences in species composition, local nutrient inputs at sampling sites, and the sampling period. it is possible that our small sample sizes were not entirely representative of the average crude protein or c:n ratios for certain species (table 1). summer diets of moose in the region of isle royale and northeastern minnesota include measurable amounts of aquatic plants (13–40%) as estimated via stable isotope analysis (berini 2019, tischler et al. 2019). further, moose inhabiting relatively warmer areas of northeastern minnesota consumed higher amounts of aquatic vegetation (berini 2019). consumption of aquatic plants certainly provides highly nutritional forage based on the crude protein and c:n ratios we measured, and moose simultaneously address other nutritional requirements including sodium balance. however, presumably moose diets are necessarily balanced with aquatic and terrestrial vegetation, in part, because aquatic foraging is believed limited by gut fill due to the high water content of aquatic macrophytes and incidental water consumption (belovsky 1978). fig. 3. correlation between crude protein content (nitrogen [n] × 6.25) and carbon:nitrogen (c:n) ratio among moose forage types (a), and among types of aquatic macrophytes (b) from isle royale national park, michigan, usa, 2002. alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 83 the efficiency with which ingested c is converted into heterotrophic biomass is negatively correlated with forage c:n ratios (elser et al. 2000). in general, the c:n stoichiometry of freshwater aquatic autotrophs is lower and less variable than that of terrestrial autotrophs up to a magnitude of three between freshwater seston and terrestrial plants (elser et al. 2000). at isle royale, moose feeding on aquatic plants during summer would acquire 1.5 times more n per c atom consumed than acquired through terrestrial foraging during summer, suggesting that the assimilation efficiency of aquatic macrophytes is greater than that of terrestrial plants. this pattern may largely be attributable to physiological constraints that obligate terrestrial plants to a large structural c investment rather than differences in n content per se (sterner and elser 2002). indeed, we found that terrestrial plants had a high and relatively fixed (46–49%) c content while the c content of aquatic macrophytes was highly variable (11–48%), perhaps reflecting that submerged and emergent aquatic macrophytes were pooled for analysis. furthermore, we found that the c:n stoichiometry of all terrestrial forage types (summer leaves, winter twigs, lichens) followed a tight pattern of exponential decay with increasing crude protein content, whereas aquatic macrophyte c:n ratios were comparatively low and much less predictable across a wide range of crude protein contents. interestingly, the crude protein-c:n relationship of emergent aquatic macrophytes, which require more structural support than submerged plants, was intermediate that of terrestrial and submergent aquatic plants. our results suggest that higher c:n ratios of terrestrial plants are due to greater structural c allocation and lower n content. thus, even if the difference in crude protein content is not biologically significant, aquatic macrophytes are a higher quality forage than terrestrial plants due to lower concentration of structural c which hinders digestibility. in support, belovsky and jordan (1978) reported higher digestibility for aquatic plants (94%) than deciduous leaves (72%) on isle royale; albeit, digestibility of both is considered high and deciduous leaves are the principle component of the summer diet of moose. furthermore, n content was negatively correlated with the table 2. comparison of crude protein content (sd) of summer terrestrial and aquatic forage among studied moose populations. source % crude protein location terrestrial† aquatic this study michigan 12 (0.3) 14 (0.6)* maccracken et al. (1993) alaska 13 (1)# 17 (1) fraser et al. (1984) ontario 13 (1)# 16 (1) crete and jordan (1982) quebec 14 (0.3)‡ na renecker and hudson (1985) alberta 13 (0.4)§ na oldemeyer et al. (1977) alaska 13 na note: samples collected between 30 june and 2 august unless otherwise noted. †samples include deciduous leaves and exclude twigs unless otherwise noted. *submergent species only. #samples include both leaves and twigs. ‡only beaked hazelnut and mountain maple were sampled; does not represent principal summer diet of moose. §a composite sampled to reflect diet. aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 84 content of phenolics (anti-herbivory compounds) (jones and hartley 1999), although the degree to which aquatic macrophytes have evolved chemical defense against herbivory has been little studied (but see parker et al. 2006). pond lilies have historically been identified as an important aquatic forage for moose in north america (peterson 1955, cobus 1972). murie (1934) provided anecdotal evidence of near extirpation of abundant pond lilies in the 1930s by an irrupting moose population on isle royale, which may be due to their preference by moose or sensitivity to disturbance (fraser and hristienko 1983). more recently, cover of watershield (brasenia schreberi), a previously abundant aquatic macrophyte, has declined in many of isle royale’s water bodies during periods of high beaver and moose density that coincided with low wolf abundance (hoy et al. 2019). it is not surprising that we found pond lilies to be a high-quality moose forage (mean crude protein content = 22.9%; table 1). among terrestrial species, our results suggest that sugar maple is a low-quality summer moose forage. sugar maple appears to be an important species in the spring diet of moose on isle royale (ackerman 1987), perhaps due to early leaf emergence and high calcium concentration relative to other terrestrial leaves, but its use declines as forage and diet diversity increase through spring and summer (belovsky et al. 1973, krefting 1974, miquelle and jordan 1979, belovsky 1981, ackerman 1987). as with northern ungulates, the protein content of the winter diet of moose is insufficient to meet maintenance protein requirements (5–7%, fig. 3) (schwartz et al. 1988). compensation of this “deficit” is achieved principally through catabolizing fat and lean body mass stored during late summer and autumn when forage is up to 3 x more nutritious than in winter (renecker and hudson 1986), recycling urea (van hoven and boomker 1985), and limiting fetal growth and gestational costs during early-mid winter (schwartz 2007). thus, the abundance of high quality forage consumed during summer influences pre-winter body condition and survival (parker 2003). it follows that consumption of high quality aquatic forage used throughout summer aids post-winter recovery, pre-winter nutritional condition, and winter survival of isle royale moose. surprisingly, in nearby northeastern minnesota moose in relatively warmer areas consumed poorer diets characterized as high in aquatic forage and low in high-preference terrestrial forage. further, moose dying overwinter consumed diets higher in aquatic forage than surviving moose (berini 2019). the spatial patterns we identified in forage quality among aquatic and terrestrial plant species is consistent with the largescale spatial (e-w) differences in soil richness and plant species composition on isle royale (see study area). spatial heterogeneity in resource quality is widely known to influence browsing behavior and the distribution of herbivores across landscapes (crawley 1983, renecker and hudson 1985, 1986, mcnaughton 1988, fryxell 1991, wallis devries 1996, parker 2003). many ungulate populations “track the pulse of production” and “green waves” (mattson 1980) via seasonal migration, thereby exploiting nutritious forage and maximizing the time period to access such forage when available (festa-bianchet 1988, merkle et al. 2016). assuming a similar distribution of macrophytes among aquatic habitats on isle royale, ~ 75% of aquatic biomass occurs on the east half of the island based on the length of shoreline available for aquatic foraging. unfortunately, little is known about or whether moose migration is common on isle royale; however, only 2 of 22 radio-collared moose migrated between the east and alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 85 west ends of the island in the late 1980s (unpublished data of author, r. o. peterson). perhaps this lack of migration can be explained by our observation that while overall forage quality appears to be better on the west end of the island, the biomass of aquatic macrophytes is higher at the east end. because no large patches of landscape are noticeably devoid of vegetation, except where fire removed it, we recognize that individuals meet their summer-autumn nutritional requirements through a varied diet. forage quality (protein content), although variable among species in summer, is on a continuum where the majority of terrestrial leaves are nutritious (protein content), highly digestible, and a mainstay of the spring-summer diet across moose range. the evolution of muzzle anatomy in moose is believed a morphological adaptation for efficient underwater feeding, a behavior unique to moose among cervids (hofmann 1989, geist 1998, clifford and witmer 2004). aquatic feeding is undoubtedly an important source of sodium for many moose populations (botkin et al. 1973, jordan et al. 1973, fraser et al. 1984); however, moose on isle royale (and elsewhere) can meet sodium requirements at mineral licks which contain much higher sodium concentrations (on a wet-weight basis) than aquatic plants (risenhoover and peterson 1986). we propose that aquatic foraging by moose at isle royale is also a mechanism to exploit relatively n-rich microsites (aquatic habitats) in an otherwise n-limited landscape (white 2012). regardless, aquatic habitats provide moose summer forage high in digestible protein critical to physical recovery and growth, while reducing their post-winter sodium deficit, insect harassment, risk of predation, and thermal stress (morris 2014). given the patchy spatial distribution of aquatic habitats on isle royale, the positive influence of aquatic feeding on the pre-winter nutritional condition of moose could affect spatial dynamics of winter population density, mortality, and predation rate. we suggest that increased consumption of high quality aquatic macrophytes on the east end of isle royale might supplement the lower quality winter forage, thereby foregoing the need to migrate. from this perspective, it is understandable that aquatic foraging by moose is prevalent (tischler et al. 2019); however, it is unknown if the pre-winter condition of moose differs at the island ends or the time associated with developing a migratory strategy. we encourage further research to test such assumptions and to better understand the relative use and role of aquatic plants on moose population dynamics. acknowledgements we thank k. pregitzer, l. vucetich, j. kaplan, d. mccormick, m. romanski, c. lawler, j. deutsch, and n. hambel for field assistance, and k. raisanen-schourek, b. allshouse, b. baibak, and d. donaldson for assistance with sample preparation. j. marr provided expertise in aquatic macrophyte identification. forage plant carbon and nitrogen content was determined by jennifer eikenberry at the school of forest resources and environmental science, michigan technological university. j. oelfke and m. romanski of the national park service provided logistical support in the field. c. giardina, p. hurley, and l. kruger, provided valuable comments on earlier versions of this manuscript. t. drummer and j. pickens provided useful advice on data analysis. this research was funded by the ecosystem science center at michigan tech and the national science foundation, and support to r. p. from the robbins chair in sustainable management of the environment at michigan technological university. j.k.b. was supported by grants nsf id#1545611 and nsf id#1556676. all aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 86 necessary permits from the national park service were obtained for the described field studies. we additionally thank 2 anonymous reviewers for comments that improved the clarity of the manuscript. references ackerman, t. n. 1987. moose response to summer heat on isle royale. m. s. thesis, michigan technological university, houghton, michigan, usa. bartels, p., j. cucherousset, k. steger, p. eklöv, l. j. tranvik, and h. hillebrand. 2012. reciprocal subsidies between freshwater and terrestrial ecosystems structure consumer resource dynamics. ecology 93: 1173–1182. doi: 10.1890/11-1210.1 belovsky, g. e. 1978. diet optimization in a generalist herbivore: the moose. theoretical population biology 14: 105–134. doi: 10.1016/ 0040-5809 (78)90007-2 _____. 1981. a possible population response of moose to sodium availability. journal of mammalogy 62: 631–633. doi: 10.2307/1380412 _____, and p. a. jordan. 1978. the time-energy budget of a moose. theoretical population biology 14: 76–104. doi: 10.1016/0040-5809(78)90006-0 _____, _____, and d. b. botkin. 1973. summer browsing by moose in relation to preference and animal density: a new quantitative approach. alces 9: 101–122. berini, j. l. 2019. evaluating how spatial heterogeneity in forage chemistry and abundance influences diet and demographics in a declining moose (alces alces) population in northeast minnesota. ph. d. dissertation. university of minnesota, saint paul, minnesota, usa. botkin, d. b., p. a. jordan, a. s. dominski, h. s. lowendorf, and g. e. hutchinson. 1973. sodium dynamics in a northern ecosystem. proceedings of the national academy of sciences 70: 2745–2748. doi: 10.1073/pnas.70.10.2745 bump, j. k. 2018. fertilizing riparian forests: nutrient repletion across ecotones with trophic rewilding. philosophical transactions of the royal society b: biological sciences 373: 20170439. doi: 10.1098/rstb.2017.0439 clifford, a. b., and l. m. witmer. 2004. case studies in novel narial anatomy: 2. the enigmatic nose of moose (artiodactyla: cervidae: alces alces). journal of zoology 262: 339–360. doi: 10.1017/s0952836903004692 cobus, m. 1972. moose as an aesthetic resource and their summer feeding behaviour. alces 8: 244–275. crete, m., and p. a. jordan. 1982. production and quality of forage available to moose in southwestern quebec. canadian journal of forestry research. 12: 151–159. de jager, n. r., j. j. rohweder, and m. j. duveneck. 2020. climate change is likely to alter future wolf – moose – forest interactions at isle royale national park, united states. frontiers in ecology and evolution 8: 543915. doi: 10.3389/ fevo.2020.543915 delgiudice, g. d., r. o. peterson, and w. m. samuel. 1997. trends of winter nutritional restriction, ticks, and numbers of moose on isle royale. journal of wildlife management 61: 895–903. doi: 10.2307/3802198 _____, b. a. sampson, m. s. lenarz, m. w. schrage, and a. j. edwards. 2011. winter body condition of moose (alces alces) in a declining population in northeastern minnesota. journal of wildlife diseases 47: 30–40. doi: 10.7589/0090-3558-47.1.30 edwards, j. 1983. diet shifts in moose due to predator avoidance. oecologia 60: 185–189. doi: 10.1007/bf00379520 elser, j. j., w. f. fagan, r. f. denno, d. r. dobberfuhl, a. folarin, a. huberty, s. alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 87 interlandi, s. s. kilham, e. mccauley, k. l. schulz, e. h. siemann, and r. w. sterner. 2000. nutritional constraints in terrestrial and freshwater food webs. nature 408: 578–580. doi: 10.1038/35046058 festa-bianchet, m. 1988. seasonal range selection in bighorn sheep: conflicts between forage quality, forage quantity, and predator avoidance. oecologia 75: 580–586. doi: 10.1007/bf00776423 fraser, d., and h. hristienko. 1983. effects of moose, alces alces, on aquatic vegetation in sibley provincial park, ontario. canadian field-naturalist 97: 57–61. _____, d. arthur, j. k. morton, and b. k. thompson. 1980. aquatic feeding by moose alces alces in a canadian lake. ecography 3: 218–223. doi: 10.1111/ j.1600-0587.1980.tb00728.x _____, e. r. chavez, and j. e. palohelmo. 1984. aquatic feeding by moose: selection of plant species and feeding areas in relation to plant chemical composition and characteristics of lakes. canadian journal of zoology 62: 80–87. doi: 10.1139/z84-014 fryxell, j. m. 1991. forage quality and aggregation by large herbivores. the american naturalist 138: 478–498. doi: 10.1086/285227 gasaway, w. c., r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs: 1–50. geist, v. 1998. deer of the world: their evolution, behaviour, and ecology. stackpole books, mechanicsburg, pennsylvania, usa. hofmann, r. r. 1989. evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. oecologia 78: 443–457. doi: 10.1007/bf00378733 hoy, s. r., r. o. peterson, and j. a. vucetich. 2019. ecological studies of wolves on isle royale. michigan technological university, houghton, michigan, usa. https://isleroyalewolf. o r g / s i t e s / d e f a u l t / f i l e s / a n n u a l r e port-pdf/wolfreport_pages_2019_ final_apr29.pdf huber, n. k. 1973. glacial and postglacial geologic history of isle royale national park, michigan. usgs numbered series 754-a (professional paper). u.s. government printing office, washington d.c., usa. (accessed january 2022). hutchinson, g. e., and e. s. deevey. 1949. ecological studies on populations. pages 325–358 in g. s. avery jr., e. c. auchter, g. w. beadle, h. b. creighton, w. u. gardner, g. e. hutchinson, l. pauling, f. o. schmitt, w. m. stanley, c. b. van niel, and d. whitaker, editors. survey of biological progress 1: 325–59. doi: 10.1016/ b978-1-4832-0000-2.50014-7 janke, r. a., d. mckaig, and r. raymond. 1978. comparison of presettlement and modern upland boreal forests on isle royale national park. forest science 24: 115–121. doi: 10.1093/ forestscience/24.1.115 johnston, c. a. 2017. beavers: boreal ecosystem engineers. springer international publishing. (accessed january 2022). _____, and s. k. windels. 2015. using beaver works to estimate colony activity in boreal landscapes. journal of wildlife management 79: 1072–1080. doi: 10.1002/jwmg.927 jones, c. g., and s. e. hartley. 1999. a protein competition model of phenolic allocation. oikos 86: 27–44. doi: 10.2307/3546567 jordan, p. a. 1987. aquatic foraging and the sodium ecology of moose: a review. swedish wildlife research supplement 1: 119–137. _____, d. b. botkin, a. s. dominski, h. s. lowendorf, and g. e. belovsky. 1973. https://isleroyalewolf.org/sites/default/files/annual-report-pdf/wolfreport_pages_2019_final_apr29.pdf https://isleroyalewolf.org/sites/default/files/annual-report-pdf/wolfreport_pages_2019_final_apr29.pdf https://isleroyalewolf.org/sites/default/files/annual-report-pdf/wolfreport_pages_2019_final_apr29.pdf https://isleroyalewolf.org/sites/default/files/annual-report-pdf/wolfreport_pages_2019_final_apr29.pdf http://pubs.er.usgs.gov/publication/pp754a http://pubs.er.usgs.gov/publication/pp754a http://link.springer.com/10.1007/978-3-319-61533-2 http://link.springer.com/10.1007/978-3-319-61533-2 http://link.springer.com/10.1007/978-3-319-61533-2 aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 88 sodium as a critical nutrient for the moose of isle royale. proceedings of the 9th north american moose conference and workshop, quebec city, quebec, canada. _____, r. o. peterson, and k. a. ledoux. 2010. swimming wolves, canis lupus, attack a swimming moose, alces alces. canadian field-naturalist 124: 54–56. doi: 10.22621/cfn.v124i1.1030 krefting, l. w. 1974. the ecology of the isle royale moose with special reference to the habitat. agricultural experiment station technical bulletin no. 297. university of minnesota, minneapolis, minnesota, usa. maccracken, j. g., v. van ballenberghe, and j. m. peek. 1993. use of aquatic plants by moose: sodium hunger or foraging efficiency? canadian journal of zoology 71: 2345–2351. doi: 10.1139/ z93-329 mattson jr, w. j. 1980. herbivory in relation to plant nitrogen content. annual review of ecology and systematics 11: 119–161. doi: 10.1146/annurev. es.11.110180.001003 mcnaughton, s. j. 1988. mineral nutrition and spatial concentrations of african ungulates. nature 334: 343–345. doi: 10.1038/334343a0 mech, l. d. 1966. the wolves of isle royale. fauna of the national parks of the united states. fauna series 7. u. s. government printing office, washington, d.c., usa. http://npshistory.com/series/fauna/7.pdf merkle j. a., k. l. monteith, e. o. aikens, m. m. hayes, k. r. hersey, a. d. middleton, b. a. oates, h. sawyer, b. m. scurlock, and m. j. kauffman. 2016. large herbivores surf waves of green-up during spring. proceedings of the royal society b: biological sciences 283. doi: 10.1098/rspb.2016.0456 miquelle, d. g., and p. a. jordan. 1979. the importance of diversity in the diet of moose. alces 15: 54–79. moen, r., j. pastor, and y. cohen. 1997. a spatially explicit model of moose foraging and energetics. ecology 78: 505– 521. doi: 10.1890/0012-9658(1997)078 [0505:asemom]2.0.co;2 morris, d. m. 2014. aquatic habitat use by north american moose (alces alces) and associated richness and biomass of submersed and floating-leaved aquatic vegetation in north-central minnesota. m. s. thesis. lakehead university, thunder bay, ontario, canada. murie, a. 1934. the moose of isle royale. miscellaneous publication 25. museum of zoology, university of michigan, ann arbor, michigan, usa. oldemeyer, j. l., a. w. franzmann, a. l. brundage, p. d. arneson, and a. flynn. 1977. browse quality and the kenai moose population. journal of wildlife management 41: 533–542. doi: 10.2307/3800528 parikh, g. l., j. s. forbey, b. robb, r. o. peterson, l. m. vucetich, and j. a. vucetich. 2017. the influence of plant defensive chemicals, diet composition, and winter severity on the nutritional condition of a free-ranging, generalist herbivore. oikos 126: 196–203. doi: 10.1111/oik.03359 parker, j. d., d. o. collins, j. kubanek, m. c. sullards, d. bostwick, and m. e. hay. 2006. chemical defenses promote persistence of the aquatic plant micranthemum umbrosum. journal of chemical ecology 32: 815–833. doi: 10.1007/s10886-006-9038-7 parker, k. l. 2003. advances in the nutritional ecology of cervids at different scales. ecoscience 10: 395–411. doi: 10.1080/11956860.2003.11682788 peek, j. m. 2007. habitat relationships. pages 351–375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, second edition. university press of colorado, boulder, colorado, usa. http://annurev.es http://annurev.es http://npshistory.com/series/fauna/7.pdf http://npshistory.com/series/fauna/7.pdf alces vol. 58, 2022 aquatic areas and high nitrogen forage – tischler et al. 89 peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. qvarnemark, l. m., and s. p. sheldon. 2004. moose grazing decreases aquatic plant diversity. journal of freshwater ecology 19: 407–410. doi: 10.1080/02705060.2004.9664913 renecker, l. a., and r. j. hudson. 1985. estimation of dry matter intake of free-ranging moose. journal of wildlife management 49: 785–792. doi: 10.2307/3801712 _____, and ____. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322–327. doi: 10.1139/ z86-052 _____, and _____. 1990. behavioral and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspen-dominated forests. alces 26: 66–72. risenhoover, k. a. 1987. winter foraging strategies of moose in subarctic and boreal forest habitats. ph. d. dissertation. michigan technological university, houghton, michigan, usa. risenhoover, and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17: 357–364. doi: 10.1139/x87-062 risenhoover, and r. o. peterson. 1986. mineral licks as a sodium source for isle royale moose. oecologia 71: 121–26. doi: 10.1007/bf00377330 schwartz, c. c. 2007. reproduction, natality and growth. pages 141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. ________, m. e. hubbert, and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. journal of wildlife management 52: 26. doi: 10.2307/3801052 schwartz, and l. a. renecker. 2007. nutrition and energetics. pages 141–171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. university press of colorado, boulder, colorado, usa. severud, w. j., j. l. belant, s. k. windels, and j. g. bruggink. 2013. seasonal variation in assimilated diets of american beavers. american midland naturalist 169: 30–42. doi: 10.1674/0003-0031-169.1.30 sterner, r. w., and j. j. elser. 2002. ecological stoichiometry: the biology of elements from molecules to the biosphere. princeton university press, princeton, new jersey, usa. timmermann, h. r., and a. r. rodgers. 2017. the status and management of moose in north america-circa 2015. alces 53: 1–22. tischler, k. b., w. j. severud, r. o. peterson, and j. k. bump. 2019. aquatic macrophytes are seasonally important dietary resources for moose. diversity 11: 209. doi: 10.3390/d11110209 van hoven, w. and e. a. boomker. 1985. pages 103–120 in r. j. hudson and r. g. white, editors. bioenergetics of wild herbivores. crc press, boca raton, florida, usa. vucetich, j. a., and r. o. peterson. 2004. grey wolves – isle royale. pages 285–296 in d. w. macdonald and c. sillero-zubiri, editors. biology and conservation of wild canids. oxford university press, new york, new york, usa. wallis devries, m. f. 1996. effects of resource distribution patterns on ungulate foraging behaviour: a modelling approach. forest ecology and management 88 (ungulates in temperate forest ecosystems): 167–177. doi: 10.1016/s0378-1127(96)03822-4 white, t. c. r. 2012. the inadequate environment: nitrogen and the abundance of animals. springerverlag, berlin, germany. aquatic areas and high nitrogen forage – tischler et al. alces vol. 58, 2022 90 supplemental table comparison of anova test results for the effect of forage type on plant crude protein content or c:n ratio when aquatic macrophytes are either pooled or separated by sample year (2002, 2003) and type (submergent and emergent, submergent only). aquatic dataset n† crude protein c:n ratio f p f p 2002 26 149.2 <0.0001 229.0 <0.0001 2003 62 118.1 <0.0001 289.8 <0.0001 2002, 2003 combined 88 122.4 <0.0001 367.9 <0.0001 2002, 2003 combined; submergent only 73 124.6 <0.0001 373.7 <0.0001 †sample size of aquatic macrophyte dataset used for analysis. f:\alces\vol_38\pagemaker\3814. alces vol. 38, 2002 spaeth et al. nutritional quality of willows 143 nutritional quality of willows for moose: effects of twig age and diameter douglas f. spaeth1,4, r. terry bowyer1, thomas r. stephenson2,5 , perry s. barboza1, and victor van ballenberghe3 1institute of arctic biology, and department of biology and wildlife, university of alaska fairbanks, fairbanks, ak 99775-7000, usa; 2kenai moose research center, alaska department of fish and game, 43961 kalifornsky beach road, soldotna, ak 99669, usa; 38941 winchester street, anchorage, ak 99507, usa abstract: alaskan moose (alces alces gigas) consume willow (salix spp.) as a fundamental component of their winter diet. we collected barclay willow (s. barclayi) from 5 nearby sites (1580 m apart) on the kenai peninsula, alaska, usa, during winter 1999-2000. we tested effects of diameter and age of twigs on nutritional quality of willows for moose. smaller-diameter twigs had higher in vitro dry matter digestibility (ivdmd), and protein content, but lower fiber content (p < 0.001) than larger twigs. an inverse relationship occurred between the age of twigs and protein content (p < 0.001), with older-aged twigs containing less protein. accordingly, age of twigs was negatively related to fiber content (p = 0.002). conversely, no relation existed between age of twigs and ivdmd (p = 0.34). tannin content (p < 0.001) and age of twigs (p = 0.04) varied among sites, with older twigs possessing more tannins than younger ones. no difference in tannins, however, occurred between diameter categories of twigs (p = 0.48). digestible energy differed between diameter categories (p = 0.02) and among ages of twigs (p = 0.02), as well as among collection sites (p < 0.001). thus, structural components of the twig to support growth were more important in affecting digestibility, whereas age of the twig was more influential in determining nitrogen and tannin content. the relation between twig age and tannin content, however, was the inverse of that expected. more research is needed to understand how quality of winter browse interacts with additional factors, such as predation risk, population density, and allometric differences between sexes, to affect diet selection and foraging behavior of moose and other large herbivores. alces vol. 38: 143-154 (2002) key words: alaskan moose, alces alces gigas, digestibility, digestible energy, kenai peninsula, nitrogen, nutrition, salix barclayi, structural carbohydrates, tannin, twig age, twig diameter, willows browse is an important element in the winter diet of moose (alces alces) inhabiting boreal forests (peek 1974, 1998; ludewig and bowyer 1985; renecker and schwartz 1998). indeed, the diet of alaskan moose (a. a. gigas) is composed principally of willows (salix spp.), which may be eaten throughout the year (van ballenberghe et al. 1989; miquelle et al. 1992; van ballenberghe 1992; bowyer et al. 1998, 1999a). further, diameter of twigs available to moose for consumption may be a crucial aspect of diet selection by this large browser (vivas et al. 1991, bowyer and bowyer 1997). most nutrients used by moose are contained in the surface of woody twigs, with hard-to-digest carbohydrates (cellulose and 4present address: u.s. forest service, coconino national forest, blue ridge ranger district, hc 31, box 300, happy jack, az, 86024, usa 5present address: california department of fish and game, 407 west line street, bishop, ca 93514, usa nutritional quality of willows spaeth et al. alces vol. 38, 2002 144 hemicellulose) composing the core (schwartz and renecker 1998). therefore, as twig diameter of browse increases (i.e., the diameter at the point of browsing), the ratio of surface nutrients to the core declines, as does the nutritional value of such forage for moose (hjeljord et al. 1982, schwartz and renecker 1998). in winter, adult moose eat forage that contains levels of crude protein below maintenance, and dry-matter intake necessary to meet nitrogen requirements is difficult to attain (schwartz and renecker 1998). further, the role that tannins play in forage selection is complex, and may affect foraging by herbivores (reid et al. 1974, bryant and kuropat 1980, leslie and starkey 1987, robbins et al. 1987, bryant et al.1991). for instance, leaders of new growth in birch (betula sp.) were heavily defended by secondary compounds, which altered foraging behavior by snowshoe hares ( l e p u s americanus; bryant et al. 1994). during winter, moose may be protein as well as energy limited; hence, forage selection should favor young twigs with smaller diameters. moose eat twigs older than first-year growth, but data on the nutritive value of those older twigs are sparse (cowan et al. 1950). indeed, diet quality for herbivores likely involves a preference for species of plants, as well as specific parts of plants (janzen 1979). there is increasing evidence that moose play a fundamental role in the structure and function of boreal ecosystems (pastor and naiman 1992, molvar et al. 1993, bowyer et al. 1997, berger et al. 2001, kie et al. 2003); however, much remains to be learned about their foraging ecology. gaining insights into why moose forage on a particular plant or select specific twigs, or diameters of twigs, from that plant is critical to understanding the mechanisms controlling foraging behavior. we tested for differences in forage quality as affected by diameter of twigs, age of twigs, tannin content, collection site, and their interactions. we also examined the digestible energy content (de) of willows, and tested for differences between age classes and diameter categories of twigs. we hypothesized that larger twigs would have a lower nitrogen content, be less digestible, have more fiber, and have a lower tannin content than smaller twigs. likewise, we also predicted that older twigs would have lower nitrogen content, be less digestible, have more fiber, and possess lower tannin content than younger twigs. further, we hypothesized that dietary energy and protein would change with size and age of browse, and that small changes in browse chemistry might alter availability of protein and energy for moose. study area we sampled twigs of willow (salix barclayi) at an elevation of 275 m along a roadside located on the kenai peninsula, near ninilchik, alaska, usa, (60° n, 149° w) during winter 1999-2000. the kenai peninsula is characterized by a maritime climate influenced by its proximity to the pacific ocean (weixelman et al. 1998). annual precipitation ranges from 40 to 50 cm with most falling as snow in winter and rain in spring or autumn (schwartz and franzmann 1991). annual snowfall ranges from 140 to 165 cm (oldemeyer and regelin 1987). mean annual temperature is 1º c, and mean monthly temperatures range from – 30 to 21º c (schwartz and franzmann 1991). we began sampling in early december after willows had become dormant and lost their leaves. moose migrated from higher e l e v a t i o n s a c r o s s o u r s t u d y s i t e t o lower-elevation valleys as winter snowfall accumulated. thus, moose use of the study area was limited, and much of the willow in this area was unbrowsed, or only lightly browsed. sampling was completed in late alces vol. 38, 2002 spaeth et al. nutritional quality of willows 145 february, and samples were stored between 0 and -25º c until analyzed. our study area was a plateau along the sides of an unpaved road that ran east from ninilchik for approximately 21 km. the roadside was surrounded by boreal forest dominated by white spruce (picea glauca). willows ranged in size from 1-3 m in height. our sampling site was located adjacent to the road (3-20 m from the snowburm) about 16 km from ninilchik. there was no overstory cover, and patches of dense growth of willows characterized the understory. shading affects nutritional quality of willows (hjeljord et al. 1990, bø and hjeljord 1991, molvar et al. 1993); however, willows we selected were shaded only slightly by a few trees, thereby minimizing that complication. likewise, this area exhibited little variation in slope, exposure, or drainage. finally, easy access allowed us to sample large quantities of willow in a relatively small area. methods we sampled an area along a roadside that encompassed 155 m, which included 5 distinct patches of willows located 15–80 m apart ( ± sd = 38.8 ± 28.69 m). all stems with abundant twigs (> 15 leaders) were cut from a plant at snow level, labeled, and transported to the laboratory for subsequent analyses. three stems (containing numerous leaders) from each of 5 sites were selected haphazardly for nutritional analyses; the remainder of branches was withheld for a related experiment on feeding behavior of moose. current annual growth (1-year-old), 2-year-old growth, and 3-year-old growth were measured with dial calipers to the nearest 0.1 mm at the bud scale scar, and pooled according to diameter and age classes. twigs were categorized according to diameter: small (0.8 2.9 mm) and large (3.0 4.9 mm). this classification was based on previous studies of twig selection by foraging moose (molvar and bowyer 1994, bowyer and bowyer 1997, stephenson et al. 1998, weixelman et al. 1998), diameter and age classes of twigs available to us for sampling, and the need to obtain sufficient material in a particular age and diameter category for nutritional analyses. samples of twigs from each site were pooled by age class and diameter category, oven dried to constant mass at 55º c, and then ground with a wiley mill (1-mm screen). all nutrients were assayed on the basis of dry mass (dm). in vitro dry matter digestibility (ivdmd; tilley and terry 1963) was determined for each sample. fresh rumen inoculum for the digestion trial was obtained f r o m 1 c a p t i v e r e i n d e e r ( r a n g i f e r tarandus) that was fistulated, and held at the robert g. white large animal research station of the university of alaska fairbanks (uaf). we conditioned the reindeer to a diet of willow by adding a mixture of approximately 12 g ground willow and 500 ml water directly into the rumen (via canula) every 2-3 days for 18 days. the forage quality analysis laboratory at uaf performed ivdmd, nutrient analyses, and tannin assays, with duplicates for selected samples. detergent analysis (van soest et al. 1991) was used to determine structural composition of plant cells (percentage dry weight of neutral-detergent fiber [ndf], acid detergent fiber [adf]), ash of acid extracted fiber, and lignin). fiber fractions were used to derive estimates of cell contents (dm ndf), hemicellulose (ndf adf), and cellulose (adf lignin). nitrogen was determined with an elemental analyzer (model # cns 2000, leco, st. joseph, mi, usa) and expressed as crude protein based on the assumption of 6.25 g protein per 1 g nitrogen (robbins 1993). soluble carbohydrates such as starch were estimated as the difference between cell contents and crude protein, with the asx alces vol. 38, 2002 spaeth et al. nutritional quality of willows 147 compared cellulose and ivdmd between large and small diameter twigs (fig. 2). conversely, considerable overlap occurred between ages of twigs when cellulose was examined in relation to ivdmd (fig. 2). these results confirm that cellulose in the core of stems strongly affected ivdmd. additional measures of forage quality followed a similar pattern with significant differences occurring among age classes and diameter categories of willow twigs, except for ash, which differed neither in twig age nor diameter, and lignin, which did not vary with age (table 1). variation in mean tannin concentration of willow twigs among sites ranged from 167.30 mg/g to 209.32 mg/g. similarly, tannin content ( ± sd) varied among ages of twigs (1-year-old = 185.4 ± 41.63 mg/g; 2-year-old = 206.4 ± 42.86 mg/g; 3-year-old table 1. forage quality (% dry mass) of 1-year-old, 2-year-old, and 3-year-old growth, and of small (0.08 – 2.9 mm) and large (3.0 – 4.9 mm) categories of twig diameter for barclay willow (salix barclayi), kenai peninsula, alaska, usa, winter 1999-2000. composites of 15-25 twigs were included in each sample. sample sizes for age and diameter categories were: 1-year-old, small (n = 27); 1-year-old, large (n = 9); 2-year-old, small (n = 13); 2-year-old, large (n = 10); 3-year-old, small (n = 5); and 3-year-old, large (n = 15). age 1-year-old 2-year-old 3-year-old variable1 x (sd) x (sd) x (sd) acid-detergent fiber small 39.68 (2.44) 40.60 (2.37) 42.33 (2.24) large 32.39 (2.78) 42.60 (3.37) 44.54 (3.00) neutral-detergent fiber small 47.98 (2.93) 49.97 (2.16) 52.97 (2.24) large 53.54 (3.39) 54.81 (3.98) 55.82 (3.33) ash of acid extracted fiber small 0.35 (0.12) 0.30 (0.08) 0.33 (0.05) large 0.31 (0.11) 0.30 (0.09) 0.31 (0.09) derived lignin small 21.64 (1.57) 21.47 (1.98) 20.98 (1.06) large 19.28 (2.01) 18.09 (1.03) 19.56 (1.73) derived hemicellulose small 8.31 (1.01) 9.37 (0.67) 10.64 (0.24) large 11.15 (1.16) 12.21 (0.60) 11.28 (0.73) derived cellulose small 17.68 (1.62) 18.83 (1.05) 21.02 (1.89) large 22.79 (2.69) 24.22 (2.79) 24.67 (2.40) 1manova indicated that significant differences in forage quality occurred among different age classes and between diameter categories (p < 0.01) for all variables, except for ash of acid extracted fiber (age: p = 0.56; diameter: p = 0.15) and derived lignin (age: p = 0.27). x alces vol. 38, 2002 spaeth et al. nutritional quality of willows 149 twigs relative to age and diameter were not large (fig. 1), such variation may be important to foraging herbivores as they accumulate nutrients over time (white 1983). importance of winter forage for moose should be viewed in a broad perspective (weixelman et al. 1998); several factors likely affect foraging behavior. browse consumed by moose during winter is composed largely of willow twigs that have a low content of crude protein (5-7 %), which will not meet maintenance requirements (schwartz 1992), or fully support reproduction (schwartz et al. 1988). northern ungulates are in a negative energy balance during winter, and foraging activities principally slow the rate of loss of body reserves (mautz 1978, barboza and bowyer 2001). some losses of body reserves, however, may be physiologically regulated, because moose voluntarily reduce their metabolic rate and food intake during winter to conserve energy (schwartz et al. 1988). if nitrogen levels are below maintenance requirements, then ivdmd may become increasingly important for survival of moose in winter. shorter retention times in the rumen are correlated with higher-quality diets and longer retention times with lower-quality forage (schwartz et al. 1988). rumen microbes ferment soluble sugars and cell solubles rapidly; however, cell walls require much longer to process (spalinger 2000, russell and rychlik 2001). lignin content also reduces digestibility of forages, as can tannins and other plant secondary compounds (bryant et al. 1991, 1994). secondary plant compounds (i.e., tannins) may play a role in food choice, because browsing vertebrates avoid consuming plant tissues that contain high concentrations of secondary metabolites (bryant and kuropat 1980, palo et al. 1985). further, tannins are thought to negatively affect digestibility of browse for moose during winter (bryant and kuropat 1980, palo et al. 1985). estimations of digestibility of woody forage, however, may not need to be adjusted for tannins, because there may be some benefits to ruminants from ingesting forages containing tannins (kumar and singh 1984, leslie and starkey 1987, hagerman and robbins 1993). reid et al. (1974) postulated that the presence of tannins provided partial protection of proteins from degradation in the rumen, thereby enhancing assimilation of nitrogen. robbins et al. (1987) suggested that reduction of protein digestion caused by tannins may not result from gastrointestinal adaptations, but may be because of the small amounts of tannins in winter browse. the saliva of moose contains large amounts of prolinerich proteins, which may bind tannins and thereby reduce their effects on moose (hagerman and robbins 1993, juntheikki 1996). further, many tannins in willow are linear-condensed tannins that moose bind well, in contrast to other tannins in lowerquality foods, which moose saliva does not bind (barry and mcnabb 1999). weixelman et al. (1998) suggested that reduced food availability, quality, and digestibility, combined with the increased energetic costs of foraging during severe weather, should force animals to maximize caloric return per unit energy expended. in addition, there may be twigs that are too small, or too widely dispersed to provide sufficient nutritional value for moose. relationships between forage digestibility, retention time in the rumen, and rate of intake (owen-smith 1982, van soest et al. 1991), indicate digestibility is likely an important factor in forage selection by ruminants. those relationships probably affect the size of a bite for moose foraging in winter, because larger bites have poorer nutritional quality (schwartz et al. 1988, molvar and bowyer 1994). decreases in digestible-energy content nutritional quality of willows spaeth et al. alces vol. 38, 2002 150 of willow twigs with age and diameter reflect declining proportions of crude protein and cell contents as the matrix of the plant cell wall increases in concentration. differences in digestible-energy content of twigs may be directly related to food intake required in winter. schwartz and renecker (1998) calculated a daily intake of digestible energy in moose during november as 975 kj/kg0.75. based on our calculations, consumption of 1-year-old twigs with small diameters would require a mean (± sd) daily intake of 124 ± 9 gdm/kg0.75 body mass, whereas intakes of 3-year-old twigs with large diameter subtend intakes that are 15% greater (141 ± 8 gdm/kg0.75). that increment in digestive load would increase gut fill and influence passage rate. changes in digestive function associated with energy demand may feedback on processes of forage selection at the level of plant and twig. the pattern of nutrients and secondary metabolites across ages and between diameter classes of willow twigs did not conform to some of our initial predictions, especially a lack or variation in ivdmd with increasing age. nonetheless, our results support the hypothesis that moose should alter their foraging behavior to respond to variation in plant nutrients (and perhaps secondary compounds), at fine scales that include nearby foraging sites and differences among twigs on the same plant. the forgoing arguments clearly indicate that quality of forage should be a critical component in diet selection by large herbivores, but such relationships have been notoriously difficult to demonstrate in free-ranging moose (weixelman et al. 1998). those difficulties likely relate to effects of predation risk, including influences of group size, distance from concealment cover, and differential vulnerability of sex and age classes to predators, on foraging behavior and diet selection by moose (edwards 1983, molvar and bowyer 1994, weixelman et al. 1998, white et al. 2001). in addition, variation in population density with respect to carrying capacity (k) of the environment (bowyer et al. 1999b, kie 1999, kie et al. 2003) undoubtly alters foraging behavior of large mammals. likewise, allometric differences between sexes of ruminants may also affect assimilation of nutrients and consequently foraging behavior (schwartz et al. 1987; barboza and bowyer 2000, 2001; spaeth et al. 2001). moreover, the propensity of sexes to partition space outside the mating season in heterogeneous habitats (miquelle et al. 1992, bowyer et al. 2001) has a strong influence on habitats selected and, in consequence, the manner in which moose forage. we believe our descriptions of nutrients in willows and how they varied with respect to site, as well as age and diameter of twigs, is an important first step in clarifying diet selection by moose. we contend, however, that a more complete understanding of foraging dynamics in this large herbivore ultimately will require a better integration of the life-history characteristics of moose with nutritional composition and abundance of their forage. acknowledgements we thank the u.s. forest service for providing funds for nutritional analysis, and the alaska department of fish and game for logistic support. we are grateful to the institute of arctic biology, and the department of biology and wildlife, at the university of alaska fairbanks for their support and funding. b. wendling also provided useful support. we thank l. k. duffy for use of his laboratory, and r. kedrowski for technical advice and analysis of forage samples. we are indebted to l. e. emerick for her assistance with fieldwork. we thank f. w. weckerly for assistance with statistical analyses. alces vol. 38, 2002 spaeth et al. nutritional quality of willows 151 references barboza, p. s., and r. t. bowyer. 2000. sexual segregation in dimorphic deer: a new gastrocentric hypothesis. journal of mammalogy 81:473-489. , and . 2001. seasonality of sexual segregation in dimorphic deer: extending the gastrocentric model. alces 37:275-292. barry. t. n., and w. c. mcnabb. 1999. the implications of condensed tannins on the nutritive value of temperate forages fed to ruminants. british journal of nutrition 81:263-272. berger, j., p. b. stacey, l. bellis, and m. p. johnson. 2001. a mammalian predator-prey imbalance: grizzly bear and wolf extinction affect avian neotropical migrants. ecological applications 11:229-240. blaxter, k. l. 1989. energy metabolism in animals and man. cambridge university press, cambridge, uk. bø, s., and o. hjeljord. 1991. do continental moose ranges improve during cloudy summers? canadian journal of zoology 69:1875-1879. bowyer, j. w., and r. t. bowyer. 1997. effects of previous browsing on the selection of willow stems by alaskan moose. alces 33:11-18. bowyer, r. t., m. c. nicholson, e. m. molvar, and j. b. faro. 1999b. moose on kalgin island: are density-dependent processes related to harvest? alces 35:73-89. , b. m. pierce, l. k. duffy, and d. a. haggstrom. 2001. sexual segregation in moose: effects of habitat manipulation. alces 37: 109-122. , v. van ballengerghe, and j. g. kie. 1997. the role of moose in landscape processes: effects of biogeography, population dynamics, and predation. pages 265-287 in j. a. bissonnette, editor. wildlife and landscape ecology: effects of pattern and scale. springerverlag, new york, new york, usa. , , and . 1998. timing and synchrony of parturition in alaskan moose: long-term versus proximal effects of climate. journal of mammalogy 79:1332-1344. , , , and j. a. k. maier. 1999a. birth-site selection in alaskan moose: maternal strategies for coping with a risky environment. journal of mammalogy 80:1070-1083. bryant, j. p., and p. j. kuropat. 1980. selection of winter forage by subarctic browsing vertebrates: the role of plant chemistry. annual review of ecology and systematics 11:261-285. , f. d. provenza, j. pastor, p. b. reichardt, t. p. clausen, and j. t. de toit. 1991. interactions between woody plants and browsing mammals mediated by secondary metabolites. annual review of ecology and systematics 22:431-436. , r. k. swihart, p. b. reichardt, and l. newton. 1994. biogeography of woody plant chemical defense against snowshoe hare browsing: comparison of alaska and eastern north america. oikos 51:385-394. chapin, f. s., iii. 1983. direct and indirect effects of temperature on arctic plants. polar biology 2:47-52. , g. r. shavers, a. e. giblin, k. j. nadelhoffer, and j. a. laundre. 1995. responses of arctic tundra to experimental and observed changes in climate. ecology 76:694-711. cowan, i. mct., w. s. hoar, and j. hatter. 1950. the effect of forest succession upon the quantity and upon the nutritive values of woody plants used as food by moose. canadian journal of research d 28:249-271. edwards, j. 1983. diet shifts of moose due to predator avoidance. oecologia nutritional quality of willows spaeth et al. alces vol. 38, 2002 152 60:185-189. hagerman, a. e., and c. t. robbins. 1993. specificity of tannin-binding salivary proteins relative to diet selection by mammals. canadian journal of zoology 71:628-633. hjeljord, o., n. hovik, and h. b. pedersen. 1990. choice of feeding sites by moose during summer: the influence of forest structure and plant phenology. holarctic ecology 13:333-343. , f. sundstol, and h. haagenrud. 1982. the nutritional value of browse to moose. journal of wildlife management 46:333-343. janzen, d. h. 1979. new horizons in the biology of plant defenses. pages 331348 in g. a. rosenthal and d. h. janzen, editors. herbivores: their intera c t i o n s w i t h s e c o n d a r y p l a n t metabolites. academic press, new york, new york, usa. johnson, r. a., and d. w. wichern. 1982. applied multivariate statistical analysis. second edition. prentice hall, englewood cliffs, new jersey, usa. juntheikki, m.-r. 1996. comparison of tannin-binding proteins in saliva of scandinavian and north american moose (alces alces). biochemical systematics and ecology 24:595-601. kie, j. g. 1999. optimal foraging and risk of predation: effects on behavior and social structure in ungulates. journal of mammalogy 80:1114-1129. , r. t. bowyer, and k .m. stewart. 2003. ungulates in western forests: habitat requirements, population dynamics, and ecosystem processes. pages 296-340 in c. j. zabel and r. g. anthony, editors. mammal community dynamics: management and conservation in the coniferous forests of western north america. the johns hopkins university press, baltimore, maryland, usa. kumar, r., and m. singh. 1984. tannins: their adverse role in ruminant nutrition. journal of agricultural food chemistry 32:447-453. lenart, e. a., r. t. bowyer, j. ver hoef, and r. w. ruess. 2002. climate change and caribou: effects of summer weather on forage. canadian journal of zoology 80:664-678. leslie, d. m., jr., and e. s. starkey. 1987. fecal indices to dietary quality: a reply. journal of wildlife management 51:321325. ludewig, h. a., and r. t. bowyer. 1985. overlap in winter diets of sympatric moose and white-tailed deer in maine. journal of mammalogy 66:390-392. martin, j. s., and m. m. martin. 1982. tannin assays in ecological studies: lack of correlation between phenolics, proanthocyanidins and protein-precipitating constituents in mature foliage of 6 oak species. oecologia 54:205-211. mautz, w. w. 1978. sledding on a brushy hillside: the fat cycle in deer. wildlife society bulletin 6:88-90. mcgarigal, k., s. cushman, and s. stafford. 2000. multivariate statistics for wildlife and ecology research. springer, new york, new york, usa. miquelle, d. g., j. m. peek, and v. van ballenberghe. 1992. sexual segregation in alaskan moose. wildlife monographs 122. molvar, e. m., and r. t. bowyer. 1994. costs and benefits of group living in a recently social ungulate: the alaskan moose. journal of mammalogy 75:621630. , , and v. van ballenberghe. 1993. moose herbivory, browse quality, and nutrient cycling in an alaskan treeline community. oecologia 94:472-479. neter, j., m. h. kutner, c. j. nachtsheim, and w. wasserman. 1996. applied alces vol. 38, 2002 spaeth et al. nutritional quality of willows 153 linear statistical models: regression, analysis of variance and experimental d e s i g n s . f o u r t h e d i t i o n . i r w i n , homewood, illinois, usa. oldemeyer, j. l., and w. l. regelin. 1987. forest succession and habitat management, and moose on the kenai national wildlife refuge. swedish wildlife research supplement 1:163-180. owen-smith, n. 1982. factors influencing the transfer of plant products into large herbivore populations. pages 359-404 in b. j. huntley and b. h. walker, editors. the ecology of tropical savannas. springer-verlag, berlin, germany. pa l o, r. t., k. s u n n e r h e i m , and o. theander. 1985. seasonal variation of phenols, crude proteins and cell wall content of birch (betula pendula) in relation to ruminant in vitro digestibility. oecologia 65:314-318. pastor, j., and r. j. naiman. 1992. selective foraging and ecosystem processes in the boreal forests. american naturalist 139:690-705. peek, j. m. 1974. a review of moose food h a b i t s t u d i e s i n n o r t h a m e r i c a . naturaliste canadien 101:131-141. . 1998. habitat relationships. pages 351-401 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. post, e., and n. c. stenseth. 1999. climate variability, plant phenology, and northern ungulates. ecology 80:13221339. reid, c. s. w., m. j. ulyatt, and j. m. wilson. 1974. plant tannins, bloat and nutritive value. proceedings of the new zealand society of animal production 34:82-93. renecker, l. a., and c. c. schwartz. 1998. food habits and feeding behavior. pages 403-440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. robbins, c. t. 1993. wildlife feeding and nutrition. second edition. academic press, san diego, california, usa. , t. a. hanley, a. e. hagerman, o. hjeljord, d. l. baker, c. c. schwartz, and w. w. mautz. 1987. role of tannins in defending plants against ruminants: reduction in protein availability. ecology 68:98-107. russell, j. b., and j. l. rychlik. 2001. factors that alter rumen microbial ecology. science 292:1119-1122. schwartz, c. c. 1992. physiological and nutritional adaptations of moose to northern environments. alces supplement 1:139-155. , and a. w. franzmann. 1991. interrelationship of black bears to moose and forest succession in the northern coniferous forest. wildlife monographs 113. , m. e. hu b b e r t , and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. journal of wildlife management 52:26-33. , w . l . r e g e l i n , a n d a . w . franzmann. 1987. seasonal weight dynamics of moose. swedish wildlife research supplement 1:301-310. , and l. a. renecker. 1998. nutrition and energetics. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. spaeth, d. f., k. j. hundertmark, r. t. b o w y e r , p . s . b a r b o z a , t . r . stephenson, and r. o. peterson. 2001. incisor arcades of alaskan moose: is nutritional quality of willows spaeth et al. alces vol. 38, 2002 154 dimorphism related to sexual segregation? alces 37: 217-226. spalinger, d. e. 2000. nutritional ecology. pages 108-139 in s. demarais and p. r. krausman, editors. ecology and management of large mammals in north america. prentice hall, upper saddle river, new jersey, usa. stephenson, t. r., v. van ballenberghe, and j. m. peek. 1998. response of moose forage to mechanical cutting on the copper river delta. alces 34:479494. tilley, j. m. a., and r. a. terry. 1963. a two-stage technique for the in vitro digestion of forage crops. journal of the british grassland society 18:104111. van ballenberghe, v. 1992. behavioral adaptations of moose to treeline habitats in subarctic alaska. alces supplement 1:193-206. , d. g. m i q u e l l e , and j. g. maccracken. 1989. heavy utilization of woody plants by moose during summer in denali national park, alaska. alces 25:31-35. van soest, p. j., j. b. robertson, and b. a. lewis. 1991. methods for dietary fiber, neutral detergent fiber and nonstarch polysaccharides in relation to animal nutrition. journal of dairy science 74:3583-3597. v i v a s , h. j., b.-e. s æ t h e r , and r. anderson. 1991. optimal twig-size selection of a generalist herbivore, the moose alces alces: implications for plant-herbivore interactions. journal of animal ecology 60:395-408. weixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces 34:213-238. white, k. s., j. w. testa, and j. berger. 2001. behavior and ecologic effects of differential predation pressure on moose in alaska. journal of mammalogy 82:422-429. white, r. g. 1983. foraging patterns and their multiplier effects on productivity of northern ungulates. oikos 40:377384. f:\alces\vol_39\p65\3917.pdf alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 65 does rectal palpation of pregnant moose cows affect preand neo-natal mortality of their calves? erling johan solberg1, morten heim1, jon martin arnemo2, bernt-erik sæther3, and øystein os4 1norwegian institute for nature research, no-7485 trondheim, norway; 2department of arctic veterinary medicine, the norwegian school of veterinary science, no-9292 tromsø, norway and department of forestry and wilderness management, hedmark university college, evenstad, no-2480 koppang, norway; 3norwegian university of science and technology, department of zoology, realfagbygget, no-7491 trondheim, norway; 4wildlife veterinary consultants, ligard, no-2580 folldal, norway abstract: chemical immobilization and handling of animals may increase the risk of mortality and reduce reproductive output and survival of offspring born to immobilized mothers. we examined to what extent winter chemical immobilization using etorphine and subsequent handling affected the immediate risk of mortality and subsequent calving success and early calf mortality in a norwegian moose population. following 227 immobilizations of 136 different moose, we experienced no mortality during the capture process or any mortality in the 6 weeks following immobilization. similarly, there were no significant differences in calving success (presence or absence of calf or calves) or summer mortality (birth – mid-september) of calves from cows drugged the preceding winter versus undrugged cows. however, splitting the material into age groups, we found significantly lower calving success among drugged cows aged 6-15-years compared to un-drugged cows of the same age group, and higher mortality of calves born to 4-year-old drugged cows compared to those from 4-year-old un-drugged cows. cows that were rectally palpated for pregnancy determination had a higher fetal and calf loss than those that were not rectally palpated. this led us to conclude that rectal palpation, and not the immobilization process per se, was the most likely cause of reduced calving success and calf survival. the presence of a palpation effect may have been influenced by increased stress involved with the weighing process to obtain body mass, but nearly all cows were weighed after immobilization, so we were unable to determine the separate effect of this procedure. alces vol. 39: 65-77 (2003) key words: alces alces, calf mortality, chemical immobilization, fetal loss, moose, rectal palpation providing estimates of big game population parameters is often necessary for wildlife research and sound wildlife management (williams et al. 2002). as a consequence, animals are routinely captured and marked in order to provide information on survival and reproductive rates (williams et al. 2002). many capture methods depend on chemical immobilization to facilitate the capturing and handling process. both chemical immobilization and physical handling, however, may be stressful events for the animal and, depending on the drug, handling method, and species, may have unwanted immediate or long-term effects. these effects may involve direct capture-related mortality (e.g., valkenburg et al. 1983) or increased risk of mortality after immobilization and marking (e.g., gasaway et al. 1978). there may also be reproductive effects following immobilization as well as increased mortality rates of newborns. for instance, chemical immobilization before the rut decreased kid production the following year in mountain goats (oreamnos americanus; côte et al. 1998), and postnatal calf mortalrectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 66 ity rates increased after winter immobilization of pregnant moose cows (alces alces) in canada (larsen and gauthier 1989). in other species, no negative effect of chemical immobilization has been reported on either reproduction or infant survival (e.g., wild horse, equus caballus, berger et al. 1 9 8 3 ; c a r i b o u , r a n g i f e r t a r a n d u s , valkenburg et al. 1983; white-tailed deer, odocoileus virginianus, delgiudice et al. 1986). since the early 1980s, moose have been regularly captured by chemical immobilization from helicopter during winter as part of ongoing research and management projects in norway. recently, arnemo et al. (2001) analyzed the outcome of 1,347 etorphine immobilizations of 1,149 free-ranging moose in norway during the period 1984-2000. seven animals (0.5%) died or were euthanized during the capture process. follow-up radiotelemetry was done for at least 1,119 of the animals (97.4%), from which no mortality caused by the capture (residual drug effects, stress, myopathy, or predation) was observed. in contrast to the short-term effects on survival, information is still lacking on the possible reproductive effects of etorphine when immobilizing pregnant moose. chemical immobilization is previously suggested to have negative immediate and long-term impact on fetal and/or early calf survival in moose (e.g., larsen and gauthier 1989), and similarly, different handling procedures, such as rectal palpation for pregnancy determination, can potentially affect the reproductive outcome (franco et al. 1987, thurmond and picanso 1993). rectal palpation is one of the most commonly used procedures to diagnose pregnancy in domestic cattle (e.g., arthur et al. 1989) and is increasingly used to determine pregnancy in immobilized moose cows (e.g., ballard and tobey 1981, haigh et al. 1982, larsen and gauthier 1989, gasaway et al. 1992). however, while this method is considered a safe method with respect to fetal survival in domestic cows (arthur et al. 1989), less information is available on the possible negative effect on reproduction in moose. in this paper, we examine the possible effects that etorphine immobilization and rectal palpation has on moose cows, subsequent calving success, and early calf mortality in a norwegian moose population. based on the assumption that chemical immobilization and handling only affect reproduction the following spring, we examined to what extent fetal and neonatal mortality differed between years of immobilization and other years. we hypothesized that neither chemical immobilization nor rectal palpation of moose cows increase the probability of fetal or neonatal loss. study area the study area was the island of vega (fig. 1) located off the coast of nordland county (65°40‘n, 11°55‘e) in northernnorway. the island was colonized by moose in 1985 when two yearling cows and one yearling bull swam over from the nearby mainland (4-8 km) (sæther et al. 2001). since then, the population has increased in number both due to immigration and reproduction, peaking at 43 animals during the winter 1993/1994 (sæther et al. 2001). since 1995, harvesting has maintained the winter population size at about 30-35 animals (0.250.30 moose/km2). the climate on vega is highly oceanic, with mild winters (fig. 2) and only short periods of continuous snow cover during most winters. one exception occurred in 1994, when relatively deep snow covered the island during the period january-march (fig. 2). large carnivores are absent from the island. methods in 1992, we started a study on the island of vega to study the dynamics in a moose alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 67 population subject to different hunting regimes (sæther et al. 2001). that included radio-collaring all animals present on the island as well as all calves and immigrating animals in subsequent years. as part of the study, we manipulated the adult sex and age structure to examine to what extent a low proportion of males and/or low male age influenced the reproductive performance of the cows. the results indicated that biased adult sex ratio during the rutting season led to delayed parturition the following spring, whereas no effects were observed on calving success (sæther et al. 2003). thus, we have no reason to believe that this manipulation confounded the results of the present study. moose were captured by the use of a helicopter and a remote drug delivery system (dan-inject, børkop, denmark) during winter from january to march. an initial dose of 7.5-9.0 mg etorphine per adult was used for immobilization (kreeger et al. 2002). in animals not recumbent within 15-20 minutes, another full dose of etorphine was administered. diprenorphine at 12 mg per 9 mg etorphine was used for reversal. after immobilization, all moose were radio-collared and ear-tagged, and a number of morphological measurements, such as leg-length and shoulder-hoof-length, were taken. body mass was measured using net and helicopter in 98% of the cases. in addition, 67 % of the moose cows were rectally palpated on one or several occas i o n s b y o n e o f t w o e x p e r i e n c e d veterinarians (coauthors jon martin arnemo and øystein os) to detect pregnancy (haigh et al. 1982, arthur et al. 1989). pregnancy fig. 2. mean snow depth (closed circles) and temperature (open circles) at vega during january-march in the study period 1992-2002. fig. 1. the location of the study area, the island of vega (65° 40’ n, map based on ©statens kartverk. mad12002-r123230). rectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 68 was determined by slipping fetal membranes (the main method employed in the first trimester) or by palpation of the uterine artery or the fetus itself (in the second trimester). other field procedures included sampling of blood, feces, and skin biopsy. during the study period a total of 44 immobilizations were carried out on 31 different cows. age of each individual was determined by the time elapsed since radio-collared as calf or yearling, or post mortem by the tooth replacement pattern (yearlings) or number of cementum annulli of the incisors (haagenrud 1978). three cows radio-collared as adults were not aged, either because they emigrated before they died or because their jaws were not collected. these animals were given a minimum age based on body size, tooth wear, and reproductive status when first handled. the number of calves was determined by locating and approaching the radio-collared moose cows on foot every 3-4 days throughout the calving season (may 15 to the end of june) (sæther et al. 1996), and more infrequently during july and august to follow up the non-calving cows. in the summer of 2002, we did not check any cows for calving after july15, assuming that the 2 individuals not observed with calves were barren this year. occasionally moose may give birth late in the summer, following late conception during the previous rutting season (see schwartz 1998 for a review), but in the present study no cow was found with calf/calves after the 8th of july. calving success was defined as the proportion of cows that were observed with a calf or calves in the spring and twinning rate was the proportion of calving cows that produced twins. we also located and approached radiocollared cows on foot in the autumn just prior to the hunting season (hunting starts between 25 september and 10 october) to determine calf survival. seven calves that disappeared during summer (3 and 4 calves from drugged and un-drugged mothers, respectively) were excluded from the sample because their deaths were probably caused by human interference. data analyses and predictions to determine the effect of immobilization and handling on reproductive performance, we compared the outcome of calving for immobilized cows (treatment group) to that of non-immobilized cows (control group). if immobilization has a negative effect, we expected lower calving success and fewer calves born per cow in years of immobilization compared to other years. fetal and/or early calf mortality may depend on the developmental stage of the fetus at the time of immobilization (24 january – 30 march) and whether rectal palpation was conducted (e.g., franco et al. 1987, thurmond and picanso 1993). to test the importance of rectal palpation and immobilization date on fetal and neonatal mortality, we compared the calving success and calf mortality for those palpated with those not, and before and after the median immobilization date (23 february), respectively. because the rectal palpated cows were immobilized 13 days after the non-palpated cows (21 february versus 6 march) we also examined to what extent there was a relationship between palpation date and fetal and neonatal mortality. we further examined to what extent variation in body mass, as an index of body condition, influenced calving success and calf mortality in the year the cows were immobilized and weighed. to control for the higher probability of maturity with increasing body mass among primiparous cows (e.g., sæther and haagenrud 1983, 1985), we split the data set into low and high body mass within age classes; i.e., we examined if small cows were more affected by immobilization and alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 69 handling than large cows of the same age. moreover, based on data from 15 palpated cows from which we also collected blood samples, we examined to what extent calf loss was associated with low serum progesterone levels (franco et al. 1987). because reproductive rates and calf mortality were recorded earlier in the study period for palpated (in the years 1992-95, 2000, and 2001) than non-palpated cows (1992, 1999, and 2002), annual variation in living conditions may have caused spurious differences in fetal and neonatal mortality between palpated and non-palpated cows. for instance, both abortion rate and early calf mortality may depend on the severity of the winter. we therefore checked whether higher fetal or calf loss occurred in 1994 when the average snow depth was higher than normal at vega (fig. 2). similarly, we tested for any trends in fetal and neonatal mortality over years as the effect of palpation may depend on the experience of the examiners. the two veterinarians involved in pregnancy determination had 8 and 17 years experience, respectively, as dairy cattle veterinarians prior to the start of the project. this regularly involved palpation of dairy cows to determine pregnancy. however, they had no previous experience in palpating moose. we analyzed the effect of immobilization on the response variables by fitting logistic models (proc genmod, sas institute 1996) to the data by maximum likelihood estimation. the factor ‘immobilization’ (yes = 1, no = 0), rectal palpation (yes, no), immobilization date (before, after median date), and body mass (high, low) were included as independent variables. because the response variables calving success (calves/calf = 1, no calf = 0), twinning rate (twins = 1, single calf = 0), and calf mortality (alive = 1, dead = 0) were assumed to follow a binomial distribution, the models were run with a logit link function (sas institute 1996). to control for the lack of independence among observations in cases where the same individual contributed several observations to the same age and study group (treatment/control), we randomly drew one observation for each individual cow. thus, for cows immobilized or checked for calving and calf mortality several times, only one observation was used in the treatment group (immobilization) and one in the control group. a possible alternative method using mean values for cows that contributed several observations within ageand study group and testing for differences in a contingency table analysis (e.g., g-test, sokal and rohlf 1981) was not employed because of the many cells with low frequencies (i.e., zero). however, for comparison, we have shown the percentage calving success, twinning rate, and calf mortality based on mean values from the different moose cows in tables 1 and 2. results during the study period, 19922002, we immobilized and handled 62 moose bulls, 45 moose cows, and 120 calves distributed over 136 different individuals. no individual died during the capturing or handling process. similarly, no captured individuals were found dead during the following 6 weeks after capturing or were observed to have any problems that could be related to the capturing process. indeed, for moose found dead after 6 weeks, we found no indications that this was related to immobilization. in general, calving success and twinning rate was high (table 1), and based on the total data set there were no significant differences in reproductive rates between cows in the year they were drugged and in years they were not immobilized. however, splitting the data set into age-groups rectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 70 table 1. differences in calving success (% with calves), twin-production (% with twins), and calf mortality (% calf mortality) between drugged and un-drugged moose cows at vega in the period 1992-2002. samples in pooled age-groups are based on randomly selecting one observation from those cows that contribute more than one observation to the study group (drugged or undrugged). for comparison, calving success, twinning rate, and calf mortality (%) based on mean values from each cow are shown in the table (mean %). statistics in bold indicate significant differences. cow drugged the previous year yes no calves twins calves calves twins calves cow /no calves / single alive /no calves / single alive age (mean %) (mean %) /calves (mean %) (mean %) /calves χ 2 p dead dead (mean %) (mean %) 2 6/5 9/10 0.14 0.70 2 2/4 3/6 0.00 1.00 2 8/0 10/1 1.22 0.27 3 4/0 16/3 1.24 0.27 3 2/2 8/8 0.00 1.00 3 4/0 20/2 0.70 0.40 4 5/0 16/0 0.00 1.00 4 4/1 12/4 0.05 0.82 4 6/3 28/0 9.37 0.002 5 9/0 10/0 0.00 1.00 5 7/2 9/1 0.54 0.46 5 16/0 19/0 0.00 1.00 6-15 7/4 (64) 14/0 (100) 7.56 0.006 6-15 7/0 (100) 12/2 (76) 1.73 0.19 6-15 13/1 (7) 24/2 (10) 0.00 0.94 2-15 24/7 (78) 20/10 (81) 0.88 0.35 2-15 16/8 (71) 11/9 (62) 0.63 0.43 2-15 38/4 (8) 38/2 (4) 0.63 0.43 showed that older-aged cows (6-15 yearolds) had lower calving success after immobilization than in other years (p = 0.006, table 1). no such effect was present in the twinning rate (table 1). the variation in calving success or twinning rate was not related to the body mass of the cows (mean = 352 kg and range = 249-438 kg) in the winter of immobilization, or to immobilization date (table 2). in alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 71 a immobilization in relation to median immobilization date before after cow calves twins calves alive calves twins calves alive age /no calves / single /calves dead /no calves / single /calves dead χ2 p (mean %) (mean %) (mean %) (mean %) (mean %) (mean %) 2-15 18/7 11/1 (92) 2.10 0.15 2-15 (70) 12/6 (69) 9/2 (77) 0.82 0.37 2-15 23/3 (9) 19/1 (5) 0.77 0.38 b above or below age-specific mean body mass below above 2-15 16/4 14/2 (84) 0.92 0.34 2-15 (75) 13/3 (78) 9/5 (68) 0.90 0.34 2-15 25/1 (3) 19/1 (4) 0.08 0.77 c cow handled by rectal palpation yes no 2-15 14/6 (67) 14/1 (97) 3.25 0.07 2-15 11/3 (77) 9/5 (70) 0.52 0.47 2-15 19/4 (15) 25/0 (0) 10.40 0.001 contrast, calving rate, but not twinning rate, was lower in cows in which we had determined pregnancy by rectal palpation compared to those without palpation (p = 0.072, table 2c). indeed, of all observations of cows not recorded with calf/calves in the spring following immobilization (10 of 43 observations), 9 were of previously palpated cows and only one was from a nonpalpated cow. three of 26 cows palpated (6 handled twice) were not found pregnant, and accordingly, none of them was seen in company with calf/calves in the spring. of the remaining 23 cow observations, all were diagnosed pregnant, but only 17 (74%, sd = 44.9) were observed with calf/calves in the spring. one of the cows not observed with calves was diagnosed to have a poorly developed fetus, possibly because of late conception, or because the fetus was dead and in the process of being reabsorbed. summer mortality of calves was analyzed based on 204 calves from 120 cow-observations. eleven (5.4%) of the calves disappeared during summer, 4 (7.5%) from cows that were immobilized and 7 (4.7%) from cows that were not. after randomly selecting 1 observation per cow table 2. variation in calving success, twinning rate, and summer calf mortality in relation to: (a) immobilization date (before and after median date); (b) cow body mass (above and below average); and (c) between cows checked for pregnancy by rectal palpation or not when immobilized in winter at vega during the period 1992-2002. samples are based on randomly selecting one observation from those cows that contribute more than one observation to the study groups. for comparison, calving success, twinning rate, and calf mortality (in %) based on mean values from each cow are shown in the table (mean %). statistics in bold indicate significant differences. rectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 72 and study group, no significant difference in calf mortality existed within the complete sample between calves produced by mothers immobilized the preceding winter and calves from mothers not immobilized (table 1). however, after splitting by age, there was significantly higher mortality among calves produced by drugged 4-year-old mothers compared to those that were not drugged (table 1). in this case, 2 cows lost 3 out of 4 calves. no difference was found in any other age group (table 1). no effect of body mass or immobilization date was found on calf mortality (table 2). however, as for calving success, there was a significant association between rectal palpation and calf mortality (p = 0.001, table 2), as all 4 calves that died were from palpated mothers. thus, of 26 palpations of 20 cows, 3 cows were not found pregnant, 6 (26%) lost their calf/calves during pregnancy or just after calving, while another 3 cows (18%) lost 4 out of 6 calves during summer. this contrasts with 17 observations of 15 cows that were not palpated, of which all but one were found with 1 (5) or 2 calves (11) in the spring and none lost their calves during summer. hence, despite the low sample size, the pattern of higher fetal/ calf loss after palpation appears quite consistent. indeed, excluding the cows that were palpated from the sample, calving success (χ2 = 4.51, p = 0.034), but not calf survival (χ2 = 2.35, p = 0.12) was significantly higher in drugged compared to undrugged cows. three yearling cows that were immobilized in the winter of 1994 were all found pregnant by palpation, but none of them was observed with calves in the spring. in contrast, 10 cows (9 adults, 1 yearling) not drugged that year were all seen with calf/ calves in the spring, although one lost a calf during summer. hence, immobilization and/ or palpation, and not winter severity (fig. 2), appear to be the most likely reason for fetus loss in immobilized cows this year. when we examined the 23 observations from 18 cows found pregnant by rectal palpation, no significant relationship existed between fetus/calf loss and year of immobilization (b = -0.442, χ2 = 1.069, p = 0.30). similarly, no significant relationship existed between fetus/calf loss and immobilization date (b = -0.014, χ2 = 0.41, p = 0.52), but there was a tendency for cows palpated early in the winter to experience higher calf loss than those palpated later. this was particularly apparent for the 4 individual cows palpated in january, as all lost their fetus, whereas only 5 of 19 palpations (26%) from 16 cows immobilized in february and march were associated with fetal/calf loss (χ2 = 4.51, p = 0.034). no relationship existed between fetus/calf loss of palpated cows and cow age (b = -0.065, χ2 = 0.13, p = 0.72) or cow body mass (b = -0.000, χ2 = 0.00, p = 0.99). based on progesterone levels in blood serum from 15 cows (mean = 16 nmol/l, range = 7-24) that were found pregnant by rectal palpation (2 cows contributed twice), 3 cows that were not observed with calf/ calves had values at average or below (mean = 12 nmol/l), but this association was not significant (χ2 = 1.74, p = 0.19). discussion a l t h o u g h a l a r g e n u m b e r o f immobilizations (n = 227) were carried out, no capture-related mortalities occurred in the present study. this contrasts with several studies from north america, where the reported mortality rates following chemical immobilization of moose typically range from 6 to 19% (roffe et al. 2001). the lack of immobilization-related mortality at vega may be due in part to the exceptionally good condition of moose on the island, following the low moose density and mild winters (sæther et al. 2001, fig. 2). carcass mass of harvested moose on the island is higher alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 73 than what is normally observed on mainland populations and accordingly, both calving success and the twinning rate is high (e.g., calving success/percent with twins: 52/33% and 94/75% for 2-year olds and older cows, respectively; sæther et al. 2001, unpublished data) compared to most norwegian (sæther et al. 1992) and north american (e.g., gasaway et al. 1992, schwartz 1998) moose populations. however, even in other, p r e s u m a b l y m o r e f o o d c o n s t r a i n e d populations in norway, capture-related mortality following chemical immobilization by etorphine during winter has been found very low (i.e., below 0.5 %; arnemo et al. 2001). etorphine has no major clinical side effects in moose and has a wide safety margin; i.e., the same dose can be used in all adults irrespective of body mass (arnemo et al. 2001, kreeger et al. 2002). in addition, a skilled and experienced capture team will significantly reduce the risk of anesthetic mortality by use of proper remote drugdelivery systems and well-established capturing methods. thus, immobilization of moose from helicopter with etorphine in winter can be considered a safe procedure in this species (arnemo et al. 2001). despite the lack of immediate mortality following moose immobilization, we found significantly lower calving success and calf survival within some age groups in years of immobilization (table 1). this supports several other studies reporting that immobilization and handling can increase fetal loss or early calf mortality in different large herbivores (e.g., mountain goats, côté et al. 1998; black rhino, diceros bicornis, alibhai et al. 2002), including moose (e.g., ballard and tobey 1981, larsen and gauthier 1989), although following the use of other drugs than etorphine. for instance, côté et al. (1998) found significant immobilization effects on kid production among young primiparous mountain goats, but not among older animals. this contrasts with the present study where the effect of drugging on calving success was only apparent in older cows. young animals may be particularly vulnerable to various types of stress because of lower than average body mass and condition, but so may older, senescent individuals (e.g., loison et al. 1999). in moose, senescence in mortality and reproduction have been reported to occur in cows, particularly after 10-12 years of age (e.g., ericsson and wallin 2001, ericsson et al. 2001), possibly because of increasing tooth wear and subsequent decreasing body condition with age (ericsson and wallin 2001). among the cows that were observed without calves in our study, however, neither were older than 9 years or had lower body mass than cows that were observed with calf/calves, and none of the older cows were immobilized during the winter of 1994 (fig. 2), when a temporary decrease in the accessibility of food may have occurred. as an alternative explanation, we suggest that the lower calving success and survival of calves born from drugged cows were due to rectal palpation. all cows that experienced calf loss were previously rectally palpated in winter, whereas only 1 of 15 cows (17 observations) immobilized, but not palpated, were observed without calf/calves or lost 1 or 2 calves during summer (table 2). the use of palpation in domestic cows is usually regarded as a safe method with respect to embryonic or fetal mortality (arthur et al. 1989), and quite aggressive handling has been assumed necessary to cause abortion (paisley et al. 1978). more recent studies, however, indicate that rectal palpation may affect the abortion rate in cattle (franco et al 1987, thurmond and picanso 1993, but see alexander et al. 1995), although at a relatively modest level. franco et al. (1987) found about 10% higher abortion rate in dairy cows that were palpated compared to those rectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 74 that were not. this contrasts with the 26% of palpated moose cows that experienced calf loss in the present study. although the cause of increasing calf loss after palpation is unknown, we speculate that palpation conducted under field conditions may increase the chance of damaging the fetus, which increases the risk of abortion or early calf mortality. in cattle, rectal palpation is carried out with the cow standing and physically restrained, whereas in moose, rectal palpation has to be performed during chemical immobilization with the cow in sternal or lateral recumbency. these differences in handling conditions may increase the possibility of trauma to the fetal membranes or the fetus. the weighing of the pregnant cows in a net beneath the helicopter may have contributed to this effect, but as nearly all cows (98%) were weighed immediately following immobilization, we were unable to determine the separate effect of this procedure. accordingly, we cannot exclude the possibility that the additional stress caused by weighing decreased the threshold for rectal palpation to have an effect. there may also exist individual differences among cows in their tolerance of palpation due to variation in age or hormonal imbalances (arthur et al. 1989). franco et al. (1987) found in dairy cows that later experienced embryonic or fetal death had lower than average milk progesterone, possibly making them less fit to maintain pregnancy following a stressful event. in the present study, no significant relationship existed between fetal mortality and blood serum progesterone concentration, although there was a trend for lower concentrations to be associated with cows experiencing fetal loss. moreover, of 11 cows that were immobilized twice, 4 experienced fetal/calf loss, of which 3 lost fetus/calves on both occasions. all these cows were palpated. thus, certain individuals seem to be more prone to fetal or calf loss than others following immobilization and palpation. another factor that may affect the reproductive outcome is the timing of palpation in relation to conception date. domestic cows are often palpated before fetal attachment (45 days after conception) to allow pregnancy diagnosis following insemination (thurmond and picanso 1993). at this stage the embryo may be particularly vulnerable to trauma following palpation and consequently it has been proposed that palpation before day 45 may increase the risk of abortion (franco et al. 1987, but see thurmond and picanso 1993). in our study, palpation occurred after placental attachment, as all immobilizations occurred between day 116 and 181 (assuming conception at october 1) at a stage where the fetus is relatively well developed (schwartz and hundertmark 1993) and presumably more resistant to fetal trauma following palpation. because of the smaller size of the fetus in the early pregnancy, however, pregnancy at this stage is often determined by the slipping fetal membrane method (arthur et al. 1989), which in our opinion is more difficult and invasive than palpation of the fetus or uterine artery as is usually done at a later stage of pregnancy. this may explain the higher incidence of fetal/calf loss among those cows that were palpated in early winter. few other studies have directly tested to what extent rectal palpation induces fetal loss in moose. ballard and tobey (1981) found that moose cows immobilized and palpated in march experienced significantly lower calving success than cows immobilized, but not palpated, in october. this could be due to an effect of palpation, but as the probability of abortion due to chemical immobilization could also vary with the age of the embryo or fetus, this is impossible to d e t e r m i n e . i n c o n t r a s t , s m i t h a n d franzmann (1979) were cited in ballard and alces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 75 tobey (1981) to find no effect of palpation on reproduction after immobilization by the use of m99 and rompun. whether this was based on extensive data is unknown as we have not been able to retrieve this technical report. in other studies, rectal palpation has often been conducted on all, or the majority of cows immobilized, making it impossible to determine whether the drug or the handling method is the main cause of fetal or calf loss (e.g., larsen and gauthier 1989). we conclude that rectal palpation was the main reason for the fetal/calf loss observed in the present study, possibly in combination with the weighing procedure to obtain body mass. we find it unlikely that the effect was due to improper palpation by the examiners as, to our experience, palpating moose does not differ extensively from palpating dairy cows (except for the position). the lack of any experience effect was also supported by the non-significant relationship between fetal/calf loss and palpation-year. moreover, we did not find it more difficult to palpate and determine pregnancy in older compared to younger cows, which could possibly have explained the age class disparity in the observed fetal and calf loss; i.e., by erroneously classifying nonpregnant old cows as pregnant. indeed, of all 4-15 year-old cows (17) palpated, 94% (16) were found pregnant, which is about the pregnancy rate to expect for this age class (schwartz 1998). based on the present results, we encourage other studies to evaluate the potential effect of rectal palpation on fetal/calf loss before uncritical use of this method for pregnancy determination in moose. this may not be a great loss as alternative noninvasive methods are available, such as pregnancy diagnosis based on progesterone levels in blood serum and fecal samples (e.g., haigh et al. 1982, monfort et al. 1993, schwartz 1998, arnemo et al. unpublished data) or by using serum pregnancy tests based on radioimmunoassays (ria) specific for moose pregnancy-specific protein b (pspb; huang et al. 2000). the former two methods cannot be used to determine the number of fetuses, but neither can, in our opinion, rectal palpation (but see haigh et al. 1982, larsen and gauthier 1989). by the use of ria and pspb in sera, however, huang et al. (2000) found that both pregnancy and the number of fetuses could be determined with high accuracy in immobilized free-ranging moose. acknowledgements we are grateful to the directorate for nature management, the county governor of nordland county and the research council of norway (changing landscape) for f i n a n c i a l s u p p o r t a n d t o t . b ø , b . aleksandersen, and o. a. davidsen for support and help. we appreciated the constructive comments from b. vereijken and her help with the english. the comments from two anonymous referees greatly improved the manuscript. references alexander, b. m., m. s. johnsen, r. o. guardia, w. l. vandegraaf, p. l. senger, and r. g. sasser. 1995. embryonic loss from 30 to 60 days post breeding and the effect of palpation per rectum on pregnancy. theriogenology 43:551-556. alibhai, s. k., z. c. jewell, and s. s. towindo. 2002. effects of immobilisation on fertility in female black rhino (diceros bicornis). journal of zoology 253:333-345. arnemo, j. m., e. o. øen, and m. heim. 2001. risk assessment of etorphine immobilization in moose: a review of 1,347 captures. page 179 in proceedings of the society for tropical veterinary medicine and wildlife disease association international joint conferrectal palpation of pregnant moose cows solberg et al. alces vol. 39, 2003 76 ence, pilanesberg national park, south africa, 22-27 july 2001. arthur, g. h., d. e. noakes, and h. pearson. 1989. veterinary reproduction and obstetrics. baillière tindall, london, u.k. ballard, w. b., and r. w. tobey. 1981. decreased calf production of moose immobilized with anectine administered from helicopter. wildlife society bulletin 9:207-209. berger, j., m. kock, c. cunningham, and n. dodson. 1983. chemical restraint of wild horses: effects on reproduction and social structure. journal of wildlife diseases 19:265-268. côté, s. d., m. festa-bianchet, and f. fournier. 1998. life-history effects of chemical immobilization and radiocollars on mountain goats. journal of wildlife management 62:745-752. delgiudice, g. d., l. d. mech, w. j. paul, and p. d. karns. 1986. effects on fawn survival of multiple immobilizations of captive pregnant white-tailed deer. journal of wildlife diseases 22:245248. ericsson, g., and k. wallin. 2001. agespecific moose (alces alces) mortality in a predator-free environment: evidence of senescence in females. ecoscience 8:157-163. , , j. p. ball, and m. borberg. 2001. age-related reproductive effort and senescence in freeranging moose, alces alces. ecology 82:1613-1620. franco, o. j., m. drost, m.-j. tatcher, v. m. shille, and w. w. tatcher. 1987. fetal development in the cow after pregnancy diagnosis by palpation per rectum. theriogenology 27:631-644. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife monographs 120. , a. w. franzmann, and j. b. faro. 1978. immobilizing moose with a mixture of etorphine and xylazine hydrochloride. journal of wildlife management 42:686-690. haagenrud, h. 1978. layers of secondary dentine in incisors as age criteria in moose (alces alces). journal of mammalogy 59:857-858. haigh, j. c., e. h. kowal, w. runge, and g. wobeser. 1982. pregnancy diagnosis as a management tool for moose. alces 18:45-53. h u a n g , f . , d . c . c o c k r e l l , t . r . stephenson, j. h. noeyes, and r. g. sasser. 2000. a serum pregnancy test with a specific radioimmunoassay for moose and elk pregnancy-specific protein b. journal of wildlife management 64:492-499. kreeger, t. j., j. m. arnemo, and j. p. raath. 2002. handbook of wildlife chemical immobilization. international edition. wildlife pharmaceuticals, fort collins, colorado, usa. larsen, d. g., and d. a. gauthier 1989. effects of capturing pregnant moose and calves on calf survivorship. journal of wildlife management 53:564-567. loison, a., m. fe s t a-bianchet, j.-m. gaillard, j. t. jorgenson, and j.-m. jullien. 1999. age-specific survival in five populations of ungulates: evidence of senescence. ecology 80:2539-2554. monfort, s. l., c. c. schwartz, and s. k. wasser. 1993. monitoring reproduction in captive moose using urinary and fecal steroid metabolites. journal of wildlife management 57:400-407. paisley, h. w., w. d. mickelsen, and o. l. prost. 1978. a survey of the incidence of prenatal mortality in cattle following pregnancy diagnosis by rectal palpaalces vol. 39, 2003 solberg et al. rectal palpation of pregnant moose cows 77 tion. theriogenology 9:481-491. roffe, t. j., k. coffin, and j. berger. 2001. survival and immobilizing moose with carfentanil and xylazine. wildlife society bulletin 29:1140-1146. sæther, b.e., r. andersen, o. hjeljord, and m. heim. 1996. ecological correlates of regional variation in life history of the moose alces alces. ecology 77:1493-1500. , and h. haagenrud. 1983. life history of moose (alces alces): fecundity rates in relation to age and carcass weight. journal of mammalogy 64:226232. , and . 1985. life history of the moose alces alces: relationship between growth and reproduction. holarctic ecology 8:100-106. , m. he i m , e. j. so l b e r g, k. jacobsen, r. olstad, j. stacy, and m. sviland. 2001. effekter av rettet avskytning på elgbestanden på vega. nina-fagrapport 049. (in norwegian with english summary). , e.j.solberg, and m. heim. 2003. effects of altering adult sex ratio and male age structure on the demography of an isolated moose population. journal of wildlife management 67: 455-466. , k. solbraa, d. p. sødal, and o. hjeljord. 1992. sluttrapport elg-skogsamfunn. nina forskningsrapport 28. (in norwegian with english summary). sas institute. 1996. sas/stat software. changes and enhancements. sas institute, cary, north carolina, usa. schwartz, c. c. 1998. reproduction, natality and growth. pages 141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. , and k. j. hundertmark. 1993. reproductive characteristics of alaskan moose. journal of wildlife management 57:454-468. smith, c. a., and a. w. franzmann. 1979. productivity and physiology of yakutian forelands moose. alaska department of fish and game. pittman robertson final report w-17-10 and w-17-11. sokal, r. r., and f. j. rohlf. 1981. biometry. second edition. w. h. freeman, new york, new york, usa. thurmond, m. c., and j. p. picanso. 1993. fetal loss associated with palpation per rectum to diagnose pregnancy in cows. journal of american veterinary medical association 203:432-435. valkenburg, p., r. d. boertje, and j. l. davis. 1983. effects of darting and netting on caribou in alaska. journal of wildlife management 47:1233-1237. williams, b. k., j. d. nichols, and m. j. conroy. 2002. analysis and management of animal populations: modeling, estimation and decision making. academic press, new york, new york, usa. alces34(1)_157.pdf alces37(1)_163.pdf alces35_125.pdf alces35_203.pdf alces34(2)_279.pdf alces36_41.pdf alces34(1)_201.pdf alces34(2)_339.pdf 4007.p65 alces vol. 40, 2004 crichton et al. response to access management 87 response of a wintering moose population to access management and no hunting – a manitoba experiment vince crichton1, trevor barker2, and doug schindler3 1wildlife and ecosystem protection branch, manitoba conservation, box 24, 200 saulteaux crescent, winnipeg, mb, canada r3j 3w3; 2regional operations, manitoba conservation, box 4000 lac du bonnet, mb, canada r0e 1a0; 3 university of winnipeg, 515 portage avenue, winnipeg, mb, canada r3b 2e9 abstract: we report on an experiment undertaken in eastern manitoba beginning in 1996, in which a moose population wintering in 62 km2 (24.2 mi2) was protected from hunting until september 2003. at the time of closure, it is speculated that about 37 (0.6/km2 (1.5/mi2)) moose wintered in the area based on aerial surveys and considering visibility bias. the closure was supported by the eastern region committee for moose management, which is comprised of manitoba conservation staff, first nation representatives from local communities, local hunting organizations, and other interest groups such as tembec manitoba incorporated and the manitoba model forest. road access to the area was curtailed by using locked gates, millstones, and v-plowing a portion of the road in 2002. the area was surveyed from a helicopter on march 4, 2003, and 107 moose were counted in the closed area and again, based on visibility bias, it is speculated that about 142 moose (2.3/km2 (5.8/mi2 )) were present. this experiment clearly demonstrates that moose will respond positively to access management and no hunting, and that v-plowing roadbeds is a useful technique for controlling access. the cost associated with such plowing varies from about $500-$1,500/km depending on material contained in the roadbed. alces vol. 40: 87-94 (2004) key words: access, financially feasible, moose, no hunting, positive response both positive and negative impacts of roads on wildlife values in forested areas have been documented. from a positive perspective, access provides opportunities for public use of many resources and offers other recreational opportunities. on the other hand, roads and associated public access are a component of increased land use which can be a threat to the sustainability of wildlife populations. these threats are manifested as: (1) direct loss of habitat along cleared rights-of-way; (2) a potential for increased hunting, subsequent harvest, and associated disturbance; (3) decreased habitat utilization adjacent to roads due to motorized traffic and other human activities (bjorge 1984, singer and beattie 1985, thiel 1985, ellison et al. 1986, shideler et al. 1986, cameron et al. 1992, nellemann and cameron 1996, jalkotzy et al. 1997); (4) habitat fragmentation; and, (5) displacement by exotic species (e.g., cowbirds displacing warblers). to minimize the negatives (e.g., destruction of important habitats, unsustainable harvests, displacement of species into less preferable habitat, etc.), the location and public use of roads must be addressed early in the forest management planning process. human activities on roads can impact wildlife in variable ways ranging from subtle energetic costs to animals constantly exposed to such disturbances, to being killed. wildlife, on the other hand, have the option of moving to avoid the disturbance or with some territorial species, existing within the response to access management crichton et al. alces vol. 40, 2004 88 zone of influence. with the former, animals may be forced into less preferred habitats and become exposed to factors such as increased predation, nest parasitism, disease, etc. reducing foraging efficiency and altering energetic costs can further affect activity budgets. these incremental energetic costs are less pronounced in undisturbed habitats. the cumulative impact of roads and associated activities must be understood and acceptance given that roads significantly alter landscape dynamics. road density is an excellent indicator of the location and the potential pressure placed on wilderness areas. with this in mind, road development proposals must be closely scrutinized with the view to maintaining the lowest road density possible and approving locations, which will minimize impacts on important habitats, wildlife sustainability, and reduce fragmentation. each proposal must also include a proactive access management and road retirement program recognizing that once a tradition of access has been established on such roads the chances of closing it are greatly reduced. manitoba conservation (mc) in 1996 initiated the eastern region committee for moose management, which functioned in an advisory capacity to mc on issues applicable to moose (alces alces andersoni) and moose habitat. this committee is comprised of mc staff, representatives from local first nation communities, local game and fish clubs, tembec manitoba inc., the manitoba model forest, and various other stakeholders. in 1996, the committee recommended to mc that the study area be closed to moose hunting from 1996 until 2000 inclusive, at which time moose hunting would again be permitted. mc accepted this recommendation and although the agreement was for a 5-year closure, this was extended to september 2003, 7 years after the initial closure. this paper presents the results of the closure on the number of moose wintering in the happy lake area and what was done to close roads. although there is much written in the literature on the impacts of roads, there is a paucity of information on how to effectively deal with access and access management. it is hoped that this paper will, in a small way, begin to address these issues. study area the study area comprising 62 km2 (24.2 mi2) is located in mc’s game hunting area (gha) 26 with the centre being at approximately latitude 500 50' 26'' and longitude 950 30' 30''. it is located in eastern manitoba in the southern portion of the lac seul upland (boreal plain ecoregion) which encompasses the southern part of canada’s precambrian shield and is 25.6 km (16 mi) west of the provincial boundary between manitoba and ontario. the gha has been extensively logged over the last 80 years with softwoods being the primary species taken and wildfires have occurred periodically. the result of the aforementioned is moose habitat, which is considered high quality. previously the area was a mature mixedwood forest. this gha is a designated route area for big game hunting and all vehicles used for moose hunting by licensed hunters are restricted to designated trails and/or roads but can leave the trails/ roads for retrieval purposes. the study area has one logging road accessing it and on the north side it is accessible by water or over the ice in winter. the response of the moose population to access control and no hunting was studied. although movement studies have not been undertaken, the presumption is, based on familiarity of the area by the senior author, that all of the moose seen are not resident in the area year round but rather an unknown number reside in adjacent areas during the summer and fall and are subjected to hunting in these habitats. it is speculated that movement to this alces vol. 40, 2004 crichton et al. response to access management 89 relatively remote area occurs in early winter. concomitant with wood harvesting has been the need to develop a network of roads to access the timber resource. all-weather access to the area was developed for timber harvesting and renewal. moose hunting by licensed hunters for the last 40 years has been restricted to a 2week period in early december and since the early 1980s, the bag limit has been bulls only. previously it was any moose. mc in the late 1960s recognized the need to restrict vehicles if the moose population was to be sustainable, and at that time embarked on a system of designated routes, which is still in effect. hunting by first nation peoples is without restrictions in gha 26 and they are able to travel anywhere using vehicles except on closed roads, can harvest any moose, and there are no restrictions on numbers that can be taken. methods the study area, which was delineated based on roads, a hydro electric transmission line, lakes, creeks, and rivers, was initially surveyed for moose using a 206b helicopter in 1996 and each subsequent year (excluding 1998/99) up to and including 2002/03. it had been logged prior to the closure. the protocol for each survey was to have 2 experienced observers, a navigator who directed the pilot to follow specified flight lines which were spaced 0.4 km (0.25 mi) apart, record all sightings on a computer using a global positioning system (gps), classify all animals as either adults or calves, and sex each adult using the presence of antlers or, in the case of animals without antlers, the presence or absence of a vulva patch. it was estimated that about one third of the moose were missed. crête et al. (1986) in mixedwood forest of quebec found that 27% of the moose were missed during early winter counts using helicopters. the access road to this study area was open at periodic intervals from 1996 to 2003 to allow logging trucks to remove wood. outside of these occasions, the road was closed using millstones and a locked gate (fig. 1) and, in 2002, bridges and culverts were pulled to further restrict access. the access road split at the southwest corner of the study area into a north and south section. a portion of the south road was vplowed in 2002. the equipment necessary to do this was attached to the back of a fiat hd 21 tractor and the roadbed ripped (figs. 2, 3, and 4). results the results of the aerial surveys are presented in table 1. the initial survey conducted in 1996 yielded a total of 28 moose. based on one-third being missed, approximately 37 moose were present in the closed area at the time of the 1996 survey (table 2). the estimated moose density was 0.6/km2 (1.5/mi2). during the 2003 survey, 107 moose were observed and again assuming one-third were missed, it is estimated that approximately 142 moose were present in march, 2003, which is an estimated density of 2.3/km2 (5.8/mi2). fig. 1. millstones and gate used to control access on happy lake road. response to access management crichton et al. alces vol. 40, 2004 90 during the survey, the remains of one moose were observed which appeared to be a poacher’s kill and 3 animals were seen that were not fully mobile and may have been wounded. natural resource officers have documented poaching in this area over the period of closure but it has not been extensive. access by poachers was done by breaching the gate (breaking locks) and millstones or by cutting trails through the bush adjacent to the gates. a few offenders (those shooting moose in the closed area) have been charged with hunting illegally within this area and these cases are currently before the courts. there have been transgressions of the gates during holiday seasons particularly at christmas when officers are on annual leave. in one case, hunting along the road to the happy lake area by a group of unidentified persons resulted in 21 moose being taken in 1 week. v-plowing is effective in prohibiting table 1. results of happy lake study area moose surveys – 1996/97 to 2002/03. access by trucks, snow machines and all terrain vehicles (atvs) (fig. 4). the cost of operating the tractor along with the attached v-plow was can $125/hour and a km of road can be done in anywhere from 4 to 12 hours depending on what is contained in the roadbed. the presence of large rocks will slow progress. therefore, the cost of doing a km will vary from can $500$1,500. this activity includes criss-crossing fig. 2. v-plow attached to back of fiat hd-21 tractor. note: survey not done in 1998/99. table 2. estimated wintering moose population and density in the happy lake study area. note: survey not done in 1998/99. year estimated population estimated density 1996/97 37 0.6/km2 (1.5/mi2) 1997/98 49 0.8/km2 (2.0/mi2) 1999/00 76 1.2/km2 (3.1/mi2) 2000/01 114 1.8/km2 (4.7/mi2) 2001/02 126 2.0/km2 (5.2/mi2) 2002/03 142 2.3/km2 (5.8/mi2) year bulls (%) cow s (%) calves (% ) u nknow n (% ) t ot al bulls/ calves/ 100 cow s 100 cow s 1996/97 4 (14.3) 7 (25) 8 (28.6) 9 (32.1) 28 57.1 114.3 1997/98 9 (24.3) 18 (48.6) 10 (27.0) 37 50 55.6 1999/00 18 (31.6) 19 (33.3) 15 (26.3) 5 (8.8) 57 94.7 78.9 2000/01 23 (26.7) 41 (47.7) 17 (19.8) 5 (5.8) 86 56.1 41.5 2001/02 40 (42.1) 41 (43.2) 14 (14.7) 95 97.6 34 2002/03 40 (38.0) 49 (46.7) 16 (15.2) 2 (1.9) 107 81.6 32.7 alces vol. 40, 2004 crichton et al. response to access management 91 fig. 3. fiat hd-21 tractor ‘ripping’ road with use of a v-plow. back and forth. the purchase price of such v-plows is about can $15,000 and adaptations had to be made so that it dug properly and did not skip. the operator advised that once perfected, there has been no breaking of bolts which he anticipated, and the equipment has functioned smoothly in all cases. discussion access to forested areas can be controlled to varying degrees by existing manitoba legislation namely, the crown lands act (chapter c340, regulation 145/91, section 3), the wildlife act (chapter w130, section 3 and section 2.1 of 351.87), the provincial parks act (chapter p20, section 27), as well as the workplace health and safety act (102/88 r). the effectiveness of this legislation can vary and the application is subject to various criteria. it is not a catch-all for access management fig. 4. results of ripping road bed to control access. and legal challenges may affect government’s ability to apply access controls in the future. wildlife values of access control access management on roads, i.e., controlling use of cars and trucks, must be given one of the highest priorities in order to minimize the impacts of forest harvesting and associated roads on wildlife, particularly big game such as moose, elk (cervus elaphus), and other species which may be impacted. this will enable manitoba (and other jurisdictions) to adhere to the conservation of biodiversity, to secure renewable resources for future generations, i.e., sustainability, and to meet the province’s fiduciary obligations to first nation peoples. the aforementioned necessitate the preparation of road management plans early in the overall planning process and must address issues such as location, type, longevity, mitigative measures, road retirement and rehabilitation, and resource values at risk. it is also important to evaluate access in adjacent operating areas. this information will permit an evaluation of the cumulative impact of road development and provide opportunity for mitigation of these impacts (i.e., road retirement in adjacent response to access management crichton et al. alces vol. 40, 2004 92 addressing only a subset of the multiple ecological impacts of roads, and is less satisfactory than outright closure and complete rehabilitation. options for consideration include minimizing road densities, locating roads away from important habitats, and controlling vehicular access as it relates to management of all recreational activities including hunting and trapping. further, the legal penalties are not severe enough and it does create tension between mc and first nation users. closures, which look good on paper, may not function as such on the ground – in some situations, the only effective technique may be to “rip” and wait for natural re-vegetation. this approach has the added advantage of returning the roadbed to productive forest at minimal cost. it is suggested that by exposing soil and reducing compaction this will facilitate re-vegetation. to enhance the success of road closures, an effective public education and communication program (developed by government and industry) relating to the rationale for closures along with effective enforcement must be part of any program. at the very least, a minimum of 1.6 km (1 mi) should be ripped as well as removing culverts and bridges. it is not a deterrent to rip only a few hundred metres of road as, although inconvenient, users of atvs will navigate over such obstacles for short distances. it also is not a deterrent to rip a few hundred metres of road and then leave the road intact and again rip another short section at some further distance. also, such vplowing should not be done on inclines where the potential for erosion is greater. it is speculated that the use of gates and millstones curtails approximately 95% of the traffic, however they are not always successful in restricting access to those ‘die hard’ individuals who view areas behind such obstacles as places where moose can be easily killed and/or their own private areas). alternatives such as the use of existing roads or portions thereof must be examined along with the need for all weather versus seasonal roads. there are 4 issues requiring attention namely: (1) access to cutting blocks which deals with location and road density; (2) public access to the roads constructed; (3) road closure, retirement, and roadbed/rightsof-way reclamation; and ( 4) access to logged areas for silvicultural purposes. these issues are not unique to manitoba. mc acknowledges that directives are required to control vehicular access to wildlife, to protect wildlife values, to ensure that unsustainable use patterns are not established, and that traffic prohibits regeneration on roadbeds. such directives will promote the department’s commitment to the goals of biodiversity and ecosystem management while at the same time sustain a viable timber industry. a first priority in managing access on forest roads must be to identify wildlife values at risk, mitigative measures to be employed, and to implement a timely road retirement program. such strategies must balance resource conservation against the need for legitimate use. travel may have to be restricted and/or prohibited on roads, which traverse the habitats of endangered species, species rich (including plants, neotropical migrants, etc.) areas, and/or which may result in over-exploitation of resident wildlife populations. the latter will enable some species (e.g., moose) to maximize their response to rejuvenated habitats as witnessed in the happy lake study area. wintering moose densities of 2.3/km2 (5.8/mi2), although only a portion of the entire gha, are the highest in any of manitoba’s ghas that are hunted and demonstrates what can be achieved with the co-operation of interested stakeholders and first nation peoples. methods to control access when evaluating options to deal with road-wildlife issues, each is a compromise alces vol. 40, 2004 crichton et al. response to access management 93 hunting grounds. the v-plowing curtails access by trucks and makes it extremely difficult for those on all terrain vehicles and snow machines to navigate such disturbed areas. making the disturbed corridor at least a kilometer long will function as a deterrent to those ‘die hards’. attempts to control vehicular access through legislation and gates do not always work and there will be those who have little or no intention of working co-operatively with government and other concerned stakeholders. costs and benefits some will suggest that the financial costs of ripping roads for long distances may be prohibitive, however, this must be balanced against the resource values that require protection over the long term and their contribution to ecosystem health and to the cultural and economic well being of each jurisdiction. we suggest that a cost of $500-$1,500/km is manageable when compared with the resource values which will be lost. the effectiveness of no hunting and road closures such as v-plowing, as well as the co-operation of all interest groups and communities, clearly illustrates what can be achieved in terms of moose density. however, it is acknowledged that certain elements of society do not appreciate the need for such proactive measures if these resources are to remain sustainable for future generations. an appreciation of this by all would not require management agencies to undertake these control measures, which can be expensive in terms of direct costs and staff time. although there has been pressure to open the happy lake area to hunting and, this will occur in 2003 as per the original agreement, efforts are being made to ensure that access to the area is rigidly controlled. some argue that the area should remain closed and being such a small area, it will have little impact on hunting and opportunity in the entire gha. it is our belief that those who support an opening of hunting clearly see it as an enhanced opportunity to harvest a moose with little effort and lack a long-term vision for moose management in this gha. in contrast, those seeking to maintain the area closed see it as an investment over the long term with the wintering moose population in the closed area functioning as the principle and those moose taken outside representing the interest to be used. what is not known at this time is how widely the wintering moose disperse during the snow free period; thus the consequences of protecting moose here may have wider implications. the committee for moose management clearly illustrates what can happen when first nations and stakeholders from different walks of life put aside real and perceived philosophical differences and work in a co-operative spirit for the wildlife resource and for future use of these resources. acknowledgements we acknowledge the futuristic thinking of members of the eastern region committee for moose management in proposing these closures and for their dedication to the work of the committee; those members of local first nation communities who have been supportive of the project and who view these management actions as a necessity to securing the moose resource for future use; and licensed hunters and environmental groups who have worked diligently to make the project successful. we also acknowledge the financial and in-kind support of the manitoba model forest and manitoba conservation. appreciation is extended to tembec pine falls operation for their financial and in-kind contribution to the eastern region committee for moose management and for paying for the v-plowing. response to access management crichton et al. alces vol. 40, 2004 94 references bjorge, r. r. 1984. habitat use by woodland caribou in west central alberta, with implications for management. pages 335-342 in w.r. meehan, t. r. merrell, jr., and t.a. hanley, editors. fish and wildlife relationships in old growth forests: proceedings of a symposium held in juneau, alaska, april 1982. american institute of fisheries research biologists. cameron, r. d., d. j. reed, j. r. dau, and w. t. smith. 1992. redistribution of calving caribou in response to oil field development on the arctic slope of alaska. arctic 45:338-342. crête, m., l.-p. rivest, h. jolicoeur, j.m. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose with observations made from fixed-wing aircraft in southern quebec. journal of applied ecology 23:751-761. ellison, g. w., a. g. rappoport, and g. m. reid. 1986. report of the caribou impact analysis workshop, arctic national wildlife refuge. report prepared for united states department of interior, united states fish and wildlife service. jalkotzy, m. g., p. i. ross, and m. d. nasserden. 1997. the effects of linear developments on wildlife: a review of selected scientific literature. report prepared for canadian association of petroleum producers, calgary, alberta, canada. nellemann, c., and r. d. cameron. 1996. effects of petroleum development on terrain preferences of calving caribou. arctic 49:23-28. shideler, r. t., m. h. robus, j. f. winters, and m. kuwada. 1986. impacts of human development and land use on caribou: a literature review. volume 1: a worldwide perspective. technical report 86. alaska department of fish and game, juneau, alaska, usa. singer, f. j., and j. r. beattie. 1985. the controlled traffic system and associated wildlife responses in denali national park. arctic 39:195-203. thiel, r. p. 1985. relationship between road densities and wolf habitat suitability in wisconsin. american midland naturalist 113:404-407. alces vol. 45, 2009 makarova and khokhlov – moose in murmansk 13 the status and management of moose in the murmansk region, russia olga a. makarova and anatoly m. khokhlov pasvik state nature reserve 184424, raiakoski, petchenga rajon, murmansk region, russia. abstract: the moose population in the murmansk region has changed considerably in the past century. moose appeared in the forest-tundra zones in the 1950s, occupied the ponoy river area in the 1960-1970s, and population growth occurred to the north of the forest zone along the tributaries and rivers flowing into the barents sea. some wintered in open tundra, but more commonly moose migrated between tundra and forested winter habitat. official harvests began in the 1950s and were managed by murmanskiy, a state owned company. a 5-year harvest ban was initiated in 1982 to recover the population; however, current harvest remains about a third of previous levels and the proportional harvest of calves and yearlings is higher. the current population is in good condition based upon weight and productivity data, occupies suitable winter habitat, and is not impacted by severe winter conditions. because the murmansk region is at the northern extent of moose range, management should focus upon regulated harvests, adequate population surveys, seasonal habitats and migratory corridors, the impact of harvest quotas and poaching, and the possible influence of global warming. alces vol. 45: 13-16 (2009) key words: alces alces, climate change, harvest, history, moose, murmansk, population dynamics, population recovery. the murmansk region is located in northwest russia on the kola peninsula that borders the barents sea and northernmost areas of finland and norway (fig. 1). it is characterized by unique ecosystems and habitats including tundra, belts of tundra-forest, and forest that largely influence the optimal habitat conditions and distribution of large mammals, specifically moose (alces alces). although moose were common in the murmansk region during the 20th century, their population in northwest russia has changed considerably in the past 100 years. moose were not common at the end of the 19th century, and only a small population was retained into the 1920s. they appeared in the area of the future lapland reserve in 1910, and their northern distribution was in the vicinity of the notozero, pulozero, and ponoy rivers in the 1940s. moose appeared in the forest-tundra zones in the 1950s and occupied the ponoy river area in the 1960-1970s. strong population growth was also documented to the north of the forest zone along the tributaries and rivers flowing into the barents sea. moose were reported to winter in open tundra, but typically used forested habitats. seasonal migrations to and from the tundra and forest habitats were noted at that time. moose were distributed on the kola peninsula near the settlements of borisoglebskiy, pechenga, the towns of kola and lyavozero, along the iokanga rover to the mouth of the sukhoy river, further north to the settlement of kanevka, and south to the white sea. in general, the area is “lacy” in character and in some places practically coincides with the southern border of the tundra (makarova 1996). all major moose wintering areas are along the northern border of the forest zone except the ponoy area. the area occupied by moose changes sharply twice a year; it is reduced by about half during winter and covers nearly the entire region during snowless seasons. such seasonal shifts are adaptive moose in murmansk – makarova and khokhlov alces vol. 45, 2009 14 for species occupying northern areas where resource availability and mobility are restricted seasonally. the murmansk region might be the largest area occupied by moose on the kola peninsula; this may be related to habitat and environmental changes associated with the recent, milder climate in this area. it is known that no moose were north of 64-65˚ northern latitude in the first half of the 19th century, but moose moved gradually to 69-70˚ northern latitude over the past 15 decades. similar northward expansion in moose range has occurred in other areas as well; however, the reason for such northern expansion is unclear but is probably not due to warming entirely. it is possible that moose will eventually reoccupy areas associated with their prehistoric range (vereshchagin et al. 1995). the moose of the kola peninsula are quite large with adult moose weighing >500 kg (makarova 1981, 1990, 1991) and harvested moose tend toward the largest in europe. investigation of osteological material indicated that a rather vital population of moose has formed in the northern range over a relatively short period of time; there is little evidence that habitat conditions are restrictive in this northern area (korablev and makarova 1993). the moose population number has varied measurably over the past 50 years in the murmansk region. aerial surveys were used initially to document population status in the middle of the 20th century. subsequent methods included both aerial surveys and winter track counts that provided a better estimation of population size and distribution. the largest population occurred in the late 1950s and early fig. 1. location of the murmansk region on the kola peninsula, russia. alces vol. 45, 2009 makarova and khokhlov – moose in murmansk 15 1960s, and an official harvest was initiated in 1952. murmanskiy, a state owned industrial company, was established in 1966 to manage this activity. on average, harvests were not less than 7% of the estimated population, or about 450 moose annually from 1967-1982. because of a sharp reduction in the estimated population, a ban on the official moose harvest was instituted for 5 years. this ban resulted in subsequent growth of the population within 2-3 years, and the harvest was reinstituted in 1988. despite the quota reduction, the population failed to return to the levels realized in the 1950-60s. the harvest rate dropped to only 2-3% of the population estimate, with annual harvests of about 150 moose. the official harvest was stopped again and murmanskiy was closed in the 1990s. licensed hunting continues, mainly for trophy bulls in the rutting period. the official harvest data provided an opportunity to compare the age structure of the population from the 1960-70s with that in the 2000s; substantial changes occurred over that period. harvests remained relatively high and consistent in both periods. calf moose were 1.6%, yearlings 6.3% and adults 92.1% of the harvest in 1966-1977; corresponding ratios were 16, 10, and 74% in 2000-2006. however, neither the fertility rate, remaining at 1.18 fetus/pregnant moose, or the proportion of cows with 1 (81%) or 2 fetuses (19%) have changed. conversely, the proportion of barren cows has risen from 37 to 50% for unknown reason and warrants investigation. bull moose have always been predominant in the harvest averaging 59% (range = 53-62%) in the period 1966-1977, and are now 62%. bulls also prevail in all age groups with 63% calves, 57% yearlings and 52% adults in 1966-1977, and 69, 57, and 62%, respectively, in 2000-2006. these ratios presumably reflect the population as a whole, although hunter preference for adult moose is likely. no substantial change in harvest weights has been detected. the current, average dressed weights of bulls and cows are 167 and 159 kg; corresponding weights from 1966-1977 were 168 kg and 163 kg. although dressed weights reported on licenses are not official, they should provide reasonable and reliable data for tracking weight trends, and be characteristic of the moose population and associated indices. the murmansk region is not uniform in its flora; forest and forest-tundra belts cover >50% of the region. moose migrate from the tundra to their typical winter habitat in forests during the hunting season (15 november-15 january). mass migration usually starts in late november ending in late january. harvest documents contain the most reliable information about harvest time and location of harvest. most harvest occurs toward the end of the season in january when moose have already lost some weight; the prime harvest areas are in southern forested areas. about half of the harvest quota was filled in the kandalaksha and terskiy districts in the 20th century, but central districts started playing a larger role in the harvest at the beginning of the 1970s, especially the lovozero district by 1976. the largest number of licenses was allotted in the lovozero, kola, and terskiy districts, and the highest harvests were registered there as well; 34, 32, and 15 moose, respectively. ironically, a major part of the tundra zone and part of the forest and forest-tundra belt are in the northwestern murmansk region. one speculation is that moose migration to, and occupation of that area, may be indicators of change in the northern forest zone associated with global warming. moose have been hunted continuously in the murmansk region for >100 years. according to all data, the population is in a good condition based upon weight and productivity, occupies suitable winter habitat, and is not impacted by severe winter conditions. the most important element of the management of this population is a regulated harvest based on accurate population estimates and moose in murmansk – makarova and khokhlov alces vol. 45, 2009 16 related harvest quotas, particularly in light of improved access and hunting techniques. there is increased concern about the impact of poaching and its effect on population management. we recommend banning the official moose harvest because no discernable increase in the population (4500-5000) has resulted from the minimum harvest quotas set in recent years. we also believe that protected natural areas in the murmansk region play an important role in maintaining the core of the moose population. this “net” of protected natural areas includes 3 state nature reserves and 11 protected territories that cover 7% of the region. because poaching occurs in the region, these areas provide sanctuary for moose, and anecdotal information indicates that moose move to these areas during hunting season. importantly, official inspectors on site not only protect these areas, but also promote study of moose and perform population surveys regularly, including winter track surveys. these data provide reliable information about the population size and distribution of moose, as well as the presence of predators and other management considerations. nature reserves play a special role in the ecology and management of moose in the murmansk region. they are located in the south, northwest, and north and include major migratory routes of moose. their regional distribution provides more and regular winter population surveys invaluable for managers responsible for establishing harvest quotas and moose hunting licenses. cooperation among scientists at different reserves enriches wildlife research resulting in improved ecological knowledge, better information for moose management and hunting, and conservation of the hunting fauna in general. given that moose in the murmansk region exist at the northern range of moose, it is necessary to embrace a management strategy that continues to adequately survey the population and its characteristics, its seasonal habitats and migratory corridors, the impact of harvest quotas and poaching, and the possible influence of global warming. references korablev, p. n., and o. a. makarova. 1993. ecological and morphogenetic analysis of moose populations in biosphere reserves. pages 200-220 in theory and practice of nature reserve management. moscow, russia. (in russian). makarova, o. a. 1981. moose in the murmansk region. pages 160-166 in ecology of ground vertebrates of the north-west ussr. petrozavodsk, russia. (in russian). _____. 1990. morphological characteristics of moose in the northern area. 3rd international moose symposium, syktyvkar, russia, 27 august-5 september. abstract only. _____. 1991. craniometrical characteristics of moose (alces alces l.) in northern areas. pages 143-147 in ecology of ground vertebrates. pertozavodsk, russia. (in russian). _____. 1996. seasonal moose migration in murmansk oblast, northwest russia. polar geography 20: 78-83. vereshchagin, n., i. kuzmina, and o. a. makarova. 1995. regarding formation of the moose area in kola peninsula. in proceedings of the first international mammoth conference, st. petersburg, russia, 16-21 october. abstract only. alces36_133.pdf alces37(2)_483.pdf alces vol. 46, 2010 shipley moose as a model herbivore 1 fifty years of food and foraging in moose: lessons in ecology from a model herbivore lisa a. shipley department of natural resource sciences, washington state university, pullman, washington 99164-6410, usa abstract: for more than half a century, biologists have intensively studied food habits and foraging behavior of moose (alces alces) across their circumpolar range. this focus stems, in part, from the economic, recreational, and ecosystem values of moose, and because they are relatively easy to observe. as a result of this research effort and the relatively simple and intact ecosystems in which they often reside, moose have emerged as a model herbivore through which many key ecological questions have been examined. first, dietary specialization has traditionally been defined solely based on a narrow, realized diet (e.g., obtaining >60% of its diet from 1 plant genus). this definition has not been particularly useful in understanding herbivore adaptations because >99% of mammalian herbivores are thus classified as generalists. although moose consume a variety of browses across their range, many populations consume 50-99% of their diets from 1 genus (e.g., salix). like obligatory herbivores, moose have demonstrated adaptations to the chemistry and morphology of their nearly monospecific diets, which precludes them from eating large amounts of grass and many forbs. new classifications for dietary niche suggest that moose fit on the continuum between facultative specialists and facultative generalists. second, moose have been the subject of early and influential models predicting foraging behavior based on the tradeoffs between quality and quantity in plants. subsequent models have predicted the size of stems selected by moose based on the tradeoffs between fast harvesting (large twigs) and quick digestion (small twigs). because of their size, moose require many hours to harvest food, often selecting large bites as browse density declines. finally, long-term monitoring of moose populations has provided evidence of how populations and communities are regulated. low reproductive rates and long-term population trends shaped by moose density and forage availability on isle royale suggest a strong bottom-up effect on moose populations. empirical data and simulation models suggest that moose may shape their own forage supply, influencing their community and their own populations, especially when large predators are scarce. likewise, predation is the primary factor affecting calf survival and thus moose populations in alaska, demonstrating the important role of top-down factors. moose will continue to provide a model for examining ecological questions such as tolerances for plant chemistry, what governs animal movements over landscapes, and reciprocal interactions between predation and reproduction. alces vol. 46: 1-13 (2010) key words: alces alces, diet selection, foraging behavior, moose, niche breadth, specialization, population regulation. for more than half a century, biologists have intensively studied food habits, foraging behavior, and nutrition of moose (alces alces) (table 1, fig. 1-3, reviews by gasaway and coady 1974 and schwartz 1992). this large body of research has arisen both because moose are of global interest occupying a circumpolar range spanning northern parts of north america, europe, and asia, and their large size, magnificent antlers, and fascinating behavior make them valuable for consumptive and non-consumptive recreation and subsistence (storaas et al. 2001, timmerman and rodgers 2005.) for example, a recent economic analysis estimated the annual harvest value of moose in alaska alone at nearly $364 million moose as a model herbivore shipley alces vol. 46, 2010 2 and viewing value at $62 million (northern economics, inc. 2006). besides research aimed directly at better management of wild and captive populations of moose, moose have emerged in the ecological literature as a model herbivore through which many key broad ecological processes have been examined. here, i will review examples from the past 50 years of how and why moose have been used to test and demonstrate ecological theory related to food and foraging. specifically i will highlight how research with moose has contributed to 3 dichotomies in herbivore ecology: 1) how to define niche breadth relative to specialist and generalist herbivores, 2) how herbivores trade off food quality and quantity when selecting diets, and 3) the roles of top-down and bottomup processes in regulating populations. dietary niche: specialist or generalist? the question of what governs the dietary niche has been the focus of community ecology for decades (e.g., hutchinson 1957, levins 1962, futuyma and moreno 1988, kawechi 1994, julliard et al. 2006). many articles have examined the question of why some herbivores consume a diverse diet consisting of many plant species, whereas others consume a very narrow diet (e.g., freeland and janzen 1974, westoby 1978, sorensen and dearing 2003). because dietary specialization is so common in herbivorous insects, specialists have been traditionally been defined as an animal consuming only 1 plant species (i.e., monophagous, crawley 1983). however, dietary specialization is rare in vertebrate herbivores, thus dearing et al. (2000) relaxed the definition of a specialist for herbivores to include animals consuming at least 60% of their diet from 1 plant genus. regardless, <1% of mammalian herbivores can be classified as a specialist using this definition (dearing et al. 2000, shipley et al. 2009), which limits its usefulness for understanding dietary strategies in most herbivores. moose provide a good example of the difficulties in assigning herbivores to a specialist-generalist category. in the literature they have been referred to as both a “generalist herbivore” (belovsky 1978) and a “specialist browser” (hagerman and robbins 1993). across much of the moose’s range in western north america, summer and winter diets consist of 75-91% willow (salix spp., fig. 1ae). likewise, diets in parts of eastern north america and sweden consist of primarily of 1 species/genus specific to location (fig. 2a-d). in contrast, in other areas moose consume a more diverse diet in which no single genus comprises >60% of their diet (fig. 3a-c). therefore, how moose are classified according to the traditional definition of specialization depends on the location and scale that their diet is measured. because assigning herbivores like moose to a specialization category based on their realized diet alone can be problematic, many have suggested characteristics of food plants and the forager that are consistent with dietary specialization. specialist herbivores are expected to compete well in habitats where large, predictable, mono-specific patches of chemically or physically defended foods occur that location season feeding (hr) ruminating (hr) reference central alberta winter 9.8 11 renecker and hudson 1989 centra alberta summer 10.3 6.5 renecker and hudson 1989 denali national park, alaska winter 4.8 9.6 risenhoover 1986 denali national park, alaska summer 7.2 7.2 van ballenberghe and miquelle 1990 rocky mountain national park, colorado summer 8.9 9.1 dungan et al. 2010 table 1. time spent foraging and ruminating by moose. alces vol. 46, 2010 shipley moose as a model herbivore 3 are avoided by other animals (freeland and janzen 1974, westoby 1978, crawley 1983, lawler et al. 1998, dearing et al. 2000, moore et al. 2004). because this food is normally less nutritious, these animals tend to be small with lower absolute energy requirements, or have relatively low mass-specific metabolic rates. the food often offers conspicuous stimuli for easy detection, which requires lower neural sophistication when selecting a diet, thus less energy invested towards brain tissue (smith 1979, bernays and funk 1999, martin and handasyde 1999). specialists are expected to have specialized anatomical, physiological, or behavioral adaptations for consuming their primary food, especially advanced and less expensive detoxification systems for a narrow range of plant chemicals (freeland and janzen 1974, crawley 1983, dearing et al. 2000, sorensen et al. 2005a). however, these adaptations reduce the range of plants, especially novel plants, that they can consume (futuyma and moreno 1988, berenbaum and zangerl 1994, sorensen et al. 2005b). as a result, specialists are difficult to feed in captivity (pahl and hume 1991), respond poorly to a changing environment, and are most likely to become vulnerable to extinction (fisher et al. 2003, smith 2008.) moose generally conform with many, but not all of these criteria. in particular, moose are adapted to foods that form the bulk of their diet. willow and certain other hardwood browse species contain linear condensed tannins that reduce protein digestibility (hagerfig. 1. examples of moose consuming a “specialist” diet of > 60% willow (salix spp.) in the summer in a) wyoming (mcmillan 1953), b) colorado (dungan and wright 2005), c) alaska (van ballenberghe et al. 1989), and during winter in alaska (risenhoover 1987) and british columbia (poole and stuart smith 2005). fig. 2. examples of moose consuming a “specialist” diet of > 60% of a plant genus on a variety of browse species in winter across its range, including a) scots pine (pinus sylvestris, shipley et al. 1998), b) balsam fir (abies balsamea, ludewig and bowyer 1985), c) 3 species of maple (acer spp., routledge and roese 2004), and d) birch (betula spp., especially b. papyrifera, thomas 1990). moose as a model herbivore shipley alces vol. 46, 2010 4 man and robbins 1993). moose produce specific salivary binding proteins particularly efficient at binding linear condensed tannins, whereas herbivores like mule deer (odocoileus hemionus) that consume a more diverse diet produce salivary binding proteins aimed at both branched and linear tannins (hagerman and robbins 1993). willows also produce salicylates, a bitter phenolic glycoside, that deters feeding by some herbivores, for example brushtail possums (trichosurus vulpecula) (markham 1971, edwards 1978, degabriel et al. 2010). likewise, conifers consumed in large amounts contain monoterpenes and other plant secondary metabolites that deter feeding by herbivores such as snowshoe hares (lepus americanus; rodgers et al. 1993) and rodents (murphy and linhart 1999) although little is known about how moose deal with large amounts of these plant chemicals, other animals that specialize on terpenes such as woodrats (neotoma spp.) and arboreal marsupials have adapted efficient and less expensive detoxification enzymes in the liver (boyle et al. 1999, sorensen et al. 2005a). moose have a relatively large liver for their size (hofmann and nygren 1992), thus an increased capacity to detoxify conifer browse with cytochrome p-450 enzymes (macarthur et al. 1991). the moose has adapted an unusually large amount of room between the nasals and premaxillae which has allowed the development of a long, muscular, prehensile nose with widely spaced nostrils (bubenik 2007). presumably this anatomy aids stripping of willow leaves to increase bite size and harvest rate. however, this long nose is a liability when consuming small bites of forbs. shipley et al. (1994) found that moose have the longest cropping time (min/bite) of 13 herbivores ranging from 0.01-500 kg. the anatomy of their nose allows moose to swallow underwater, enabling efficient consumption of aquatic plants that are avoided by most north american cervids (geist 1999). moose also resemble specialist herbivores because they are difficult to feed in captivity. the typical herbivore diet consisting of grainbased pellets supplemented with alfalfa or grass hay causes diarrhea, enteritis, and wasting in moose (schwartz et al. 1985, shochat et al. 1997). although general diets of herbivores are much higher in starch than a typical moose diet dominated by browse, moose do not lack enzymes (e.g., pancreatic alpha amylase and intestinal maltase) for digesting starch (schwartz et al. 1996, shochat et al. 1997). however, moose only thrive in captivity when fed large amounts of supplemental browse and aspen-based herbivore pellets (schwartz et al. 1985, shochat et al. 1997). therefore, other components of browse diets such as lignin, tannins, and salicin may contribute to the digestive health of moose. because many other herbivores fall in the gray area between a specialist and generalist, shipley et al. (2009) developed a specialization key designed to more accurately place a herbivore along the specialist-generalist continuum, thus accounting for a variety of fig. 3. examples of moose consuming a “generalist” diet where no plant genus constitutes > 60% of the winter diet in alaska (wixelman et al. 1998), b) ontario (cummings 1987), and british columbia (poole and stuart smith 2005). alces vol. 46, 2010 shipley moose as a model herbivore 5 dietary strategies and forming a framework for comparative studies. the key assigns the modifiers “obligatory” and “facultative” to the terms “specialist” and “generalist” based on 1) relative breadth of the animal’s realized niche and diet (what it eats), 2) relative breadth of the fundamental niche and available diet (what it could eat), 3) the extent of chemical or physical characteristics, termed “difficulty”, that make food items either low in value or unpalatable to most herbivores, and 4) relevant temporal and spatial scales at which diets and niche breadth are measured. obligatory specialists always consume a narrow diet of a difficult plant. they have unique adaptations that allow them to consume this plant that is generally abundant in their habitat, but these adaptations also tend to prevent them from expanding their diet as environmental conditions change. like the obligatory specialist, facultative specialists always have a consistently narrow, realized niche for difficult foods during at least 1 spatial or temporal scale such as winter, but because their fundamental niche is broader, they can expand their diet to include less difficult foods when environmental conditions allow, such as summer. like facultative specialists, the realized niche of the facultative generalist can change depending on the local conditions. however, they differ from facultative specialists in that their diet is more commonly broad, they focus on different plant species in different seasons or locations, and when their diet becomes narrow, they tend to focus on less difficult plants that are also consumed readily by other herbivores. finally, obligatory generalists always consume a mixed diet because they have a a low tolerance for difficult foods that precludes them from eating much of any difficult plant. therefore, based on these criteria, moose would fall on the continuum between the facultative specialist and the facultative generalist, because their diets consist of only 1 species and genus of moderately difficult plants in many areas, but their diet can expand or change across their range. diet selection: quality or quantity? the fibrous cell walls of plants are difficult for herbivores to digest, thus the nutritional quality and biomass of plants are usually inversely related (van soest 1984). therefore, herbivores must make tradeoffs when selecting diets, a process which forms the backbone of most models predicting diet selection for herbivores. many innovations in optimal foraging models have been designed for and tested with moose. one of the first and best-known optimal foraging models for mammalian herbivores was belovsky’s linear programming model (belovsky 1978). this model was based on the simple tradeoffs moose in northeastern usa make when choosing whether to consume deciduous leaves or aquatic plants. deciduous leaves are less fibrous and easier to harvest and digest than aquatic plants, but because many boreal forests are depauperate in sodium (belovsky and jordan 1981), forest plants typically have less sodium than aquatic plants. to effectively meet their daily sodium requirement, moose need to consume a minimal amount of aquatic plants. based on simple intake and digestion models, belovsky (1978) suggested that moose needed to consume either a large amount of aquatic plants or a moderate amount of deciduous plants to meet their energy requirements, but because of digestion limitations, could only consume a moderate amount of aquatic plants or a large amount of deciduous plants. his model, therefore, predicted that moose must consume a mixture of aquatic and terrestrial plants within a narrow range of possibilities. he suggested that the exact mixture a moose should consume depends on its goal – whether to maximize energy by consuming the minimal amount of aquatic plants possible with the maximum amount of deciduous plants subject to digestion limitations, or to minimize time spent feeding by consuming the minimum of moose as a model herbivore shipley alces vol. 46, 2010 6 both types of plants to provide more time for other life requisites. his data on diet composition of moose from isle royale national park in michigan fit most closely with the energy maximizing strategy. two decades later, moose were also the subject of several models designed to examine diet selection on a finer scale – what twig diameter a moose should select when foraging on deciduous browse in winter (vivås et al. 1991, kielland and osborne 1998, shipley et al. 1999). these models incorporated more mechanistic intake and digestion models to examine tradeoffs between consuming bites of a larger or smaller stem diameter. cropping stems at larger diameters allows moose to take larger bites (vivås et al. 1991, shipley et al. 1998, 1999), and in turn, taking larger bites allows moose to harvest food faster (risenhoover 1987, shipley and spalinger 1992, gross et al. 1993). however, smaller stem diameters have less fiber making them easier to crop, chew, and digest (shipley and spalinger 1992, kielland and osborne 1998, shipley et al. 1998). keilland and osborne (1998) and shipley et al. (1999) predicted the twig size that herbivores should select to maximize digestible energy/d based on mass-specific constraints on consumption and digestion based on specific architecture and chemistry of browse species. in sweden, moose selected twig diameters very consistent with the predictions of the optimal bite size model when fed 5 deciduous browse species varying in structure and chemistry in concentrated patches (shipley et al. 1999). likewise, the twig diameters that moose selected from feltleaf willow (salix alaxensis) in alaska (keilland and osborne 1998) and pubescent birch (betula pubescens) in norway (vivås et al. 1991) were predictable from tradeoffs in quality and quantity. further experiments also showed that moose perceive these tradeoffs between harvesting and digesting plants quickly, and modify their harvesting behavior as plant density changes. for example, moose consumed larger stem diameters as the size of patches of red maple (acer glabrum) stems declined and the distance between patches increased (shipley and spalinger 1995), and consumed proportionately more birch as density declined (vivås and sæther 1987). fast harvesting and digestion is particularly important for large herbivores like moose that spend on average >8 h/d each feeding and ruminating throughout the year (table 1). therefore diet choices that reduce the time spent in these activities allow moose more time for other life requisites such as raising young, avoiding predators, and thermoregulation. finally, many of the first spatially explicit individual-based foraging models (ibm) for large herbivores were built for moose (roese et al. 1991, moen et al. 1997, 1998). for example, moen et al. (1997) used ibm to examine how foraging rules affect emergent properties such as body mass and movement pathways of moose foraging across patchy and seasonally changing landscapes. landscapes consisted of grids with 1 m2 feeding stations containing bites of deciduous browse. quantity and quality of browse was updated seasonally and with herbivory, and animals moved in nested time steps according to foraging rules. moen et al. (1997) validated the emergent properties of their model with field and pen data, and found that simulated moose using optimal rules based, in part, on quantity and quality of forage had higher body mass and survival at the end of the year than moose foraging randomly. although different in approach, these linear programming, optimal bite size, and ibm models indicate that when choosing diets, moose seem to weigh the value of fast harvesting and fast digesting and select the diet that gives them the highest digestible energy per day, and in many cases are more sensitive to the effects of plant morphology on intake rate than plant chemistry on digestion (keilland and osborne 1998). alces vol. 46, 2010 shipley moose as a model herbivore 7 population regulation: top-down or bottom-up? moose populations provide excellent case studies for investigating the classic question in population ecology of whether populations are regulated top-down by predators or bottom-up by food. the theoretical implications of this question have been recognized and debated for at least 100 years since the beginning of the field of ecology (pimm 1991); however, the practical implications have become increasingly important as many large carnivores are either disappearing [e.g., lynx (lynx canadensis) in northwestern usa; koehler et al. 2008] or reappearing [e.g., successful recolonization of wolves (canis lupus) in the northern rocky mountains of the usa; oakleaf et al. 2006] on the landscape. almost 50 years ago, hairston et al. (1960) argued that “the world is green” because plants are more abundant than animals, thus herbivore populations must be controlled top-down by carnivores. they suggested that carnivores should compete for food because they lack predators to limit their populations, whereas because herbivore populations are limited by predators, they should not compete for food. assuming that herbivore populations are limited by predators, they should be unable to limit plant populations that in turn would compete for resources. therefore, removing carnivores should have a strong effect on herbivores, but removing herbivores should have little effect on plant densities. murdoch (1966) later suggested that the world is not green – instead plants are mostly “prickly and taste bad” – arguing that ecosystems are regulated bottom-up because physical and chemical defenses make plants largely inedible. therefore, herbivores are scarce and compete intensely for limited nutrients available in plants. as a consequence, predators are limited by competing for scarce herbivores. thus, removing either predators or herbivores has little effect on their food supply. finally, oksanen et al. (1981) suggested that whether ecosystems are limited from the top down or bottom up depends on the productivity of the ecosystem. they argued that extremely unproductive systems like deserts and tundra do not produce enough forage to support herbivores, let alone carnivores. in moderately productive ecosystems, plants are limited by herbivores, but plants support insufficient herbivore populations to support large numbers of predators. in very productive systems, such as rainforests, herbivores are limited by predators, but plants are not limited by herbivores as hairston et al. (1960) proposed originally. moose are considered a “classic textbook case” by ecologists for examining this debate (peckarsky et al. 2008), both because many moose populations live in remote areas with relatively intact ecosystems in which their large predators are still present, and because of the availability of unique long-term datasets such as that collected in isle royale national park for the last 50 years. on isle royale, bottom-up effects seem to predominate in the unique ecosystem of single plant (balsam fir, abies balsamea), large herbivore (moose), and large predator (gray wolves) largely unaltered by humans. vucetich and peterson (2004) found that annual production of balsam fir was 3 times more important in models predicting moose density over the last 50 years than was wolf density. in turn, wolf density was predicted both by moose density and balsam fir production. in areas with more diverse plants, herbivores, and predators and those that have been modified by humans through harvest, both bottom-up and top-down effects have been observed. for example, moose range in south-central alaska contains willow and other forage plants, an alternative large herbivore, and at least two abundant large carnivores [bears (ursus americanus, u. arctos) and wolves; testa 2004]. in a 4-year study, testa (2004) found female moose in poor body condition, with low twinning rates and delayed age of first reproduction, and a negative relationship between raising a calf moose as a model herbivore shipley alces vol. 46, 2010 8 successfully and producing twins the following year. these characteristics suggest that food (willow) plays a role in limiting the moose population. however, in the same study, testa (2004) found high calf and adult mortality from predation, resulting in low recruitment indicating that predation had a greater influence on population dynamics than nutrition. whether a large herbivore might control an ecosystem as a top-down influence may depend on its ability to escape higher top-down control by carnivore populations through large size, migration, or availability of alternative prey or predators (sinclair 2003). a vast body of literature (e.g., bergström and danell 1987, lozinov and kuznetsov 2002, morris 2002, persson et al. 2005, siipilento and heikkilä 2005, stolter 2008) suggests that in areas where populations of large carnivores are naturally or artificially low, moose can control their food sources from the top-down. a review (pastor and danell 2003) of moose across their circumpolar range concluded that as moose select the most nutritious parts of hardwood browse, they damage or remove the photosynthetic and meristematic tissues. many species of browse respond by growing more quickly and with less fiber and plant secondary metabolites. this, in turn, provides more high quality moose forage and an incentive for moose to re-browse the same plant. however, over the long term these plants grow more slowly, have lower survival, and are less competitive. furthermore, stolter (2008) found that browsed willows had fewer catkins and reduced reproductive output. it follows that in situations of overpopulation, preferred moose forages could be replaced with less nutritious plants, such as conifers, that grow and decompose more slowly and ultimately change the composition and reduce productivity of an ecosystem. overview moose are widespread, charismatic, and popular for recreation and subsistence, and thus have been the subject of much research and even a dedicated scientific journal (alces). additional characteristics of moose and the habitats in which they reside have caused moose to emerge as a model herbivore for testing broad ecological principles. first, moose often reside in simple and/or intact ecosystems that facilitate basic research. for example, in many boreal systems moose select from fewer than 10 species of available plants during winter (shipley et al. 1998, vucetich and peterson 2004). in addition, many habitats moose occupy have been protected by parks and reserves (e.g., yellowstone national park, rocky mountain national park, isle royale, denali national park), or are protected de facto by their remoteness. therefore, much of moose range still contains large carnivores and natural vegetation. second, compared with many small and wary herbivores, moose have proven to be surprisingly easy to observe both in the field and captivity. for example, risenhoover (1987), van ballenberghe and miquelle (1990), and dungan and wright (2005) were able to count bites from wild moose that tolerated their presence, and researchers such as renecker and hudson (1986) and shipley and spalinger (1992, 1995) hand-reared moose for foraging studies in semi-natural conditions. moose also leave conspicuous remnants of their foraging activity such as easy-to-see browsed twigs and fecal pellets. finally, insights into food and foraging are often most lucrative when studying large herbivores like moose that must spend most of their time foraging to satisfy their high forage requirements. with these desirable characteristics, moose will continue to provide a model for examining ecological questions such as tolerances for plant chemistry, what governs animal movements over landscapes, and reciprocal interactions between predation and reproduction. acknowledgments thanks to t. bowyer, k. danell, n. t. hobbs, j. forbey, b. moore, and d. e. spalalces vol. 46, 2010 shipley moose as a model herbivore 9 inger for their helpful insight on ecology and moose. references belovsky, g. e. 1978. diet optimization in a generalist herbivore, the moose. journal of theoretical population biology 14: 105-134. _____, and p. a. jordan. 1981. sodium dynamics and adaptations of a moose population. journal of mammalogy 62: 613-621. berenbaum, m. r., and a. r. zangerl. 1994. facing the future of plant-insect interaction research: le retour á la “raison d’être”. plant physiology 146: 804-811. bergström, r., and k. danell. 1987. effects of simulated winter browsing by moose on morphology and biomass of two birch species. journal of ecology 75: 533544. bernays, e. a., and d. j. funk. 1999. specialists make faster decisions than generalists: experiments with aphids. proceedings of the royal society of london b 266: 151-156. boyle, r. t, s. mclean, n. davies, w. foley, and b. moore. 1999. folivorous specialization: adaptations in the detoxification of the dietary terpene, p-cymene, in australian marsupial folivores. american zoologist 39: 120a. bubenik, a. b. 2007. evolution, taxonomy, and morphophysiology. pages 77-123 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. second edition. university press of colorado, boulder, colorado, usa. crawley, m. j. 1983. herbivory: the dynamics of animal-plant interactions. studies in ecology volume 10. blackwell scientific publications, oxford, england. cumming, h. g. 1987. sixteen years of browse surveys in ontario. alces 23: 125-155. dearing, m. d., a. m. mangione, and w. h. karasov. 2000. diet breadth of mammalian herbivores: nutrient vs. detoxification constraints. oecologia 123: 397-405. degabriel, j. l., b. d. moore, l. a. shipley, a. k. krockenberger, i. r. wallis, c. n. johnson, and w. j. foley. 2010. interpopulation differences in the tolerance of a marsupial folivore to plant secondary metabolites. oecologia, in press. dungan, j. d., and r. g. wright. 2005. summer diet composition of moose in rocky mountain national park, colorado. alces 41: 139-146. _____, l. a. shipley, and r. g. wright. 2010. activity patterns, foraging ecology, and summer range carrying capacity of moose in rocky mountain national park, colorado. alces 46: ##-##. edwards, w. r. n. 1978. effect of salicin content of palatability of populus foliage to opossum (trichosurus vulpecula). new zealand journal of science 21: 103-106. fisher, d. o, s. p. blombert, and i. p. e. owens. 2003. extrinsic versus intrinsic factors in the decline and extinction of australian marsupials. proceedings of the royal society of london b 270: 1801-1808. freeland, w. j., and d. h. janzen. 1974. strategies in herbivory by mammals: the role of plant secondary compounds. american naturalist 108: 269-289. futuyma, d. l., and g. moreno. 1988. the evolution of ecological specialization. annual review of ecology and systematics 19: 207-233. gasaway, w. c., and j. w. coady. 1974. review of energy and rumen fermentation in moose and other ruminants. naturaliste canadien 101: 227-262. geist, v. 1999. moose: behavior, ecology, conservation. voyageur press, stillwater, minnesota, usa. gross, j. e., l. a. shipley, n. t. hobbs, d. e. spalinger, and b. a. wunder. 1993. foraging by herbivores in food-concenmoose as a model herbivore shipley alces vol. 46, 2010 10 trated patches: tests of a mechanistic model of functional response. ecology 74: 778-791. hagerman, a., and c. t. robbins. 1993. specificity of tannin-binding proteins relative to diet selection by mammals. canadian journal of zoology 71: 628-633. hairston, n. g., f. e. smith, and l. b. slobodkin. 1960. community structure, population control and competition american naturalist 94: 421-425. hofmann, r. r., and k. nygren. 1992. mor-1992. morphophysiological specialization of the moose digestive system. alces supplement 1: 91-100. hutchinson, g. e. 1957. concluding remarks. cold spring harbor symposium on quantitative biology. 22: 415-427. julliard, r., j. clavel, v. devictor, f. jiguet, and d. couvet. 2006. spatial segregation of specialists and generalists in bird communities. ecology letters 9: 1237-1244. kawechi, t. j. 1994. accumulation of delete-1994. accumulation of deleterious mutations and the evolutionary cost of being a generalist. american naturalist 144: 833-838. kielland, k., and t. osborne. 1998. moose browsing on feltleaf willow: optimal foraging in relation to plant morphology and chemistry. alces 34: 149-155. koehler, g. m., b. t. maletzke, j. a. vaon kienast, k. b. aubry, r. b. wielgus, and r. h. naney. 2008. habitat fragmentation and the persistence of lynx populations in washington state. journal of wildlife management 72: 1518-1524. lawler, i. r., w. j. foley, b. m. eschler, d. m. pass, and k. handasyde. 1998. interspecific variation in eucalyptus secondary metabolites determines food intake by folivorous marsupials. oecologia 116: 160-169. levins, r. 1962. theory of fitness in a heterogeneous environment. 1. the fitness set and adaptive function. american naturalist 96: 361-373. lozinov, g. l., and g. v. kuznetsov. 2002. the impact of moose on ash productivity. alces supplement 2: 81-84. ludewig, h. a., and r. t. bowyer. 1985. overlap in winter diets of sympatric moose and white-tailed deer in maine. journal of mammalogy 66: 390-392. markham, k. r. 1971. a chemotaxonomic approach to the selection of opossum resistant willows and poplars for use in soil conservation. new zealand journal of science 14: 179-186. martin, r., and k. handasyde. 1999. the koala: natural history, conservation and management. australian natural history series. second edition. unsw press, sydney, australia. mcarthur, c., a. e. hagerman, and c. t. robbins. 1991. physiological strategies of mammalian herbivores against plant defenses. pages 103-114 in r. t. palo and c. t. robbins, editors. plant defenses against mammalian herbivory. crc press, boca raton, florida, usa. mcmillan, j. f. 1953. some feeding habits of moose in yellowstone park. ecology 34: 102-110. moen, r., j. pastor., and y. cohen. 1997. a spatially-explicit model of moose foraging and energetics. ecology 78: 505-521. _____, y. cohen, and j. pastor. 1998. linking moose population and plant growth models with moose energetics model. ecosystems 1: 52-63. moore, b. d., i. r. wallis, j. palá-paúl, j. j. brophy, r. h. willis, and w. j. foley. 2004. antiherbivore chemistry of eucalyptu cues and deterrents for marsupial herbivores. journal of chemical ecology 30: 1743-1769. morris, k. 2002. impact of moose on aquatic vegetation in northern maine. alces 38: 213-218. murdoch, w. w. 1966. population stability and life history phenomena. americ alces vol. 46, 2010 shipley moose as a model herbivore 11 naturalist 100: 5-11. murphy, s. m., and y. b. linhart. 1999. comparative feeding morphology of the gastrointestinal tract in the feeding specialist, sciurus aberti and several generalist congeners. journal of mammalogy 80: 1325-1330. northern economics, inc. 2006. the value of alaska moose. northern economics, inc. anchorage, alaska, usa. oakleaf, j. k., d. l. murray, j. r. oakleaf, e. e. bangs, c. m. mack, d. w. smith, j. a. fontaine, m. d. jimenez, t. j. mei-fontaine, m. d. jimenez, t. j. meier, and c. c. niemeyer. 2006. habitat selection by recolonizing wolves in the northern rocky mountains of the united states. journal of wildlife management 70: 554-563. oksanen, l., s. d. fretwell, j. arruda, and p. niemelä. 1981. exploitation ecosystems in gradients of primary productivity. american naturalist 118: 240-262. pahl, l. i., and i. hume. 1991. preferences for eucalyptus species of the new england tablelands and initial development of an artificial diet for koalas. pages 123-128 in a. k. lee, k. a. handasyde, and g. d. sanson, editors. biology of the koala. surrey beatty & sons and the world koala research corporation, sydney, australia. pastor, j., and k. danell. 2003. moosevegetation-soil interactions: a dynamicsystem. alces 39: 177-192. peckarsky, b. l., p. a. abrams, d. i. bolnick, l. m. dill, j. h. grabowski, b. lutitbeg, j. l. orrock, s. d. peacor, e. l. preisser, o. j. schmitz, and g. c. trussell. 2008. revisiting the classics: considering nonconsumptive effects in textbook examples of predator prey interactions. ecology 89: 2416-2425. persson, i.-e., k. danell, and r. bergström. 2005. different moose densities and accompanied changes in tree morphology and browse production. ecological applications 15: 1296-1305. pimm, s. l. 1991. the balance of nature? ecological issues in the conservation of species and communities. the university of chicago press, chicago, illinois, usa. poole, k. g., and k. stuart smith. 2005. fine scale winter habitat selection by moose in interior montane forests. alces 41: 1-8. renecker l. a., and r. j. hudson. 1986. seasonal foraging rates of free-ranging moose. journal of wildlife management 50: 143-147. _____, and _____. 1989. seasonal activity budgets of moose in aspen-dominated boreal forests. journal of wildlife management 53: 296-302. risenhoover, k. l. 1986. winter activity patterns of moose in interior alaska. journal of wildlife management 50: 727-734. _____. 1987. composition and quality of moose winter diets in interior alaska. journal of wildlife management 53: 568-577. rodgers, a. r., d. williams, a. r. e. sinclair, t. p. sullivan, and r. j. anderson. 1993. does nursery production reduce antiherbivore defences of white spruce? evidence from feeding experiments with snowshoe hares. canadian journal of forest research 23: 2358-2361. roese, j. h., k. l. risenhoover, and l. j. folse. 1991. habitat heterogeneity and foraging efficiency: an individual-based model. ecological modeling 133-143. routledge, r. g., and j. roese. 2004. moose winter diet selection in central ontario. alces 40: 95-101. schwartz, c. c. 1992. physiological and nutritional adaptations of moose to northern environment. alces supplement 1: 139-155 _____, d. l. harmon, k. j. hundertmark, c.t. robbins, and b. a. lintzeňich. 1996. carbohydrase activity in the pancreas and moose as a model herbivore shipley alces vol. 46, 2010 12 small intestine of moose and cattle. alces 32: 25-29. _____, w. l. regelin, and a. w. franzmann. 1985. suitability of a formulated ration for moose. journal of wildlife management 49: 137-141. shipley, l. a., s. blomquist, and k. danell. 1998. diet choices by free-ranging moose in relation to plant distribution, chemistry, and morphology in northern sweden. canadian journal of zoology 76: 1722-1733. _____, j. s. forbey, and b. d. moore. 2009. revisiting the dietary niche: when is a mammalian herbivore a specialist? integrative and comparative biology: doi: 10.1093/icb/icp051 _____, j. e. gross, d. e. spalinger, n. t. hobbs, and b. a. wunder. 1994. the scaling of intake rate of mammalian herbivores. american naturalist 143: 1055-1082. _____, a. w. illius, k. danell, n. t. hobbs, and d. e. spalinger. 1999. predicting bite size selection of mammalian herbivores: a test of a general model of diet optimization. oikos 84: 55-68. _____, and d. e. spalinger. 1992. mechanics of browsing in dense food patches: effects of plant and animal morphology on intake rate. canadian journal of zoology 70: 1743-1752. _____, and _____. 1995. influence of size and density of browse patches on intake rates and foraging decisions of young moose and white-tailed deer. oecologia 104: 112-121. shochat, e, c. t. robbins, s. m. parish, p. b. young, t. r. stephenson, and a. tamayao. 1997. nutritional investigations and management of captive moose. zoo biology 16: 479-494. siipliento, j., and r. heikkilä. 2005. the effect of moose browsing on the height structure of scots pine saplings in a mixed stand. forest ecology and management 205: 117-126. sinclair, a. r. e. 2003. the role of mammals as ecosystem landscapers. alces 39: 161-176. smith, m. 1979. behaviour of the koala, phascolarctos cinereus goldfuss, in captivity i. non-social behaviour. australian wildlife research 6: 117-129. smith, a. t. 2008. conservation of endangered lagomorphs. pages 297-315 in p.c. alves, n. ferrand, and k.hackländer, editors. lagomorph biology: evolution, ecology and conservation. springerverlag, berlin, germany. sorensen, j. s., and m. d. dearing. 2003. elimination of plant toxins by herbivorous woodrats: revisiting an explanation for dietary specialization in mammalian herbivores. oecologia 134: 188-194. _____, j. d. mclister, and m. d. dearing. 2005a. plant secondary metabolites compromise the energy budgets of specialist and generalist mammalian herbivores. ecology 86: 125-139. _____, _____, and _____. 2005b. novel plant secondary metabolites impact dietary specialists more than generalists (neotoma spp.). ecology 86: 140-154. stolter, c. 2008. intra-individual plant response to moose browsing: feedback loops and impacts on multiple consumers. ecological monographs 78: 167-183. storaas, t., h. gundersen, h. henriksen, and h. p. anreassen. 2001. the economic value of moose in norway a review. alces 37: 97-107. thomas, d. c. 1990. moose diets and use of successional forest in the canadian taiga alces 26: 24-29. testa, j. w. 2004. population dynamics and life history trade-offs of moose (alces alces) in south-central alaska. ecology 85: 1439-1454. timmerman, h. r., and a. r. rodgers. 2005. moose: competing and complementary values. alces 41: 85-120. alces vol. 46, 2010 shipley moose as a model herbivore 13 van ballenberghe, v. and d. g. miquelle. 1990. activity of moose during spring and summer in interior alaska. journal of wildlife management 54: 391-396. _____, _____, and j. g. maccraken. 1989. heavy utilization of woody plants by moose during summer at denali national park. alces 25: 31-35. van soest, p. j. 1984. nutritional ecology of the ruminant. cornell university press, ithaca, new york, usa. vivås, h. j., and b.-e. sæther. 1987. interactions between a generalist herbivore, the moose alces alces, and its food resources: an experimental study of winter foraging behavior in relation to browse availability. journal of animal ecology 56: 509-520. _____, _____, and r. andersen. 1991. optimal twig-size selection of a generalist herbivore, the moose alces alces: implications for plant-herbivore interactions. journal of animal ecology 60: 395-408. vucetich, j. a., and r. o. peterson. 2004. the influence of top-down, bottom-up, and abiotic factors on the moose (alces alces) population of isle royale. proceedings of the royal society of london b 271: 183-189. westoby m. 1978. what are the biological bases of varied diets? american naturalist 112: 627-631. wixelman, d. a., r. t. bowyer, and v. van ballenberghe. 1998. diet selection by alaskan moose during winter: effects of fire and forest succession. alces 34: 213-238. f:\alces\vol_39\p65\3929.pdf alces vol. 39, 2003 tõnisson and randveer – monitoring of moose in estonia 255 monitoring of moose-forest interactions in estonia as a tool for game management decisions jüri tõnisson1 and tiit randveer2 1center for forest protection and silviculture, rõõmu tee 2, 51013 tartu, estonia; 2estonian agricultural university, f.r. kreutzwaldi 5, 51014 tartu, estonia abstract: this paper reviews several big fluctuations in moose (alces alces) numbers and related problems in estonia during the last century. the biggest conflict appeared during the period 1960 – 1980, when the moose population achieved its highest density. the result of the overpopulation of moose was extensive forest damage. the establishment of a monitoring system and its acceptance by game management authorities at the beginning of the 1990s contributed to the improvement of the situation. the monitoring includes both the estimation of moose population parameters and estimation of moose influence on forest regeneration. current moose numbers match the optimal population level outlined in the estonian environmental strategy, approximately 10,000 animals, and forest damage has decreased. alces vol. 39: 255-261 (2003) key words: forest damage, monitoring, moose, population parameters, wolf estonia is a small country, rich in forests and bogs, situated on the shores of the baltic sea. almost 50% of its surface is covered with forest, providing moose (alces alces) with excellent natural habitat. at the same time, forestry is of great importance to the national economy. the forest is a resource shared by both man and moose, and sometimes conflicts arise. through the ages estonia has experienced several undesirable fluctuations of its moose population, a phenomenon shared throughout the baltic region (baleishis et al. 1998). maximum populations have been pleasing to hunters but disturbing to foresters and vice versa. it is now commonly understood that the most important means for avoiding such conflict is adequate information concerning both the moose population and the condition of the young forest. until recently, moose population data (official survey data) were based only on the reports of hunters and were very subjective. since 1994, the monitoring of the moose population has been financed both by the state budget and by the state center for environmental investments. development of moose population and conflicts the first conflict between hunters and forest owners in our territories started at the end of the 19th century, when the moose population was high. the landowners split into two camps: those who received a large share of their income from the timber market, and those who were passionate hunters. although the landowners owned the hunting rights, their efforts to regulate moose numbers were not successful. instead, the population was reduced drastically due to poaching by peasants and soldiers during the revolution of 1905 and the first world war (rootsi 1998). by 1924, only about 25 moose were estimated to exist in the newborn republic of estonia (teino 1939). several decades of peace between hunters and foresters followed. after wwii, a rapid increase in the moose population started in estonia and in neighboring regions (haagenrud et al. 1987, monitoring of moose in estonia – tõnisson and randveer alces vol. 39, 2003 256 smirnov 1987, baleishis et al. 1998). in estonia, the growth of the moose herd (fig. 1) occurred due to the following conditions: (1) extensive post-war clear-cuts and reforestation of clearings; and (2) a significant decrease in the number of the main natural enemy, the wolf (canis lupus). the unprecedented increase in the number of moose was not reflected in official statistics. at the end of the 1970s, when the moose population was probably bigger than ever, hunters counted about 0 5000 10000 15000 20000 25000 1 9 4 7 1 9 5 2 1 9 5 7 1 9 6 2 1 9 6 7 1 9 7 2 1 9 7 7 1 9 8 2 1 9 8 7 1 9 9 2 1 9 9 7 h a establishing of plantations clearcuts 0 200 400 600 800 1000 1 9 4 7 1 9 5 2 1 9 5 7 1 9 6 2 1 9 6 7 1 9 7 2 1 9 7 7 1 9 8 2 1 9 8 7 1 9 9 2 1 9 9 7 2 0 0 2 w o lv e s number of wolves 0 5000 10000 15000 20000 25000 1 9 4 7 1 9 5 2 1 9 5 7 1 9 6 2 1 9 6 7 1 9 7 2 1 9 7 7 1 9 8 2 1 9 8 7 1 9 9 2 1 9 9 7 2 0 0 2 m o o se number, counted by hunters calculated number harvest fig. 1. the size of the moose population in estonia in relation to the number of wolves and to the establishment of forest plantations and area of clearcuts. alces vol. 39, 2003 tõnisson and randveer – monitoring of moose in estonia 257 9,000 individuals annually (fig. 1). according to our calculations and the opinion of several competent hunters, there were probably at least twice as many moose at that time. otherwise, it would not have been possible to have killed an average of 4,800 moose during the hunting season (fig. 1). by the mid-1970s the high number of moose was coupled with a marked decrease in the area of young pine plantations. in 1965, there were 60,000 ha of 10-20 year old pine stands and 6,000 – 8,000 moose. by 1975, the area of similar stands was 33,000 ha and the moose numbers had at least doubled. obviously, the young pine stands suffered from severe damage. at the end of the 1970s, 21,000 ha of age-class i young pine stands were damaged by moose (örd and tõnisson 1986). in this situation, the only solution could have been the decisive reduction of the moose-population, but the hunting quota remained the same. as we know, a similar situation also developed in fennoscandia in the mid-1980s. authorities there reacted quickly and the problem was more or less solved (haagenrud et al. 1987). this was not the case in estonia. no state-organized monitoring system existed in those days, and hunting officials had to rely on hunters’ observations. the hunters were not interested in or motivated to have a higher quota due to an existing policy that dictated a large portion of hunted moose had to be handed over to the state. in addition, a state restriction on rifle ownership meant most hunters had to use shotguns, thus impacting their ability to bag a moose. as a result, hunters knowingly reported lower population numbers to the hunting officials. fearing forest damage by moose, foresters began to cultivate more spruce than pine, which resulted in a decrease in the area of young pine stands (paal 1996). at the end of the 1980s, more serious damage became evident in the form of bark peeling of middle-aged spruce (randveer and heikkilä, 1996). moose had always fed on spruce bark to a certain extent, but it had never been dangerous to spruce silviculture. now, the damage reached catastrophic dimensions in some regions. according to the inventory organized by the estonian forest protection service in 1991, serious bark stripping was found in 18% (12,800 ha) of middle-aged spruce stands. in order to decrease the forest damage, moose hunting was intensified considerably at the request of the forest administration. the largest numbers (6,589) of moose were shot legally in 1992. actually, by that time, the number of moose had already decreased and “the sons were punished for the sins of their fathers.” socio-economically, the early 1990s was a difficult period in the baltic states. as always, in hard times, the number of wolves rose abruptly (fig. 1) and poaching increased. thus, the cumulative effects of several factors, including increased hunting quotas, poaching, and predation, led to an undesirable decrease in the number of moose. monitoring the estonian moose population has been studied since the beginning of the 1960s. for a decade it was done by dr. harry ling, whose work has been summed up in the monograph “the structure and dynamics of the population of moose in estonian ssr” (ling 1977 a, b). among other things, he showed the inadequacy of official survey data. unfortunately, his recommendations were not applied in game management practice. in the 1990s a new method of monitoring moose was developed based on the experiences of estonia, the other baltic states, and fennoscandia. the monitoring method includes estimating moose numbers and other population parameters as well as estimating the impact of cervidae (mainly moose) on forest regeneration. monitoring of moose in estonia – tõnisson and randveer alces vol. 39, 2003 258 the main aim of monitoring is to estimate moose population density, sex, and age structure, annual growth rate, and some other parameters. since 1991, the estimation of the number of moose is made by analyzing hunting data coming from the hunters and checking them against results from winter pellet group counts from at least 6 permanent monitoring areas, following the methods provided by v. padaiga (1970) and v. chervonnõi (1973). some local hunters’ societies use this pellet-count method as a main tool to count moose. we used the pellet group counting method for the first time in lahemaa national park in the early 1980s. there, the population density estimated by this method was 2-3 times higher than the hunters‘ estimation (randveer 1986). it is quite probable that this was the case throughout estonia. in 1991, the two estimates were nearly identical but exhibited very different growth rates. we estimated the number of moose at 12,000 after 10-15 years of decline instead of an increase, as could have been erroneously deduced from the official survey data (fig. 1). since 1991, the game management authorities have used the monitoring data to set the moose harvest quota. other population parameters, estimated annually are sex distribution and proportion of calves in the population (based on 2,000 – 8,000 observations of moose every autumn), and age distribution (by studying 1,000 – 2,000 lower jaws of hunted adult moose). these data are considered when determining the harvest quota and its structure. the population, sex and age structure, and annual increase estimates are applied to a model developed in finland (nygren and pesonen 1993). the model predicts the size and composition of the moose population the next fall and recommends a harvest quota. additionally, three ways of evaluating the effect of moose on forest regeneration are examined annually: 1. browsing pressure in preferred summer habitats of moose is evaluated by determining the percentage of deciduous trees and shrubs (apart from alder) under 2 – 2.5 m with fresh browsing traces on every monitoring area in the last week of august or in september. we adopted this simple method on the recommendation of our latvian colleague a. prieditis, who has been using it for estimating the browsing pressure on summer habitats of cervidae since 1984 (prieditis 1996). 2. rumen contents of hunted moose are examined to determine the average content and frequency of occurrence of economically important conifers (see methods in tõnisson and mardiste 1996). during the years 1990–2001, 171-1,024 rumen contents were examined annually (4,682 total). 3. the damage done by moose during the previous year is measured in permanent survey plots in 114 middle-aged spruce stands and in 94 young pine plantations. as the survey plots were established in 1998 and 2000 respectively, the collected data are preliminary. results and conclusions both the counting of winter pellet groups and the counting based on the reports of hunters indicate that the average population density of moose has increased after 1995 and remained at an average of 4-5 moose per 1,000 ha (approximately 10,000 moose in estonia) during the last 2 years. the variability is great, ranging from 2.0 – 9.6 moose per 1,000 ha in 6 monitoring areas in the spring of 2002. the browsing pressure on summer habitats varied between 14 – 32.8 % in different monitoring areas during the period 1994 – 2001 (fig. 2). this indicates a low usage of alces vol. 39, 2003 tõnisson and randveer – monitoring of moose in estonia 259 summer habitats. according to a. prieditis (1996), a critical browsing level is attained if 50% of edible trees and shrubs have been browsed. he has also shown that the browsing pressure varied between 35 – 70% in different forestry districts in latvia in the 1980s. we suppose that the same occurred in estonia during this period. the occurrence of spruce bark in the rumen contents has decreased significantly during the last years compared to the beginning of the 1990s. the content of pine twigs and needles is diminishing as well (figs. 3 and 4). it seems generally that the current moose population density does not cause problems for forestry. however, conditions exist for a rapid rise in moose numbers and an increase in forest damage. in a sense, the present situation is similar to the post-war period. first, the intensity of cuttings has multiplied during the last few years (fig. 1) and the biomass of available browse is growing. second, in recent years one could again notice a rise in browsing of young pine and peeling of spruce bark, which has also been confirmed by the data from the analysis of the survey plots (figs. 5 and 6), and last, the number of wolves has declined (fig. 1). certainly, unlike the post-war years, nobody favors extermination of the 0 10 20 30 40 1 9 9 4 1 9 9 5 1 9 9 6 1 9 9 7 1 9 9 8 1 9 9 9 2 0 0 0 2 0 0 1 year % o f tr e e s b ro w se d max min fig. 2. variation of browsing pressure in 6 monitoring areas in estonia during the years 1994 – 2001. 0 1 2 3 4 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 co n te n t, % 0 5 10 15 20 o cc u rr e n ce , % content in october content in november occurrence fig. 3. spruce (picea abies) in rumen contents of hunted moose. 0 5 10 15 20 25 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 co n te n t, % 0 10 20 30 40 50 60 70 o cc u rr e n ce , % content in october content in november occurrence fig. 4. pine (pinus sylvestris) in rumen contents of hunted moose. monitoring of moose in estonia – tõnisson and randveer alces vol. 39, 2003 260 wolves. our aim is to maintain a stable population of approximately 150 wolves in our country, although such a small population probably cannot regulate the moose herd. we expect that in the next decade our monitoring system and collaboration with game management authorities will be put to the test. if we can foresee and avoid the next undesirable fluctuation, we will be confident that we are going in the right direction. acnowledgements we thank the center for environmental investments for financial support. references baleishis, r., p. bluzma, a. ornicans, and j. tõnisson. 1998. the history of moose in the baltic countries. alces 34:339-345. chervonnõi, v. v. 1973. the estimation of the number of moose using the method of counting the winter faecal pellet groups. trudõ okskogo gosudarstvennogo zapovednika 9:104–111. haagenrud, h., k. morow, k. nygren, and f. stalfelt. 1987. management of moose in nordic countries. swedish wildlife research supplement 1:635– 642. ling, h. 1977a. põdrapopulatsiooni struktuur ja dünaamika eesti nsv-s. populatsiooni süsteemse analüüsi katse. acta et commentationes universitatis tartuensis 407:15–125. . 1977b. põdrapopulatsiooni struktuur ja dünaamika eesti nsv-s. populatsiooni süsteemse analüüsi katse. acta et commentationes universitatis tartuensis 408:1–105. nygren, t., and m. pesonen. 1993. the moose population and methods of moose management in finland 1979 – 89. finnish game research 48:46–53. ö r d , a . , a n d j . t õ n i s s o n . 1 9 8 6 . p o v r e z d e n i j e l o s e m s o s n o v õ h m o l o d n j a k o v v e s t o n s k o i s s r i vozmoznosti umenshenija povrezdenii. metsanduslikud uurimused. xxi:7-25. paal, h. 1996. metsakultiveerimine eestis. akadeemilise metsaseltsi toimetised 5:5–27. padaiga, v. 1970. metodõ regulirovanija chislennosti olenjei v intensivnom lesnom hozjaistve. kaunas. 30 pp. prieditis, a. 1996. moose population and browsing level in the summer habitats. pages 245–249 in n. botev, editor. proceedings of the international union of game biologists 22nd congress, sofia, bulgaria, 4-8 september 1995. randveer, t. 1986. ob utchjote losja v estonii. metsanduslikud uurimused. xxi:50–55. , and r. heikkilä. 1996. damage caused by moose (alces alces l.) by b a r k s t r i p p i n g o f p i c e a a b i e s . scandinavian journal of forest re0 20 40 60 80 100 1999 2000 2001 2002 % spruce pine fig. 5. percentage of survey plots with freshly damaged trees. 0 10 20 30 1999 2000 2001 2002 % spruce pine fig. 6. percentage of freshly damaged trees on the survey plots. alces vol. 39, 2003 tõnisson and randveer – monitoring of moose in estonia 261 search 11:153–158. rootsi, i. 1998. mets, ulukid, inimene – suhted läbi aegade. teaduse ajaloo lehekülgi eestist. xii:194–207. smirnov, k. a. 1987. rolj losja v biotsenozah juznoi taigi. nauka, moskva. 113 pp. teino, j. 1939. jahipidamisest 1918 – 1938. eesti mets 1:33–37. tõnisson, j., and m. mardiste. 1996. põtrade talve–eelsest toitumisest. metsanduslikud uurimused. xxvii: 155–164. monitoring of moose in estonia – tõnisson and randveer alces vol. 39, 2003 262 alces vol. 46, 2010 child et al. vulnerability and antler regulations 113 potential vulnerability of bull moose in central british columbia to three antler-based hunting regulations kenneth n. child1, daniel a. aitken2, roy v. rea3, and raymond a. demarchi4 16372 cornell place, prince george, british columbia v2n 2n7, canada; 2college of new caledonia, 3330 22nd avenue, prince george, british columbia v2n 1p8, canada; 3natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia v2n 4z9, canada; 4934 khenipsen road, duncan, british columbia v9l 5l3, canada. abstract: antlers from bull moose (alces alces andersoni) harvested in the omineca subregion of central british columbia were submitted by hunters for inspection, measurement, and comparison by age in 1982-1989. after correcting for non-reporting bias, we examined the potential vulnerability of these moose (n = 1,886) to 3 antler-based hunting regulations currently advertised in british columbia: spike/fork (s/f), tripalm (tp), and 10 point (10pt). the s/f regulation put 15.9% of all bulls at risk, and the tp and 10pt regulations put 11.1% and 12.0% at risk, respectively. bulls with cervicorn antlers were at higher risk (41.3%) to the s/f regulation than the tp (1.4%) or 10pt (<1%) regulations. by contrast, bulls with palmicorn antlers were at low risk (5.4%) to the s/f regulation, but were at high risk to the tp (19.0%) and 10pt (17.1%) regulations. the s/f regulation focused harvest on yearlings, potentially exposing 46% of yearlings to harvest. the tp regulation exposed 20-40% of bulls older than 4.5 years of age; whereas, the 10pt regulation exposed 40-60% of bulls >7.5 years of age to harvest. maximum spread and shaft circumferences of antlers were significantly smaller for yearlings at risk to the s/f regulation than for their same aged counterparts not at risk. distance between the innermost points on the brow palm was significantly larger for yearlings at risk to the s/f regulation than for yearlings not at risk. maximum spread, shaft circumference, palm height, and width were all significantly greater for bulls at risk to the tp and 10pt regulations than for those not at risk. distance between the innermost points on the brow palms was significantly smaller for bulls at risk to tp and 10pt regulations than for those not at risk. these findings suggest that yearling bulls with smallest antlers are most at risk to harvest by the s/f regulation, whereas the largest antlered bulls are most at risk to harvest by the tp and 10 pt regulations. the consequences of this directed selection of bull moose by antler-based hunting regulations on the breeding biology, population genetics, and fitness of moose requires further study. alces vol. 46: 113-121 (2010) key words: alces alces, harvest risk, hunting, social class, spike/fork, tripalm, yearling bull, 10 point. moose (alces alces andersoni) hunting in british columbia has traditionally been oriented toward the male segment of the population. in the long term, bull-only seasons may lead to age and sex imbalances that affect the growth, productivity, and ability of moose populations to sustain management and recreational objectives (baker 1975, demarchi and hartwig 2008). consequently, restrictive hunting seasons with increasing complexity of regulations and hunter dissatisfaction result (hatter 1994, child 1996, hatter 1999). traditional practices of harvesting may act as an evolutionary force that can chalvulnerability and antler regulations child et al. alces vol. 46, 2010 114 lenge conservation goals for wildlife and may impair both the health and genetic diversity of a species (boer 1991, darimont et al. 2009). selective harvesting of large antlered males over the long term can alter genetics (laurian et al. 2000) by negatively impacting those alleles that underpin fitness (hundertmark and bowyer 2004). such changes can be irreversible if harvesting systems continue to target the larger individuals in a population (van ballenberghe 2004, paquet 2009). in this study we examined the potential vulnerability of bull moose harvested in the 1980s from the omineca sub-region of the central interior of british columbia (fig. 1) to 3 antler-based hunting regulations (fig. 2) practiced in the province: spike/fork (s/f), tripalm (tp), and 10 point (10pt). we evaluated the potential vulnerability of these bull moose specific to age class, social class, and antler characteristics. methods morphometry of moose antlers in the omineca sub-region of british columbia was described by child et al. (2010). our data set was comprised of 1,686 sets of antlers from moose harvested in 1982-1989. of these, 1,586 sets were submitted by successful limited entry hunters (leh) for mandatory inspection; another 100 sets were submitted voluntarily by non-leh hunters (i.e., hunters not possessing an leh authorization who hunted in an open season for s/f bulls). we assumed that the leh harvest was taken randomly from the population (schwartz et al. 1992), whereas the non-leh harvest of s/f bulls was taken primarily from the yearling component (hatter and child 1992). concerns regarding harvest bias against s/f bulls by leh hunters, as well as under reporting by non-leh hunters (hatter and child 1992, hatter 1993), are reflected in the lack of yearling bulls in the reported age distribution (child et al. 2010). to correct for non-reporting bias, we increased the number of s/f bulls until their vulnerability to the s/f regulation was 46%. this adjustment matched the vulnerability reported by hatter (1993) and resulted in a hypothetical sample (hereafter considered to be the population) of 1,886 bull moose for study. from the population (n = 1,886), we reported age of bulls potentially at risk to harvest when subjected to s/f, tp, and 10pt regulations (fig. 2). we also report the proportions of bulls at risk by age class, social class, and antler form. for the analysis, we used the social classes described by bubenik (1971): yearlings (1.5 years), teens (2.5-3.5 years), primes (4.5-11.5 years), and seniors (>12.5 years). antler forms were described by child et al. (2010) as cervicorn (pole type) and palmicorn (split palm or full palm). we separated those with palmicorm antlers as split palm and full palm antlers, and analyzed harvest risk to the regulations for each group. proportions were calculated only if there were at least 5 bull moose in any age class, social class, or category of antler form. the maximum spread, maximum height, fig. 1. the omineca sub-region (region 7a) of the british columbia ministry of environment in central british columbia (from british columbia hunting and trapping regulations and synopsis, 2008). alces vol. 46, 2010 child et al. vulnerability and antler regulations 115 palm width, shaft circumference, and distance between the inner most points on the brow palms (child et al. 2010) of yearling bulls at risk under the s/f regulation were compared ( t-test, p = 0.05) with the same morphometrics for yearlings not at risk. similarly, we compared the same morphometrics of antlers from bulls >2.5 years old at risk to the tp and 10pt regulations to bulls of similar age not at risk. we treated yearlings separately because this is the only age class subject to high harvest risk when exposed to the s/f regulation. conversely, we separated all bulls >2.5 years old because they are at risk when exposed to the tp and 10pt regulations. we used levene’s test (milliken and johnson 1984) for equality of variances and then used the t-test for equal or unequal variances as appropriate. age-specific mean maximum antler spreads of bulls in the population were compared graphically with mean maximum antler spreads of bulls at risk to each of the regulations. age-specific mean maximum spreads were calculated if there were at least 5 bull moose in the age class. results harvest risk of bull moose exposed to s/f regulation bulls in our study (n = 1,886) ranged from 1.5-19.5 years with a mean of 3.9 ± 2.7 years (fig. 3); nearly 16% were at risk to the s/f regulation. the mean age of bulls at risk was 1.9 ± 1.2 years of age (n = 100); 81% were yearlings and the oldest was 9.5 years. agespecific vulnerability declined from 46.0% for yearlings to <5.0% for moose >2.5 years (fig. 4). by social class, 46.2% of yearlings, 6.0% of teens, and 2.4% of primes were at risk. sample size was insufficient to determine the fig. 2. antler-based regulations for bull moose in british columbia (from british columbia ministry of environment hunting and trapping regulations and synopsis, 2008). 0 100 200 300 400 500 600 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5 10.5 11.5 12.5 13.5 14.5 age (years) n u m b e r o f m o o s e fig. 3. age distribution of the adjusted population of bull moose corrected for yearling reporting bias. note: due to sample size (n <5) no data were plotted for bulls >15.5 years old. vulnerability and antler regulations child et al. alces vol. 46, 2010 116 proportion of senior bulls at risk. additionally, when considering antler form, 41.3% of bulls with cervicorn antlers and 5.4% of bulls with palmicorn antlers, including both split palm (5.4%) and full palm antlers (5.3%), were at risk (table 1). both yearling and 2.5 year old bulls at risk had mean maximum antler spreads that were smaller than the mean maximum antler spread calculated for all bulls of similar age (fig. 5). the maximum spread and shaft circumferences of antlers for yearling bulls at risk were smaller (p <0.001) than those of yearlings not at risk. maximum antler height, palm width, and distance between the inner most points on the brow palms of yearlings were not different (p >0.05) between those yearlings at risk and those not at risk (table 2). harvest risk of bull moose exposed to tp regulation of the 1,886 bull moose in the sample population, 12% were at risk to the tp regulation. the mean age of bulls exposed to the tp regulation was 6.3 ± 3.0 years (n = 227); bulls 1.5-19.5 years old were at risk. vulnerability increased linearly from 5% at 2.5 years to 35% at 7.5 years, then fluctuated between 25-45% to 13.5 years (fig. 4). sample size was insufficient to determine the proportion of yearlings at risk, but 7.0% of teens, 25.9% of primes, and 38.2% of seniors were at risk. by antler form, 1.4% with cervicorn antlers and 19.0% with palmicorn antlers were at risk, including both split palm (18.3%) and full palm antlers (25.4%, table 1). mean maximum antler spread for each age class at risk was generally larger than the mean maximum antler spread calculated for the same age class in the population (fig. 5). antlers of bulls at risk had larger (p <0.001) maximum spread, height, palm width, and shaft circumference, and smaller (p <0.001) distance between the inner most points on the brow than bulls not at risk (table 2). 0 10 20 30 40 50 60 70 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5 10.5 11.5 12.5 13.5 age (years) p ot en tia l v ul ne ra bi lit y (% ) fig. 4. potential vulnerability of bull moose by age to 3 antler-based regulations. broken line = s/f, gray line = tp, and black line = 10pt. note: no data points were plotted for s/f regulation ages >5.5 years old, for tp regulation ages 1.5, 12.5, and 14.5 years and older, and for the 10pt regulation ages 1.5, 2.5, and 14.5 years and older due to insufficient sample size (n <5). regulation social class % antler form % s/f population 15.9 cervicorn 41.3 yearling 46.2 palmicorn 5.4 teen 6.0 split palm 5.4 prime 2.4 full palm 5.3 senior nc tp population 12.0 cervicorn 1.4 yearling nc palmicorn 19.0 teen 7.0 split palm 18.3 prime 25.9 full palm 25.4 senior 38.2 10pt population 11.1 cervicorn nc yearling nc palmicorn 17.1 teen 1.1 split palm 17.5 prime 29.7 full palm 14.0 senior 44.1 table 1. potential vulnerability (%) of bull moose subjected to 3 antler-based regulations (s/f, tp, and 10pt) by social class and antler form for the population. note: nc = % not calculated (n <5). alces vol. 46, 2010 child et al. vulnerability and antler regulations 117 harvest risk of bull moose exposed to 10pt regulation of the 1,886 bull moose in the sample population, 11% were at risk to the 10pt regulation. the mean age of bulls at risk was 7.7 ± 2.7 years (n = 210), ranging from 2.5-15.5 years old. age-specific vulnerability increased linearly from <5% for bulls 3.5 years old, to about 50% at 8.5 years, then fluctuated between 40-65% to 13.5 years (fig. 4). sample size was insufficient to determine the proportion of bulls at risk that were <2.5 years or >13.5 years old. by social class, 1.1% of teens, 29.7% of primes, and 44.1% of seniors were at risk. by antler form, 17.1% with palmicorn antlers were at risk, including both split palm (17.5%) and full palm (14.0%, table 1). sample size was insufficient to determine the proportion of bulls with cervicorn antlers that were vulnerable to the 10pt regulation. the age-specific, mean maximum antler spread for each age class at risk was generally larger than the mean maximum antler spread calculated for the same age class in the population (fig. 5). bulls at risk had larger (p <0.001) sized antlers by maximum spread, height, palm width, and shaft circumference, and a smaller (p <0.001) distance between the inner most points on the brow palms than bulls not at risk (table 2). discussion assessment of the harvest risk of bull moose revealed that most bulls at risk to the s/f regulation were yearlings and those yearlings at risk had smaller antlers than yearlings not at risk. on the other hand, when subjected to the tp regulation, a large proportion of prime and senior bulls were at risk; when subjected to the 10pt regulation, risk to prime and senior bulls was higher still. importantly, bulls at risk to either the tp or 10pt regulations had larger antlers (i.e., greater spread, width, height, number of points) and narrower distance between the innermost points on the brow palms than bulls not at risk to these regulations. generally, bull moose with cervicorn antlers were at greatest risk to harvest under the s/f regulation, and bulls with palmicorn antlers were at high risk to both the tp and 10pt regulations. bull moose with split palm antlers were similarly vulnerable to the tp and 10pt regulations, whereas bulls with full palm antlers were at higher risk to the tp regulation. antler size and symmetry reflects social status and fitness in cervids (markusson and folstad 1997, pelabon and van breukelen 1998, ditchkof et al. 2001, malo et al. 2005, vanpé et al. 2007) including moose (bubenik 1983, solberg and saether 1993, 1994, bubenik 1998). prime bulls carry the largest antlers (gasaway et al. 1987) and their high numbers on rutting areas are required for optimal breeding and productivity (bubenik 1983, aitken and child 1991, 1992, solberg et al. 2002, saether et al. 2003). the combination of antler size, form, and symmetry that cows recognize when selecting mates is not fully understood (solberg and saether 1993, bowyer 400 500 600 700 800 900 1000 1100 1200 1300 1400 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5 10.5 11.5 12.5 13.5 age (years) m ea n m ax im um s pr ea d (m m ) fig. 5. age-specific, mean maximum spread of antlers of bull moose in the population compared to those subjected to the 3 antler-based regulations. thin solid line with open circles = population, broken line = s/f, gray line = tp, and thick black line = 10pt. no data points were plotted for s/f regulation for ages 3.5 years and older, for tp regulation ages 1.5, 12.5, and 14.5 years and older, and for the 10pt regulation for ages 1.5, 2.5, and 14.5 years and older due to insufficient samples (n <5). vulnerability and antler regulations child et al. alces vol. 46, 2010 118 et al. 2001). however, prolonged harvests of large antlered bulls and/or those with palmated brow structures may, over time, reduce genetic variability and cause an irreversible loss of alleles specific to antler features (hundertmark et al. 1993, hundertmark and bowyer 1998, bowyer et al. 2002, hundertmark and bowyer 2004, van ballenberghe 2004). moose hunting focused on bulls often results in age and sex imbalances that can lead to a scarcity of mature breeding bulls. hunting regimes should ideally produce sex and age-specific mortality patterns similar to those occurring naturally, and should maintain demographic structures conducive to natural breeding patterns (harris et al. 2002) in order to ensure social well-being (bubenik 1971, 1983), genetic variability (ryman et al. 1981, hartl et al. 1991, hundertmark et al. 1993, coltman et al. 2003), and high productivity regulation morphometric mean value p at risk not at risk s/f ms 569 ± 122, 241 667 ± 95, 65 <0.001 mhl 289 ± 140, 116 324 ± 97, 27 0.122 mhr 287 ± 143, 98 302 ± 102, 24 0.567 pwl 111 ± 63, 199 104 ± 39, 22 0.610 pwr 114 ± 67, 203 102 ± 37, 22 0.442 scl 110 ± 16, 270 124 ± 17, 75 <0.001 scr 110 ± 18, 266 125 ± 18, 75 <0.001 dipb 431 ± 60, 189 414 ± 82, 28 0.295 tp ms 1030 ± 181, 244 831 ± 194, 1536 <0.001 mhl 676 ± 161, 154 489 ± 187, 769 <0.001 mhr 658 ± 160, 146 474 ± 183, 670 <0.001 pwl 227 ± 63, 241 149 ± 61, 1118 <0.001 pwr 222 ± 57, 242 149 ± 61, 1113 <0.001 scl 168 ± 21, 251 144 ± 26, 1634 <0.001 scr 167 ± 21, 250 144 ± 26, 1566 <0.001 dipb 323 ± 86, 216 383 ± 85, 1303 <0.001 10pt ms 1142 ± 154, 235 815 ± 174, 1545 <0.001 mhl 771 ± 123, 124 481 ± 175, 799 <0.001 mhr 756 ± 109, 107 469 ± 173, 709 <0.001 pwl 259 ± 52, 196 147 ± 56, 1163 <0.001 pwr 254 ± 46, 197 146 ± 56, 1158 <0.001 scl 180 ± 18, 241 142 ± 24, 1644 <0.001 scr 179 ± 18, 238 142 ± 24, 1578 <0.001 dipb 319 ± 92, 213 384 ± 84, 1306 <0.001 table 2. summary of morphometric measurements (mean ± sd, n) and statistical significance of differences between bull moose at risk and those not at risk to the 3 antler-based regulations. comparisons (t-tests) for the s/f regulation were only made for 1.5 year-old bulls, whereas comparisons for both the tp and 10pt regulation were made for bulls 2.5 years and older (see methods). ms = maximum spread, mhl = maximum height left side, mhr = maximum height ride side, pwl = palm width left side, pwr = palm width right side, scl = shaft circumference left side, scr = shaft circumference right side, and dipb = distance between the innermost points on brow. alces vol. 46, 2010 child et al. vulnerability and antler regulations 119 (aitken and child 1991, timmerman 1991, aitken and child 1992, schwartz 1998). moreover, since a high proportion of mature breeders in the population prevents declines in population fitness (ferer et al. 2003), a harvest strategy that reduces pressure on older, larger antlered males may be the most prudent. open seasons or limited entry hunting (leh) systems without antler restrictions are generally thought to randomize bull harvests (child and aitken 1989, schwartz et al. 1992) and thereby ensure a normal age distribution. antler-based regulations, on the other hand, direct hunters to selectively harvest bulls by antler characteristics that may have either beneficial or harmful consequences depending on the particular antler restriction (harris et al. 2002). the results of this study suggest that the s/f regulation targets mainly young bulls with the smallest antlers whereas the tp and 10pt antler regulations target bulls with the largest antlers across all age classes. it is important to understand the harvest risk of bull moose to antler-based regulations because genetic effects are suspected, if not likely (hartl et al. 1991, hundertmark et al. 1993, coltman et al. 2003), and normal behavior (bubenik 1987, 1998) and reproductive patterns (schwartz 1998, timmermann 1991) may be disrupted. because of these negative consequences associated with over harvest of the largest bulls in a population, we advocate further monitoring and study of harvest impacts associated with antler-based hunting regulations. acknowledgements we thank sean barry for his meticulous attention to detail in measuring and recording the forms of all inspected antlers, and to ken fujino and sean barry of the wildlife branch who prepared and aged the tooth samples. a special thanks to gerry kuzyk for release of the antler data records and to the many hunters who willingly submitted their antlers for inspection. we also thank nic larter and an anonymous reviewer for their comments on an earlier draft of the manuscript. references aitken, d. a., and k. n. child. 1991. gross productivity of moose in the central interior of british columbia. proceedings of the 1991 moose harvest management workshop, kamloops, british columbia wildlife branch, british columbia ministry of environment, victoria, british columbia, canada. _____, and _____. 1992. relationship between in-utero productivity of moose and population sex ratios: an exploratory analysis. alces 28: 175-187. baker, r. a. 1975. biological implications of a bull moose-only hunting regulation in ontario. proceedings of the north american moose conference and workshop 11: 464-476. boer, a. h. 1991. hunting: a product or tool for wildlife managers? alces 27: 74-78. bowyer, r. t, k. m. stewart, j. g. kie, and w. c. gasaway. 2001. fluctuating asymmetry in antlers of alaskan moose: size matters. journal of mammalogy 82: 814-824. _____, _____, b. m. pierce, k. j. hundertmark, and w. c. gassaway. 2002. geographical variation in antler morphology of alaskan moose: putative effects of habitat and genetics. alces 38: 155-165. bubenik, a. b. 1971. social well-being as a special agent of animal sociology. international conference on the behavior of ungulates and its relation to management. calgary, alberta, canada. _____. 1983. the behavioral aspects of antlerogenesis. pages 389-449 in r. d. brown, editor. antler development in cervidae. caesar kleberg wildlife research institute. texas a&i university, kingsville, texas, usa. _____. 1987. behaviour of moose (alces alces) of north america. swedish wildlife vulnerability and antler regulations child et al. alces vol. 46, 2010 120 research supplement 1: 333-366. _____. 1998. behavior. pages 173-221 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d.c., usa. child, k. n. 1996. moose harvest management in british columbia: regulation simplification and strategy harmonization. wildlife branch, ministry of environment, lands and parks, victoria, british columbia, canada. _____, and d. a. aitken. 1989. selective harvests, hunters and moose in central british columbia. alces 25:81-97. _____, d. a. aitken, and r. v. rea. 2010. morphometry of moose antlers (alces alces andersoni) in central british columbia. alces 46: 123-134. coltman, d. w., p. o’donougue, j. t. jorgensen, j. t. hogg, c. strobeck, and m. festa-blanche. 2003. undesirable evolutionary consequences of trophy hunting. nature 426: 655-658. darimont,c. t., s. m. carlson, m. t. kinnison, p. c. paquet, t. e. reimchen, and c. c. wilmers. 2009. human predators outpace other agents of trait change. proceedings national academy of sciences, usa. 106: 952-954. demarchi, r. a., and c. l.hartwig. 2008. towards an improved moose management strategy for british columbia. habitat conservation trust fund report cat070-0325. victoria, british columbia, canada. ditchkof, r. l., r. l. lochmiller, r. e. masters, w. r. starry, and d. m. leslie jr. 2001. does fluctuating asymmetry of antlers in white-tailed deer (odocoileus virginianus) follow patterns predicted for sexually selected traits? proceedings biological sciences 268: 891-898. ferrer, m., v. penteriani, j. balbontín, and m. pandolfi. 2003. the proportion of immature breeders as a reliable early warning signal of population decline: evidence from the spanish imperial eagle in doñana. biological conservation 114: 463-466. gasaway, w. c., d. j. preston, d. j. reed, and d. d. roby. 1987. comparative antler morphology and size of north american moose. swedish wildlife research supplement 1: 311-325. harris, r. b., w. a. wall, and f. w. allendorf. 2002. genetic consequences of hunting: what do we know and what should we do? wildlife society bulletin 30: 634-643. hartl, g. b., g. lang, f. klein, and r. willing. 1991. relationship between allozyme heterogeneity and morphological characters in red deer (cervus elaphus), and the influence of selective hunting on allele frequency distribution. heredity 66: 343-350. hatter, i. w. 1993. yearling moose vulnerability to spike-fork regulation. memo dated jan. 26, 1993. wildlife branch, british columbia environment, victoria, british columbia, canada. _____. 1994. moose harvest regulations review-peace/liard subregion. wildlife conservation and management section, wildlife branch, british columbia environment, victoria, british columbia, canada. _____. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35: 91-103. _____, and k. n. child. 1992. an evaluation of a spike-fork bull moose antler regulation in central british columbia. proceedings of the 1991 moose harvest workshop, kamloops, bc. wildlife branch, british columbia environment, victoria, british columbia, canada. hundertmark, k. j., and r. t. bowyer. 1998. effects of population density and selective harvest on antler phenotype in simulated moose populations. alces 34: 375-383. alces vol. 46, 2010 child et al. vulnerability and antler regulations 121 _____, and _____. 2004. genetics, evolution, and phylogeography of moose. alces 40: 103-122. _____, t. h. thelen, and c. c.schwartz. 1993. population and genetic effects of selective harvest strategies in moose: a modeling approach. alces 29: 225-234. laurian, c., j-p. ouellet, r. courtois, l. breton, and s. st-onge. 2000. effects of intensive harvesting on moose reproduction. journal of applied ecology 37: 515-531. malo, a. f., e. r. s. roldan, j. garde, a. j. soler, and m. gomendio. 2005. antlers honestly advertise sperm production and quality. proceedings: biological sciences 272: 149-157. markusson, e., and i. folstad. 1997. reindeer antlers: visual indicators of individual quality? oceologia 110: 501-507. milliken, c. a., and d. e. johnson. 1984. analysis of messy data. volume i. wadswoah inc., belmont, california, usa. paquet, p. 2009. humans as ‘super-predators’ driving evolution. utoday, university of calgary news, calgary, alberta, canada. pélabon, c., and l. van breukelen. 1998. asymmetry in antler size in roe deer (capreolus capreolus): an index of individual and population conditions. oecologica 116: 1-8. ryman, n., r. baccus, c. reuterwall, and m. h. smith. 1981. effective population size, generation interval, and potential loss of variability in game species under different hunting regimes. oikos 36: 257-266. saether, b. e., e. j. solberg, and m. heim. 2003. effects of altering sex ratio structure on the demography of an isolated moose population. journal of wildlife management 67: 455-466. schwartz, c. c. 1998. reproduction, natality, and growth. pages 141-171 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institution press, washington, d.c., usa. _____, k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bullmoose harvest on the kenai peninsula, alaska. alces 28: 1-13. solberg, e. j., loison, a., ringsby, t. h., sæther, b-e., and m. heim. 2002. biased adult sex ratio can affect fecundity in primiparous moose alces alces. wildlife biology. 8: 117-128. _____, and b-e. saether. 1993. male traits as life-history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammalogy 75: 1069-1079. _____, and _____. 1994. fluctuating asymmetry in antlers of moose (alces alces): does it signal male quality? proceedings: biological sciences 354: 251-252. timmermann, h. r. 1991. moose sociobiology and implications for harvest. proceedings of the 1991 moose harvest management workshop, kamloops, british columbia, wildlife branch, british columbia environment, victoria, british columbia, canada. van ballenberghe, v. 2004. in the company of moose. stackpole books, mechanicsburg, pennsylvania, usa. vanpé, c., j-m.gaillard, p. kjellander, a. mysterud, p. magnien, d. delorme, g. van lafere, f. klein, o. liberg, and a. j. m. hewison. 2007. antler size provides an honest signal of male phenotypic quality in red deer. the american naturalist 4: 481-493. 71 status and management of moose in the parkland and grassland natural regions of alberta ronald r. bjorge1, delaney anderson2, emily herdman2, and scott stevens3 135 ansett crescent, red deer, alberta, canada t4r 2l9; 2alberta environment and parks, 250 diamond avenue, spruce grove, alberta, canada t7x 4c7; 3alberta environment and parks, 304 4920 51 street, red deer, alberta, canada t4n 6k8 abstract: moose (alces alces) naturally colonized the parkland natural region of alberta during the 1980s and early 1990s, and later colonized the grassland natural region by the early 2000s. we summarize population data during 1996–2016 for these regions, examining density, population trends, productivity, distribution, management, and moose-human conflicts to determine population status and sustainability. within the parkland, aerial surveys from one frequently monitored wildlife management unit (wmu) indicated a significant increase (r2 = 0.7476, p < 0.001) in density, representing an annual rate of change of 1.07. pooled data from an additional 21 parkland wmus indicated a mean annual rate of change of 1.11. mean density for the 22 parkland wmus over the study period was 0.19 ± 0.06 moose/km2, and aerial surveys indicated a mean of 74.4 ± 3.6 calves/100 cows and 51.9 ± 2.9 bulls/100 cows. within the grassland, winter aerial survey data from 4 wmus indicated a mean density of 0.05 ± 0.01 moose/km2, and 72.5 ± 6.75 calves/100 cows and 108.8 ± 34.4 bulls/100 cows. hunting in these regions has been managed with a limited entry hunt. resident rifle hunting opportunity for moose in the parkland and grassland increased 4.2-fold between 1996 and 2015. opportunity in this region also represented an increasing proportion of that available province-wide, from 3.4% in 1996 to 19.8% in 2015. alces vol. 54: 71–84 (2018) key words: alberta, alces alces, density, inventory, management, moose, population status. moose (alces alces) have a circumpolar distribution and are typically associated with forested boreal ecosystems (peterson 1974, reeves and mccabe 2007). they are renowned as a reliable source of food and recreation and for their cultural and economic significance (franzmann 1978). recent population declines in some regions of canada and the united states are causing growing concerns among wildlife managers and the public (murray et al. 2006, lenarz et al. 2009, crichton et al. 2015). in contrast, this paper documents the success of a population of moose during a 20-year period (1996– 2016) following establishment at low density in the agriculturally-dominated parkland and grassland natural regions (hereafter parkland and grassland) of alberta (bjorge 1996). the parkland and grassland has extensive agricultural and human development, making these regions appear unlikely to support growing moose populations. in the parkland, about 90% of the natural vegetation has been removed and in the grassland, excluding riparian areas, there is limited woody vegetation. in forested ecosystems moose select habitats that provide forage, cover, and security from predators (telfer 1984). although agricultural areas with limited woody cover may provide moose with adequate forage resources (laforge et al. 2016), these habitats may be associated with moose population status in alberta – bjorge et al. alces vol. 54, 2018 72 increased risk of heat stress (dussault et al. 2004) due to lack of cover. in addition, infrastructure including roads, highways, farms, towns, cities, and energy development is abundant and, along with associated human activity, may pose additional challenges for moose. since the time of european settlement, moose were not commonly observed here by residents (dwier 1969, stelfox and stelfox 1993) nor were they often observed during aerial inventory of the parkland prior to the early 1980s or in the grassland prior to 2000. here we document density, population trends, productivity, harvest management, and public complaints of a moose population in the parkland and grassland of alberta from 1996–2016. we also discuss biological and social factors contributing to utilization of this agricultural and human-dominated landscape, and future concerns regarding population dynamics and management. study area the study area included the parkland (primarily alberta 200 series wildlife management units [wmus], plus wmus 728, 730, and 936) and grassland wmus (primarily alberta 100 series wmus) in southeastern alberta (dowling and pettapiece 2006; fig. 1). wmu 224 was excluded from analysis because of the high proportion that fell within the boreal natural region. wmu 166 was treated as a parkland wmu because of the high proportion of the area that fell within the parkland. wmus 728 and 730 (canadian division support base, edmonton detachment, wainwright) were treated as a single wmu because they abutted each other and were managed as one unit. the parkland is a broad transitional zone between the warmer, drier grass-dominated grassland to the south and the more heavily treed boreal natural region to the north and west (strong and leggat 1992). in alberta, the central and foothills parkland cover almost 9% (57,627 km2) of alberta and extend into saskatchewan and manitoba (riley et al. 2007); the outlying peace parkland in northwest alberta was not part of this study. the dominant tree species in the parkland was trembling aspen (populus tremuloides), although they were less common prior to european settlement when wildfires were more frequent (strong and leggat 1992). scattered pockets of white spruce (picea glauca) and balsam poplar (populus balsamifera) also occur. common shrubs included willow (salix spp.), chokecherry (prunus virginiana), saskatoon berry (amelanchier alnifolia), red osier dogwood (cornus stolonifera), and canada buffaloberry (shepherdia canadensis). the dominant natural grass in the area was rough fescue (festuca scabrella). common large mammal species included white-tailed deer (odocoileus virginianus), mule deer (o. hemionus), and coyote (canis latrans), all of which are found in the grassland. about 4% of the land area was covered by water (dowling and pettapiece 2006) including thousands of small wetlands. major rivers with valleys and tributaries included the north saskatchewan, red deer, and battle. average total annual precipitation was ~400 mm (strong and leggat 1992). more than 90% of the parkland was privately owned (bjorge et al. 2004) and it includes the major urban centers of edmonton, calgary, red deer, wetaskiwin, camrose, and lloydminster. the grassland is the warmest and driest natural region in alberta. water comprises 1–2% of the land base (dowling and pettapiece 2006), consisting primarily of major rivers (red deer, south saskatchewan, oldman, bow), and shallow lakes and wetlands. native grass species included needle and thread grass (hesperostipa comate), wheat grass (agropyron spp.), and rough fescue. narrow leaf cottonwood alces vol. 54, 2018 moose population status in alberta – bjorge et al. 73 (populus angustifolia), western plains cottonwood (p. deltoids), and balsam poplar were the dominant trees and found primarily in riparian areas. shrubs included willow, buck brush (ceanothus cuneatus), silverberry (elaeagnus commutate), silver sage (artemisa cana), and saskatoon berry. private ownership was estimated at 70% (prairie conservation forum 2016) and major urban centers included calgary, lethbridge, and medicine hat. land in both natural regions has been heavily modified for agriculture, industrial development (primarily oil and gas and renewable energy), infrastructure, and urban development. bjorge et al. (2004) estimated ~10% of the central parkland remained as native vegetation, mostly as woody vegetation. remaining native vegetation in the grassland was ~40% (abmi 2015) with native grass dominant. the current human population in the combined area was fig. 1. wildlife management units (wmus) in the parkland and grassland (prairie), alberta, canada. moose population status in alberta – bjorge et al. alces vol. 54, 2018 74 estimated at >3 million (statistics canada 2017). methods we examined data collected during aerial surveys conducted for ungulate management in the parkland and grassland during 1996–2016. prioritization of wmus for aerial surveys was based on the following criteria: 1) time interval since most recent survey, 2) local budgets, 3) density of target species in the wmu, 4) prevalence of chronic wasting disease, and 5) stakeholder interests including hunter concerns and public complaints. surveys of individual wmus were intended to occur once every 3–6 years; however, this was often not achieved due to budgetary and/or weather constraints. the notable exception to this was wmu 728/730 where the objective was to fly the unit a minimum of once every 2 years. surveys occurred during prime snow conditions, generally in december, january, and early february. two types of aerial survey methods were utilized during the study period: 1) stratified random block surveys (1996–2010) and 2) strip-transect surveys (all surveys in wmu 728/730 and other wmus in 2011–2016). the change to strip-transect surveys was to implement a more efficient methodology by eliminating pre-flight stratification surveys and minimizing time travelling between study blocks. comparison of the 2 survey methods produced similar results on the same areas (j. allen, alberta environment and parks, pers. commun.). therefore, we assumed that the 2 survey methods provided data suitable for direct comparison between years in the same wmu. stratified random block surveys followed gasaway et al. (1986) and were modified according to lynch and schumaker (1995). the survey area was broken into degree blocks, (such as 3 min latitude by 5 min longitude) and stratified into 3 or 4 strata (high, medium, low, very low) based on pre-flights for the target species or interpretation of aerial photos. survey blocks within each strata were randomly selected for inventory (hofman and grue 2012). surveys were conducted from bell 206 helicopters on flight lines spaced 15 sec apart, with the objective of complete coverage of each block. flight crews consisted of a pilot, a navigator/recorder seated beside the pilot, and observers seated behind the pilot and the navigator/recorder. survey speed was 80–120 km/h and height was 80–120 m agl. all observed ungulates were counted and classified to age and sex when possible. female moose were identified by their white vulva patch visible from the air and calves by their small size (timmermann 1993). population estimation spreadsheets (i.e., quad6.xls files, microsoft excel, redmond, washington) adapted from gasaway (1986) were used to calculate population size, confidence limits, density, and sex/age classifications. strip-transect surveys (jolly 1969, alberta environment and sustainable resource development 2014) were conducted by flying transect lines at 1.6 km intervals with 25% coverage (400 m-wide survey strip); some variability in transect spacing occurred depending on tree cover, size of the wmu, and study objectives. in parkland wmu 728/730, the most frequently surveyed wmu, transects were flown at 800-m intervals, with moose observed within 400 m on either side of the flight line. surveys were conducted from a bell 206 helicopter with flight speed, altitude, and survey crew as described for the stratified random block surveys (see above). because transects varied in length (jolly 1969, alberta environment and sustainable resource development 2014), beginning in 2011, the average density (r; moose/km2) alces vol. 54, 2018 moose population status in alberta – bjorge et al. 75 was calculated by summing the total animals counted per transect (∑x), and dividing by the total area searched (length of transects multiplied by width of survey strip [∑z]). we calculated population estimates (unequal sized units, sampling without replacement) by multiplying the average density (r) by the overall area of the wmu (z). we estimated 90% confidence intervals by multiplying the t statistic for the left-tailed inverse of the student’s t-distribution, (t 0.05,df=n–1 ) by standard error (se; without replacement) of the abundance estimate, where se = square root of variance, and variance=n*(n–n)/ (n*(n–1))*(∑x2+r2*∑z2–2*r*∑xz) with n as the total number of possible transects given 100% coverage, and n as the number of transects sampled. we estimated the mean overall density of moose in the parkland over the study period by establishing the mean density in each wmu and then calculating the overall mean for the 22 wmus. where appropriate, means were presented as ± standard error of the mean (se). we used population estimates to estimate population growth rates and trend during the study period. in wmu 728/730, where 12 population estimates were available for the study period, we used log-linear regression (harris 1986) of moose/km2 against year to test for evidence of a significant population trend. in the other 21 wmus (166, 200, 202, 203, 204, 206, 208, 220, 228, 230, 232, 234, 236, 240, 242, 246, 248, 250, 258, 260, 963) where fewer inventories were conducted, as well as wmu 728/730, we calculated annual population growth rates (λ) after hatter (1999, 2001). growth rates were estimated as λ = (n t /n 0 )1/t, where n t is the number of moose/km2 in year t, and n 0 is the number moose/km2 in the initial survey year. we then estimated the mean population growth rate as the average of λ estimates per wmu, recognizing that specific study periods varied by wmu. harvest data were also available from compulsory reporting at a wmu check station in wmu 728/730. in all other wmus, harvest statistics were estimated from data collected by an annual telephone questionnaire (lynch and birkholz 2000); beginning in 2011, an online questionnaire was distributed to all licenced hunters. hunter success was calculated based on the success of respondents that held licences and hunted. we used log-linear regression to test for a significant trend in the harvest of moose in wmus 728/730 over the study period. harvest rates (%; harvest/preseason population estimate) were estimated for wmu 728/730 for the 12 years that winter population estimates were available. preseason estimates were calculated by applying an annual winter mortality rate of 5% (an estimate) to the winter population estimate and adding calf production as indicated from aerial surveys the previous winter. management history was derived from a review of annual alberta guide to hunting regulations, alberta hunting draw annual publications, other provincial summaries, and from personal communication with provincial wildlife management staff. data summarizing public complaints (concerns expressed by the public and recorded by district fish and wildlife officers) about moose were summarized for parkland wmus and were available for 1999–2015. categories of complaint included vehicle collision/unspecified injury (including injury of all types, many of which were from moose-vehicular collisions), human conflict (specific concerns such as human safety and nuisance that did not fit into other categories), sighting (usually close to human activity and of concern to humans), agricultural conflict (garden damage, tree damage, crop damage, stack damage, damage to game farms and harassment of livestock, plus other unspecified damage), disease, orphaned moose population status in alberta – bjorge et al. alces vol. 54, 2018 76 moose, and harassment of wildlife. all categories of public complaint were summarized, except for those subject to enforcement actions which were unavailable due to privacy concerns. we treated sightings as legitimate complaints within these summaries because they demonstrated marked concerns from the public about the presence of moose. wmus 212, 220, and 248 were classified as urban wmus because they included alberta’s 3 largest cities and associated urban sprawl. results the mean density of moose in the 22 parkland wmus was 0.19 ± 0.06/km2 (range = 0.05 – 1.28/km2). the most consistent population data were available from wmu 728/730, a military base inventoried with strip-transect surveys 12 times between 1998 and 2015. moose were first observed in wmu 728/730 during aerial surveys in 1983 when the density was estimated at 0.02/km2 (bjorge 1996). from 1998–2015, the population growth rate was λ = 1.07 and significantly increasing (r2 = 0.74, f = 28.8, p < 0.001; fig. 2). compulsory registration of all hunting indicated substantial and increasing harvest (r2 = 0.81, f = 79.1, p < 0.001; fig. 3); the mean harvest rate was 17.9% (range = 13.3–22.4%) for the 12 years following the winter population estimates. the mean rate of population increase for 21 additional parkland wmus was λ = 1.11 (range = 0.94–1.41); only 2 wmus (220 and 250) had declining populations. sex and age classifications were available from 60 aerial surveys from parkland wmus in winters 1996–2016. the mean number of calves/100 cows was 74.4 ± 3.5 (range = 27–150). the mean number of bulls/100 cows was 51.9 ± 2.90 (range = 5–97). parkland moose populations began expanding into grassland wmus during their growth phase in the midto late 1990s. multiple aerial surveys conducted in 4 grassland wmus (151, 152, 162, 163 fully within the grassland area) indicated the pattern of population establishment and growth (table 1). in these wmus, no moose were observed in aerial inventories during the 1990s, very low numbers were observed in the early 2000s, but by 2014–2016, populations were well established at low density, ranging from 0.02 to 0.07 moose/km2. populations were large enough to warrant establishment of hunting seasons in 5 grassland wmus adjacent to parkland wmus (156, 158, 160, 163,164) by 1999, and by 2015, hunting seasons were established in 16 of 26 grassland wmus. productivity was fig. 2. the increase in moose population density estimates derived from natural log-linear regression analysis of aerial survey data in wmu 728/730, 1998-2015, alberta, canada. zero values indicate years when surveys were not conducted. y = 0.065x – 131.23 r² = 0.7424 –1.8 –1.6 –1.4 –1.2 –1 –0.8 –0.6 –0.4 –0.2 0 1995 2000 2005 2010 2015 lo g e (p op ul a� on d en si ty ) year y = 4.504x – 8969.1 r² = 0.8064 0 20 40 60 80 100 120 140 160 1995 2000 2005 2010 2015 to ta l m o o se h ar ve st ed year fig. 3. the increase in moose harvest (linear regression) in wmu 728/730 from 1996 to 2016, alberta, canada. alces vol. 54, 2018 moose population status in alberta – bjorge et al. 77 72.5 ± 6.7 calves/100 cows, with an adult sex ratio of 109 ± 34 bulls/100 cows during recent winter surveys in 4 wmus (151, 152, 162, 163). since inception, hunting in the parkland and grassland was through limited entry antlered and antlerless special licences using a draw process. in wmu 728/730, special licences for calves were available until 2013 when they were amalgamated with antlerless licences. the exception was for archery hunts in wmus 212 and 248 that surround the cities of edmonton and calgary, in which there was no restriction on licences issued for archery hunting. there was a substantial increase in moose hunting opportunity in the parkland and grassland during the study period as 3,555 special licences were granted in the parkland and grassland in 2015 compared to 852 in 1996. specifically, there was a 4.1-fold increase in antlered special licences and a 4.3-fold increase in antlerless and calf special licences combined (table 2). the proportion of provincial rifle hunting opportunities for residents also increased in the parkland and grassland from 3.4% of the provincial total in 1996 to 19.9% in 2015. this pattern was driven by increasing opportunity in the parkland and grassland as provincial opportunity declined when all moose hunting in alberta was placed on limited entry hunting during this period. resident hunter success rates in the parkland and grassland (excluding archery-only hunts) were estimated at 74.5 ± 7.3% in the grassland and 79.3 ± 3.1% in the parkland in 2015, compared to 48.0 ± 2.5% for alberta as a whole. a similar pattern was observed 20 years earlier in 1996 when hunter success rates were 76.5 ± 2.5% in the parkland and 36.7% for alberta, which still had general resident moose seasons over much of the province. table 1. number of moose counted per survey during aerial surveys in 4 wmus in the grassland natural region of alberta (1990s-2016), including population estimates (± se) and moose density during the most recent period, 2011–2016. na = no survey conducted. wmu 1990s 2000–2005 2006–2010 2011–2016 population estimate density (moose/km2) 151 0 3 na 50 98 ± 22 0.07 152 0 5 49 79 154 ± 46 0.04 162 0 10 na 32 64 ± 17 0.02 163 0 5 na 72 186 ± 29 0.05 table 2. moose hunting opportunity (# and %) for residents of alberta in the parkland and grassland natural regions (pgnr) compared to province-wide totals, 1996 and 2015. special licences were managed through a draw system where the number of licences available for a given group (e.g., antlered) was limited. by 2015, general licences which had previously been issued with no restriction in number were no longer available, and special calf licences were considered as special antlerless licences. licence province 1996 pgnr 1996 province 2015 pgnr 2015 special antlered 11,800 378 (3%) 12,114 1515 (12%) special antlerless 1435 439 (31%) 4603 2040 (44%) special calf 562 35 (6%) 1155 0 general 11,549 0 0 0 total 25,346 852 (3%) 17,872 3555 (20%) moose population status in alberta – bjorge et al. alces vol. 54, 2018 78 a total of 5,653 public complaints about moose were registered at provincial fish and wildlife enforcement branch offices and recorded into provincial enforcement databases during 1999–2015 for the parkland (fig. 4, table 3). complaints more than doubled from a low of 219 in 1999 to a high of 482 in 2015, but varied substantially over the study period. overall, the most common complaints were vehicle collisions/injury (42%), human conflict (27%), and sightings (19%). agricultural damage (6%), disease (3%), orphaned moose (3%), and harassment of wildlife (<1%) were minor complaints. only 14 of 330 (4%) agricultural complaints were attributed to crop damage. damage to trees, livestock including harassment, feed stacks, gardens, game farms, and unspecified damages (125 complaints) comprised the other agricultural complaints; some of the unspecified damage could have been crop-related. the majority of complaints (60%) were from wmus 212, 220, and 248 (fig. 1 and 4, table 3) which include alberta’s 3 largest cities—calgary, edmonton, and red deer. here the majority of complaints related to human conflict (34%), vehicle collision/injury (31%), and sightings (28%). these wmus were estimated to hold < 20% of the parkland moose population, but supported an estimated 2.8 million people (statistics canada 2017). among rural wmus, the most common complaints were vehicle collisions/injury (56%), human nuisance conflict (16%), and agricultural damage (10%). fig. 4. number of public moose complaints from rural and urban areas within the parkland natural region as registered by the alberta fish and wildlife enforcement branch, 1999-2015, alberta, canada. 0 100 200 300 400 500 600 1999 2001 2003 2005 2007 2009 2011 2013 2015 c o m p la in ts rural urban year alces vol. 54, 2018 moose population status in alberta – bjorge et al. 79 discussion moose populations that established in the alberta parkland during the 1980s and mid-1990s (bjorge1996) have continued to grow and expand over the last 20 years. in the early 2000s moose expanded into grassland wmus and established at low density, eventually providing an increasing proportion of moose hunting opportunity in alberta. overall, production and survival have been greater than the combined influences of natural and human-induced mortality, resulting in substantial population growth. the success of this population has occurred at a time when moose populations are declining in several boreal wmus within alberta (j. castle, c. found, l. vander vennen, alberta environment and parks, unpublished data), and in several other north american jurisdictions (murray et al. 2006, crichton et al. 2015, kuzyk 2016). this agriculturally-dominated study area with limited natural habitat and extensive fragmentation would seem unlikely habitat for expansion of a moose population. however, these 2 natural regions have the basic ecological and social conditions necessary for population growth, as observed in similar habitats in saskatchewan (laforge et al. 2016), manitoba (crichton et al. 2015), and north dakota (j. smith, north dakota game and fish, pers. commun.). between 2001 and 2014, the provincial moose population increased ~25% from 92,000 to 115,000 (timmermann and rodgers 2017). the estimated population increased 3-fold between 2000 and 2014 in parkland wmu 728/730, indicating much higher local growth. calf production and survival was high in the study area at >70 calves/100 cows based on winter surveys, similar to that observed earlier by bjorge (1996), and much higher than the 46 calves/100 cows estimated in boreal wmus in northwestern alberta (d. moyles, alberta environment and parks, unpublished data). the high survival of calves is presumed to reflect the paucity of moose predators throughout the study area. major predators (ballard and vanballenberghe 2007) such as wolves (canis lupus), black bears (ursus americanus), and grizzly bears (u. arctos) were essentially absent from all but the extreme western and northern perimeter of the study area. cougars (felix concolor) were also at very low density, although coyotes, which have potential to prey on moose calves (benson and patterson 2013), were considered abundant. high calf:cow ratios are not uncommon among moose populations with limited predators (rolley and keith 1980, j. smith, north dakota game and fish, pers. commun.). table 3. summary of wildlife complaints (#) regarding moose in the parkland natural region of alberta between 1999 and 2015 in urban (n = 3) and rural (n = 31) wildlife management units. complaint type urban rural total road kill/injury 1058 1330 2388 human conflict 1131 369 1500 sighting 942 148 1090 agricultural damage 104 226 313 disease 63 104 167 orphan 74 76 150 wildlife harassment 10 18 28 total 3382 2271 5653 moose population status in alberta – bjorge et al. alces vol. 54, 2018 80 although browse production and availability was not assessed in our study area, we suggest that the abundance and diversity of shrubs and vegetation in riparian habitats, and the remaining patches of forest and associated edge (schneider and wasal 2000), provide adequate browse and forage for population growth (gasaway and coady 1974). the parkland, and to a lesser extend the grassland, has an abundance of small wetlands and several major river valleys and tributaries which likely contribute measurably to available moose habitat. moose are subject to heat stress (dussault et al. 2004) during summer, and wetlands and other riparian areas presumably play a role in thermoregulation (renecker and hudson 1986, renecker and schwartz 2007). laforge et al. (2016) documented strong selection for wetlands and forest cover in farmland in southcentral saskatchewan, indicating the proportional importance of these habitats. moose also consume agricultural crops such as canola, cereals, and legumes such as alfalfa which are likely important food sources (sorenson et al. 2015), although their consumption level and nutritional quality are unknown. human-associated mortality including licenced hunting, poaching, aboriginal harvest, vehicular collisions, and infrastructure-associated injuries, in combination with natural mortality, did not prohibit population growth in the study area. harvest rates were often conservative, estimated at 13–22% of the moose population. given our observations of >70 calves/100 cows during winter and the absence of significant predators, these harvest rates would allow population growth. crichton et al. (2015) noted the expansion of moose into parkland and grassland habitats of manitoba, saskatchewan, and alberta occurred during a period of human depopulation of these areas as farms became larger, which may have reduced undocumented illegal harvest. bjorge (1996) indicated that establishment of moose populations in the parkland may have been associated with a possible change in attitude of rural residents resulting in less poaching of moose dispersing from adjacent boreal habitats. one consequence of population growth in our study area was the increasing occurrence of moose in urban environments and areas of concentrated rural residences that pose unique management issues. urban environments appear to provide several advantages to moose including unutilized browse and forage, limited or no hunting, and very few predators, albeit, high potential for moose-human conflicts. in norway, lykkja et al. (2009) observed that moose moved away from inhabited houses during periods of high human activity, suggesting they are somewhat responsive to such activity. in our experience, moose complaints in 3 urban wmus exceeded complaints in 31 rural parkland wmus combined, suggesting that urban residents have high interest and interactions in moose. although the time, labor, and costs associated with responding to urban complaints have not been quantified, it is reasonable to conclude that it is substantial. further, it requires specialized training and equipment to immobilize and transport or euthanize moose in areas with high visibility and human population. managers need to consider impacts and issues associated with urban moose populations when establishing harvest goals and management strategies in adjacent rural areas, and be prepared to address moose human conflicts in urban environments. we were surprised that only 6% of all public complaints were related to agricultural damage, given that moose in the parkland and grassland were living in an agriculturally-dominated landscape. further, only 4% of these complaints were attributed alces vol. 54, 2018 moose population status in alberta – bjorge et al. 81 specifically to crop damage. we believe that this low rate likely reflects the low density of often solitary moose spread over an extensive agricultural land base, making widespread damage attributable to moose less evident. laforge et al. (2016) indicated that moose in southcentral saskatchewan did not exhibit strong selection for crop types with the possible exception of oilseeds in summer; conversely, maskey (2008) found selection for local crops in north dakota. we also expected more complaints related to disease (3% of complaints), especially parasitism by winter tick (dermacentor albipictus) because associated hair loss (samuel et al. 2000, samuel 2007) was common in the study area. although the moose population in the parkland and grassland continuously increased during the study period, several factors could deter future population growth. for example, decline in wetlands and woody cover would negatively impact moose habitat and carrying capacity. the associated vegetative and cover resources are especially important in consideration of climate change (parmesan 2006, mcgraw et al. 2012), and that only about 10% of the parkland remains as native woody cover (bjorge et al. 2004), with even less in the grassland. further, increased impact of winter tick parasitism or disease could negatively impact moose populations. higher poaching or legal first nations and métis harvest might also reduce local populations (carmichael 2015). regadless of environmental changes, the moose population will eventually exceed its carrying capacity or some other density-dependent mechanism will curb population growth. a paradigm in the management of large herbivores is that following introduction to a new range or cessation of harvest, the population may increase to peak abundance and then crash and re-establish at a lower level (caughley 1970, forsyth and caley 2006). we suggest it is important that parkland and grassland moose populations be managed to avoid major declines due to exceeding the carrying capacity or other factors related to population density. multiple and often unique conditions influence the moose populations in the agricultural landscape of the parkland and grassland. clearly, adequate population monitoring and assessing both social and ecological carrying capacity of these populations are necessary management objectives in this human-dominated ecosystem. effective management will involve measuring these carrying capacities, stakeholder priorities, and risks to safety and property. means to determine stakeholder values and ongoing measures of public safety and property damage are required to assess social tolerance both education and preventative management must be emphasized. appropriate training and equipment to respond professionally to urban moose issues, monitoring disease and conflicts, and continued enforcement oversight are critically important for an adaptive and effective management program. acknowledgements we are grateful for assistance provided by the following: s. nadeau for providing licencing data, j. allen for reviewing the presentation and an earlier draft of this paper and providing information related to implementation of strip-transect surveys, d. prescott for providing statistical and data advice, j. unruh for gis support, r. bjorge for logistic support, r. corrigan for providing provincial moose management data, fish and wildlife enforcement branch for public complaint information, d. moyles and l. vander vennen for data from the boreal region of north-western alberta, and j. castle for discussing various aspects of alberta moose management. we especially moose population status in alberta – bjorge et al. alces vol. 54, 2018 82 thank the numerous wildlife management professionals who conducted aerial inventories and assessments in the parkland and grassland over the years including j. allen, d. cole, k. froggatt, e. hofman, d. moore, and r. russell. staff at canadian forces base, wainwright are thanked for their support in hosting the wmu 728/730 hunt and for logistic support. the canadian wildlife federation supported the attendance of r. bjorge at the 50th moose conference and workshop where an initial version of this paper was presented. references alberta biodiversity monitoring institute (abmi). 2015. the status of biodiversity in the grassland and parkland regions of alberta; preliminary assessment. a l b e r t a b i o d i v e r s i t y m o n i t o r i n g institute, edmonton, alberta, canada. alberta environment and sustainable resource development. 2014. alberta aerial ungulate survey protocol. alberta environment and sustainable resource development. edmonton, alberta, canada. ballard, w., and v. vanballenberghe. 2007. predator-prey relationships. pages 247–274 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. benson, j. f., and b. r. patterson. 2013. moose (alces alces) predation by eastern coyotes (canis latrans) and eastern coyote x eastern wolf (canis latrans x canis lycaon) hybrids. canadian journal of zoology 91: 837–841. bjorge, r. r. 1996. recent occupation of the alberta aspen parkland ecoregion by moose. alces 32: 141–147. _____, j. schieck, l. george, and g. nieman. 2004. status and management of native vegetation in the parkland natural region – central parkland. pages 34–36 in g. c. trottier, e. anderson, and m. steinhilber, editors. proceedings of the 7th prairie conservation and endangered species conference, february 2003. natural history occasional paper number 26. provincial museum of alberta, edmonton, alberta, canada. carmichael, r. 2015. exploring a rich wilderness tradition. the wildlife professional. fall 2015: 37–39. caughley, g. 1970. eruption of ungulate populations with emphasis on himalayan thar in new zealand. ecology 51: 53–77. crichton, v., k. child, and b. ranta. 2015. exploring the future of canada’s moose. the wildlife professional fall 2015: 40–44. dowling, d. j., and w. w. pettapiece. 2006. natural regions of alberta. government of alberta publication t/852. edmonton, alberta, canada. dussault, c., j. p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioral responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. dwier, m. v. 1969. the ecological characteristics and historical distribution of the family cervidae in alberta. m. sc. thesis, university of alberta, edmonton, alberta, canada. forsyth, d., and p. caley. 2006. testing the irruptive paradigm of large-herbivore dynamics. ecology 87: 297–303. franzmann, a. w. 1978. moose. pages 67–77 in j. l. schmidt and d. l. gilbert, editors. big game of north america, ecology and management. wildlife management institute, washington, d. c., usa. gasaway, w. c., and j. w. coady. 1974. review of energy and rumen fermentation in moose and other ruminants. naturaliste canadien 101: 227–262. _____, d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. alces vol. 54, 2018 moose population status in alberta – bjorge et al. 83 biological papers of the university of alaska, number 22. institute of arctic biology, university of alaska, fairbanks, alaska, usa. harris, r. b. 1986. reliability of trend lines obtained from variable counts. journal of wildlife management 50: 165–171. hatter, i. w. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 37: 91–103. _____. 2001. an assessment of catch per unit effort to estimate rate of change in deer and moose populations. alces 39: 71–77. hofman, e., and g. grue. 2012. wildlife management unit 160 mule deer. pages 46–49 in m. ranger and r. anderson, editors. delegated aerial ungulate surveys, 2010/2011 survey season. data report d-2011-009, alberta conservation association, sherwood park, alberta, canada. jolly, g. m. 1969. sampling methods for aerial censuses of wildlife populations. east african agricultural and forestry journal 34: 46–49. kuzyk, g. w. 2016. provincial population and harvest estimates of moose in british colombia. alces 52: 1–11. laforge, m. p., n. l. michel, a. l. wheller, and r. k. brook. 2016. habitat selection by female moose in the canadian prairie ecozone. journal of wildlife management 80: 1059–1068. lenarz, m. s., m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in north eastern minnesota. journal of wildlife management 73: 503–510. lykkja, o. n., e. j. soldberg, i. herfindal, j. wright, c. m. rolandsen, and m. hanssen. 2009. the effects of human activity on summer habitat use by moose. alces 45: 109–124. lynch, g. m., and s. birkholz. 2000. a telephone questionnaire to assess moose harvest. alces 36: 105–109. _____, and g. e. shumaker. 1995. gps and gis assisted moose surveys. alces 31: 145–151. maskey, j. j. 2008. movements, resource selection and risk analysis for parasitic disease in an expanding moose population in the northern great plains. ph. d. thesis, university of north dakota, grand forks, north dakota, usa. mcgraw, a. m., r. moen, and l. overland. 2012. effective temperature differences among cover types in northeastern minnesota. alces 48: 45–52. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. r. barnett, and t. k. fuller. 2006. pathogens, nutrition deficiency and climate change influences on a declining moose population. wildlife monographs 166: 1–30. parmesan, c. 2006. ecological and evolutionary responses to recent climate change. annual review of ecology, evolution and systematics 37: 637–669. peterson, r. l. 1974. moose: yesterday, today and tomorrow. naturaliste canadien 101: 1–8. prairie conservation forum. 2016. prairie conservation. prairie conservation forum, lethbridge, alberta, canada. reeves, h., and r. mccabe. 2007. of moose and man. pages 1–76 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose, 2nd edition. university of colorado, boulder, colorado, usa. renecker, l. a., and r. j. hudson. 1986. seasonal energy dynamics and thermoregulation of moose. canadian journal of zoology 64: 322–327. _____, and c. c. schwartz. 2007. food habits and feeding behavior. pages 403–440 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university of colorado press, boulder, colorado, usa. moose population status in alberta – bjorge et al. alces vol. 54, 2018 84 riley, j. l., s. e. green, and k. e. brodribb. 2007. a conservation blue print for canada’s prairies and parklands. nature conservancy of canada, toronto, ontario, canada. rolley, r. e., and l. b. keith. 1980. moose population dynamics and winter habitat use at rochester, alberta, 1969– 1979. canadian field naturalist 94: 9–18. samuel, w. m. 2007. factors affecting epizootics of winter tick and mortality on moose. alces 43: 39–48. _____, m. s. mooring, and l. aalangdong. 2000. adaptation of winter tick (dermacentor albipictus) to invade moose and moose to invade ticks. alces 36: 183–195. schneider, r. r., and s. wasel. 2000. the effect of human settlement on the density of moose in northern alberta. journal of wildlife management 64: 513–520. sorenson, a. a., f. m. van beest, and r. k. brook. 2015. quantifying overlap in crop selection patterns among three sympatric ungulates in an agricultural landscape. basic and applied ecology 16: 601–609. statistics canada. 2017. 2016 census profile. government of canada, ottawa, ontario, canada. stelfox, j. b., and j. g. stelfox. 1993. distribution. pages 45–61 in j. b. stelfox, editor. hoofed mammals of alberta. lone pine publishing, edmonton, alberta, canada. strong, w. l., and k. r. leggat. 1992. ecoregions of alberta. publication t/245. alberta forestry, lands and wildlife, edmonton, alberta, canada. telfer, e. 1984. circumpolar distribution and habitat requirements of moose. pages 145–182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. timmermann, h. r. 1993. use of aerial surveys for estimating and monitoring moose. alces 29: 35–46. _____, and a. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. 139 kris j. hundertmark distinguished moose biologist 2007 recipient the distinguished moose biologist award was presented to dr. kris j. hundertmark at the 43rd north american moose conference and workshop, held at the university of northern british columbia in prince george, british columbia, canada, 2-7 june 2007, in recognition of his numerous contributions to improving our understanding of moose biology and management. kris is a graduate of the pennsylvania state university (bs 1976), oregon state university (ms 1981), and the university of alaska fairbanks (phd 2001). he spent 21 years (1982-2002) with the alaska department of fish and game working principally on the ecology, nutrition, and management of moose; much of this work was conducted at the moose research center on the kenai peninsula. he also was employed as a conservation geneticist with the zoological society of london, stationed at the king khalid wildlife research centre in thumamah, saudi arabia from 2003-2005. since 2005, he has held the position of assistant professor in the department of biology and wildlife, and the institute of arctic biology at the university of alaska fairbanks. kris is a critical and innovative thinker who continues to be a productive scientist and whose research will aid in resolving issues related to the management of moose. indeed, kris has published >30 articles on moose (21 papers in alces), including an influential chapter in the “moose book” on home range, migration, and dispersal. his scientific interests in moose are extremely broad, and cover topics ranging from reproduction to disease. his research on genetics and evolution has revolutionized modern concepts dealing with the origins of moose and their colonization of the new world. his work on genetics extends to understanding effects of harvest on moose antlers, and the role of genetics in the management and conservation of this unique large mammal. his current collaborations in alaska will integrate genetics with the landscape ecology of moose. indeed, no other person has published as many papers on the genetics of moose as kris. kris has been an active participant in the north american moose conference and workshop as well as the international moose conference. he never failed to participate in auctions, some of which he was able to recall. he has attended seven moose conferences and helped organize one north american conference (in anchorage) and one international conference (in fairbanks). he was an invited speaker at the international conference in norway. he has published numerous papers in alces since 1990, and served as a reviewer and associate editor for the journal, as well as playing a major role for supplement 2 (the russian papers). he also has contributed articles to the moose call. in short, he has made regular and important contributions to understanding the biology of moose and clearly is a dedicated “mooser.” the north american moose conference and workshop is proud to recognize over a quarter century of professional experience, devoted mostly to moose, by dr. kris j. hundertmark, the recipient of this year’s distinguished moose biologist award. f:\alces\vol_38\pagemaker\3809. alces vol. 38, 2002 silverberg et al. – impacts of wildlife viewing 205 impacts of wildlife viewing on moose use of a roadside salt lick judith k. silverberg1, peter j. pekins2, and robert a. robertson3 1new hampshire fish and game department, 2 hazen drive, concord, nh 03301,usa; 2department of natural resources, university of new hampshire, james hall, durham, nh 03824, usa; 3department of resources and economics, university of new hampshire, james hall, durham, nh 03824, usa abstract: in northern new hampshire, we examined the use patterns of moose visiting a roadside salt lick before (1996) and after (1997-1999) a blind was built specifically to view moose at the lick. moose visitation patterns were monitored with trail monitors equipped with cameras placed on trails leading into the study and control salt licks. there was no difference in frequency of use and time of use at the study and control sites in any year. nocturnal use was higher than diurnal use; use was greatest at 2200-2400 and 0400-0600 hours at both sites. reduced use of the trail closest to the blind indicated that placement of the blind probably altered access patterns of certain moose. a trend in 1998-1999 toward more visits during early morning than peak afternoon viewing time indicated that assessment of viewing opportunity warrants further study. alces vol. 38: 205-211 (2002) key words: alces alces, behavior, salt lick, wildlife viewing even though the term “nonconsumptive wildlife users” has been applied to describe people who do not hunt or fish, recreationists such as wildlife watchers do use and disturb recreational resources along spatial, visual, and physical dimensions. disturbances may be intentional or unintentional; unintentional disturbances often occur when photographing wildlife, viewing nesting birds, or hiking into an animal’s territory (knight and cole 1991, 1995). unintentional impacts also include direct harassment of animals or alteration of habitat (kuss et al. 1990). nonconsumptive users trample and rearrange vegetative patterns, disturb wildlife behavior and activity, and are the chief distributors of refuse across the land (goldsmith 1974, wilkes 1977). moose (alces alces) are strongly attracted to supplementary sodium during spring and early summer in large parts of their north american range (fraser 1979), and commonly use roadside salt licks in new hampshire that are created from runoff of salt spread on roadways in winter (miller and litvaitis 1992). such areas provide excellent places to view moose during may, june, and july and their high visibility has created a strong interest in moose viewing. northern new hampshire and maine are well known places to view moose and the wildlife viewing programs of both states have published guides for wildlife viewing (e.g., silverberg 1997). unfortunately, many viewing opportunities occur along roadsides during summer, and traffic congestion regularly occurs in certain locations. anecdotal information from moose viewers on route 3 in pittsburg, new hampshire, a popular moose viewing area, suggested that moose shifted use of salt licks to late night to avoid disturbance from viewers. limited research has been conducted on impacts of wildlife impacts of wildlife viewing – silverberg et al. alces vol. 38, 2002 206 viewing in situations such as those associated with moose viewing in northern new hampshire. the wildlife viewing program of the new hampshire fish and game department proposed construction of a moose viewing area on route 26 in dixville notch to provide viewers with an opportunity to view moose from a blind as an alternative to viewing from their cars along the roadside. the planning phase of this project provided the opportunity to design a research project that would explore specific questions about the use of roadside salt licks by moose at a state-sanctioned wildlife viewing facility. these questions fell into 3 categories: moose visitation rate and use of the lick from preconstruction to post construction; moose responses to wildlife viewing actions and other human caused stimuli; and the characteristics, motivations, and attitudes of wildlife viewers. from previous work, it is known that there exists a wide range of intra and interspecific variation of responses to disturbance (knight and temple 1995). studies conducted by mcmillan (1954) and altmann (1958) in yellowstone national park, showed a variety of behavioral responses in moose. in sibley provincial park, ontario, cobus (1972) found that in general moose developed a tolerance towards humans. the effect of an increase in road traffic on wildlife from 1973-1983 was examined in denali national park, alaska. this elevated volume correlated with a 72% decrease in moose sightings (signer and beattie 1986). this paper specifically focuses on the impact of the facility and viewing activities that could be assessed by monitoring moose movement activity preand postconstruction. specifically to determine if the visitation rate and time of use by moose at the salt lick in dixville notch were affected by the construction and subsequent use of the wildlife viewing area. study area the viewing site was located in northern new hampshire, to the east of dixville notch on route 26. this 4 ha area, inclusive of the viewing site, was harvested (clear-cut) in 1991 and was characterized by a regenerating northern hardwood/ spruce-fir forest community. a buffer strip of mature balsam fir (abies balsamea) and red spruce (picea rubra) was left on both sides of the road. the primary salt lick, about 175 m long, was on the north side of the road, and a smaller lick about 70 m long was on the south side. a 6-car parking lot, trail, and viewing blind were built in december 1996 across from the primary salt lick. construction occurred in december because moose reduce their use of licks after the fall rut (adams 1995). a trail (125 m) led to the viewing blind located about 30 m from the primary salt lick. the viewing blind held up to 20 people and had slits that faced the main lick and a moose trail that entered the lick from the east. the control site consisted of 2 roadside salt licks (200 m and 50 m long) 1.5 km east of the viewing site. these salt licks were approximately 0.2 km from a clear-cut. both sites were frequented regularly by moose prior to the study. the similarity between the study and control sites was ascertained by comparing aerial photos which showed that both were predominately spruce-fir forests before harvest; the control site was clear-cut 1 year after the study site. the licks on the study and the control sites were approximately the same distances from the center line of the highway. methods trailmaster 1500 game monitors were used to measure the visitation rate and time alces vol. 38, 2002 silverberg et al. – impacts of wildlife viewing 207 of use of salt licks by moose. the monitors are ideal for monitoring moose and other mammal movements because measurement is continuous and potential interference from observers is eliminated (kucera and barrett 1993). these monitors were used previously to measure use at salt licks in pittsburg and milan, new hampshire during 19941995 (adams 1995). a monitor consisted of a transmitter that emitted an infrared beam to a receiver that tripped an automatic 35 mm camera. when an animal walked through the beam, the receiver recorded the date and time, and the camera took a picture. the monitor could store a maximum of 1,000 events. the sensitivity of the trigger and the length of time the beam must be broken to register an event was adjusted to 0.05 seconds. every time the beam was broken the data recorder marked the event. to prevent a photograph from being taken multiple times of the same animal, the camera was set for a photograph to be taken every 2 minutes. date and time were recorded on each photograph. the cameras had flashes and professional high-speed (asa 1600) film was used to ensure an image was recorded at night. five monitors were placed at the viewing site (#1-5) and 4 monitors were placed at the control site (#6-9) simultaneously. the 2 licks at the control site were considered as one due to their proximity and interconnected moose trails. because the location of monitors is crucial to provide maximum information (kucera and barrett 1993), they were located on major moose trails entering the licks. the monitor and receiver camera package were placed on a tree or stake on the opposite sides of a well established trail. specific placement took advantage of localized terrain, trail characteristics, and surrounding vegetation. care was taken to minimize the possibility of sunlight and blowing vegetation breaking the infrared beam, thus triggering the camera. monitors were placed at heights of 3075 cm to also record the presence of medium-sized mammals (e.g., white-tailed deer (odocoileus virginianus), bear (ursus americanus), and coyote (canis latrans)). monitors were placed in the same locations each year. data were collected from 10 june-14 july during 1996-1999. monitors were checked twice weekly when data were downloaded and recorded in a logbook; film was replaced as needed. the date and time stamp on the developed film was compared to the information recorded by the monitor. the data were entered into a spreadsheet indicating the monitor number, year, time, date, whether there was a photograph, whether an animal was in the photograph, identity of animal, and sex and age of moose (if possible). judgements were made to eliminate multiple data collected in a short period of time caused by a stopped animal, or an animal moving in and out of the lick within a 2minute period. for example, if the monitor recorded 10 passes within 2 minutes, and photographs indicated it was the same moose, only 1 visit was counted. moose were not marked, consequently, there was no way to determine how many times a particular moose entered a lick, or if the same moose used the area annually. in situations when a camera ran out of film, but events were recorded at similar frequencies as when photographs indicated single visits, these events were classified as moose visits. it was assumed that a monitor malfunctioned when it recorded hundreds of events per day. malfunction was apparent during periods of heavy rain or wind. data were analyzed using spss (spss inc., chicago, illinois). graphs and frequency distributions were used to provide an overall depiction of moose encounters. for anova, moose encounter data were aggregated on a weekly basis by year to test impacts of wildlife viewing – silverberg et al. alces vol. 38, 2002 208 for differences in the number of moose visits at the viewing and control sites annually. combining data on a weekly basis eliminated the problem of small sample size on any given day. data of visitation times were aggregated into 12, 2-hour time blocks for analysis. this aggregation eliminated potential problems with small sample sizes in any 1-hour block. time was described as 14 diurnal hours (0600-2000 h) and 10 nocturnal hours (2000-0600 h) based on daylight and times when viewers could view moose without artificial light. statistical significance was set at 0.05 a priori. results the number of annual moose encounters at the viewing site (mean ± sd= 228.0 ± 16.7) and the control site (mean ± sd = 273.5 ± 19.7) was relatively constant during the 4 years. there was no difference in the annual weekly encounter rate from year to year at the viewing site (f = 0.280; df = 3, 16; p = 0.839) or control site (f = 0.712; df = 3, 16; p = 0.559). variability occurred at individual monitors at both sites annually (fig. 1). monitors 2-4 had more encounters the last 2 years than the previous years; encounters at monitor 5 were constant. conversely, monitor 1, located < 10 m from the viewing blind, had about 50% less encounters the last 2 years (fig. 1) and the pattern of encounters was different than that at monitors 2 (χ2 = 52.63, df = 3, p = 0.000), 3 (χ2 = 18.44, df = 3, p = 0.000), 4 (χ2 = 44.19, df = 3, p = 0.000), and 5 (χ2 = 7.810, df = 3, p = 0.050). although annual variability in encounters occurred at the control site monitors, no obvious pattern was evident (fig. 1). over 3 times as many encounters occurred nocturnally (n = 661) than diurnally (n = 182) at both the viewing and control sites (figs. 2 and 3). encounters at both the viewing and control sites occurred most often at 2200-2400 h and 0400-0600 h (figs. 2 and 3). diurnal visitation was low and little variation occurred among time blocks (figs. 2 and 3). the annual pattern of visitation within a 24-hour period was not different at either site (f = 0.239; df = 3, 16; p = 0.787). there was no significant change in the diurnal or nocturnal pattern of visitation when comparing 1996 data (pre-con0 20 40 60 80 100 120 140 160 1 2 3 4 5 6 7 8 9 study monitors (1-5) control monitors (6-9) m o o se e n c o u n te r s 1996 1997 1998 1999 fig 1. annual moose encounters per monitor at the viewing site (monitors 1-5) and control site (monitors 6-9), 10 june – 14 july, 1996 (pre-construction) and 1997-1999 (post-construction), dixville notch, new hampshire, usa. impacts of wildlife viewing – silverberg et al. alces vol. 38, 2002 210 fold increase at 0600-1000 h. discussion the total number of moose encounters fluctuated little at the viewing and control sites over the 4-year time period. while there was no overall effect on encounter rates at the viewing site, the decline at monitor 1, located < 10 m from the viewing blind, indicated that the presence of the blind and wildlife viewing probably caused moose to enter the lick from other trails. this type of impact could probably be minimized by considering movement patterns on individual trails in similar projects. the most active use of salt licks by moose at the control and viewing sites was at 2000-0600 h. there was no evidence moose changed their nocturnal visitation patterns as was suggested from anecdotal information from pittsburg, new hampshire, where moose viewing has been a popular pastime since the mid-1980s. it should be noted that most viewing in pittsburg occurs at night with the use of spotlights and viewing pressure is so intense on weekends that local traffic congestion is common. the general pattern of nocturnal visitation was similar to that measured at licks in pittsburg and in milan, new hampshire, 10 june14 july 1994 (adams 1995). the overall tolerance of moose to human activity was consistent with observations on shiras moose (alces alces shirasi) in yellowstone national park, where moose behavior in an area where tourists were prevalent was compared with moose behavior in an area with few people (mcmillan 1954). moose at the tourist site showed little interest in humans and appeared to tolerate their presence. similarly, the aquatic feeding behavior of moose in sibley provincial park, ontario, was only slightly affected by viewing (cobus 1972). quiet viewing in the blind produced no measurable behavioral response by moose; conversely, cars stopping alongside the lick produced an increased fleeing response (silverberg 2000). ironically, there was a striking lack of overlap between the predominant nocturnal use of licks and potential diurnal viewing opportunities. although not statistically significant, there were several interesting changes in encounter numbers relative to diurnal moose visitation at the viewing site. these included a more than 2-fold increase in the number of encounters at 0600-1000 h in 1998 and 1999, a > 50% reduction in encounters at 1600-1800 h in 1998 and 1999, and by 1999 a 33% reduction in the 3 peak visitation times measured in 1996 (fig. 2). these reductions occurred during the most popular viewing times (silverberg 2000). it is possible that moose shifted their diurnal use to avoid consistent use of the viewing blind. because the number of encounters during all diurnal periods was relatively low, slight shifts in visitation patterns could reduce viewing opportunities. unfortunately, opportunities to view moose from the blind were relatively low from 0600 to 2000 h when most visitors were present. most viewers were well aware that the best time to view moose is early morning or late in the evening. however, viewer satisfaction levels were not affected by whether they saw a moose (silverberg 2000). perhaps wildlife viewers should be informed that the best time to view moose in natural light during june and july is shortly before and after sunrise (0400-0600 h) when moose were active at licks. considering evidence from this and other studies, the impact of increased viewing during these hours should be minimal, but may warrant further monitoring. further, it is possible that by promoting early morning viewing opportunities, expectation levels of seeing a moose would increase and affect satisfaction levels. promotion of earlier viewing should also include informaalces vol. 38, 2002 silverberg et al. – impacts of wildlife viewing 211 tion about proper viewing behavior to assure that viewing impacts remain minimal. references adams, k.p. 1995. evaluation of moose population monitoring techniques and harvest data in new hampshire. m.s. thesis, university of new hampshire, durham, new hampshire, usa. altmann, m. 1958. the flight distance in free-ranging big game. journal of wildlife management 22:207-209. cobus, m.w. 1972. moose as an aesthetic resource and their summer feeding behavior. proceedings of the north american moose conference and workshop 8:244-275. fraser, d. 1979. sightings of moose, deer and bears on roads in northern ontario. wildlife society bulletin 7:181-184. goldsmith, f.b. 1974. ecological effects of visitors in the countryside. pages 217-231 in a.warren and f.b. goldsmith, editors. conservation in practice. john wiley and sons, london, england. knight, r.l., and d.n. cole. 1991. effects of recreational activity on wildlife in wildlands. transactions of the north american wildlife and natural resources conference 56:238-247. , and . 1995. wildlife responses to recreationists. pages 51 70 in r.l. knight and k.j. gutzwiller, editors. wildlife and recreationists: coexistence through management and research. island press, washington, d.c., usa. , and s.a. temple. 1995. origin of wildlife responses to recreationists. pages 81 91 in r.l. knight and k.j. gutzwiller, editors. wildlife and recreationists: coexistence through management and research. island press, washington, d.c., usa. kucera, t.e., and r.h.barrett. 1993. the trailmaster camera system for detecting wildlife. wildlife society bulletin 21:505-508. kuss, f.r., a.r. graefe, and j.j. vaske. 1990. visitor impact management, a review of the research. volume 1 and 2, national parks and conservation association, washington, d.c., usa. mcmillan, j.f. 1954. some observations on moose in yellowstone park. american midland naturalist 52: 392-399. miller, b.k., and j.a.litvaitis. 1992. use of roadside salt licks by moose, alces alces, in northern new hamps h i r e . c a n a d i a n f i e l d n a t u r a l i s t 106:112-117. signer, f.j., and j.b. beattie. 1986. the controlled traffic system and associated wildlife responses in denali national park. arctic 39: 195-203. silverberg, j.k. 1997. new hampshire wildlife viewing guide. falcon press, helena, montana, usa. . 2000. impacts of wildlife viewing: a case study of dixville notch wildlife viewing area. ph.d. dissertation, university of new hampshire, durham, new hampshire, usa. wilkes, b. 1977. the myth of the nonconsumptive user. canadian fieldnaturalist 91:343-349. alces34(2)_453.pdf alces34(2)_395.pdf alces39_131.pdf alces vol. 39, 2003 timmermann – status of moose in north america 131 the status and management of moose in north americacirca 2000-01 h. r. timmermann r. r. # 2, nolalu, on, canada pot 2ko; e-mail: ttimoose@aol.com abstract: at the turn of the century 2000, the north american moose population was estimated at about 1 million distributed in 28 jurisdictions. populations occur in 11 canadian provinces or territories, and in at least 17 u.s. states. densities are believed to be increasing in 12, stable to increasing in 14, and stable to decreasing in only 2. moose continue to expand their range in new england and several western u.s. states. in 2000-01, an estimated 382,951 licensed moose hunters harvested 82,619 moose in 23 jurisdictions, down from 418,619 and 89,027 a decade earlier. additional harvests by native and subsistence users although largely unquantified, are believed substantial in alaska, minnesota, and all 11 canadian jurisdictions. a wide variety of active and passive harvest strategies used to manage moose are discussed. population estimates are presented for 28 of 35 national parks where moose occur, but where licensed hunting is prohibited. alces vol. 39: 131-151 (2003) key words: distribution, first nations, harvest, harvest strategies, hunter numbers, license qualifications, moose population status, national parks, seasons, subsistence the status and management of north american moose (alces alces) circa 20002001 is updated from that reported by timmermann and buss (1995). a comprehensive 10-page questionnaire similar to that used by timmermann (1987) and timmermann and buss (1995) was used to update the current status, population estimates, as well as harvest and non-harvest strategies used by 23 jurisdictions that manage an annual licensed moose harvest. an additional 5 jurisdictions where hunting is currently prohibited were contacted to determine population status. tabulated data were returned for final perusal and changes or corrections solicited. this paper reports on current (year 2000-2001) population status and strategies used to manage hunting harvest and non-harvest management. historical distribution and current status the distribution of moose in north america during the latter half of the 20th century has been described by several authors including: peterson (1955), telfer (1984), kelsall (1987), karns (1998), franzmann (2000), and rodgers (2001). four subspecies are recognized, namely a. a. gigas, andersoni, americana , and shirasi (peterson 1955). during the past 30 years, kelsall and telfer (1974), karns (1998), and peek and morris (1998), detailed expanding distributions in both western and eastern states. currently, moose (a. a. americana) appear to still be expanding and re-establishing on their former range in the states of maine, vermont, new hampshire, massachusetts, new york, and connecticut (hicks 1986, alexander 1993, bontaites and guftason 1993, morris and elowe 1993, vecellio et al. 1993, al hicks, new york state department of environmental conservation, personal communication 2002, howard kilpatrick, connecticut department of environmental protection, personal communication 2002, bill woytek, massachusetts wildlife, personal commustatus of moose in north america – timmermann alces vol. 39, 2003 132 nication 2002; fig. 1). moose in vermont have re-occupied all suitable habitat and are currently believed to be increasing (cedric alexander, vermont fish and wildlife, personal communication 2001). current moose populations in maine are considered unacceptably high and need to be reduced according to karen morris (maine department of inland fisheries and wildlife, personal communication 2002). likewise, populations of a. a. shirasi continue to increase and expand in the western states of washington (donny martorello, department of fish and wildlife, washington state, personal communication 2002), as well as idaho (compton and oldenburg 1994), utah (jim karpowitz, utah division of wildlife resources, personal communication 2002), wyoming (hnilicka and zornes 1994), and colorado (kufeld 1994, kufeld and bowden 1996, john ellenberger, colorado division of wildlife, personal communication 2002). low predator densities, reduced deer populations, reversion of farmland to forest, increased logging and fire disturbance, legal protection, and conservative harvests are believed responsible (karns 1998, peek and morris 1998). minnesota closed their northwestern moose range to harvest in 1997 due to a dramatic population decline from unknown causes (mike schrage, fond du lac band of lake superior chippewa, and gretchen mehmel, minnesota department of natural resources, personal communications 2002). the estimated population declined from 4,264 in 1983 to 1,486 in 1995 to approximately 900 in 2001. a research study is currently underway in an effort to determine the causes of this decline. mainland michigan population estimates are controversial and believed to be somewhere between 600 and 1,100, while those on isle royale were estimated at 900 (aho et al. 1996, dodge et al. 2001, mary hindelang, michigan technological university, personal communication 2002). moose regularly move in and out of northern michifig. 1. 2000-2001 post-hunt moose (alces alces) population estimates for 28 north american jurisdictions. alces vol. 39, 2003 timmermann – status of moose in north america 133 gan and minnesota into northern wisconsin and adrian wydeven (wisconsin department of natural resources, personal communication 2002) estimates the current wisconsin population at 20-40. periodic winter aerial surveys based on the gasaway method are used by most agencies to estimate populations and trends (gasaway et al. 1986, peterson and page 1993, timmermann 1993, smits et al. 1994, lynch and shumaker 1995, bisset 1996, lenarz 1998, timmermann and buss 1998, bisset and mclaren 1999, bontaities et al. 2000, ward et al. 2000). most agencies estimate total jurisdictional populations based on the cumulative total of specific management areas sampled every 3 or more years. such jurisdictional estimates are considered relatively crude and are primarily used to compare population trends over time. new hampshire, maine, and vermont rely heavily on deer hunter reported moose observations and vehicle collision incidents to estimate moose trends. new hampshire and vermont also use annual deer hunter observations and a regression formula developed from concurrent infrared aerial surveys obtained over a 3-year new hampshire study to estimate regional moose populations (bontaites et al. 2000). continental moose populations in 28 jurisdictions (circa 2000-2001) are estimated at 938,350 to 1,064,130 (fig. 1). population estimates from 23 of 28 agencies which manage an annual licensed harvest are similar to those given in 1991 (table 1). in summary, current moose densities in 23 jurisdictions are believed to be stable to decreasing in minnesota, decreasing in alaska, and relatively stable or increasing in the balance (table 1). five states where hunting is prohibited report expanding populations in four (michigan, massachusetts, connecticut, new york) while those in wisconsin are considered stable (fig. 1). managing a harvest economic impact moose hunting provides a significant annual economic impact in some jurisdictions. the value of resident moose hunting in british columbia, for example was estimated to be canadian (can) $15.8 m in 1995 (reid 1997), while legg (1995) estimated can $134.7 m in ontario for all hunters in 1993. legg and kennedy (2000) estimated moose hunting in ontario contributed can $77.0 m to the gross provincial income and sustained 1,645.8 person years of employment in 1996. regelin and franzman (1998) estimated the economic impact of 33,000 resident and 1,000 nonresident alaskan hunters to represent us $32.6 m in the late 1990s. bisset (1987) reported a gross value of can $464 m generated in 19 north american jurisdictions in 1982. harvest control objectives two territories and 9 provinces in canada, and 12 states in the usa, administered a moose hunt in 2000 (table 1). collectively, 385,569 licensed moose hunters harvested an estimated 82,466 moose in 2000-2001. a decade earlier, 417,072 licensed hunters killed 89,100 (table 1). hunting regulations continue to become more restrictive and complex as the demand on moose populations and corresponding success rates increase due in part to increased road access and use of mechanized equipment (timmermann and buss 1998). control of hunting is required to affect the desired allocation of moose harvest among licensed hunters, to secure the sustainability of moose populations, and achieve other specified management objectives for a particular area. british columbia, for example, has recently suggested future management objectives focus on maintaining appropriate adult sex ratios, provide diverse hunting opportunities, and optimize recreational days status of moose in north america – timmermann alces vol. 39, 2003 134 table 1. numbers of sport hunters, harvest, and post-hunt population estimates for 23 north american jurisdictions, 1990-91 vs 2000-01. total non-resident total estimated moose hunters hunters estimated harvest population agency 1991 2001 1991 2001 1991 2001 1991 2001 yukon territory 2,040 1,410 387 342 640 716 50,000 70,000+ northwest territories 1,300 1,300 60 65 1,400 1,400 9,000 20,000+ british columbia 39,400 31,500 1,860 2,250 13,500 9,200 175,000 165,000* 4 alberta 50,000 20,429 1,150 1,139 12,200 7,971 100,700 92,000* saskatchewan 12,000 10,000 1,170 5,260 4,100 3,412 50,000 46,000* manitoba 6,500 5,409 100 100 1,100 1,000 27,000 35,000+ ontario 92,000 100,000 2,700 3,000 11,000 11,000 120,000 100,000* 5 quebec 150,000 130,000 2,500 2,000 11,900 14,000 67,500 100,000+ 6 new brunswick 5,200 4,174 — 97 1,700 2,537 20,000 25,000+ nova scotia1 200 200 — — 113 186 3,000 6,000+ newfoundland 29,200 40,449 1,400 3,044 21,000 19,322 140,000 125,000*7 alaska 22,000 30,000 2,410 3,200 6,100 5,509 155,000 120,000washington1 8 69 — — 8 64 200 1,000+3 idaho1 500 1,011 — — 490 774 5,500 15,000+ utah 299 182 290 72 290 175 2,700 3,400+ wyoming 1,713 1,379 218 199 1,475 1,215 13,645 13,865+ montana 675 609 19 16 511 596 4,000 4,000* north dakota1 110 132 — — 107 117 550 700+ colorado1 7 74 — — 7 64 425 1,070+ minnesota1 1,820 442 — — 410 125 6,700 5,100-* maine 2,000 6,000 200 300 960 2,550 23,000 29,000+ vermont — 215 — 22 — 155 1,300 3,500+ new hampshire 100 585 20 76 89 378 4,000 5,000+ total 417,072 385,569 14,484 16,117 89,100 82,466 979,220 985,635 + increase, decrease, * no change. 1 no non-resident season. 2 plus 36 permits available for resident and non-resident hunters. 3 range 850-1,000. 4 range 130,000-200,000. 5 range 100,000-110,000. 6 range 95,000-105,000 winter 2002. 7 range 115,000-140,000, with 1,000 in labrador. alces vol. 39, 2003 timmermann – status of moose in north america 135 per harvested moose. this would replace traditional objectives associated with population size, harvest, hunter numbers, and hunter days, which are difficult to achieve or measure (hatter 1999). harvest policy is currently guided by a written approved or draft moose management policy, including goals and objectives in 15 jurisdictions, while 8 employ an unwritten or generalized wildlife policy. specific moose management plans, guidelines, or statements have been prepared or are being updated in maine (morris and elowe 1993, anonymous 2000a), vermont (alexander et al. 1998, anonymous 2001), new hampshire (anonymous 1997), utah (anonymous 2000b), colorado (kufeld 1994), wyoming (wyoming game and fish commission 1990, hnilicka and zornes 1994), idaho (idfg 1990, leege 1990), ontario (omnr 1980), québec (mlcp 1993), saskatchewan (arsenault 2000), british columbia (british columbia ministry of environment, lands and parks 1996), yukon (yukon renewable resources 1996, 1999), and alberta (alberta natural resources draft pending). alaska uses a dated 1980 moose policy (alaska department of fish and game 1980) and currently manages 45 distinct populations individually. examples of specific moose plans include those for region 1, southeastern alaska, as well as the yukon flats and koyukuk river (alaska department of fish and game 1990, 2001). several agencies have recently attempted to review and evaluate their moose harvest program and policy. these include ontario (simmons 1997, prov i n c i a l a u d i t o r 1 9 9 8 , o m n r 2 0 0 1 , timmermann et al. 2003), british columbia (hatter 1999), saskatchewan (arsenault 2 0 0 0 ) , n e w f o u n d l a n d ( m e r c e r a n d mclaren 2002), québec (courtois and lamontagne 1999, lamoureux 1999, sigouin et al. 1999), and alaska (schwartz et al. 1992, hundertmark and schwartz 1996, hundertmark et al. 1998, kovach et al. 1998, regelin and franzmann 1998). allocation of hunting opportunities moose are essentially publicly owned and held in trust by provincial, territorial, and state wildlife agencies. the first priority of most agencies is to ensure the longterm conservation of moose populations and their habitats. harvest allocation is given prime consideration to subsistence use by native people under treaty or other legal agreements in at least 10 of 23 jurisdictions that manage a harvest. resident hunters are typically favored over non-residents and nonresident aliens, in allocating harvest opportunities. added controls, such as increased license fees, resident only seasons, guide requirements, and limited permits are commonly placed on non-resident hunters, giving residents priority in allocation of hunting opportunities. in 2000-2001, non-residents were eligible to hunt 17 of 23 jurisdictions (table 1). a guide was required by 8 of 23 agencies, and at least 2 agencies required non-residents to register with a licensed tourist outfitter to stimulate local economic benefits. allowances to enable some nonresidents to hunt with resident hunters have been made. for example, a non-resident of british columbia, who is a resident of canada or a canadian citizen, may be accompanied by a resident of british columbia who holds a $40.00 permit to accompany (british columbia ministry of water, land and air protection 2001). some agencies restrict or limit moose hunting opportunities. they are limited in all u.s. states except alaska. washington, north dakota, and minnesota offer 1 moose hunt per lifetime, while colorado and utah limit hunters to 1 antlered animal per lifetime. others require a waiting period between hunts; idaho and maine 2 years, new hampshire and vermont 3 years, status of moose in north america – timmermann alces vol. 39, 2003 136 wyoming 5 years, and montana 7 years after a moose is taken. hunters in alaska and all 11 canadian jurisdictions may hunt annually within quotas whether they were successful or not the previous year. ontario has introduced a pilot study in 1 wildlife management unit that offers moose hunting opportunities for physically-challenged hunters only (armstrong and simons 1999). control concepts agencies employ a variety of strategies to regulate harvests and distribute hunting pressure (timmermann 1987). passive strategies used include season length and timing, access restrictions, weapon requirements, and license qualification prerequisites; while active measures include limiting license sales or specifying the sex, age, or number of animals taken by specific area. objectives often include the harvest of predetermined numbers to sustain, increase, or decrease populations. both new hampshire and vermont have recently applied harvest rates and antlerless quotas aimed at reducing moose densities in some areas to help reduce impacts of browsing on regenerating forests and vehicle collisions (cedric alexander, vermont fish and wildlife, personal communication 2002). in 2000-01, 10 agencies offered unlimited selective or nonselective harvest opportunities while all (23 of 23) restricted or limited harvests on a selective or non-selective basis in some management areas (fig. 2). in addition, closed seasons were employed to prevent licensed harvests in specific moose inhabited areas, including some provincial, territorial, state, and national parks. alaska, for example, has eliminated or restricted any-sex seasons and now uses regulations limiting bull harvests to specific antler shape and size in much of the road accessible portions of the state (schwartz et al. 1992, hundertmark and schwartz 1996, huntertmark et al. 1998, kovach et al. 1998, regelin and franzmann 1998). license qualifications and fees in 2000, proof of hunting proficiency, including either a previous license or completing a hunter safety education course, was required to obtain a moose hunting license in all jurisdictions. in 2001, resident license fees averaged can $35.76 in canada (range $10.00 northwest territories to $ 57.50 in nova scotia and new brunswick), while non-resident licenses averaged can $204.41 (range $20.00 northwest territories to $460.00 in new brunswick). resident fees in the u.s. averaged us $106.00, (range $20.00 in north dakota to $ 310.00 in minnesota), while non-resident fees averaged us $727.70 (range $80.00 in vermont to $1,643 in idaho). some agencies, including alaska and maine for example, charged higher fees to non-resident aliens. export permits or trophy fees are required in addition to the license fee to transport an animal out of alaska, northwest territories, alberta, and ontario (table 3). currently, no jurisdictions require moose hunters to demonstrate shooting proficiency using conventional firearms, as described by buss et al. (1989). both new brunswick and newfoundland had previously required hunters to pass a shooting and written test before qualifying for a big game hunting license (timmermann and buss 1995). alaska however, requires all archery and black powder hunters to pass a proficiency test (wayne regelin, alaska department of fish and game, personal communication 2002). seasons season length and timing are used to control the amount of hunting opportunity available, hunter success due to moose vulnerability based on behavior, and seasonal access. seasons are generally specific to alces vol. 39, 2003 timmermann – status of moose in north america 137 fig. 2. moose (alces alces) harvest strategies employed by 28 north american jurisdictions (circa 2000-2001). numbers of management areas or subdivisions under each harvest strategy in each jurisdiction are indicated. 20 00 0 1 h a r v e s t s t r a t e g ie s o p e n li m it e d u n li m it e d s e le c t iv e s e x /a g e s e x a n t le r e d a n t le r le s s a d u lt / c a lf c a lf a b 1 0 9 p q 1 8: 17 1 0 m b 2 5 b c 5 5 a k 7 7 y t 3 65 __ __ p q 3 1 1 m b 1 0 s k 1 8 2 b c 10 7 4 a b 8 b c 2 6 o n 68 s e le c t iv e s e x /a g e s e x m i w i c t m a n y 12 a n t le r e d a n t le r le s s n d 1 c o 1 1 u t 1 2 7 p q 1 3 2 m b 1 5 a k 1 8 w y 2 8 n f 5 3 m t 5 9 y k 6 0 id 9 2 b c 1 16 a b 1 45 v t 2 n h 4 n f 4 n d 5 u t 5 6 a k 6 c o 9 m e 9 id 1 6 w y 2 1 m t 2 9 b c 4 4 a b 5 7 a d u lt / c a lf c a lf n o n s e le c t iv e b c 2 1 m b 5 n f 4 6 o n 6 8 3 a b 7 n s 1 p q 4 5 a k 5 m b 6 n d 7 w a 8 w y 9 v t 1 0 n h 1 8 m e 1 8 13 s k 1 9 n b 2 5 m t 2 6 m n 2 9 n f 6 0 c lo s e d n d 1 m n 2 w a 2 p q 3 o n 3 m b 3 w y 4 u t 4 a k 7 c o 1 0 m e 1 2 v t 1 5 u t 2 0 s k 2 0 y k 2 0 b c 2 5 id 3 5 a b 3 6 n o n s e le c t iv e p q 3 1 1 n w t 6 m b 8 a k 9 a b 1 5 8 s k 3 1 8 1 2 1 . c ow o r ca lf or s pi ke b ul l 2. bu ll or c al f 3. bu ll or c al f / c ow o r ca lf, p re gu n ar ch er y 26 w m u ’s 4. tr i p al m b ul l 2 7, 2 pt . b ul l 1 07 5. w ild lif e pr es er ve s 6. pl us 2 c oo p w m u ’s 7. pl us 1 7 c oo p w m u ’s 8. pr eg un a rc he ry o nl y 9. pl us 7 8 w m u ’s pr eg un a rc he ry o nl y 10 . a lte rn at e yr . 1 bu lls & c al ve s, y r. 2 bu lls , c ow s & c al ve s 11 . a lte rn at e yr . 1 bu lls o nl y, y r. 2 bu lls , c ow s, & c al ve s 12 . e nt ire s ta te c lo se d 13 . a nt le rle ss p er m its is su ed in 9 w m u ’s status of moose in north america – timmermann alces vol. 39, 2003 138 firearm type (e.g., conventional firearms, black powder, or archery). in addition, seasons tend to be longer in more remote areas and shorter in roaded areas closer to population centers. alaska provides the most liberal season length (243 days), extending from august to march in some game management areas (table 2). season lengths for all hunts in parts of idaho, wyoming, montana, yukon, northwest territories, british columbia, alberta, manitoba, ontario, québec, and newfoundland equal or exceed 3 months, while new brunswick, vermont, and new hampshire restrict season length to 3, 4, and 9 days, respectively. special early archery seasons are offered by 9 agencies (table 2), while most offer firearm seasons beginning during the latter portion of the rut period (wilton 1995) and extending into november or december. split seasons (early vs late fall) occur in at least 8 jurisdictions. management areas and harvest strategies all agencies have subdivided their moose range into various sized areas (wildlife, game, or moose management units) to facilitate specific harvest control measures. moose management areas vary in size from 53 km2 (vermont) to 1,629,049 km2 in the northwest territories, and number 4 in washington state to 445 in the yukon (table 2). all jurisdictions except the northwest territories continue to employ either a selective or non-selective limited hunter participation strategy, or a combination of both (fig. 2). most favor some form of limited selective or limited non-selective strategy to control sex and/or age related harvests. alaska alone continues to employ registration hunts which require mandatory kill registration and season termination once a prescribed harvest is achieved. a selective harvest strategy allowing control of harvest size and composition was introduced in saskatchewan in 1977 followed by british columbia, ontario, newfoundland, and québec between 1980 and 1994 (timmermann and buss 1995). this strategy’s objective is to promote herd growth by reducing the adult female harvest while maintaining or increasing adult bull and calf harvest (stewart 1978). a selective bull harvest strategy based on antler architecture is used in alaska and british columbia (child and aitken 1989, schwartz et al. 1992, timmermann and buss 1995, hatter 1999). the objective is to increase the number of bulls in areas where low bull numbers are a concern because of low reproduction, by diverting harvest pressure to young (spike and forked antlered bulls) and old bulls (antlers with >3 brow tines on 1 antler, or larger than 106 cm spread), and away from prime (6-10 year-old) bulls. québec introduced an alternating hunting strategy in 1994 by offering combinations of bull-only, bull/calf, female draw, and either sex depending on year and location (courtois and lamontagne 1997, 1999; lamoureux 1999; sigouin et al. 1999). sharing a moose between >2 hunters optimizes hunting opportunities and accommodates hunters who wish to hunt with friends. some agencies, such as minnesota since 1971, require all eligible hunters to apply together in groups of up to 4 individuals for the chance to harvest 1 animal (judd 1972). more recently, several agencies have introduced additional limiting or sharing mechanisms. british columbia offers a “group hunt” whereby up to 4 persons can combine their applications and have them entered as 1 application. if drawn, each hunter within the group receives an authorization to shoot 1 moose (british columbia ministry of water, land and air protection 2001). in addition, british columbia introduced new limited entry shared hunts in 2001. if drawn, a group of 2 is allowed to take 1 moose and a group of 3 or 4 can take alces vol. 39, 2003 timmermann – status of moose in north america 139 table 2. characteristics of moose hunting seasons in north america, 2000-2001. number of management areas season length/timing with size (km2) with open agency moose min. max. season max days earliest latest yukon territory 445 64 2,919 431 92 aug. 01 oct. 31 northwest territories 6 57,379 1,629,049 6 153 sept. 01 jan. 31 british columbia 193 465 18,980 1691 1182 aug. 15 nov. 30 alberta 148 210 33,700 1481,4 912 aug. 25 nov. 30 saskatchewan 60 2,000 120,000 401,4 642 aug. 26 nov. 30 manitoba 46 585 139,214 431,4 1172 aug. 27 dec. 22 ontario 68 1,700 85,800 681,4 88 sept. 15 dec. 15 quebec 24 2,150 225,200 211 92 sept. 01 dec. 01 new brunswick 25 826 6,402 254 3 sept. 27 sept. 29 nova scotia 25 3,000 4,400 14 12 sept. 24 oct. 13 newfoundland 64 116 4,500 641,4 1206,7 sept. 09 jan. 06 alaska 94 290 5,300 831 2432 aug. 01 mar. 31 washington 10 680 700 10 61 oct. 01 nov. 30 idaho 88 246 6,480 53 86 aug. 30 nov. 23 utah 13 1,200 8,000 9 412 sept. 09 oct. 29 wyoming 41 450 4,100 371 76 sept. 01 nov. 20 montana 80 250 2,500 80 87 sept. 01 nov. 26 north dakota 9 1,350 17,000 81 792 sept. 01 dec. 17 colorado 21 130 1,540 11 25 sept. 08 oct. 09 minnesota 31 202 772 293 16 sept. 28 oct. 20 maine 30 4,000 13,800 184 122 sept. 23 oct. 12 vermont 21 53 203 10 4 oct. 20 oct. 23 new hampshire 22 490 1,810 22 9 oct. 15 oct. 24 1 special archery seasons in some areas. 2 split seasons. 3 nw region closed since 1997. 4 closed sundays in some or all areas. 5 cape breton island and nova scotia mainland. 6 special winter season in one area. 7 labrador 197 dayssep 08-mar 16. 2 moose only. all members must hunt together and those who apply for these special limited entry shared hunts will have an advantage in the draw over a single applicant. alberta introduced a special antlered moose partner license to increase resident hunting opportunities in 2000. residents who did not apply or were unsuccessful in the license draw could partner with a resident holder of an antlered moose special license. residents who were successful could also designate a non-resident (in an area that offered a non-resident hunt) and a resident as a partner. nova scotia allows residents who have drawn a license to designate up to 2 companions who may fully participate in the hunt; provided the designated licensee is within hailing disstatus of moose in north america – timmermann alces vol. 39, 2003 140 tance of the license holder at all times and possesses a companion moose hunting stamp. manitoba issues some moose licenses on the basis of 1 tag for 2 hunters and each must sign the other’s license. if hunting alone, the licensee must be in possession of the game tag and may not sign up with another party. manitoba resident hunters may also purchase a conservation moose license together (2 licenses / 1 tag) allowing for a shared harvest of 1 moose. québec authorizes a bag limit of 1 moose per 2 hunters in most areas and 1 moose per 3 hunters in some zecs (organizations that manage specific areas). in addition, the limit is 1 moose per group composed of 3 or 4 hunters in limited access wildlife reserves. yukon hunters who wish to hunt together may apply jointly and if the application is drawn, both applicants receive a permit for the same subzone. the voluntary “hunt with a partner” slogan encourages yukoners to share 1 moose. maine, new hampshire, and vermont all manage a short season, whereby each successful permittee may select a subpermittee to hunt with them and harvest 1 moose. several agencies have developed mechanisms to grant hunters a higher success rate in a limited random draw. wyoming hunters are given 1 moose preference point every year in which they are unsuccessful in the draw. alternatively applicants may purchase a preference point for us $7.00 instead of applying for a license. successful applicants are generally those who have the highest preference points. both ontario and newfoundland either encourage or give preference to a party over individual draw applicants. newfoundland restricts party size to 2 individuals who may hunt for 1 moose provided they are within sight of each other when both are hunting. members of a party license may hunt alone provided they carry the license and tag. in addition, newfoundland gives party applicants preference over individuals and to those who were unsuccessful in previous years in each of 5 pools, thereby maximizing hunter opportunity. ontario has offered a voluntary group application system for adult moose since 1991 (timmermann et al. 2003). this system was designed to allow a more equitable allocation of harvest opportunities among more hunters. a 2-pool preference system gives hunters who were unsuccessful in obtaining a tag the previous year a preference over those who received a tag when re-applying the following year. in 2000, for example, 42% of ontario hunters applied in groups of 2 or more. the average group size was 4.43 hunters per group and 63% of groups received a moose tag compared to only 18% of individual applicants (omnr 2001:35). in addition, a tag is guaranteed to a group of hunters when the number of pool 1 hunters in the group meets a pre-determined size. harvest assessment all sources of mortality must be assessed to monitor the effectiveness of various harvest strategies. hunters are required to report their hunting activity in 9 of 23 jurisdictions, whether successful or not, while kill registration is compulsory in the majority (16 of 23, table 3). thirteen of these 16 agencies apply a non-compliance penalty to hunters failing to report, although enforcement of these requirements varies among agencies. new brunswick has experimented with interactive voice response technology (redmond et al. 1997) and alberta has used a telephone questionnaire (lynch and birkholz 2000) to help assess moose harvests. modeling is also used to predict population changes resulting from various harvest strategies (heydon et al. 1992, schwartz 1993, mckenney et al. 1998). timmermann and buss (1998) provide a more detailed description of this subject. alces vol. 39, 2003 timmermann – status of moose in north america 141 harvest by native and subsistence u s e r s kay (1997) suggested that historically moose were extremely vulnerable to predation by natives in western north america and that native peoples had no effective conservation practices. reeves and mccabe (1998) estimated annual consumption of moose for north american indians living in moose range to be 0.142 moose per person. currently, most north american moose management agencies give primary consideration to subsistence use by canadian first nation peoples and native american peoples in recognition of obligations made under historical treaties signed by both federal governments (crichton et al. 1998). in many areas, they have unfettered access to moose year round and current regulations are considered liberal and unrestrictive given the widespread use of modern technology (courtois and beaumont 1999). the harvest by natives is difficult to quantify and unfortunately little effort has been made to measure the magnitude of this harvest, which some managers believe approaches or exceeds the licensed harvest. in the usa, 4 of 12 agencies reported formal agreements governing moose harvests have been signed with some tribal bands. they include montana, utah, maine, and minnesota. the latter state has signed agreements with 2 ojibwe bands, another is being negotiated and 2 minnesota bands have closed seasons on their reserves due to low populations. schrage (2001) reported 80 moose taken by minnesota natives compared to 125 by all non-natives in 2001. bands in montana, utah, and maine regulate harvests on tribal lands. in canada, first nations have signed a few formal agreements with 5 of 11 jurisdictions. they include the yukon and northwest territories, british columbia, ontario, and québec. in the yukon, those agreements have yet to be implemented and managers currently estimate harvest levels to equal or exceed those of the licensed harvest based on limited data. the northwest territories land claims agreement governs subsistence harvesting by first nations in the inuvialuit, gwich’in, and sahtu areas (marshal 1999). british columbia, ontario, and québec have signed several agreements, while others are being negotiated, and many jurisdictions have no agreements in place. however all current harvests by first nation peoples are poorly documented. in ontario, the only formal agreement was a 10-year history of annual agreements with the algonquins of golden lake to take moose in algonquin provincial park from 1990-2000. no agreement was signed in 2001 and documentation of kill magnitude was difficult to obtain under previous agreements. the annual moose harvest by first nation peoples is “substantial” in specific local areas of british columbia and ontario. moose managers in ontario estimate the harvest by first nation and metis peoples may approach the licensed hunting harvest for some wildlife management units in northwestern and northeastern ontario in areas adjacent to first nations communities (ted armstrong and peter davis, ontario ministry of natural resources, personal communications 2002). first nation moose harvests are believed to equal or exceed the total licensed harvest in alberta (7,971 +), at least 50% of the licensed harvest in saskatchewan (1,706+), equal to double in manitoba (1,000-2,000), slightly more than half in new brunswick (1,300+), at least double in nova scotia (360+), at least 1,000 in the northwest territories, and is thought to equal or exceed harvests by non-firstnation peoples in the yukon (743). there are no reliable statistics on harvests by first nations in québec, except in the james bay region where their harvests are considered status of moose in north america – timmermann alces vol. 39, 2003 142 table. 3. moose harvest assessment strategies used in north america, 2000-2001. hunt activity report kill registration non-compliance agency compulsory voluntary compulsory voluntary penalty4 yukon territory x x fine northwest territories x x2,3 fine british columbia x x1 fine alberta x x2 none saskatchewan x x n/a manitoba x x n/a ontario x x2,3 n/a quebec x x fine & loss of license new brunswick x x fine & loss of license nova scotia x x fine & jail & 7 yr. wait-out newfoundland x x fine alaska x1 x1,2 none washington x x none idaho x x fine, jail & loss of license utah x x n/a wyoming x x none montana x x n/a north dakota x x none colorado x x ineligible for draw minnesota x x fine & loss of license maine x x fine & jail & loss of license vermont x x fine & loss of license new hampshire x x fine, jail & loss of license 1 limited draw hunts only (british columbia – regions 3,4,5,6,7a,8-incisor and kill information). 2 export permit/trophy fee. 3 non-resident hunter only. 4 variable enforcement. of the same order or greater than that of licensed hunting (st.-pierre 2001, réhaume courtois, québec ministère de l’environment et de la faune, personal communication 2002). newfoundland has only 1 first nation reserve and currently no allowances are made for harvesting of moose by their people. first nation use of moose in labrador is limited due to the low moose population (paul saunders, newfoundland and labrador department of tourism, culture and recreation, personal communication alces vol. 39, 2003 timmermann – status of moose in north america 143 2002). moose managers generally presume that first nation peoples take a higher proportion of cows than bulls, although such data are speculative and poorly documented. both metis and non-status indians are testing their perceived rights in court. metis are considered as any persons of mixed indian and white ancestry not considered an indian (swail 1996). in ontario, self-identified metis are considered to be members of and accepted by their local metis community and organization, which retain a historic metis community connection in areas where moose hunting is considered a historic activity (richard stankiewicz, ontario ministry of natural resources, personal communication 2002). on february 23, 2001, the ontario court of appeal decided the case of r. v. powley et al. (2001), 53 o.r. (3rd) 35, ruling in favor of 2 metis who claimed moose hunting was an “integral practice, custom or tradition of that metis community”. this case is currently on appeal to the supreme court of canada (omnr 2002). another judge in an alberta case (the crown vs fergeson) ruled that metis have the right to hunt anywhere where they have right of access at any time without a license, provided they were raised in the indian culture (could speak “indian”, grew up hunting and trapping, etc.), according to lynch (alberta wildlife management consulting, personal communication 2002). in canada, first nation peoples are restricted to their treaty areas with respect to unlicensed harvest. harvest of wildlife by first nation peoples and metis, including moose, remains a controversial subject and is considered a substantial undocumented kill in most jurisdictions (crichton 1981, feit 1987, kay 1997, hatter 1999). co-management between government agencies, first nations, and metis is believed by some managers to offer the potential for local control of the moose resource, as long as hunting rights are balanced with conservation efforts (feit 1987, nepinak and payne 1988, graf 1992, messier 1996, crichton et al. 1998, marshal 1999, arsenault 2000, crichton 2001). crichton (2001) offers 4 ingredients for successful co-management. first nations must have: a decision making role in development of management programs; be supportive of partnerships; there be recognition of traditional cultural and economic values, including a removal of cultural and linguistic barriers to facilitate use; and, a dispute resolution process to resolve disagreements. regelin and franzmann (1998) reported that new laws in alaska, primarily the alaska native claims settlement act, have dictated a priority for harvest by rural citizens, using subsistence regulations to redistribute harvest among users. this law also shifted management responsibility from the state to the federal government. under this law, all alaskan residents are potentially qualified as subsistence hunters. as such, there is potential for subsistence use to increase significantly (alaska department of fish and game 2001). alaska estimated a subsistence harvest of 2,000 compared to 5,000 for the 2001 licensed harvest. regional managers in british columbia may issue a possession permit for the purpose of sustenance, while local fish and wildlife offices in alberta report issuing about 100 subsistence licenses for “those on the land” (ian hatter, british columbia wildlife branch, and gerry lynch, alberta wildlife management consulting, personal communications 2001). future sustainable harvests and population goals will largely remain elusive until the total harvest, including harvests by first nation and metis peoples and subsistence users, are agreed to and are verifiable. illegal hunting losses (timmermann and buss 1998) appear to be significant in some jurisdictions, including colorado (kufeld 1994), utah (anonymous 2000b), and onstatus of moose in north america – timmermann alces vol. 39, 2003 144 tario (harnish 2000). most agencies encourage all hunters to report illegal infractions using a toll-free telephone number. ontario has recently introduced a “moose watch” program to help reduce moose poaching (harnish 2000). managing a nonharvest parks, refuges, and special areas most north american jurisdictions where moose occur, provide for areas where hunting is not a primary management objective. currently 5 u.s. states do not have an open moose hunting season and 19 of 23 jurisdictions provide closed seasons in anywhere from 1 to 36 management areas (fig. 2). the assumed common objective of closed areas including game or wildlife reserves, national, provincial, territorial, or state parks, or nature reserves, is the preservation of moose in representative natural habitats for education and recreational enj o y m e n t . f u r t h e r m a i n t e n a n c e o f biodiversity and ecosystem function is often a stated objective. only 6 of 23 responding jurisdictions indicated management considerations or special objectives had been developed for pro-active management in such protected areas. the provision of viewing opportunities and natural history interpretations were commonly integrated in their park’s programs. a review of moose management objectives and programs in parks, refuges and special areas is detailed by timmermann and buss (1995). moose are native to at least 35 north american national parks in 16 jurisdictions (table 4). isle royale is perhaps the most famous, boasting a 43-year continuous ecological study of wolves and moose beginning in 1959 (peterson 2002). jordan et al. (2000) summarized moose related studies and provided an extensive list (150+) of research papers. a sampling includes a report on osteoporosis and other skeletal pathologies by hindelang et al. (1992), studies on tooth wear and perodontal disease by hindelang and peterson (1993, 2001), and the impact of wolves and moose on vegetative succession by mclaren and peterson (1994, 1995). several other national parks have also yielded moose related research. they include: data on 151 moose (a. a. andersoni) weights and measurements from elk island national park, alberta, following a herd reduction program (lynch et al. 1995); ecological status of moose and white-tailed deer in voyageurs national park, minnesota (gogan et al. 1997); and, a description of extreme moose demographics in gros morne national park, newfoundland (mclaren et al. 2000). population estimates vary from unknown in several alberta and alaskan national parks to as high as 7,738 in gros morne national park, newfoundland. bisset (1987) estimated that the value associated with wildlife appreciation (non-consumptive use, vicarious recreation, etc.) could have been as much as can $1,623 m in the early 1980s. acknowledgements appreciation is extended to the following individuals who provided unpublished information in response to a 10-page questionnaire survey: wayne regelin, doug larsen, and kris hundertmark, alaska department of fish and game, juneau, anchorage, and soldotna, ak; donny martorello, department of fish and wildlife, washington state, olympia, wa; bradley compton, idaho fish and game, idaho falls, id; jim karpowitz, utah division of wildlife resources, price, ut; kerry olson, justin binfet, and tim thomas, wyoming game and fish, cheyenne and sheridan, wy; jerry brown, montana fish, wildlife and parks, libby, mt; william jensen, north dakota game and fish department, bismark, nd; john ellenberger, colorado division of wildlife, grand junction, co; mike schrage, fond du lac band alces vol. 39, 2003 timmermann – status of moose in north america 145 of lake superior chippewa, conservation department, cloquet, mn; gretchen mehmel, minnesota department of natural resources, roosevelt, mn; karen morris, maine department of inland fisheries and wildlife, bangor, me; kristine bontaites, new hampshire fish and game, new hampton, nh; cedric alexander, vermont fish and wildlife, st. johnsbury, vt; howard kilpatrick, north franklin, ct; bill woytek, massachusetts wildlife, westborough, ma; rick ward, yukon department of renewable resources, whitehorse, yk; alasdair veitch, wildlife management division, department of renewable resources, yellowknife, nwt; ian hatter, british cotable 4. moose population status in north american national parks (n.p.). location population year of estimated jurisdiction/park state/province estimate survey by acadia n.p. maine 6 2000 guess banff n.p. alberta — — — beringland bridge preserve alberta — — — cape breton highlands n.p. nova scotia 2,500 2001 aerial survey denali n.p. & preserve alaska 2,000 1990 aerial survey elk island n.p. alberta 400 2002 guess forillion n.p. quebec 122 1997 aerial survey fundy n.p. new brunswick 123 1993 aerial & ground gates of arctic n.p. & preserve alaska — — — glacier n.p. montana 100 1985 — grand teton n.p. wyoming 120 1988 — gros morne n.p. newfoundland 7,738 1995 aerial survey isle royale n.p. michigan 900 2001 aerial survey ivvavik n.p. yukon territory 300 — guess jasper n.p. alberta 100-150 1992 ground survey kenai national wildl. refuge alaska — — — kejimkujik n.p. nova scotia 0-5 2001 guess kluane n.p. yukon territory 700 1997 aerial survey kootenay n.p. british columbia 75 1985 cws biologists kouchibouguac n.p. new brunswick 110 1995 aerial survey lac mauricie n.p. quebec 212 1989 — lake clark n.p. & preserve alaska — — — mt. revelstoke & glacier n.p. british columbia 15-20 1991 ground survey nahanni n.p. northwest territories 300 1979 aerial survey noatak preserve alaska — — — prince albert n.p. saskatchewan 950+ 1997 aerial survey pukaskwa n.p. ontario 411±143 1999 aerial survey riding mountain n.p. manitoba 5,000 2000 aerial survey terra nova n.p. newfoundland 150-200 2002 guess voyageur n.p. minnesota 80-100 1998 aerial survey vuntut n.p. yukon territory 875 — guess waterton lakes n.p. alberta 50 1988 aerial & ground wood buffalo n.p. alberta 1,300 1989 — wrangell st. elias n.p. alaska — — — yellowstone n.p. montana 200 1990 aerial survey status of moose in north america – timmermann alces vol. 39, 2003 146 lumbia, wildlife branch, victoria, bc; gerry lynch, alberta wildlife management consulting, sherwood park, ab; rhys beaulieu, saskatchewan environment and resources management, meadow lake, sk; vince crichton, wildlife and ecosystem protection branch, manitoba conservation, winnipeg, mb; rick stankiewicz, jim saunders, and howard smith, ontario ministry of natural resources, wildlife section, peterborough, on; gilles lamontagne and réhaume courtois, québec ministère de l’environment et de la faune, québec city, pq; gerry redmond, new brunswick maritime forest ranger school, fredericton, nb; tony nette, nova scotia department of natural resources, kentville, ns; paul saunders and brian mclaren, newfoundland and labrador department of tourism, culture and recreation, inland fish and wildlife division, corner brook, nf. in addition, the following provided written and verbal information on moose status in non hunted jurisdictions, clarification on specific questions, assistance in questionnaire mailouts, and preparing tables and figures: mary hindelang, chassell, mi; rolf peterson, michigan technological university, houghton mi; adrian wydeven, wisconsin department of natural resources, park falls, wi; al hicks, new york state; howard kilpatrick, north franklin, ct; bill w o y t e k , m a s s a c h u s e t t s w i l d l i f e , westborough, ma; marty orwig, north american moose foundation, mackay, id; keith wade, pukaskwa national park, marathon, on; al franzmann, soldotna, ak; art rodgers, centre for northern forest ecosystem research, ontario ministry of natural resources, thunder bay, on; susan rodgers, thunder bay on; bruce ranta, ted armstrong, and peter davis, ontario ministry of natural resources, kenora, thunder bay, and cochrane, on; and, brian hutchinson, parks canada, cornwall, on. references aho, r. w., s. m. schmitt, j. hendrickson, and t. r. minzey. 1996. michigan’s translocated moose population: 10 years later. michigan department of natural resources, report number 3245. wildlife division, lansing, michigan, usa. alaska department of fish and game. 1980. alaska wildlife management plans. species management policies. juneau, alaska, usa. . 1990. strategic plan for management of moose in region 1, southeast alaska. 1990-94. alaska department of fish and game, douglas, alaska, usa. . 2001. koyukuk river moose management plan 2000-2005. alaska department of fish and game, juneau, alaska, usa. alexander, c. e. 1993. the status and management of moose in vermont. alces 29:187-195. , p. fink, l. garland, and f. hammond. 1998. moose management plan for the state of vermont, 19982007. agency of natural resources, department of fish and wildlife, waterbury vermont, usa. anonymous. 1997. managing deer, moose and bear in new hampshire 1997-2005. wildlife division, new hampshire fish and game department, concord, new hampshire, usa. . 2000a. maine moose goals and objectives 2000-2010. maine department of inland fisheries and wildlife. bangor, maine, usa. . 2000b. utah division of wildlife resources statewide management plan for moose. utah department of natural resources, division of wildlife resources. salt lake city, utah, usa. . 2001. vermont moose hunter’s guide. vermont fish and wildlife dealces vol. 39, 2003 timmermann – status of moose in north america 147 partment, st. johnsbury, vermont, usa. armstrong, e. r., and r. simons. 1999. moose hunting opportunities for physicallychallenged hunters in ontario: a pilot study. alces 35:125-134. arsenault, a. a. 2000. status and management of moose (alces alces) in saskatchewan. saskatchewan environment and resource management, fish and wildlife branch. fish and wildlife technical report 00-1. saskatoon, saskatchewan, canada bisset, a. r. 1987. the economic importance of moose (alces alces) in north america. swedish wildlife research supplement 1:677-698. . 1996. standards and guidelines for moose population inventory in ontario. ontario ministry of natural resources, fish and wildlife branch, peterborough, ontario, canada. , and m. a. mclaren. 1999. moose population aerial inventory plan for ontario: 1999-2002. ontario ministry of natural resources, northwest science and technology. information report ir-004. thunder bay, ontario, canada. bontaites, k. m., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29:163-167. , , and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36:6975. british columbia ministry of environment, lands and parks. 1996. wildlife harvest strategy, improving british columbia’s wildlife harvest regulations. wildlife branch, victoria, british columbia, canada. british columbia ministry of water, land and air protection. 2001. hunting and trapping regulations synopsis / 2001-02. monday tourism publications, victoria, british columbia, canada. b u s s , m . , r . g o l l a t , a n d h . r . timmermann. 1989. moose hunter shooting proficiency in ontario. alces 25: 98-103. child, k. n., and d. a. aitken. 1989. selective harvests, hunters and moose in british columbia. alces 19:162-177. compton, b. b., and l. e. oldenburg. 1994. the status and management of moose in idaho. alces 30:57-62. courtois, r., and a. beaumont. 1999. the influence of accessibility on moose hunting in northwestern québec. alces 35:41-50. , and g. lamontagne. 1997. management system and current status of moose in québec. alces 33:97-114. , and . 1999. the protection of cows: its impact on moose hunting and moose populations. alces 35:1129. crichton, v. f. j. 1981. the impact of treaty indian harvest on a manitoba moose herd. alces 17:56-63. . 2001. co-management the manitoba experience. alces 37:163173. , w. l. regelin, a. w. franzmann, and c. c. schwartz. 1998. the future of moose management and research. pages 655-663 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. dodge, w. b., s. r. winterstein, d. e. beyer, jr., and h. r. campa. 2001. why aren’t there more moose in michigan’s upper peninsula? michigan state university, department of fisheries and wildlife. lansing, michigan, usa. feit, h.a. 1987. north american native hunting and management of moose status of moose in north america – timmermann alces vol. 39, 2003 148 populations. swedish wildlife research supplement 1:25-42. franzmann, a. w. 2000. moose. pages 578-600 in s. demarais and p. r. krausman, editors. ecology and management of large mammals in north america. prentice hall, upper saddle river, new jersey, usa. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, number 22. fairbanks, alaska, usa. gogan, j. p., k. d. kozie, e. m. olexa, and n. s. duncan. 1997. ecological status of moose and white-tailed deer at voyageurs national park, minnesota. alces 33:187-201. graf, r. p. 1992. status and management of moose in the northwest territories, canada. alces supplement 1:22-28. harnish, d. 2000. moose watch reduces illegal hunting in northeast region. news release, december 8/00. ontario ministry of natural resources, sault ste marie ontario, canada. hatter, i. w. 1999. an evaluation of moose harvest management in central and northern british columbia. alces 35:91-103. heydon, c., d. euler, h. smith, and a. bisset. 1992. modeling the selective moose harvest program in ontario. alces 28:111-121. hicks, a. c. 1986. the history and current status of moose in new york. alces 22:245-252. hindelang, m., r. o. peterson. 1993. relationship of mandibular tooth wear to gender, age, and periodontal disease of isle royale moose. alces 29:63-73. , and . 2001. skeletal integrity in moose at isle royale national p a r k : b o n e m i n e r a l d e n s i t y a n d osteopathology related to senescence. alces 36:61-68. , , and a. l. maclean. 1992. osteoporosis in moose on isle royale: a pilot study of bone mineral density using ct scans. alces 28:35-39. hnilicka, p., and m. zornes. 1994. status and management of moose in wyoming. alces 30:101-107. hundertmark, k. j., and c. c. schwartz. 1996. considerations for intensive management of moose in alaska. alces 32:15-24. , t. h. thelen, and r. t. bowyer. 1998. effects of population density and selective harvest on antler phenotype in simulated moose populations. alces 34:375-383. (idfg) idaho department of fish and game. 1990. species management plan 1991-1995: moose. idaho department of fish and game. boise, idaho, usa. jordan, p. a., b. e. mclaren, and s. m. sell. 2000. a summary of research on moose and related ecological topics at isle royale, usa. alces 36:233-267. judd, s. l. 1972. minnesota’s 1971 moose season. proceedings of the north american moose conference and workshop 8:240-243. karns, p. d. 1998. population distribution, density and trends. pages 125-139 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. kay, c. e. 1997. aboriginal overkill and the biogeography of moose in western north america. alces 33:141-164. kelsall, j. p. 1987. the distribution and status of moose (alces alces) in north america. swedish wildlife research supplement 1:1-10. , and e. s. telfer. 1974. biogeography of moose with particular refere n c e t o w e s t e r n n o r t h a m e r i c a . alces vol. 39, 2003 timmermann – status of moose in north america 149 naturaliste canadien 101:117-130. kovach, s. d., c. c. schwartz, r. l. willis, and t. h. spraker. 1998. modeling moose populations for management decision making in alaska. alces 34:125-138. kufeld, r. c. 1994. status and management of moose in colorado. alces 30:41-44. , and d. c. bowden. 1996. survival rates of shiras moose (alces alces shirasi) in colorado. alces 32:9-13. lamoureux, j. 1999. effects of selective harvest on moose populations of the bas-saint-laurent region, québec. alces 35:191-202. leege, t. a. 1990. moose management plan 1991-1995. idaho department of fish and game. boise, idaho, usa. legg, d. 1995. the economic impact of moose hunting in ontario, 1993. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. , and m. kennedy. 2000. the economic impact of moose hunting in ontario, 1996. analysis and planning section, land use planning branch, ontario ministry of natural resources, peterborough, ontario, canada. lenarz, m. s. 1998. precision and bias of aerial moose surveys in northeastern minnesota. alces 34:117-124. lynch, g. m., and s. birkholz. 2000. a telephone questionnaire to assess moose harvest. alces 36:105-109. , b. lajeunesse, j. willman, and e. s. telfer. 1995. moose weights and measurements from elk island national park, alberta, canada. alces 31:199207. , and g. e. shumaker. 1995. gps and gis assisted moose surveys. alces 31:145-151. marshal, j. p. 1999. co-management of moose in the gwich’in settlement area, northwest territories. alces 35:151158. mckenney, d. w., r. s. rempel, l. a. venier, yonghe wang, and a. r. bisset. 1998. development and application of a spatially explicit moose population model. canadian journal of zoology 76:1922-1931. mclaren, b. e., c. mccarthy, and s. mahoney. 2000. extreme moose demographics in gros morne national park, newfoundland. alces 36:217232. , and r. o. p e t e r s o n. 1994. wolves, moose, and tree rings on isle royale. science 266:1555-1558. , and . 1995. seeing the f o r e s t w i t h t h e t r e e s : u s i n g dendrochonology to investigate mooseinduced changes to a forest understory. alces 31:77-86. mercer, w. e., and b. e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose. alces 38:123141. messier, f. 1996. moose co-management in the trilateral agreement territory: principles and recommendations based on scientific knowledge and aboriginal rights. report to the algonquins of barriere laketrilateral secretariat, hull, québec, canada. morris, k., and k. elowe. 1993. the status of moose and their management in maine. alces 29:91-97. (mlcp) ministere du loisir, de la chasse et de pêche. 1993. plan de gestion de l’orignal, 1994-1998. objectifs de gestion et scénarios d’exploitation. éditeur officiel du québec, québec, canada. nepinak, h., and h. payne. 1988. the hunting rights of indian people in manitoba: an historic overview and a contemporary explication towards enhanced conservation through joint management. status of moose in north america – timmermann alces vol. 39, 2003 150 alces 24:195-200. (omnr) ontario ministry of natural resources. 1980. moose management policy. queen’s printer for ontario, toronto, ontario, canada. . 2001. 2000 review of moose population objectives in ontario. draft report of 3 regional workshops. ontario ministry of natural resources, wildlife branch, peterborough, ontario, canada. . 2002. metis and non-status rights requests. febuary 2002 news release, ontario ministry of natural resources, peterborough, ontario, canada. peek, j. m., and k. i. morris. 1998. status of moose in the contiguous united states. alces 34:423-434. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. peterson, r. o. 2002. ecological studies of wolves on isle royale. 2001-2002. school of forestry and wood products, michigan technological university, houghton, michigan, usa. , and r. e. page. 1993. detection of moose in midwinter from fixed-wing aircraft over dense forest cover. wildlife society bulletin 21:80-86. provincial auditor. 1998. audit of the ministry of natural resources, fish and wildlife program. toronto, ontario, canada. (http://www.gov.on.ca/ opa/english/e98/309.htm). redmond, g. w., a. arseneault, and c. lanteigne. 1997. using technology to survey moose hunters in new brunswick. alces 33:75-83. reeves, h. m., and r. e. mccabe. 1998. of moose and man. pages 1-74 in a. w. franzmann and c. c. schwartz, editors. ecology and management of t h e n o r t h a m e r i c a n m o o s e . smithsonian institution press, washington, d.c., usa. regelin, w. l., and a. w. franzmann. 1998. past, present, and future moose management and research in alaska. alces 34:279-286. reid, r. 1997. the economic value of resident hunting in british columbia. 1995. ministry of environment, lands and parks, wildlife branch, victoria, british columbia, canada. rodgers, a. r. 2001. moose. voyageur press, stillwater, minnesota, usa. schrage, m. 2001. status of minnesota’s moose populations, seasons and harvest: 2001. fond du lac resource management division, 1720 big lake road, cloquet, minnesota, usa. schwartz, c. c. 1993. constructing simple population models for moose management. alces 29:235-242. , k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula, alaska. alces 28:1-13. sigouin, d., s. st.-onge, r. courtois, and j.-p. ouellet. 1999. change in hunting activity and hunters’ perceptions following the introduction of selective harvest in québec. alces 35:105-123. simmons, g. 1997. independent review of the moose and deer tag allocation for ontario. recommendations from ontario’s hunters. queens printer, toronto, ontario, canada. smits, c. m. m., r. m. p. ward, and d. g. larsen. 1994. helicopter or fixedwing aircraft; a cost-benefit analysis for moose surveys in yukon territory. alces 30:45-50. stewart, r. r. 1978. introduction of sex and age specific hunting licenses for the moose harvest in saskatchewan. proceedings of the north american moose conference and workshop 14:194-208. s t . p i e r r e , d . 2 0 0 1 . a n a l y s e e t interprétation des résultats de la saison alces vol. 39, 2003 timmermann – status of moose in north america 151 de chasse 1999 pour la zone 17 et 22. pages 201-204, 221-225 in c. daigle, editor. compte rendu de l’atelier sur la grande faune – 2000 et bilan de récolte des grands gibiers 1999-2000. société de la faune et des parcs du québec, québec city, québec, canada. swail, p. j. 1996. blais vs the queen. a conviction under the criminal code of canada on august 22, 1996, for hunting big game out of season contrary to section 26 of the manitoba wildlife act. http://www.canlii.org/ca/cas/scc/ 2003/2003scc44.html. telfer, e. s. 1984.circumpolar distribution and habitat requirements of moose (alces alces). pages 145-182 in r. olson, r. hastings, and f. geddes, editors. northern ecology and resource management. university of alberta press, edmonton, alberta, canada. timmermann, h. r. 1987. moose harvest strategies in north america. swedish wildlife research supplement 1:565579. . 1993. use of aerial surveys for estimating and monitoring moose populationsa review. alces 29:35-46. , and m. e. buss. 1995. the status and management of moose in north americaearly 1990s. alces 31:1-14. , and . 1998. population and harvest management. pages 559615 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. , r. gollat, and h. a. whitlaw. 2003. reviewing ontario’s moose management policy— 1980-2000: targets achieved, lessons learned. alces 38: 11-45. vecellio, g. m., r. d. deblinger, and e. cardoza. 1993. status and management of moose in massachusetts. alces 29:1-7. ward, m. p., w. c. gasaway, and m. m. dehn. 2000. precision of moose density estimates derived from stratification survey data. alces 36:197-203. wilton, m. l. 1995. the case against calling and hunting dominant moose during the main rut period a viewpoint. alces 31:173-180. wyoming game and fish commission. 1990. a strategic plan for the comprehensive management of wildlife in wyoming 1990-1995. cheyenne, wyoming, usa. yukon renewable resources. 1996. moose management guidelines. fish and wildlife branch, department of renewable resources, whitehorse, yukon territory, canada. . 1999. yukon moose. fish and wildlife branch, department of renewable resources, whitehorse, yukon territory, canada. f:\alces\vol_38\pagema~1\3802.pdf alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 11 reviewing ontario’s moose management policy 1980-2000 targets achieved, lessons learned h.r. timmermann1, r. gollat2, and h. a. whitlaw3 1r.r. # 2 nolalu, on, canada pot 2ko; e-mail: ttimoose@aol.com; 2ontario ministry of natural resources, thunder bay, on, canada p7c 5g6; 39305 winston ave., lubbock, tx 79424, usa abstract: we examine progress made in meeting the 1980, 20year ontario moose management policy (mmp) directive. specific interim (1985, 1995) and final (year 2000) provincial program targets, including population, harvest, hunting, and viewing opportunities, particularly those in the nw region, are reported. in addition to mmp guidelines, other management policy achievements and shortfalls pertaining to harvest control, population management, enforcement, habitat management, inventory and assessment, research, and hunter education are discussed. provincially, moose numbers have increased only 7-20% throughout the 1990s plateauing at 100,000120,000 while the number of adult tags has almost been halved. hunter numbers during this period have increased by about 4% and total harvest has remained fairly constant. adult tag draw success has declined and success in filling a tag has increased while harvest remained similar in absolute numbers. this suggests that factors other than hunting pressure are limiting further population growth. knowledge gained since 1980 suggests overall population and harvest targets are unattainable and should be revised using adaptive management principles, to more closely reflect land capability and societal demands. reduced hunter reporting rates in recent years have jeopardized the quality of harvest estimates and diminished overall hunter confidence. recommendations for policy changes, including revisions to program direction and targets, are made based on lessons learned. alces vol. 38: 11-45 (2002) key words: enforcement, habitat management, harvest targets, hunter education, inventory and assessment, management policy, population targets, predator control, program targets specific goals and objectives to guide the development of management plans are employed by most jurisdictions that manage moose in north america (timmermann and buss 1995). policy objectives generally relate to maintaining or increasing moose populations and the recreational, social, and economic benefits associated with a harvest, as well as gaining new knowledge about moose ecology. management policies with specific goals and objectives typically guide moose management plans over a 5 year period (timmermann and buss 1998). ontario introduced a comprehensive moose management policy in 1980 following a series of public meetings (omnr 1980a) and in response to declining moose populations and related recreational and economic benefits (omnr 1980b). the cabinet approved policy included a broad program objective, 4 program targets, 14 policy guidelines, and 15 management goals spanning a 20 year horizon (omnr 1980b). specific program targets included: (1) increase the herd from 80,000 to 100,000 by 1985, 140,000 by 1995, and 160,000 by the year 2000; (2) harvest 10,000 moose per year after 1985, 18,000 per year by 1995, and 25,000 per year by 2000; (3) provide for ontario’s moose management policy timmermann et al. alces vol. 38, 2002 12 750,000875,000 hunter days annually by 2000; and (4) create sites by the year 2000 where more than 1 million people annually have the opportunity to observe moose. the new policy prompted many changes in moose management including the requirement of sharing a moose during1980-82 (timmermann and gollat 1984) and establishing parameters for selective harvesting of moose beginning in 1983 (euler 1983, g o l l a t a n d t i m m e r m a n n 1 9 8 3 , timmermann and gollat 1986, smith 1990, heydon et al.1992, timmermann and whitlaw 1992). the selective harvest system introduced in 1983 proportioned the allowable harvest among the tourist industry on a 90/10 percent (non tourist / tourist industry) basis provincially (bisset and timmermann 1983). this allocation was based on historic use and capacity of individual outfitters and may vary (above or below 10%) by wildlife management unit (wmu). year 2000 moose population density targets for each wmu were originally set out in the northwestern ontario strategic land use plan (omnr 1982). three population densities were targeted; 0.05-0.11/ km2 (wmus 1c and 17), 0.15-0.35/km2 (wmus 16a, 16b, 16c, 18a, 18b), and 0.39/km2 for the remaining 21 wmus (fig.1). these population targets were based on the belief that moose populations with good habitat (i.e., quetico provincial park and the chapleau crown game reserve) supported densities of about 0.40/km2 (bisset 1992). likewise, year 2000 sport harvest rates were established at 17.5% per year for all but 4 northern wmus, where they varied between 12.3 and 15.0% 1c 17 16a 16c 18a 18b 16b 4 3 5 15a 15b 19 21a 21b 1413 8 9a 12a 7b 9b 2 6 7a 12b 11b 11a fig. 1. location of 28 wildlife management units (wmus) in ontario's northwest region (shaded), in relation to provincial wmus used to manage moose in ontario. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 13 (omnr 1982). this provincial policy guided field offices, which developed more specific approaches that were appropriate for their area. some regions targeted a faster rate of population increase by limiting adult hunting opportunities (number of adult validation tags — avts), while others, opted for a stable, or slower rate of increase, and provided more avts. each wmu harvest quota was apportioned into a specific number of bulls, cows, and calves beginning in 1983. the applied 1983 ratio of 50:20:30 was refined to a baseline 60:20:20 in 1984, based on a computer simulation population model (omnr 1984, gollat et al. 1985). annual harvest quotas were set for each wmu based on the number of adult bulls and cows that could be taken and still allow the herd to increase or remain stable, depending on population status relative to year 2000 targets. these were calculated based either on a percentage of adult cows in the population or a harvest rate applied to the total pre-hunt population estimate (i.e., 10-15%; greenwood et al. 1984). adult bull and cow harvest opportunities, governed by the number of avts, were restricted during most of the 18 year period. a 3 year moving average of hunter success rates was used initially to govern the number of bull and cow avts. a regression-based success rate projection was used more recently to calculate the number of avts. managers also targeted 67 bulls /100 cows as a minimum post-hunt adult bull/cow sex ratio in an attempt to increase populations and optimize harvests to year 2000 targets (crête et al. 1981). the objective of this paper is to exami n e p r o g r e s s m a d e t o w a r d s m e e t i n g the 1980 moose management policy (mmp) (omnr 1980b). it is not a detailed analysis of mmp objectives. our intent is to provide an overview of accomplishments and shortfalls that hopefully will lead to a comprehensive program review. we report on specific targets achieved and lessons learned, particularly in ontario’s northwest region (nwr); an area that contains more than one half of the provincial moose population and annual harvest (whitlaw et al. 1993). in addition, we chronicle a partial list of published material relating to meeting the 20year (1980-2000) provincial moose management policy objectives. methods population and harvest estimates were compiled from a variety of published and unpublished ontario ministry of natural resources (omnr) sources, generally spanning the period 1980-2000. detailed data for 28 of 67 provincial wmus from the current nwr (fig. 1) were examined. moose population density estimates were obtained from aerial inventories conducted approximately every 3 years in most wmus a n d c o m p a r e d t o t a r g e t e d d e n s i t i e s (timmermann and whitlaw 1992; bisset and mclaren 1995,1999; bisset 1996; bisset et al. 1997, 1998, 2000). estimated densities were derived from moose observed plus those not sighted but believed to have been missed, based on a track aggregate method described by bergerud and manuel (1969), or a resurvey of random plots at a higher search intensity (bisset and rempel 1991). harvests were largely determined from annual (1980-96) district post-hunt mail surveys (dms) as described by gollat and timmermann (1987) as well as those conducted centrally beginning in 1997 (bisset et al.1999, 2001). trends in overall provincial hunter numbers and total harvests were obtained from the annual centrally conducted provincial mail survey (barbowski 1972, cumming 1974). the number of avts and harvest quotas assigned to each wmu were retrieved from regional files in kenora and thunder bay (whitlaw et al. 1993), annual omnr moose ontario’s moose management policy timmermann et al. alces vol. 38, 2002 14 table 1. ontario provincial moose population estimates and targets, 1953-2000. year population source comments estimate 1953 42,000 lumsden (1953) first province-wide estimate 1956-57 70,548 lumsden (1958) based on 21 district areas 1957-58 80,325 lumsden (1958) based on 21 district areas 1958 125,000 lumsden (1959) based on 21 district areas 1978 75,000 bisset (1993) 1982 80,000 timmermann (1987) stefanski, personal communication, 1984 1990-91 120,000 timmermann & buss (1995) bisset, personal communication, 1994 1990 92,883 whitlaw et al. (1993) 1991 91,100 euler (1994) e.a. decision 1992 104,500 bisset (1993) review of survey data 1975-92 1993 120,000 bisset (1993) “in the order of” 1994 120,000 simmons (1997) province-wide independent review 1997 120,000 simmons (1997) province-wide independent review 1997 100,000 provincial auditor (1998) audit report for f&w program 1999 120,000 bisset & mclaren 1999 aerial inventory plan 1999-2002 targets 1985 100,000 omnr (1980a) year 1985 province-wide target 1995 140,000 omnr (1980a) year 1995 province-wide target 2000 160,000 omnr (1980a) year 2000 province-wide target hunter fact sheets, and wildlife branch files. we had difficulty in readily obtaining population and harvest information for some wmus and not all databases were similar. achievement of population targets was assessed using wmu mean trend estimates and/or applying a ± 20 % at the 90% confidence interval (c.i.) minimum level of precision (bisset and mclaren 1995) where the pattern is questionable. in reality c.i.s were higher and averaged ± 28.4% (range ±15-73%) based on 107 surveys (1980 93) conducted in nwr's 28 wmus (whitlaw et al. 1993). based on these criteria, we then categorized each wmu as to: (1) target achieved; (2) unclear; and (3) target not achieved. likewise, we compared planned harvests to estimated harvests using dms data, realizing that calf harvests were underestimated, as “few districts attempt to s a m p l e c a l f o n l y l i c e n c e d h u n t e r s " (timmermann and rempel 1998:25). our review follows the 1980 mmp outline structure, respecting headings and subheadings (omnr 1980a). program targets population trends historical provincial post-hunt moose population estimates peaked at 125,000 in 1958 and this estimate was used until the early 1970s when densities in some areas were thought to have declined due to indications of over-hunting (cumming 1974). no comparable published estimates of nwr densities are available for this period. bisset (1993:abstract) believed that provincial populations “hit a low point” about 1978 at an estimated 75,000 (table 1). reduced harvests following mmp restrictions alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 15 begining in 1980 were thought largely responsible for population recovery to near previous levels by the 1990s. the most currently used population estimate (1999), derived from aerial inventories, places the nwr population at 51,047. this is slightly higher than the 1975 estimate of 49,806 but far short (59%) of the 2000 mmp target of 86,483 (table 2). population targets were achieved in only 8 of 28 wmus, 5 were unclear, and 15 were judged to have failed to have met expected target densities. similarly, the “official” provincal population estimate (1997) of 120,000 fell short (25%) of the year 2000 target of 160,000 (table 1). the accuracy of the 120,000 (1993-97) estimate is suspect, particularly when actual estimates (including missed moose) were nearly 15,000 less in 1992 (bisset 1993) and 20,000 lower in 1997 (provincial auditor 1998). “a 1996 ministry study found that 93% of all wmus within core range were below their population target levels” (provincial auditor 1998:6). incomplete or poor quality aerial surveys carried out by untrained and inexperienced survey crews during mild, low snow winters, may have contributed to lower density estimates (see timmermann1993). population swings of about 20% or more are required before changes can be detected (gasaway and dubois 1987). further, most wmu survey estimates for nwr reported by whitlaw et al. (1993) were higher than the “minimum acceptable level” of ± 20% at the 90% c.i. suggested by bisset and mclaren (1995:6) “to provide reliable trend through time information”. sightability estimators are used to calculate estimates of actual numbers. those used in ontario have varied from a low of 1.04-1.06 (novak and gardner 1975) to a high of 1.75 2.60 reported by thompson (1979). wmu densities are assumed to remain unchanged between survey years. hence, provincial population estimates are based on the cumulative total of all wmus, even though most are only sampled every 3 or more years. several explanations are plausible as to why mmp population targets (table 1) were not achieved. overhunting resulting from ineffective harvest control may be the single most likely cause of density shortfall (a. bisset, ontario ministry of natural resources, personal communication). we believe factors affecting population growth are complex, variable, and poorly understood. additional factors that may have impacted carrying capacity and density response include: parasites and diseases, predation, subsistence harvest by native people under treaty, poaching losses, winter severity, green period length, summer heat, and lower than expected land capability (timmermann and whitlaw 1992; g. lynch, alberta wildlife consultant, personal communication 2001). increased road access resulting from accelerated timber harvests has raised success rates in many wmus. unfortunately, few solid data exist on levels of subsistence hunting or predation losses, even though they undoubtedly play a role (see predator control section). parasite and disease studies suggest a plausible link to population declines in some areas (see research section).winter ticks (dermacentor albipictus) have been implicated as a mortality factor in highdensity moose areas (samuel and barker 1979, lankester 1987). recently (1998-99) a major winter moose dieoff was reported in algonquin park ( n. quinn, park biologist, ontario ministry of natural resources, personal communication 2001). interspecific c o m p e t i t i o n w i t h w h i t e t a i l e d d e e r (odocoileus virginianus) and transmission of the brain-worm (parelaphostrongylus tenuis ) from deer may impact local populations (whitlaw and lankester 1994a). in addition, e. addison, (ontario ministry of natural resources, personal communication 2001) suggested a possible ontario’s moose management policy timmermann et al. alces vol. 38, 2002 16 table 2. year 2000 moose population density targets and population estimates (1975-99) for 28 wmus in ontario’s northwest region. year 2000 targets1 population estimates target achievement wmu density population 19752 19802 1985 1990 1995 1999 (/km2) 1c3 0.10 13,228 7,000 7,000 7,000 7,000 7,000 7,000 no 2 0.39 4,575 1,230 1,130 1,050 1,600 3,203 1,562 no 3 0.39 4,536 3,360 1,640 1,900 1,500 1,700 1,408 no 4 0.39 3,825 2,360 1,250 1,600 1,900 1,550 2,031 no 5 0.39 3,340 2,130 1,459 3,050 3,050 2,601 3,521 yes? 4 6 0.39 1,393 720 490 330 1,100 1,300 1,740 yes 7a 0.39 285 612 116 200 660 977 660 yes 7b 0.39 2,370 468 800 1,350 1,600 1,751 no 8 0.39 1,748 1,133 1,040 1,700 1,950 2,485 2,819 yes 9a 0.39 1,425 1,756 952 750 1,000 1,450 1,525 yes 9b 0.39 1,230 200 500 940 1,100 no? 5 11a 0.39 1,010 2,520 2,886 541 638 1,097 550 unclear 11b 0.39 581 414 529 775 600 yes 12a 0.39 1,340 3,232 2,750 1,207 1,209 1,100 1,400 unclear 12b 0.39 2,088 2,042 2,188 2,463 2,450 yes 13 0.39 4,373 2,889 2,129 3,385 4,966 4,013 3,894 unclear 14 0.39 463 160 203 585 721 325 319 no 15a 0.39 3,701 5,980 6,600 1,200 2,800 2,500 3,600 no? 5 15b 0.39 5,844 4,091 4,091 5,440 6,120 yes? 4 16a 0.33 4,837 3,090 745 769 950 825 825 no 16b 0.15 1,325 388 550 800 847 1,615 unclear 16c 0.15 1,447 590 1,480 1,480 1,134 1,000 unclear 17 0.11 2,965 1,000 1,000 1,914 1,914 1,914 1,332 no 18a 0.33 2,580 1,510 1,360 845 615 657 657 no 18b 0.15 1,649 666 770 770 528 no 19 0.39 4,015 1,210 1,154 1,602 1,485 1,349 1,690 no 21a 0.39 5,320 7,914 6,180 6,347 3,895 2,976 3,205 no 21b 0.39 4,990 3,942 6 2,474 3,105 6 3,105 6 no total nwr 86,483 49,806 41,530 43,160 43,661 49,096 51,047 no 1 omnr (1982). 2 bisset (1991). 3 bisset (1992). 4 not clear; pattern suggests target achieved. 5 not clear; pattern suggests target likely not achieved. 6 virginia thompson, ministry of natural resources, personal communication, 2001. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 17 link between canine parvovirus that swept through domestic canids and apparently also wolf populations in the late 1970s and early 1980s and hence may have increased survival of moose. illegal hunting or poaching losses appear to be significant in some areas of ontario (d. harnish, ontario ministry of natural resources, personal communication 2000). additionally, above average winter severity and lower than expected land capability where density targets in some wmus may have been unrealistically high, could help explain some target shortfalls (peterson and allen 1974, timmermann and whitlaw 1992, rempel et al. 1997b). elevated calf harvests in some wmus due to an unlimited calf harvest strategy may be impacting growth potential (timmermann and rempel 1998). drought and warmer summer temperatures that lead to reduced net energy intake may result in lower fertility levels and pregnancy rates as recently suggested for isle royale moose by r.o. peterson (michigan tech university, personal communication 2000). conversely, an extended summer green period may have extended the period of positive energy balance in some years to enhance production as reported by stewart et al. (1977). ferguson et al. (2000) studied the influence of density on growth and reproduction in northwestern ontario moose and concluded that populations living in areas of low primary productivity and low natural predation show less persistence and require greater conservation efforts. recently mckenney et al. (1998) developed a spatially explicit moose population model to help increase understanding of the myriad of factors regulating ontario moose populations. bisset (1992) proposed revisions to some of the 14 “original” nwr population targets downwards in 3 , no change in 2, and upwards in 9 wmus, while retaining the original total population target. he believed where targets were excessive, they should be reduced, and raised in most areas in core range to a population density of at least 0.70 moose/km2, similiar to those acheived on the aulneau peninsula (wmu 7a). unfortunately no similiar province-wide exercise was undertaken, and hence “official”year 2000 population targets for the province (table 1) remained as those generated in the northwestern ontario strategic land use plan in 1982 (table 2). a new set of draft population targets was recently generated, following an internal province-wide review (omnr 2001). a further public review of these targets including a full review of the mmp is intended in the near future (e.r. armstrong, ontario ministry of natural resources, personal communication 2001). we conclude that moose populations are regulated by a host of factors, and not necessarily by any single factor such as overhunting. targeting a specific moose density and attempting to manage at that density level over time may be unrealistic. witness the dramatic reductions of swedish moose populations from 1982-92 that demonstrate the difficulties involved in managing wild populations at a predetermined density (sylvén 1995). moose shot by hunters declined from 175,000 to 99,000 during this 10 year period. sizeable, but unquantified effects of global warming, if occurring, may also negatively influence population growth. we agree with morris (1959), that often the major, common mortality factors may not be as important in influencing population fluctuations as those variable factors that operate inconsistently and over which managers have no control. we support the current target review exercise in re-examining population and harvest targets, to determine their relevance and achievability. further reducing avts and hunting opportunities is, in our opinion not necessarily ontario’s moose management policy timmermann et al. alces vol. 38, 2002 18 the answer in all cases. we suggest the overall population target of 160,000 be reduced. lower moose densities can provide proportionately more recreation per kill than higher densities (crête 1987,1989). lower densities (e.g., half of k) also provide a greater sustained yield if hunting is assumed to be the major mortality factor. the question then arises whether it is really necessary to increase populations in all wmus by actively restricting participation t o s o c i a l l y u n a c c e p t a b l e l e v e l s . timmermann and gollat (1986) suggested managers consider offering a mix of hunting qualities high but limited in some w m u s a n d l o w e r a n d m o r e l i b e r a l in others, coupled with reducing hunter efficiency. harvest trends provincial harvests increased incrementally from 1,456 in 1953, 2 years after seasons were re-opened (cumming 1974), to around 12,000 in 1962 (table 3). they averaged 12-13,000 for the next decade, peaking at 14,610 in 1965, following a period of long, liberal nonselective hunting seasons. substantial harvest reductions occured 1980-81 when seasons were shortened, and opening dates were delayed and hunters were required to share a moose (timmermann and gollat 1984). harvests were lowest (7,971) in 1983, following introduction of the current selective harvest program. they rebounded and varied little (10,00011,000) from 1984 98 (table 3), even though hunter numbers (1980 98) increased by 20,000 (table 4). nwr harvests however increased from 39.2% (4,188) of provincial harvest in 1982, to 51.3 % (5,611), in 1998, while hunter numbers increased by about 7,000 (table 5). table 3. total estimated ontario provincial moose harvests, 1953-1998. year estimated year estimated year estimated year estimated harvest1 harvest1 harvest harvest 1953 1,456 1965 14,610 1977 1989 10,7712 1954 1,781 1966 14,517 1978 1990 1955 2,867 1967 13,207 1979 1991 11,0003 1956 4,540 1968 12,050 1980 8,3612 1992 1957 5,943 1969 12,332 1981 8,0922 1993 1958 6,787 1970 11,918 1982 10,6912 1994 10,0004 1959 8,925 1971 13,072 1983 7,9712 1995 1960 10,048 1972 13,114 1984 10,3462 1996 1961 11,830 1973 1985 10,1622 1997 9,8135 (10,500) 1962 12,147 1974 1986 10,7902 1998 10,9296 1963 13,113 1975 1987 10,7632 1964 11,924 1976 1988 1 cumming (1974) for the period 1953-1972. 2 omnr moose hunter fact sheets, ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. 3 timmermann and buss (1995). 4 simmons (1997). 5 bisset et al. (1999); 10,500 estimate includes hunters who did not apply for avts. 6 bisset et al. (2001). alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 19 table 4. ontario provincial moose hunter numbers and harvests. resident tourist industry all hunters year number of harvest number of harvest number of harvest hunters hunters hunters 19801 71,666 7,669 3,927 692 75,593 8,361 1981 66,852 7,626 3,105 466 69,957 8,092 1982 82,678 9,747 5,841 944 88,519 10,691 1983 65,062 7,971 6,794 1,082 65,062 7,971 1984 72,194 6,828 6,949 1,143 79,143 10,346 1985 71,408 8,781 7,191 1,381 78,599 10,162 1986 72,959 9,378 7,217 1,412 80,176 10,790 1987 76,918 9,438 7,971 1,325 84,889 10,763 1989 82,600 9,104 8,091 1,667 90,691 10,771 19972 95,004 9,813 19983 89,006 10,929 1 1980-89 data from omnr hunter fact sheets, ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. 2 bisset et al. (1999). 3 bisset et al. (2001). higher harvests for tourist industry based hunters in the 1990s have contributed to increased overall nwr harvests (table 5). planned harvests (# of bulls, cows, and calves) in the nwr 1984-99 remained reasonably constant during the 16 year period (4,6295,372; fig. 2). they were highest in 1988-90 when populations were thought to have increased (timmermann and whitlaw 1992) and returned to around 5,000 thereafter. estimated nwr bull harvests exceeded planned harvests in 9 of 15 years, but were generally within10% of planned harvests whereas cow harvests were almost always higher (up to 36%) (fig. 3). hence the impact of heavy cow harvests may have curtailed herd increase in some areas. estimated total harvests ranged from 3,711 in 1984 to a high of 5,587 in 1998. however, an increased effort in assessing the calf kill beginning in 1997 is believed responsible for elevated estimates in recent years (fig. 4). if only the adult harvest estimates are examined, data suggest peak harvests of bulls and cows occurred in 1988-90. a provincial harvest target of 10,000 moose was achieved by 1985. however, harvests stalled at 10,000-11,000 per year thereafter (table 3), hence the higher harvest targets of 18,000 and 25,000 by 1995 and 2000, respectively, were grossly underachieved. likewise nwr harvests (48.8% of provincial harvests in 1985, rising to about 55% by 1997) fell far short of mmp harvest target expectations, even though no specific interim nwr wmu specific harvests were defined. hunting opportunities the mmp targeted an increase in hunting opportunities from 350,000 400,000 user-days after 1985 to 750,000-875,000 by the year 2000 (omnr 1980a). published data suggest that these user-day targets were largely met (713,000 in 1993 and 817,000 ontario’s moose management policy timmermann et al. alces vol. 38, 2002 20 fig. 2. planned and estimated harvests for bulls, cows, and calves combined in 28 wmus of the northwest region of ontario, 1984-1999 (source: district mail surveys). 0 1000 2000 3000 4000 5000 6000 7000 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 year h a r v e st planned harvest estimated harvest table 5. ontario nw region (28 wmus – wmu 1c through 21b) hunter numbers and harvests. percent of resident tourist industry all hunters provincial harvest year number of harvest number of harvest number of harvest hunters hunters hunters 19821 22,286 3,643 2,688 545 24,974 4,188 39.2 1983 20,896 3,875 3,041 649 23,937 4,524 56.8 1984 19,995 3,902 2,967 726 27,590 5,354 51.7 1985 22,368 4,166 3,527 798 25,895 4,964 48.8 1986 22,238 4,473 3,565 786 25,803 5,259 48.7 1987 23,773 4,483 4,169 937 27,942 5,420 50.4 1989 26,942 4,636 4,215 1,089 31,157 5,725 53.2 19972 32,319 5,422 55.3 19983 31,957 5,611 51.3 1 1980-89 data from omnr hunter fact sheets, ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. 2 bisset et al. (1999). 3 bisset et al. (2001). alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 21 in 1997; table 6). the large provincial and regional increase in hunter numbers (tables 4 and 5), and relatively long seasons, combined with the introduction of legal party hunting in 1988, the introduction of a group application for an adult moose tag in 1992 (omnr 1991), and unlimited calf hunting opportunities for all licenced hunters, are believed largely responsible. archery seasons were also introduced to provide additional hunting opportunities in 6 wmus in 1984 and increased to 25 wmus by 2000. the nwr offers the majority (6070%) of provincial adult archery tags in 20 fig. 3. planned and estimated bull and cow harvests in 28 wmus of the northwest region of ontario, 1984-99 (source: district mail surveys). 0 500 1000 1500 2000 2500 3000 3500 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 year b u ll h a r v e st planned bull harvest estimated bull harvest 0 200 400 600 800 1000 1200 1400 1600 1800 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 year c o w h a r v e st planned cow harvest estimated cow harvest ontario’s moose management policy timmermann et al. alces vol. 38, 2002 22 of 25 wmus (figs. 5 and 6). on the other hand a significant reduction in resident gun avts occurred (fig. 7). combined provincial bull and cow avts fell stepwise from 55,886 in 1983 to 15,994 in 2000, 29% of the original allocation. likewise tag numbers in the nwr in 2000 represented about 35% of those available in 1983 (i.e., from 22,291 [39.9% of total provincial tags] in 1983 to 7,894 [64.4 % of the provincial total] in 2000; fig. 8). fig. 4. planned and estimated calf harvests in 28 wmus of the northwest region of ontario, 19841999 (source: district mail surveys). 0 200 400 600 800 1000 1200 1400 1600 1800 2000 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 year c a lf h a r v e st planned harvest estimated harvest area closures following logging operations were suggested as a management strategy to protect vulnerable moose populations and limit hunting opportunities in some areas. (timmermann and gollat 1984). racey et al. (2000) reported hunting opportunities were likely reduced in these areas but that quality viewing opportunities were provided in areas with good access and relatively high moose densities in a closed area case study. table 6. ontario provincial moose hunting opportunities and targets, 1973-1997. year number of user days source comments 1973 460,000 cumming (1974) economic impact of $13m per year 1993 713,000 legg (1995) economic impact of $134.7m per year 1997 817,000 bisset et al. (1999) average of 8.6 days per hunter targets 1985 350,000-400,000 omnr (1980a) year 1985 target 1995 630,000-720,000 omnr (1980a) year 1995 target 2000 750,000-875,000 omnr (1980a) year 2000 target alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 23 viewing opportunities the mmp targeted the development of specific interim (1985, 1995) and year 2000 moose viewing opportunities (omnr fig. 6. number of available ontario resident archery adult validation tags (avts) for the northwest region (1983-2000) and percent of total available provincial archery avts. 0 500 1000 1500 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 year n u m b e r o f a v a il a b le a v t 's 0 20 40 60 80 100 p e rc e n t available bull avt's available cow avt's per cent of provinicial avts nw region total 1980a). little or no effort was made to identify moose viewing sites, hence it remains unclear how much progress was made in reaching this goal. currently the best viewing opportunities appear to be in lightly 0 500 1000 1500 2000 2500 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 year n u m b e r o f a v a il a b le a v t 's available cow avt's available bull avt's fig. 5. number of available ontario provincial archery resident adult validation tags (avts), 1983-2000. ontario’s moose management policy timmermann et al. alces vol. 38, 2002 24 0 5000 10000 15000 20000 25000 30000 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 year n u m b e r o f a v a il a b le a v t 's 0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 p e rc e n t available bull avt's available cow avt's per cent of provinicial avts nw region total fig. 8. number of available ontario resident gun adult validation tags (avts) for the northwest region (1983-2000) and percent of total available provincial avts. 0 10000 20000 30000 40000 50000 60000 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 year n u m b e r o f a v a il a b e a v t 's available cow avt's available bull avt's fig. 7. number of available ontario provincial resident gun adult validation tags (avts), 1983-2000. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 25 hunted or non-hunted areas with high land capability such as portions of algonquin park, lake superior park, quetico provincial park, chapleau crown game preserve, and several local closures near thunder bay (timmermann and whitlaw 1992) where population densities approach or exceed 1.0/ km2. policy guidelines the mmp listed 14 policy guidelines to provide a framework for planning and management towards development of more specific program policies (omnr 1980a). these included, the development of harvest plans to meet targets, use of wmus as the basic unit for planning and management, close coordination of habitat management with forest management, and public participation in management planning. most, if not all, of these policy guidelines were followed during the 20-year mmp period. public reviews of the program included consultations held across ontario in 1987 using a consultant organized mail questionaire (omnr 1987, wedeles et al. 1989), a public review conducted by the ministry of natural resources in 1991 (omnr 1991), and an independent review conducted by george simmons, december 1996-march 1997 (simmons 1997). all 3 reviews added new components to the original mmp. included were the introduction of legal party hunting in 1988, development of an extensive moose hunter education manual (omnr 1990), development of a group application system begining in 1992 (omnr 1991), introduction of a sportsman’s card to reduce draw application errors in 1992 (hhhf 2000), introduction of a special mobility-impaired hunt for disabled hunters in 1992 (omnr 1991, armstrong and simons 1999), and in 2000, the first review of moose population targets since their inception in 1980 (hhhf 2000). population assessment (policy item #12) was identified in the mmp, along with the need to consider harvest and habitat management. however, no further policy details concerning how and when population assessments were to be made have been provided since the inception of the mmp. standards and guidelines for moose aerial inventory in ontario were drafted in 1980 (omnr 1980c). a survey schedule of once every 3 years (omnr 1993) was suggested for core wmus and bisset and mclaren (1995) provided criteria for establishing survey priority. oswald (1982, 1997) produced a detailed moose aerial observation manual. inventory surveys were carried out in most core wmus during the 1980s (timmermann and whitlaw 1992, bisset 1993; fig. 1). in the early to mid 1990s, population assessments were not conducted “frequently enough to enable managers to make informed decisions”(provincial auditor 1998:7). standards and guidelines for moose aerial inventory were revised in 1991 and again in 1996 (bisset 1991, 1996). a plan to restore the 3-year inventory cycle across moose range was first implemented in the winter of 1995-96 and later updated to cover the period 1999-2001 (bisset and mclaren 1995, 1999). bisset et al. (1997, 2000) published a comprehensive report on province-wide population surveys for 199596, 1996-97, and 1998-99. a pilots’ manual aimed at helping pilots to better understand the survey process and increase consistency was issued in 1998 (bisset et al.1998). as a result of all these measures, “a more regular schedule of aerial moose surveys is being carried out across the province and allowable harvests are being recalculated to reflect new information” (provincial auditor 1998:9). survey crew experience, training, and adherence to survey guidelines will remain a challenge to managers in future years. ontario’s moose management policy timmermann et al. alces vol. 38, 2002 26 management policy harvest control the mmp directed a limit on harvest using age and sex specific licencing (omnr 1980a). the selective harvest system introduced in 1983 accommodated this policy by limiting the number of bull and cow avts issued by wmu (omnr 1984), yet allowing all unvalidated licence holders the opportunity to hunt for calf moose.the number of avts issued was directly related to the wmu allowable harvest, past hunter success rates, and overall herd status (euler 1983). the mmp further proposed regulating harvests by influencing hunter access to moose by controlling use of aircraft, snowmobiles, and all-terrain vehicles. this policy was not implemented; currently there are relatively few restrictions on firearm use and no specific regulations that limit or control the use of allterrain vehicles (mcmillan et al. 1993). the harvest control policy included season closures for short periods in specific local areas such as recently logged areas where moose are especially vulnerable to harvest (timmermann and gollat 1982). one such area in the nwr was closed for a period 1977-89 and then re-opened to hunting (racey et al. 2000). results of a monitoring study suggested that closures not be considered an alternative to moose habitat guideline application even though such closures may enhance moose densities to meet population targets or provide alternate recreational opportunities. despite the former, social pressures make support for closures very difficult. access and hunting pressure are probably more significant than the moose habitat guidelines in regulating moose densities (racey et al. 2000:21). hunting pressure was distributed according to the mmp (omnr 1980a), beginning in 1983 by wmu specific licences for avt holders (omnr 1984). however, all unvalidated licence holders were able to hunt calves in any wmu, as well as legally party hunt (since 1988) for adult moose with an avt holder (wedeles et al. 1989). finally, the harvest control policy recommended increasing recreation from moose hunting by requiring hunters to hunt in groups. this policy was implemented for 3 years beginning in 1980, but abandoned because it failed to include a mechanism for predictable area specific control of the harvest (timmermann and gollat 1984). a voluntary group application system for adult moose tags was introduced in 1992 (omnr 1991). this system allowed a fairer allocation of avts and spread tags amongst more hunters. in 1999, for example, 37.7% of hunters applied in groups of 2 or more. the average group size was 4.23 hunters per group and 64.7% of groups received an avt compared to only 19% of individual applicants (omnr 2000:36). predator control the mmp targeted a limited predator control program to allow moose numbers to increase where gray wolves (canis lupus) a r e s i g n i f i c a n t l y d e p r e s s i n g m o o s e populations (omnr 1980a). no efforts were made to implement this policy, nor were studies implemented to assess the impact of black bear (ursus americanus) predation on moose, even when other jurisdictions identified both predators as capable of limiting or regulating moose populations (gasaway et al. 1983, wilton 1983, schwartz and franzmann 1991, ballard 1992, van ballenberghe and ballard 1994, ballard and van ballenberghe 1998). kolenosky (1981) reviewed the status and management of wolves in ontario, while bergerud (1981) and bergerud et al. (1983) suggested wolf predation limited moose populations particularly in the pukaskwa n a t i o n a l p a r k a r e a o f n o r t h c e n t r a l ontario. thompson and peterson (1988) argued that wolf predation alone did not alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 27 limit moose populations in the park, while bergerud and snider (1988) provided additional arguments supporting their position. closure of the ontario spring black bear season in 1999 (hhhf 2000) potentially may increase moose calf losses to this predator, assuming bear populations increase (see ballard 1992). newfoundland, on the other hand, has bears which predate some moose calves but no wolves, white-tailed deer, or moose ticks, and currently has an estimated pre-hunt population of 150,000 moose despite liberal hunting regulations. the 1999 legal hunter kill was estimated at 19,500 by mercer and mclaren (2002). enforcement hunters commonly believe there are insufficient conservation officers afield to enforce moose hunting regulations (bottan 1999). the provincial auditor reported a decrease in the amount of time spent on general deterrent patrols by conservation officers and in the number of charges laid under the game and fish act 1996-98 (provincial auditor 1998:3). the mmp proposed increased enforcement of legislation and regulations to control illegal hunting and to suppress poaching (omnr 1980a). few moose enforcement data concerning hunter compliance have been analysed or published. timmermann and gollat (1986) provided the only known published information for the former northcentral region, which contained 14 of the current 28 nwr wmus. they indicated that moose related charges increased from 358 (1980-82) to 511 during the first three years of the selective harvest (1983-85), averaging about 171 per year. c.j.w. todesco, (ontario ministry of natural resources, personal communication 2001) reported 549 illegally killed moose in the northeastern region of ontario during the period 1997-2000 of which 224 were related to abandoned animals. northeastern region managers identified a growing problem of illegal moose hunting with 472 moose-hunting related charges laid in 1999. consequently they launched a high profile enforcement campaign prior to the 2000 fall season. "moose watch 2000" was designed prior to the 2000 fall season to combat a perceived increase in illegal moose hunting, including those shot and abandoned. by november 30th 2000, almost 500 charges relating to illegal moose hunting were laid and a further 170 charges were pending with 65 investigations under way. mnr officers seized 82 illegally killed moose, and investigators found 53 moose shot and abandoned, based on 126 tips received. bob stewart, an experienced thunder bay district ontario ministry of natural resources conservation officer, reported more hunters were charged and more moose seized in 2000 than ever before (personal communication 2001). he believed several factors contributed, including better enforcement tools, use of dna analysis, improved officer training, and better communications. hunter dissatisfaction with recent federal firearm legislation and avt reductions are believed by stewart and others to have contributed to increased illegal activities. avt manipulation is considered to be common (i.e., hunters using other family member’s tags and abusing party hunting regulations). moose tag transfers and new tags issued provincially totaled 841(5.5% of total tags) with the majority (568 or 68%) occurring in the nwr (table 7). the number of hunters applying in a group (>2) fluctuated from a high of 45,447 in 1992 when the program was introduced to a low of 32,884 in 1997 (table 8). approximately 61-65% of groups applying receive an avt, compared to 18-20% of individual applicants. large hunting parties of 8 or 10 who have one adult tag issued between them, may shoot more moose than ontario’s moose management policy timmermann et al. alces vol. 38, 2002 28 they are licenced for without the knowledge of all party members (dave harnish, ontario ministry of natural resources, personal communication 2001). habitat management the mmp directed maintenance of moose habitat, by recommending wildlife and forest managers work closely to modify cutovers, especially around aquatic feeding areas, mineral licks, and winter concentration areas (omnr 1980a). production of irregularly shaped cuts, scattered shelter patches, and high age class diversity among species was the prime objective. significant progress in meeting this target has been made beginning with the release of timber management guidelines for the provision of moose habitat (omnr 1988a), which provided direction regarding forest access, harvest operations, site preparation, regeneration, and maintenance. racey et al. (1989a) studied the application of the moose guidelines and their impact on forest industry investment. a review of habitat planning was provided by payne et al. (1988), while management tools regarding habitat interpretation were provided by racey et al. (1989b), jackson et al. (1991), and timmermann (1998). an inventory manual for use in timber management planning was recently issued (ranta 1998 ). in addition to identification of habitat, mcnicol and baker (1998) devised a “ranking” for both early and late winter habitat from 1 (low potential) to 4 (very high potential).w.b. ranta (ontario ministry of natural resources, personal communication 2001) is currently updating the “forest management guidelines for the provision of moose habitat” originally issued in 1988. (omnr 1988a). inventory funding is currently provided to districts who are involved in preparing forest management plans (m. sobchuck, ontario ministry of natural resources, personal communication 2001). inventory and assessment the effectiveness of a harvest control system depends on a reasonably accurate assessment of hunter kill (timmermann 1987, timmermann and buss 1998). in the early 1990s, 16 of 21 north american jurisdictions that actively manage moose practiced compulsory harvest registration (timmermann and buss 1995). the mmp directed an improved program of voluntary reporting by hunters and a phased-in mandatory registration and reporting system. this policy failed to deliver as ontario hunters are currently neither required to register their kill nor provide data to managers, except on a voluntary basis. in our opinion, current assessment techniques are ineffective in providing managers with a timely and accurate assessment of the annual moose harvest. this program currently relies on a centralized mail survey of licenced hunters to assess harvest (barbowski 1972; omnr 1997; bisset et table 7. ontario moose tag transfers and new tags issued in the northwest region and provincewide, 2000. number of area number of tags total number percent of wmus transferred of tags total 26 nw region 568 9,722 5.8 33 other regions 273 5,488 5.0 59 province 841 15,210 5.5 alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 29 t ab le 8 . o n ta ri o a d u lt m o o se v al id at io n t ag (a v t ) d ra w – p ro v in ci al s u m m ar y ( s o u rc e: o n ta ri o a n n u al h u n ti n g r eg u la ti o n s s u m m ar y , 1 9 9 2 2 0 0 1 ). t o ta l y ea r n o . o f n o . o f n o . o f a v er ag e % o f t o ta l % o f t o ta l n u m b er % o f a v t s a p p li ca n ts g ro u p s g ro u p h u n te rs n u m b er o f g ro u p o f h u n te rs in d iv id u al s s iz e a p p ly in g h u n te rs i n r ec ei v in g a p p ly in g a s r ec ei v in g in g ro u p s g ro u p s t ag s in d iv id u al s t ag s 1 9 9 2 2 8 ,7 9 2 9 8 ,0 2 6 1 1 ,7 5 1 3 .9 4 6 .4 4 5 ,4 4 7 5 2 ,5 7 9 1 9 9 3 2 7 ,7 2 6 1 9 9 4 2 6 ,4 0 2 1 0 3 ,2 4 4 1 0 ,8 9 6 3 .9 4 0 .8 4 2 ,1 6 0 6 1 ,0 8 4 1 9 9 5 2 4 ,0 4 7 1 0 6 ,0 1 3 9 ,9 5 4 4 .2 3 9 .1 4 1 ,4 3 9 6 4 ,5 7 4 1 9 9 6 2 2 ,8 0 2 1 9 9 7 2 0 ,5 9 2 1 0 0 ,7 3 1 7 ,3 4 1 4 .5 3 2 .6 3 2 ,8 8 4 6 1 .1 6 9 ,0 3 2 1 7 .7 1 9 9 8 2 0 ,3 5 1 1 0 2 ,4 4 3 9 ,1 4 1 4 .4 3 8 .8 3 9 ,7 0 1 6 2 .0 6 9 ,7 3 2 2 0 .0 1 9 9 9 1 9 ,5 2 0 1 0 3 ,4 9 9 9 ,0 7 3 4 .2 3 7 .7 3 8 ,4 2 5 6 4 .7 6 5 ,0 7 4 1 9 .0 2 0 0 0 1 7 ,5 4 0 1 0 3 ,8 3 5 9 ,8 5 2 4 .4 4 2 .0 4 3 ,6 5 1 6 3 .0 6 0 ,1 8 4 1 8 .0 ontario’s moose management policy timmermann et al. alces vol. 38, 2002 30 al. 1999, 2001) replacing the system of post-card surveys previously conducted by f i e l d d i s t r i c t s 1 9 8 4 9 6 ( g o l l a t a n d timmermann 1987). response rates from those former district-conducted mail surveys were considered high (80+%) when prepaid return postage and a follow-up mailing to non-respondents was included. response rates to centralized mailed questionnaires has declined noticeably since introduction in1997, thus lowering confidence in wmu harvest estimates (g.eason, b. ranta, and r.hartley, ontario ministry of natural resources, personal communication 2001). factors believed responsible include lack of return postage, no second follow-up mailings, and reduced hunter support. the provincial auditor (1998:2) indicated “the ministry did not have adequate procedures in place to provide the information necessary for measuring and reporting on the program’s effectiveness in sustaining fish and wildlife resources”. timmermann and rempel (1998) examined moose age and sex structure from 38,870 hunter submitted samples from northcentral ontario over the period 197291. their analysis demonstrated the value of hunter-submitted kill records to help evaluate changes in population structure and assess the effectiveness of management strategies. findings suggest a significant change in age and sex structure and an elevated calf harvest that may have impacted growth potential since introduction of a selective harvest strategy in 1983. the voluntary jaw collection program was eliminated in the early 1990s when government downsizing occured and hunters were informed that managers no longer needed this information for management purposes. hence, hunters are not responsible for providing data and appear more skeptical than ever in supporting the program even though repeated program reviews and published reports have recommended mandatory reporting (timmermann and whitlaw 1992, timmermann et al. 1993, hansen 1995, hansen et al. 1995, timmermann and rempel 1998, bottan 1999). stewart (2000) in a discussion document towards a hunting management strategy for ontario recommended that the required knowledge for effective big game management includes a measure of licenced hunter and aboriginal harvest (sex, age, numbers). simmons (1997:33) recommended mnr develop a harvest reporting system that would account for all moose harvests including those by “natives who hold treaty and aboriginal rights.” bisset (1999) provided an overview of mandatory reporting, including cost estimates of quality information and relative importance. he argued that assessing and improving systems designed to manage the current voluntarily provided data should be completed before introducing a more expensive mandatory reporting program. simmons (1997:56) recommended “an accurate data base” be established to monitor populations and harvest numbers. the ontario ministry of natural resources has stated that enhanced data management is a ministry priority (omnr 1999). research research on habitat management was a major focus of the mmp. early efforts focused on gaining basic knowledge. mcnicol and gilbert (1980) studied late winter use of upland cutovers by moose north of lake superior. mcnicol et al. (1980) reported the effects of heavy browsing pressure while mcnicol and timmermann (1981) reviewed the effects of forestry practices on moose populations in the boreal mixedwood forest. thompson et al. (1981) studied the traditional use of early-winter concentration areas in northeastern ontario, while thompson and vukelich (1981) described the use of logged habitat in winter by moose alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 31 cows with calves in northeastern ontario. cumming (1980) related moose track counts to cover types in northcentral ontario. euler (1981) proposed a moose habitat strategy and thompson and euler (1987) discussed the changing perception of moose habitat in ontario. eason (1985,1989) reported on hunting vulnerability in recently logged areas. effects of hunting closures and timber harvest on local moose densities and hunting opportunities were examined by racey et al. (2000). several related habitat studies originated from lakehead university in thunder bay. they included those on the impact of glyphosate on moose by cumming ( 1 9 8 5 ) , c o n n o r ( 1 9 8 6 ) , c o n n o r a n d mcmillian (1988), and kelly and cumming (1994). moose vegetative preferences resulting from 16 years of browse surveys were detailed by cumming (1987). winter use by moose of modified strip cuts compared to clearcut use were reported by todesco et al. (1985) and todesco (1988). mastenbrook and cumming (1989) examined moose use of residual strips of timber left within cutovers, while dalton (1989) detailed moose use of partially and totally l o g g e d c l e a r c u t s . t i m m e r m a n n a n d mcnicol (1988) provided a literature review of overall moose habitat needs and timmermann (1991), completed a review of ungulate and aspen management. descriptive studies of moose access routes to an aquatic feeding area and moose cratering for equisetum sp. were provided by timmermann and racey (1989) and timmermann et al. (1990). moose use of aquatic plants, road salt, and natural mineral springs was reported by fraser and reardon (1980), fraser and thomas (1982), fraser and hristienko (1983), and fraser et al. (1980, 1982, 1984). the mmp recommended an increased effort to research the effectiveness of management policies, moose productivity, and diseases of moose. the effectiveness of the 1980-82 2tag harvest system was examined in the former northcentral region (timmermann and gollat 1984). effectiveness of the selective harvest system was reported by timmermann and gollat (1986, 1994), timmermann and whitlaw (1992), bisset (1992, 1993), heydon et al. (1992), t i m m e r m a n n e t a l . ( 1 9 9 3 ) , a n d timmermann and rempel (1998). research on moose diseases and parasites included reports on pathological anomalies by lankester and bellhouse (1982); studies on the moose fly by lankester and sein (1986), on gastro-intestinal helminths by kennedy et al. (1985), snider and lankester (1986) and fruetel and lankester (1988), on the brainworm (p. tenuis) by whitlaw (1993), whitlaw and lankester (1994a, b), and on t h e m o o s e t i c k ( d . a l b i p i c t u s ) b y t i m m e r m a n n a n d l a n k e s t e r ( 1 9 8 0 ) , addison and mclaughlin (1988, 1993), addison and smith (1981), addison et al. (1998a, 1998b), and wilton and garner (1993). cadmium levels in ontario moose and potential sources of contamination were investigated by glooschenko et al. (1988) and kronberg and glooschenko (1994). studies on the structure and composition of calving sites in algonquin park were reported by addison et al.(1990) and wilton and garner (1991). in 1994, under term and condition 80 of the environmental assessment board decision (oeab 1994), ontario was directed to undertake long-term scientific studies to assess the efficacy of the timber management guidelines for the provision of moose habitat. the moose guidelines evaluation program (mgep), originally established after the introduction of guidelines in 1988, was modified and expanded to comply with term and condition 80 (oeab 1994, rodgers et al. 2000). the mgep was designed to study ontario moose population dynamics, habitat use, condiontario’s moose management policy timmermann et al. alces vol. 38, 2002 32 tion and productivity, characterization of moose calving sites and aquatic feeding a r e a s , a n d t o d e l i n e a t e m o o s e morphometrics and genetics (rodgers et al. 2000). mgep publications to date include those on: ecosystem management (hénault et al. 1999); home range size (lawson and rodgers 1997, rodgers and carr 1998); global positioning system (gps) (rodgers and anson 1994; rempel et al. 1995; rodgers et al. 1995, 1996, 1997, 1998; rempel and rodgers 1997); timber management and natural disturbance effects on moose habitat (rempel et al. 1997a); moose browse production (rempel et al.1997b); sensitivity of harvest data to changes in aerial population estimates (timmermann et al. 1993); and calving site fidelity (welch et al. 2000). hunter education the mmp targeted the introduction of a voluntary moose hunter education course and firearm proficiency test, and a phasedin mandatory course and test for new moose hunters (omnr 1980a). this policy was supported by 79% of hunters who attended 72 public meetings that attracted 7,350 hunters across the province in 1979 (omnr 1980b). a great deal of information was prepared and circulated to hunters. included were a moose hunter handbook (omnr 1984), instructional magazine articles, annual moose hunter fact sheets distributed at all licence sales locations and government offices, and a moose identification quiz (timmermann 1992). in addition, an extensive 78 page moose hunter educational manual (omnr 1990) along with a draft instructor’s manual were published as the core curriculum for a mandatory moose hunter education course. two videos entitled "moose hunt, a guide to success" (interesting services inc., emsdale, ontario, canada, 1989) and "firearms for the moose hunter" (interesting services inc., emsdale, ontario, canada, 1988) were circulated to hunter education instructors. in addition, a standardized shooting skill scoring sheet along with a life-sized target, shooting instructions, and an illustrated moose anatomy pamphlet were prepared to test shooting skills (omnr 1988c, buss et al. 1989, timmermann and buss 1998). repeated studies by rollins (1987), romano (1988), rollins and romano (1988,1989), and wedeles et al. (1989) recommended an expanded hunter education effort to strengthen hunter understanding and compliance. hansen et al. (1995) reported only a third (597 of 2,007) of hunters responding to a survey of satisfaction with the ontario moose management system had reported attending a voluntary moose hunter seminar. stewart (2000) identified hunter education as a critical component toward development of an effective hunting management strategy for ontario. the strategy made 26 recommendations, the majority dealing with various aspects of hunter education. recommendations included that new hunters be encouraged to take advanced species specific courses (waterfowl, moose, etc.) and that voluntary training does not result in significant advances in the knowledge, skill, and conduct of new hunters. they concluded advanced courses should continue to be voluntary for existing licenced hunters. resource allocation the mmp gives primary consideration to subsistence use by first nations people in recognition of obligations made under historical treaties. this policy has been largely honored. however, little effort has been made to measure the magnitude of this harvest. otherwise, the allocation policy which provides for all residents to be treated equally, and favours residents over nonresidents, and resident canadians over non alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 33 resident aliens, has largely been followed. in addition, a process was estabished that required non residents of ontario to use established tourist facilities that were issued a special quota of licences (10% of total allocation) (bisset and timermann 1983). licence allocation to the tourist industry must be made a year in advance of up-to-date data, hence a lag in quota adjustment occurs. in 1998, the policy restricting non residents from hunting in resident only wmus was relaxed. they now may hunt in those wmus as long as they both obtain their avt from and use designated tourist outfitting facilities. resident hunter resentment over this change is believed to stem from misconceptions derived from lack of information such as hunting as a group on only one adult tag (bottan 1999). management implications and recommendations targets achieved, lessons learned a significant number of year 2000 population targets proposed in 1980 for many wmus were not achieved. we believe the 20 year policy timeframe lacked a feedback mechanism to allow periodic program review and target adjustment (i.e., adaptive management) as added information and experience was gained. why, for example were targets not adjusted when it became apparent that those set for1995 (140,000) would not be met? in future, we suggest wmu-specific population targets be tailored to more closely reflect land productivity as well as a host of mortality factors which managers are unable to control, or in most cases measure, as reported by mckenney et al. (1998). we suggest a 5 year policy time-frame in which population status is reviewed and targets are adjusted if necessary. further consideration should be made to target a population range for each wmu, reflecting a minimum density, below which hunter harvest would be curtailed. such a target range would reduce the frequency of downward avt adjustments and work toward increasing hunter confidence in the program. decision support tools, including the use of models, need to be employed to ensure integration of all factors affecting populations. finally, managers need to remember that aerial surveys used to generate density estimates, nearly always underestimate the number of animals present and that population data thus obtained are best treated as trend indicators and not as absolute numbers (gasaway et al. 1986; timmermann 1974,1993). hence, we suggest managing for a density range, not a specific density. in addition, managers need to recognize the time-lag in quota adjustments as population targets are directly influenced by annual allocation decisions. annual provincial harvest targets of 10,000 moose by 1985 were met, however they levelled out thereafter at 10-11,000 per year (table 3). projected harvest targets of 18,000 by 1995 and 25,000 per year by 2000 were grossly underachieved. these targets were not translated to various regions or wmus, hence there is no way of judging achievement. we do know, however, that the proportion of hunter-killed moose taken in the nwr increased from about 40% of provincial total in 1982 to over 50% beginning in1987 (table 5). failure to increase overall provincial harvests suggests that other mortality factors play a more important but ill-defined role than previously realized. timmermann and rempel (1998) also suggested changes in harvest structure and a growing calf harvest resulting from the selective harvest system may be impacting future growth potential in some wmus. harvest assessment was centralized in 1997 and subsequent lower response rates to mailed questionnaires is believed to have affected quality of data. in addition, an ontario’s moose management policy timmermann et al. alces vol. 38, 2002 34 accurate assessment of calf harvest remains an elusive challenge. we suggest a huntersupported harvest assessment program be given a high priority and that hunters be encouraged to provide harvest information and become responsible partners in the moose management program. serious consideration should be given to returning to a provincially coordinated dms harvest assessment system including return postage, and a second reminder mailing. alternatively, consideration should be given to requiring all hunters to complete and remit a simple questionnaire, attached to each licence as is required by many jurisdictions (timmermann and buss 1995). failure to comply would trigger a penalty or default on a predescribed reward. consideration could also be given to using a telephone questionnaire carried out by resource user groups to assess moose harvest as has been employed successfully in alberta since 1985 (lynch and birkholz 2000). without such support, the program cannot properly function. data collected over the past 20 + years suggest that a provincial population target of 100-120,000 and an annual harvest of 10-12,000 moose per year is a sustainable target (tables 1 and 3). population estimates have never exceeded 125,000, and harvests (13-14,000) were only exceeded for a short period of liberal any sex hunting in the 1960s and early 1970s. controlling adult cow harvests is considered essential in maintaining huntable populations. long-term data sets should be consulted to guide realistic future wmu population and harvest targets. the economic impact of moose hunting and the annual number of user days is substantial. legg (1995) estimated a total sales impact (gross output) of $134.7 million in 1993, while bisset et al. (1999) reported 817,000 user days in 1997 (table 6). recent efforts to increase the emphasis on regional enforcement suggest the level of illegal hunting has been underestimated and may be significantly impacting population densities in some wmus. extending the “moose watch” enforcement effort to all regions in the fall of 2001 is a positive step in controlling illegal moose hunting activities. we suggest a full analysis of past charges be made to help identify areas in which hunter education is deficient and target additional education efforts where needed. we further recommend a review of the guaranteed group size option be conducted to determine if large groups licenced for 1 or 2 adult moose contribute significantly to the number of moose found shot and abandoned. much progress has been made towards m e e t i n g t h e m m p t a r g e t s r e g a r d i n g habitat management and research. continued support for ongoing habitat research projects and sufficient funding to complete ongoing studies is essential. consistent funding to evaluate connectivity between habitats prior to forest management planning is essential.we suggest further research is needed to help determine the magnitude of non-hunting mortality factors especially predation as well as the role of parasites and diseases, removal of moose by poaching losses, and first nation harvest. in addition, managers need to more closely examine changes in harvest sex/age composition, especially the magnitude of calf harvests. the value of periodic season closures and access control mechanisms to r e d u c e h u n t e r e f f i c i e n c y s h o u l d b e re-examined. all regions should closely e x a m i n e m o o s e r e l a t e d e n f o r c e m e n t charges to determine trends and motivating causes. hunter education was a major component of the mmp, and results 20 years later suggest the omnr failed to deliver on this component of the policy. voluntary moose hunter education courses were few alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 35 and far between and generally poorly attended. no mandatory course for new moose hunters was phased-in, nor was a firearm proficiency test initiated. evidence from hunter opinion surveys indicated a majority o f h u n t e r s i n i t i a l l y s u p p o r t e d t h e selective harvest program (rollins 1987, rollins and romano 1989), but more recently they have “lost faith in the system”(simmons 1997:56). we believe lack of hunter support and understanding can be partly traced to a failure to effectively communicate the program and better educate new hunters. voluntary courses simply did not attract enough hunte r s t o m a k e a d i f f e r e n c e . c u r r e n t communications efforts directed to hunters should be reviewed and re-evaluated. hunt quality, including realistic hunter expectations, needs to be re-examined and recognized as an essential component in overall hunter satisfaction when a new policy is drafted. managers must improve hunter communications related to setting and acheiving realistic population targets. we fully support the major emphasis p l a c e d o n h u n t e r e d u c a t i o n i n t h e recent discussion document toward a hunting management strategy for ontario (stewart 2000). we recommend a mandatory course for all new moose hunters, similiar to that offered by the ontario fede r a t i o n o f a n g l e r s a n d h u n t e r s f o r all new turkey hunters since 1987 (hhhf 2 0 0 0 ) . u n t i l h u n t e r s b e c o m e m o r e involved, knowledgeable, and responsible, their understanding and support will be lacking. further, results of population and harvest surveys should be published in the annual hunting regulations summary. viewing opportunities remain an under utilized component of the 1980 mmp who’s full potential remains unrealized. the majestic moose is a wilderness symbol and a much sought after species, especially for tourists to view and photograph. in future, moose viewing opportunities need to be better identified and funding made available to develop specific sites in a variety of wmus across moose range. the mmp has generated a host of numerical data which need to be carefully assessed and interpreted to evaluate the l e v e l o f t a r g e t a c h i e v e m e n t . s o m e values are suspect, especially when they deviate from long-term patterns. hence, managers need to recognize their inherent limitations when recommending management action. such a process should be reflected in a revised set of standards and guidelines that provide a unified approach to data interpretation and away from the past focus on chasing numbers. the importance of testing management policies and strategies in the field while monitoring their effects on wmu populations is emphasized. hunters, hunting organizations, and non-hunters all must be directly involved and support the soundness of recommended strategies. current target review exercises should involve all user groups and social interests. in the absence of such support, it is doubtful that any regulated harvesting concept would survive long enough to allow a clear response and evaluation of that response. we recommend a “basic program”, which includes linked population and harvest evaluation components. such a program needs to be properly staffed, funded, and coordinated. data collection, compilation, assessment, and central reporting should become a district/provincial priority. prompt data analysis and development of a simple user-friendly computer based reporting and data access program is essential to rebuilding program confidence. we believe the current québec moose management system which adopted a multi-harvest scenario in close cooperation with hunters has merit (courtois and lamontagne 1997). harvest control in québec varies ontario’s moose management policy timmermann et al. alces vol. 38, 2002 36 f r o m a l i b e r a l a p p r o a c h ( a n y s e x / age) in hunting zones with few problems, to very restrictive strategies including complete protection of cows for 5 years in zones where moose are scarce or where hunters wanted a rapid increase. such a varied, hunter-supported approach, if closely monitored, lends itself to adaptive management policies which retain the best, and reject strategies that prove ineffective or unsuccessful. finally we strongly recommend that all past and current wmu specific population and harvest data be compiled and published. consideration should also be given to carrying-out an independent biological evaluation of the current selective harvest program, before major changes are considered. such a review should compare ontario’s management policies with those of other jurisdictions to ensure moose continue to provide sustained benefits based on a sound biological rationale, consistent with recreational and economic objectives and program targets. acknowledgements we would like to thank the following for providing unpublished population, harvest, and enforcement data: ted armstrong, northwest regional wildlife biologist, thunder bay; art rodgers ungulate research scientist, thunder bay; howard smith, senior biologist, large mammals, wildlife section, fish & wildlife branch, peterborough; al bisset, retired leader, big game management information system, provincial assessment unit, kenora; bruce ranta, area biologist, kenora; charlie todesco, wawa distirct enforcement supervisor; and bob stewart, conservation officer, thunder bay district. we thank: mike buss, retired wildlife specialist, dorset; gord eason, area biologist, wawa; john mcnicol, retired wildlife specialist, forest policy section, thunder bay; harold cumming, retired lakehead university, forestry faculty, thunder bay; murray lankester, lakehead biology department, thunder bay; rosemary hartley, area biologist, nipigon; mark sobchuk, nw region fisheries biologist, thunder bay; dave euler, retired lakehead university, forestry faculty, thunder bay; ed addison, retired wildlife researcher, ontario ministry of natural resources, maple ontario; and gerry lynch, wildlife consultant, sherwood park, alberta, for suggestions on an earlier draft of this paper. references a d d i s o n , e . m . , d . g . j o a c h i m , r . f . mclaughlin, and d.j.h. fraser. 1998a. ovipositional development and fecund i t y o f d e r m a c e n t o r a l b i p i c t u s (acari:axodidae) from moose. alces 34: 165-172. , and r.f. mclaughlin. 1988. growth and development of winter tick, dermacentor albipictus, on moose, alces alces. journal of parasitology 74: 670-678. , and . 1993. seasonal variation and effects of winter ticks (dermacentor albipictus) on consumption of food by captive moose (alces alces) calves. alces 29: 219-224. , , and j.d. broadfoot. 1998b. effects of winter ticks (dermacentor albipictus) on blood characteristics of captive moose (alces alces). alces 34: 189-199. , j.d. smith, r.f. mclaughlin, d.j.h. fraser, and d.g. joachim. 1990. calving sites of moose in central ontario. alces 26: 142-153. , and l.m. smith. 1981. productivity of winter ticks (dermacentor albipictus) collected from moose killed on ontario roads. alces 17: 136-146. armstrong, e.r., and r. simons. 1999. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 37 m o o s e h u n t i n g o p p o r t u n i t i e s f o r physicallychallenged hunters in ontario: a pilot study. alces 35: 125-134. ballard, w.b. 1992. bear predation on moose: a review of recent north american studies and their management implications. alces supplement 1:1-15. , and v. van ballenberghe. 1998. predator/prey relationships. pages 247273 in a.w. franzmann and c.c. s c h w a r t z , e d i t o r s . e c o l o g y a n d management of the north american moose. smithsonian institution press, washington, d.c, usa. barbowski, j. 1972. mail surveys of moose hunters in ontario. proceedings of the north american moose conference and workshop 8:326-339. bergerud, a.t. 1981. the decline of moose in ontario a different view. alces 17: 3043. , and f. manuel. 1969. aerial census of moose in central newfoundland. journal of wildlife management 33: 910-916. , and j.b. snider. 1988.predation in the dynamics of moose populations: a reply. journal of wildlife management 52:559-564. , w. wyett, and j.b. snider. 1983. the role of wolf predation in limiting a moose population. journal of wildllife management 47: 977-988. bisset, a.r. 1991. standards and guidelines for moose population inventory in ontario. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. . 1992. moose management in the northwestern region: towards a new strategy. ontario ministry of natural resources, kenora, ontario. unpublished report, may 5, 1992. . 1993. the moose population of ontario revisited— a review of survey data, 1975-1992. ontario ministry of natural resources, kenora, ontario, canada. unpublished report. . 1996. standards and guidelines for moose population inventory in ontario. ontario ministry of natural resources, fish and wildlife branch, peterborough, ontario, canada. . 1999. mandatory reporting: an implementation strategy. northwest science and technology unit, thunder bay, ontario, canada. information report ir-003. , b. crowell, and c. hansson. 1998. moose aerial inventory pilot’s manual. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tm001. , l . d i x -g i b s o n , a n d m . a . mclaren. 2001. 1998 deer and moose harvest in ontario. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tr128. , , , j.mellor, and c. davies. 1999. 1997 deer and moose harvest in ontario. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tr117. , and m.a. mclaren. 1995. moose population inventory plan for ontario, 1996-1998. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. information report ir-002. , and . 1999. moose popul a t i o n a e r i a l i n v e n t o r y p l a n f o r ontario: 19992002. ontario ministry of natural resources, northwest science and technology unit, thunder ontario’s moose management policy timmermann et al. alces vol. 38, 2002 38 bay, ontario, canada. information report ir-004. , , c . l . e d m o n d s , w . s a w y e r , w . g l u h u s h k i n , k . morrison, and c. davies. 1997. report on the 1995-96 and 1996-97 moose population surveys: with considerations for future surveys. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tr113. , , , and . 2000. report on the 19981999 moose population surveys. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tr-127. , and r.s. rempel. 1991. linear a n a l y s i s o f f a c t o r s a f f e c t i n g t h e accuracy of moose aerial inventories. alces 27: 127-139. , and h.r. timmermann. 1983. resource allocation: an ontario solution. alces 19: 178-190. bottan, b j. 1999. exploring the human dimension of thunder bay moose hunters with focus on choice behaviour and e n v i r o n m e n t a l p r e f e r e n c e s . m . a . thesis, faculty of forestry and forest environment, lakehead university, thunder bay, ontario, canada. b u s s , m . , r . g o l l a t a n d h . r . ti m m e r m a n n. 1989. moose hunter shooting proficiency in ontario. alces 25: 98-103. connor, j. 1986. early winter utilization b y m o o s e o f g l y p h o s a t e t r e a t e d cutovers. b.sc.f. thesis, lakehead university, thunder bay, ontario, canada. , and l. mcmillan. 1988. winter utilization by moose of glyphosate treated cutoversan interm report. alces 24: 133-142. courtois, r., and g. lamontage. 1997. management system and current status of moose in quebec. alces 33: 97-114. crête, m. 1987. the impact of sport hunting on north american moose. swedish wildlife research supplement 1: 553-563. . 1989. approximation of k carrying capacity for moose in eastern quebec. canadian journal of zoology 67: 373-380. , r.j. taylor, and p.a. jordan. 1981. optimization of moose harvests in southwestern québec. journal of wildlife management 45:598-611. cumming, h.g. 1974. moose management i n o n t a r i o f r o m 1 9 4 8 t o 1 9 7 3 . naturaliste canadien 101: 673-687. . 1980. relation of moose track counts to cover types in north-central ontario. proceedings of the north american moose conference and workshop 16: 444-462. . 1985. first year effects on moose browse from two silvicultual applications of glyphosate in ontario.alces 25:118-132. . 1987. sixteen years of moose browse surveys in ontario. alces 23: 125-156. dalton, w.j. 1989. use by moose (alces alces) of clearcut habitat where 100% or 50% of the production forest was logged. cofrda project 32001, ontario ministry of natural resources, toronto, ontario, canada. eason, g. 1985. overharvest and recovery of moose in a recently logged area. alces 21: 55-75. . 1989. moose response to hunting and one km2 block cutting. alces 25: 63-74. euler, d. 1981. a moose habitat strategy for ontario. alces 17:180-192. . 1983. selective harvest, compensatory mortality and moose in ontario. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 39 alces 19: 148-161. . 1994. page 179 in environmental a s s e s s m e n t b o a r d , r e a s o n s f o r decision and decision. class environmental assessment by the ministry of natural resources for timber management on crown lands in ontario. queen’s printer for ontario, toronto, ontario, canada. f e r g u s o n, s.h., a.r. bi s s e t , and f. messier. 2000. the influence of density on growth and reproduction in moose alces alces. wildlife biology 6: 31-39. fraser, d.j. , d. arthur, j.k. morton, and b.k. thompson. 1980. aquatic feeding by moose alces alces in a canadian lake. holarctic ecology 3: 218-223. , e.r. chavez, and j.e. paloheimo. 1984. aquatic feeding by moose: selection of plant species and feeding areas in relation to plant chemical composition and characteristics of lakes. canadian journal of zoology 62: 80-87. , and h. hristienko. 1983. effects of moose alces alces, on aquatic vegetation in sibley provincial park, ontario, canada. canadian field-naturalist 7: 57-61. , and e. reardon. 1980. attraction of wild ungulates to mineral-rich springs in central canada. holarctic ecology 3: 36-40. , and e.r. thomas. 1982. moosevehicle accidents in ontario: relation to highway salt. wildlife society bulletin 10: 261-265. , b.k. thompson, and d. arthur. 1982. aquatic feeding by moose: seasonal variation in relation to plant chemical composition and use of mineral licks. canadian journal of zoology 60: 3121-3126. fruetel, m., and m.w. lankester. 1988. nematodirella alcidis (nematoda: trichostrongyloidea) in moose of northwestern ontario. alces 24: 159163. gasaway, w.c., and s.d. dubois. 1987. estimating moose population parameters. swedish wildlife research supplement 1: 603-617. , , d.j. re e d, and s.j. h a r b o . 1 9 8 6 . e s t i m a t i n g m o o s e population parameters from aeial surveys. biological papers of the university of alaska, number 22. fairbanks, alaska, usa. , r.o. stephenson, j.l. davis, p.e.k. shepherd, and o.e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska.wildlife monographs 84. glooshenko, v., c. downes, r. frank, h.e. braun, e.m. addison, and j. hickie. 1988. cadmium levels in ontario moose and deer in relation to soil sensitivity to acid precipitation. science of the total environment 71: 173186. gollat, r., and h.r. timmermann. 1983. determining quotas for a moose selective harvest in north central ontario. alces 19: 191-203. , and .1987. evaluating ontario moose harvests using a postcard questionnaire. alces 23: 157-180. , , and j.mcnicol. 1985. a r e v i e w o f m e t h o d o l o g y u s e d t o formulate moose quotas, northcentral region. ontario ministry of natural resources, thunder bay, ontario, canada. greenwood, c., d. euler, and k. m orrison. 1984. standards and guidelines for the determination of allowable moose harvest in ontario. ontario ministry of natural resources, toronto, ontario, canada. unpublished report. hansen, s. 1995. moose hunter opinion survey:perceived satisfaction with the ontario moose management system. unontario’s moose management policy timmermann et al. alces vol. 38, 2002 40 published dissertation, department of outdoor recreation, parks and tourism, lakehead university, thunder bay, ontario, canada. , w.j. dalton, and t. stevens. 1995. an overview of a hunter opinion survey of satisfaction with the ontario moose management system. alces 31: 247-254. hénault, m., l. belanger, a.r. rodgers, g. redmond, k. morris, f. potvin, r. courtois, s. morel, and m. mongeon. 1999. moose and forest ecosystem management: the biggest beast but not the best. alces 35: 213-225. heydon, c., d. euler, h. smith, and a. bisset. 1992. modelling the selective moose harvest program in ontario. alces 28:111-121. (hhhf) hunting heritage, hunting futures. 2000. wildlife chronicle — a history of ontario’s wildlife legacy. hunting heritage, hunting futures. huntsville, ontario, canada. jackson, g.l., g.d. racey, j.g. mcnicol, and l.a. godwin. 1991. moose habitat interpretation in ontario. ontario ministry of natural resources, northwest ontario forest technology development unit, thunder bay, ontario, canada. technical report 52. kelly, c.p., and h.g. cumming. 1994. effects of vision application on moose winter browsing and hardwood vegetation. alces 30: 173-188. kennedy, m.j., m.w. lankester, and j.b. snider. 1985. paramphistomum cervi and paramphistomum liorchis (digenea: paramphistomatidae) in moose, alces alces, from ontario. canadian journal of zoology 63:1207-1210. kolenosky, g.b. 1981. status and management of wolves in ontario. pages 35-40 in l. carbyn, editor. wolves in canada a n d a l a s k a . c a n a d i a n w i l d l i f e service, ottawa, ontario, canada. report series no. 5. kronberg, b.i., and v. glooschenko. 1994. investigating cadmium bioavailability in nw ontario using boreal forest plants. alces 30: 71-80. lankester, m.w. 1987. pests, parasites, and diseases of moose (alces alces) in north america. swedish wildllife research supplement 1:461-489. , and t.j. bellhouse. 1982. pathol o g i c a l a n o m a l i e s i n m o o s e o f northwestern ontario. alces 18: 1724. , and r.d. sein. 1986. the moose fly haematobosca alcis (muscidae) and skin lesions on alces alces. alces 22: 361-376. lawson, e.j.g., and a.r. rodgers. 1997. differences in home-range size computed in commonly used software programs. wildlife society bulletin 25: 721-729. legg, d. 1995. the economic impact of moose hunting in ontario, 1993. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. lumsden, h.g. 1958. ontario moose inventory 1958. department of lands and forests, wildlife branch, toronto, ontario, canada. unpublished report. . 1959. ontario moose inventory winter, 1958-59. ontario department of lands and forests, wildlife branch, toronto, ontario, canada. unpublished report. lynch, g.m., and s. birkholz. 2000. a telephone questionnaire to assess moose harvest. alces 36: 105-109. mastenbrook, b., and h. cumming. 1989. use of residual strips of timber by moose within cutovers in northwestern ontario. alces 25: 146-155. mc ke n n e y , d.w., r.s. re m p e l, l.a. venier, y onghe wang, and a.r. bisset. 1998. development and application of alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 41 a spatially explicit moose population model. canadian journal of zoology 76: 1922-1931. mcmillian, l.m., h.r. timmermann, and m.e. buss. 1993. access management relative to the vulnerability of moose to recreational harvest a discussion paper. ontario ministry of natural resources, north central region, thunder bay, ontario, canada. mcnicol, j.g , and j. baker. 1998. identification of early winter and late winter moose habitat.pages 1 to 9 in w. b. ranta, editor. selected wildlife and habitat features: inventory manual for use in forest management planning. version 1.0. queen’s printer, toronto, ontario, canada. , and f.f. gilbert. 1980. late wint e r u s e o f u p l a n d c u t o v e r s b y moose. journal of wildlife management 44: 363-371. , and h.r. timmermann. 1981. eff e c t s o f f o r e s t r y p r a c t i c e s o n ungulate populations in boreal mixed forest. pages 141-154 in proceedings of the boreal mixedwood symposium. canadian forestry service, ottawa, ontario, canada. , , and r. gollat. 1980. the effects of heavy browsing pressure over eight years on a cutover in quetico park. proceedings of the north american moose conference and workshop 16:360-373. mercer, w.b. and b.e. mclaren. 2002. evidence of carrying capacity effects in newfoundland moose.alces 38:123141. morris, r.f. 1959. single factor analysis in population dynamics. ecology 40: 580-588. novak, m., and j. gardner. 1975. accur a c y o f m o o s e a e r i a l s u r v e y s . proceedings of the north american moose conference and workshop 11: 154-180. (oeab) ontario environmental assessment board. 1994. class environmental assessment by the ministry of natural resources for timber management on crown lands in ontario. queen’s printer for ontario, toronto, ontario, canada. (omnr) ontario ministry of natural resources. 1980a. moose management in ontario: a report of open house public meetings. queen's printer for ontario, toronto, ontario, canada. . 1980b. moose management policy. queen’s printer for ontario, toronto, ontario, canada. . 1980c. standards and guidelines f o r m o o s e a e r i a l i n v e n t o r y i n ontario. ontario ministry of natural resources, toronto, ontario, canada. . 1982. northwestern ontario strategic land use plan. ontario ministry of natural resources,toronto ontario, canada. . 1984. moose hunter’s handbook 1984. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. . 1987. improving the quality and e n j o y m e n t o f m o o s e h u n t i n g i n ontario: background report. wildlife branch, ontario ministry of natural resources, toronto, ontario, canada. . 1988a. timber management guidelines for the provision of moose habitat. ontario ministry of natural resources, toronto, ontario, canada. . 1988b. moose anatomy for the hunter. ontario ministry of natural resources, wildlife branch, toronto ontario, canada. . 1990. the moose in ontario. book #1moose biology, ecology and management, chapters 17, book 2 moose hunting techniques, hunting ethics and the law, chapters 8-14. onontario’s moose management policy timmermann et al. alces vol. 38, 2002 42 tario federation of anglers and hunters, peterborough, ontario, canada. . 1991. review of adult moose validation draw. final report. june 1991. queeen’s printer for ontario, toronto ontario, canada. . 1993. a minimum wildlife program. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. . 1997. 1996 moose hunt summary. queen’s printer for ontario, toronto, ontario, canada. . 1999. ontario government response to the consolidated recommendations of the boreal west, boreal east and great lakes-st. lawrence round tables. queen’s printer for ontario, toronto, ontario, canada. . 2000. 2000 hunting regulations summary. fall 2000 spring 2001. ontario ministry of natural resources, toronto, ontario, canada. . 2001. 2000 review of moose population objectives in ontario. draft report of 3 regional workshops. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. oswald, k . 1982.moose aerial observation manual. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. . 1997. moose aerial observation manual. ontario ministry of natural resources, northeast science and technology unit, timmins, ontario, canada. technical manual tm008. payne, d., j.g. mcnicol, g. eason, and d. abraham. 1988. moose habitat management planning: three case studies. forestry chronicle 64: 270-276. peterson, r.o. and d.l. allen. 1974. snow conditions as a parameter in moosewolf relationships. naturaliste canadien 101: 481-492. provincial auditor. 1998. audit of the ministry of natural resources, fish and wildlife program. toronto, ontario, canada.(http://www.gov.on.ca/ opa/english/e98/309.htm). r a c e y , g.d., l.m. m c m i l l a n , h.r. timmermann, and r. gollat. 2000. effects of hunting closures and timber harvest on local moose densities and hunting opportunities in northwestern ontario: a case study. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. technical report tr-85. , j . d . m c n i c o l , a n d h . r . timmermann. 1989a. application of moose and deer habitat guidelines: impact on investment. pages 119-131 in forest investment: a critical look. canadian forest research centre symposium proceedings. o-p-17. , t.s. whitfield and r.a. sims. 1989b. moose habitat. pages 4-3 – 4-4 in northwestern ontario forest ecosystem interpretations. ontario ministry of natural resources, northwest ontario forest technology develoment unit, thunder bay, ontario, canada. technical report 46. ranta, w.b., editor. 1998. selected wildlife and habitat features: inventory manual, for use in forest management planning. version 1.0. queen’s printer, toronto, ontario, canada. rempel, r.s., p.c. elkie, a.r. rodgers, and m.j. gluck. 1997a. timber management and natural disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61: 517-524. , g.d. racey and k.a. cumming. 1997b . predicting moose browse production using the northwestern ontario forest ecosystem classification. alces 33: 19-31. alces vol. 38, 2002 timmermann et al. ontario’s moose management policy 43 , and a.r. rodgers. 1997. effects o f d i f f e r e n t i a l c o r r e c t i o n o n accuracy of a gps animal location system. journal of wildlife management 61: 525-530. , , and k.f. abraham. 1995. p e r f o r m a n c e o f a g p s a n i m a l location system under boreal forest canopy. journal of wildlife management 59: 543-551. rodgers, a.r., and p. anson. 1994. animal-borne gps: tracking the habitat. gps world 5: 20-32. , and a.p. carr. 1998. hre: the home range extension for arcview. users manual. ontario ministry of natural resources, thunder bay, ontario, canada. , r.s. rempel, and k.f. abraham. 1995. field trials of a new gps-based telemetry system. pages 173-178 in c. cristalli, c.j. amlaner, jr., and m.r. neuman, editors. biotelemetry xiii, proceedings of the 13th international symposium on biotelemetry, march 2631,1995, williamsburg, virginia, usa. , , and . 1996. a g p s b a s e d t e l e m e t r y s y s t e m . wildlife society bulletin 24: 559-566. , , and b. allison. 2000. moose guidelines evaluation project. centre for northern forest ecosystem research, ontario ministry of natural resources, thunder bay, ontario, canada. , , r. moen, j. paczkowski, c.c. schwartz, e.j. lawson, and m.j. gluck. 1997. gps collars for moose telemetry studies: a workshop. alces 33: 203209. , s.m. tomkiewicz, e.j. lawson, t.r. stephenson, k.j.hundertmark, p.j. wilson, b.n. white, and r.s. rempel. 1998. new technology for moose management: a workshop. alces 34: 239-244. rollins, r. 1987. hunter satisfaction with t h e s e l e c t i v e h a r v e s t s y s t e m f o r moose in northern ontario. alces 23: 181-193. , and l. romano. 1988. hunter attit u d e s t o t h e s e l e c t i v e h a r v e s t system in northern ontario: a longitudinal study. unpublished dissertation, s c h o o l o f o u t d o o r r e c r e a t i o n , lakehead university, thunder bay, ontario, canada. , and . 1989. hunter satisfaction with the selective harvest system for moose management in ontario. wildlife society bulletin 17: 470475. romano, l.a. 1988. a comparative study: hunter’s attitudes to the selective harvest system for wildlife managem e n t u n i t s i n n o r t h e r n o n t a r i o . unpublished dissertation, school of outdoor recreation, lakehead university, thunder bay, ontario, canada. samuel, w.m. and m.j. barker. 1979. the winter tick, dermacentor albipictus (packard, 1869) on moose, alces alces (l.), of central alberta. proceedings of the north american moose conference and workshop 15: 303-348. schwartz, c.c., and a.w. franzmann. 1991. interrelationship of black bears to moose and forest succession in the northern coniferous forest. wildlife monographs 113. simmons, g. 1997. independent review of the moose and deer tag allocation for ontario. recommendations from ont a r i o ’ s h u n t e r s . q u e e n ' s p r i n t e r , toronto, ontario, canada. smith, h. 1990. managing a moose population. pages 25-39 in m. buss and r. truman, editors. the moose in ontario, book 1. ontario ministry of natural resources, wildlife branch, toronto, ontario, canada. snider, j.b., and m.w. lankester. 1986. ontario’s moose management policy timmermann et al. alces vol. 38, 2002 44 rumen flukes (paramphistomum spp.) in moose of northwestern ontario. alces 22: 323-344. stewart, a. 2000. discussion document towards a hunting management strategy for ontario. hunting heritage, hunting futures, huntsville, ontario, canada. stewart, r.r., r.r. maclennan, and j.d. kinnear. 1977. the relationship of plant phenology to moose. saskatchewan department of tourism and renewable resources. technical bulletin number 3. sylvén, s. 1995. moose harvest strategy to maximize yield value for multiple goal managementa simulation study. agricultural systems 49: 277-298. thompson, i.d. 1979. a method of correcting population and sex and age estimates from aerial transect surveys for moose. proceedings of the north american moose conference and workshop 15: 148-168. , and d.l. euler. 1987. moose habitat in ontario: a decade of change in perception. swedish wildlife research supplement 1:181-193. , and r.o. peterson. 1988. does wolf predation alone limit the moose population in pukaskwa park? a comment. journal of wildlife management 52: 556-559. , and m.f. vukelich. 1981. use of logged habitats in winter by moose cows with calves in northeastern ontario. canadian journal of zoology 59: 21032114. , d.a. welsh, and m.f. vukelich. 1981. traditional use of early-winter concentration areas by moose in northeastern ontario. alces 17: 1-14. timmermann, h.r. 1974. moose inventory methods: a review. naturaliste canadien 101: 615629. . 1987. moose harvest strategies in north america. swedish wildlife research supplement 1 :565-579. . 1991. ungulates and aspen management. pages 99-110 in s. navratil and p.b. chapman, editors. aspen management for the 21st century. forestry canada, edmonton, alberta,canada. . 1992. moose hunter education in north america. alces supplement 1: 65-76. .1993. use of aerial surveys for estimating and monitoring moose populationsa review. alces 29: 35-46. . 1998. importance and use of mixedwood sites and forest cover for moose (alces alces). boreal mixedwood technical note. queen's printer, toronto, ontario, canada. , and m.e. buss. 1995. the status and management of moose in north americaearly 1990s. alces 31: 1-14. , and . 1998. population a n d h a r v e s t m a n a g e m e n t . p a g e s 559-615 in a.w. franzmann and c.c. s c h w a r t z , e d i t o r s . e c o l o g y a n d management of the north american moose. smithsonian institution press, washington, d.c., usa. , and r. gollat. 1982. age and sex structure of harvested moose related to season manipulation and access. alces 18: 301-328. , and . 1984. sharing a moose in north central ontario. alces 20:161-185. , and . 1986. selective moose harvest in north central ontarioa progress report. alces 22: 395417. , and . 1994. early winter social structure of hunted vs unhunted moose populations in n. central ontario. alces 30: 117-126. , and m.w. lankester. 1980. studi e s o f w i n t e r t i c k , d e r m a c e n t o r albipictus, on the bell of moose in northalces vol. 38, 2002 timmermann et al. ontario’s moose management policy 45 w e s t e r n o n t a r i o . p r o c e e d i n g s o f the north american moose conference and workshop 16: 137-151. , and j.g. mcnicol. 1988. moose habitat needs. forestry chronicle 64: 238-245. , and g.d. racey. 1989. moose access routes to an aquatic feeding site. alces 25: 104-111. , , and r. gollat. 1990. moose cratering for equisetum in early winter. alces 26: 86-90. , and r.s. rempel. 1998. age and sex structure of hunter harvested moose u n d e r t w o h a r v e s t s t r a t e g i e s i n n o r t h c e n t r a l o n t a r i o a l c e s 3 4 : 21-30. , and h.a.whitlaw. 1992. selective moose harvest in north central ontario. alces 28: 137-163. , , and a.r. rodgers. 1993. testing the sensitivity of moose harvest data to changes in aerial population estimates in ontario. alces 29: 4753. todesco, c.j. 1988. winter use of upland c o n i f e r a l t e r n a t e s t r i p c u t s a n d clearcuts by moose in the thunder bay district. m.sc.f. thesis, lakehead university, thunder bay, ontario, canada. , h.g. cumming, and j.g. mcnicol. 1985. winter moose utilization of alternate strip cuts and clearcuts in northwestern ontario: preliminary results. alces 21: 447-474. van ballenberghe, v., and w. ballard. 1994. limitation and regulation of moose populations: the role of predation. canadian jounal of zoology 72: 2071-2077. wedeles, c.h.r., h. smith, and r. rollins. 1989. opinions of ontario moose hunters on changes to the selective harvest system. alces 25: 15-24. welch, i.d., a.r. rodgers, and r.s. mckinley. 2000. timber harvest and calving site fidelity of moose in northwestern ontario. alces 36: 93-103. whitlaw, h. a. 1993. an evaluation of the effects of parelaphostrongylosis on moose populations. m.a. thesis, lakehead university, thunder bay, ontario, canada. , and m.w. lankester. 1994a. a r e t r o s p e c t i v e e v a l u a t i o n o f t h e effects of parelaphostongylosis on moose populations. canadian journal of zoology 72: 1-7. , and . 1994b. the co-occ u r r e n c e o f m o o s e , w h i t e t a i l e d deer, and parelaphostrongylus tenuis in ontario. canadian journal of zoology 72: 819-825. , h.r. timmermann, p. pernsky, and a.r. bisset. 1993. northwest region moose population and harvest profile. ontario ministry of natural resources, northwest science and technology unit, thunder bay, ontario, canada. report number 75. wilton, m.l. 1983. black bear predation on young cervids a summary. alces 19: 136-147. , and d.l. garner. 1991. preliminary findings regarding elevation as a major factor in calving site selection in south central ontario, canada. alces 27: 111-117. , and . 1993. preliminary observations regarding mean april temperature as a possible predictor of tick-induced hairloss on moose in south central ontario, canada. alces 29: 197200. alces 46 (2010) contents in memoriam albert w. franzmann ...................................................................... i albert w. franzmann and distinguished colleagues memorial award.................................................................................................. iv fifty years of food and foraging in moose: lessons in ecology from a model herbivore ....................................................... lisa a. shipley 1 optimal harvesting of moose in alberta ............................................... ................................................................................. cailin xu and mark s. boyce 15 a history of moose management in utah ................................................. ...................................... michael l. wolfe, kent r. hersey, and david c. stoner 37 understanding the impact of meningeal worm, parelapho strongylus tenuis, on moose populations ...... murray w. lankester 53 activity patterns, foraging ecology, and summer range car rying capacity of moose (alces alces shirasi) in rocky mountain national park, colorado ............. jason d. dungan, lisa a. shipley, and r. gerald wright 71 modeling seasonal distribution and spatial range capacity approximations of moose in southeastern wyoming ....................... ...... phillip e. baigas, richard a. olson, ryan m. nielson, scott n. miller, and frederick g. lindzey 89 potential vulnerability of bull moose in central british columbia to three antler-based hunting regulations .................... ... kenneth n. child, daniel a. aitken, roy v. rea, and raymond a. demarchi 113 morphometry of moose antlers in central british columbia ..... ............................................. kenneth n. child, daniel a. aitken and roy v. rea 123 the impact of moose (alces alces andersoni) on forest regeneration following a severe spruce budworm outbreak in the cape breton highlands, nova scotia, canada .......................... ..................... craig smith, karen beazley, peter duinker, and karen a. harper 135 nutritonal condition of adult female shiras moose in northwest wyoming .. scott a. becker, matthew j. kauffman, and stanley h. anderson 151 reducing non-target moose capture in wolf snares ...................... ..................................................................................................... craig l. gardner 167 youtubetm insights into moose-train interactions ........................... ............................................ roy v. rea, kenneth n. child, and daniel a. aitken 183 an advisory committee process to plan moose management in minnesota ............... amanda m. mcgraw, ron moen, grant wilson, andrew edwards, rolf peterson,louis cornicelli, mike schrage, lee frelich, mark lenarz, and dennis becker 189 (continued on inside back cover) 44th north american moose conference and workshop ................ 201 previous meeting sites................................................................................. 203 distinguished moose biologist kenneth n. child ............................ 204 distinguished moose biologist past recipients............................... 205 distinguished moose biologist award criteria ............................... 206 editorial review committee........................................................................ 207 additional copies available from: lakehead university bookstore, thunder bay, ontario, canada p7b 5e1 alces 39-46 price $40.00 canadian or u.s. each (including supplementary issues) alces 24-38 price $38.00 canadian or u.s. each (including supplementary issues) alces 17-23 price $20.00 canadian or u.s. each proceedings of the north american moose conference and workshop 8-16 price $20.00 canadian or u.s. each make cheques, money orders or purchase orders payable to lakehead university bookstore. all prices include 5% g.s.t., mailing and handling costs. prices are subject to change. acknowledgements brooke pilley worked long hours formatting and typesetting manuscripts. alces home page further information on contents of past issues, prices, ordering, as well as instructions to submitting authors, are available at our website: http://bolt.lakeheadu.ca/~alceswww/alces.html alces32_41.pdf alces32_85.pdf alces32_141.pdf alces37(2)_411.pdf alces37(2)_435.pdf alces 31_1.pdf 43 demographic status of moose populations in the boreal plain ecozone of canada a. alan arsenault1, arthur r. rodgers2, and kent whaley3 1wood canada limited, 4015 millar avenue, saskatoon, saskatchewan, canada s7k 2k6; 2ontario ministry of natural resources and forestry, centre for northern forest ecosystem research, 103-421 james street south, thunder bay, ontario, canada p7e 2v6; 3government of manitoba, manitoba sustainable development (retired), box 2550, the pas, manitoba, canada r9a 1m4. abstract: broad scale analyses of winter population survey data collected between 1985 and 2015 were conducted to provide a synthesis of the current status and historical performance of 14 moose (alces alces) populations residing in the boreal plain ecozone of saskatchewan and western manitoba. population time series models indicated a broad scale decline averaging 30% in moose populations across the boreal plain ecozone since 2000 relative to the long-term (1985 to 2015) cumulative mean population size. demographic patterns and rates of population change were variable among and within populations across years. we found an inverse relationship between adult sex ratio (bull:cow) and population density (r² = 0.48, p < 0.001), which suggests negative population growth (λ < 1.0) when the adult sex ratio falls below a density-dependent threshold for population growth. winter calf recruitment (calves/cow) was positively correlated (r² = 0.12, p = 0.027) with adult sex ratio. stable or increasing populations (λ ≥ 1.0) tended to have lower adult sex ratios relative to winter calf recruitment ratios than declining populations. population state and vital rate relationships are useful to assess population performance and guide science-based moose management strategies in a management-byobjective decision-analytic framework. alces vol. 55: 43–60 (2019) key words: alces alces, boreal plain ecozone, demography, moose, population, management-by objective moose (alces alces) population density in north america varies geographically and temporally (messier 1994, timmermann and rodgers 2017). divergent trends in abundance include an apparent decline across much of the continental moose range (laliberte and ripple 2004, murray et al. 2006, delgiudice 2013, mccann et al. 2013, kuzyk 2016) that is in contrast with increase of certain populations along the southern periphery of moose range and on both coasts (foster et al. 2002, darimont et al. 2005, faison et al. 2010, musante et al. 2010, murray et al. 2012, laforge et al. 2016, timmermann and rodgers 2017). however, there is limited detailed information regarding the magnitude and trend of population change at regional scales including in the boreal plain ecozone. demography of a given moose population is strongly influenced by metrics of population state (abundance, age/sex structure), vital rates (annual finite rate of population change [λ], survival, and recruitment), and movement dynamics (immigration, emigration). adult female survival and calf recruitment are well studied in ungulate populations and have a dynamic influence on ungulate population demography, λ, and abundance (gaillard et al. 2000, eberhardt moose population demography – arsenault et al. alces vol. 55, 2019 44 2002, raithel et al. 2007, environment canada 2012, monteith et al. 2015). landscape (configuration, dynamics), habitat (condition, availability), and temporal (seasonal) effects also contribute to the complexity of interacting variables that govern population performance. consequently, the dynamics of one population can substantially differ from those characterizing another. reliable information on regional moose population dynamics is central to identifying drivers of population change and informing management and conservation action (taber and raedeke 1979). surveying winter populations across multiple years provides a sequential time series of population state and vital rate metrics useful to model abundance trends and demographic changes that provide inference about population performance (taber and raedeke 1979, eberhardt 2002). moose populations in the boreal plain ecozone of saskatchewan and manitoba are more exploited by hunting and incur greater harvest levels relative to boreal shield ecozone populations further north (arsenault 2000, government of manitoba 2014). however, neither province monitors hunting mortality consistently or accurately at the local scale because of low response rate to harvest questionnaires; manitoba recently ceased using these questionnaires. in addition, harvest by rights-based hunters has never been monitored in either province. poorly managed harvest mortality can have substantial effects on the sex and age structure of a population, especially if age or sex classes are selectively harvested (slalski et al. 2005). for example, selective and excessive harvest can influence productivity through skewed sex ratio and age class distribution. therefore, management strategies should consider maintenance of appropriate adult sex ratios to ensure maximum reproductive efficiency in hunted moose populations (raedeke et al. 2002). given the numerous factors that affect the performance of moose populations, the challenge for managers is to interpret the various relationships identified with survey data and other information in a structured, decision-making process to provide science-based recommendations for population management. management-by objective is a results-based performance appraisal approach accomplished through strategic planning and population modeling (strickland 1985, arsenault 2000, thiele 2007). the establishment of numerical population metrics (i.e., abundance, population structure, recruitment, and λ) from longterm data sets and subsequent demographic modelling helps to develop appropriate, biologically sustainable management strategies and a means to evaluate management prescriptions through population performance monitoring. management-byobjective requires development of area specific management goals and numerical population objectives that are biologically and ecologically sound. this requires systematic population data collection, analysis, and evaluation of population performance relative to goals and objectives. as a prerequisite to management actions, it is important to understand the relationships among these various aspects to apply biological principles appropriately in a structured decision-making process (sauer and knutson 2008, artelle et al. 2018). the objectives of our analyses in the absence of reliable harvest data for the boreal plain ecozone were to: 1) estimate and evaluate long-term demographic trends of our study populations, 2) present an overview of moose population status in the boreal plain ecozone, and 3) provide suggestions for evaluating population demographic performance within a management by-objective framework. alces vol. 55, 2019 moose population demography – arsenault et al. 45 study area our study area was within the boreal plain ecozone of saskatchewan and western manitoba (fig. 1) that lies south of the precambrian shield (boreal shield ecozone) and north of the aspen parkland ecoregion of the prairie ecozone (ecological stratification working group 1995, padbury et al. 1998, smith et al. 1998, marshall et al. 1999), and where continuous moose population survey data were available and hunting occurs. the local topography is influenced by underlying glacial deposits and characterized by closed-crown mixed wood and coniferous forest, interspersed with peatland complexes, riparian watercourses, wetlands, and lakes. the forested landscape incurs substantial timber harvesting and wildfire suppression. the temperate climate is characterized by long, cold (x̅ january = −15°c) and snowy winters and shorter warm (x̅ july = +15°c) and moist summers; average annual precipitation is 450 mm. the boreal plain ecozone is more productive moose range than the adjacent boreal shield ecozone and can support higher population densities (arsenault 2000), but it also contains more linear access development for resource extraction and recreation. the 14 discrete moose populations in our study area were subject to predation fig. 1. study area and moose management units (mmu) delineated within the boreal plain ecozone of saskatchewan and western manitoba. the primary range represents core moose distribution and high-quality habitats; secondary range represents lower quality habitats and/or discontinuous moose distribution. moose population demography – arsenault et al. alces vol. 55, 2019 46 from wolves (canis lupus) and black bears (ursus americanus). population survey data were available for all, and each incurred harvest mortality from licensed and rights-based hunting. three populations had portions of their range that provided refuge from hunting within a national park (sled-prince albert national park, riding mountain) or an air-weapons range (meadow). four populations (candle-cub, duck mountain, meadow, and porcupine) had areas of restricted harvest in portions that were provincial parks. all populations were subject to the effects of landscape disturbance and fire suppression. these are open populations that likely mix along the periphery of adjacent boundaries, but each is unique in the configuration, quality, and amount of moose habitat, as well as magnitude or type of landscape scale, anthropogenic disturbances including agriculture, forestry, mining and exploration, recreation, linear development, and urban development. methods survey data all population survey data were acquired from publicly funded wildlife survey programs implemented by the saskatchewan and manitoba governments. prior to 1984, population structure and trend in saskatchewan were monitored with a strip transect sampling design. a habitat-stratified random block quadrat sampling design (stewart 1983) was used to obtain winter population structure and density estimates from 1984 to 1995; a >10% sightability error was maintained 95% of the time (i.e., estimates within a 95% ci). in 1996, saskatchewan transitioned to the gasaway et al. (1986) sampling design, as modified by lynch and shumaker (1995), that used 5.0 × 5.0 km survey units (utm grid system) to assess winter population structure and density corrected for sightability (90% ci). manitoba monitored moose populations with a strip transect method prior to 1992, and subsequently adopted the modified gasaway survey method that used 3.5 × 5.5 km survey units based on 3-minute grid cells (wgs 84) to obtain winter population structure and density estimates (95% ci; knudsen 2007). the gasaway surveys were conducted when snow conditions (>30 cm snow cover) and timing (january to early february) optimized sightability. one exception was the annual cervid survey in riding mountain national park and adjacent farmland that employed a strip transect design with ~25% coverage of the area (tarleton 1992). population trend and demographic analyses survey data were acquired for all moose populations regularly sampled in both provinces within the boreal plain ecozone over a 31-year period (1985 to 2015). because the survey data were collected with different methods, our analyses were constrained to those that produced winter population structure and density estimates with known confidence limits (i.e., modified gasaway surveys conducted between 1992 and 2015 inclusive; n = 41). abundance estimates with known confidence limits obtained from habitat-stratified random block quadrat sampling (1985 to 2005 inclusive; n = 12) were used only to inform population trend models, and were excluded from demographic analyses because sample size of classified animals was limited. all survey data were spatially partitioned into 14 moose management units composed of combinations of saskatchewan wildlife management zones (wmzs) and manitoba game hunting areas (ghas) (fig. 1) that are the administrative units used to allocate hunting licenses, but do not (in most cases) delineate individual moose populations. moose management alces vol. 55, 2019 moose population demography – arsenault et al. 47 units present an ideal landscape scale at which population assessment, management strategies, and policies are implemented (arsenault 2000, funk et al. 2012) because they share contiguous geophysical landscapes and similar ecological characteristics. importantly, they encompass the core and fringe distribution of local populations based on spatial distribution from interpolation of aerial survey data and landscape features. each management unit population (n = 14) was reconstructed in microsoft excel© from survey data spanning the 31-year period to discern patterns of population structure, abundance, and trend as a time series model (white 2000). model construction involved linear interpolation of abundance and demographic structure data between survey years. a third-degree polynomial was used to fit a long-term population trend line to the 3-year moving average of abundance estimates for each management unit. the polynomial was used because it is more sensitive to fluctuations in population size than a linear or log-linear trend line (kuzyk 2016). the objective of model fitting was to identify periods of population increase and decline within each management unit and to enable assessment of population performance metrics both within and among management units with respect to changes over time in: 1) population state (winter abundance and demographic composition), 2) vital rates (λ, calf recruitment, sex ratio), and 3) historical range of variability. the time series population reconstruction models were used to generate annual estimates of λ in each management unit to determine whether each gasaway population estimate occurred during a period of increase, stability, or decline as well as to determine if the estimate was above or below the long-term mean for that population. we calculated the annual finite rate of population change as λ = n t+1 /n t . investigation of demographic relationships between population structure and density relied solely on survey data collected using the gasaway method to minimize potential for confounding effects of multiple survey methods in the analyses. linear regression analyses of winter survey data (n = 41) were used to examine the relationships between population composition metrics (calf:cow and bull:cow ratios) and population density (moose/km²) relative to λ. we used multiple regression with calf:cow and bull:cow ratios as covariates to examine the relative effects of these variables on moose density in the same model. annual harvest data by licensed or rights-based hunters were not available from either province for our analyses that were performed in excel (microsoft corporation, redmond, washington, usa) or r (r foundation for statistical computing, vienna, austria). results population trend the time series modelling revealed a common trend of general decline in all 14 management units, although the rates and temporal patterns varied by unit (fig. 2). cumulative abundance estimates obtained by combining the model results for all 14 management units indicated a steady decline since the early 1990s in the absolute annual winter population (fig. 3) that is currently estimated as ~30% less than the 31-year average. no cumulative estimate was above this average since the winter of 2010–2011 (fig. 4). it is important to note that although surveys are designed to achieve population estimates with a precision level of ±10–20% (i.e., within the manitoba 95% and saskatchewan 90% ci), a >10–20% difference is required to detect a significant change in abundance between surveys (gasaway and dubois 1987, lenarz et al. 2010). moose population demography – arsenault et al. alces vol. 55, 2019 48 moose demography no significant relationship was detected between population density and winter calf recruitment that varied between 0.3 and 0.6 calves/cow (fig. 5a), indicating that winter calf recruitment was not density dependent. a significant negative relationship was found between the adult sex ratio fig. 2. annual estimates of population abundance for 14 moose populations in the boreal plain ecozone in saskatchewan and manitoba. eight populations in saskatchewan (top 3 rows – meadow to porcupine; 90% ci) and 6 populations in manitoba (bottom 2 rows – duck mountain to red deer bog; 95% ci) are illustrated relative to the average abundance estimate across 1985–2015. alces vol. 55, 2019 moose population demography – arsenault et al. 49 fig. 3. the annual change in moose population abundance in 14 populations (pooled) in the boreal plain ecozone in saskatchewan and manitoba, 1985–2015. fig. 4. the annual population change (%) in 14 moose populations in the boreal plain ecozone in saskatchewan and manitoba relative to their respective 31-year (1985 to 2015) average abundance. moose population demography – arsenault et al. alces vol. 55, 2019 50 fig. 5. the relationship of population structure, population density, and λ for moose populations residing in the boreal plain ecozone in saskatchewan and manitoba. solid symbols = gasaway surveys where λ ≥ 1.0 and the population estimate was above the long-term average (1985–2015). open symbols = gasaway surveys where λ ≤ 1.0 and the population estimate was below the longterm average (1985–2015). (a) winter density versus winter recruitment (calves/cow); (b) winter density versus winter sex ratio (bulls/cow); and (c) winter recruitment (calves/cow) versus winter sex ratio (bulls/cow). alces vol. 55, 2019 moose population demography – arsenault et al. 51 (bull:cow) and population density (fig. 5b). for a given bull:cow ratio, population density was higher when λ was increasing than when declining (fig. 5b). a significant positive linear relationship was detected between the calf:cow and bull:cow ratios (fig. 5c), suggesting that calf production and recruitment increase with an increasing adult sex ratio. when included as covariates in the same model (r² = 0.478, p < 0.001), winter calf recruitment (β = −0.018) had no effect on moose density relative to the adult sex ratio (β = −0.685). discussion our modelling indicates a broad scale, 30% average decline since 2000 in the study populations across the boreal plain ecozone. low moose density can influence population demographics (e.g., bimodal parturition reducing calf survival, fitness, population growth) and cause genetic effects (e.g., reduced heterozygosity, bottlenecks, founder effects) that influence longterm population viability (broders 1998, gaillard et al. 2000, eberhardt 2002). further, persistence of isolated and small, low-density populations is particularly susceptible to demographic stochasticity (skalski et al. 2005, broms et al. 2010) that increases the probability of extinction from the amplified effects of random annual fluctuations in vital rates of small populations (snaith and beazley 2002). environmental stochasticity occurs less frequently and may cause decline in populations of any size (lande et al. 2003). valuable reference information is provided from consistent and repetitive population surveys that estimate abundance and sex and age structure of moose populations. comparisons with longterm averages and historical variances are useful to evaluate current conditions, assess potential of and threats to a population, and determine population performance relative to management decisions or habitat change (haufler et al. 2002). a multitude of factors and pathways potentially affect long-term viability, demographic trends, and range occupancy of moose. the functional pathways of drivers of population change occur at spatial and temporal scales that affect habitat suitability (karns 1998, van beest and milner 2013, monteith et al. 2015), habitat selection (schwab and pitt 1991, fahrig and rytwinski 2009, herfindal et al. 2009, bjorneraas et al. 2012, van beest et al. 2012), population demography (murray et al. 2006, 2012, brown 2011), abundance (van ballenberghe 1983, timmermann 1992, sylvén 2003), mortality risk (hebblewhite 2008, laurian et al. 2008, wasser et al. 2011), behaviour patterns (dussault et al. 2004, bjorneraas et al. 2011, broders et al. 2012, street et al. 2015), fitness (renecker and hudson 1990, crichton 1992, wilton 1992, lowe et al. 2010, mccann et al. 2013), predator-prey dynamics (stewart et al. 1985, messier 1994, rayl et al. 2015), pathogen burdens (murray et al. 2006, lenarz et al. 2009, doak and morris 2010), and population viability (popescu et al. 2016). although there is a paucity of empirical data to quantify the relative effects of these drivers of population change (e.g., comprehensive hunter harvest statistics) in the boreal plain ecozone, there are common probable causes of moose population decline (table 1). we consider hunting mortality and to a lesser extent predation to be proximate (immediate) drivers of decline in these populations. we consider habitat alteration from anthropogenic effects of linear and polygonal disturbance, wildfire suppression, sensory disturbance, and climate change effects (e.g., shorter winters, increased temperature, extreme weather events, or drought-altered wetland phenology) to be ultimate (critical/ definitive) drivers of population change in the boreal plain ecozone. moose population demography – arsenault et al. alces vol. 55, 2019 52 predation and harvest mortality (e.g., sex-selective harvest) can have significant effects on adult sex ratios and calf survival which influence demographic parameters including production, recruitment, abundance, λ, and overall population performance (haufler et al. 2002, skalski et al. 2005). because hunting (licensed and rightsbased) is likely additive to natural mortality, its effect on population state and structure is potentially exacerbated by a harvest strategy not linked to population performance measures which should inform biologically sustainable, license allocation strategies. this management shortcoming has likely contributed to the long-term declines observed in the study populations (fig. 2–4). various forms of sex and age selective harvest strategy were implemented in each of the management units during recent decades despite the unknown stochastic effects of rights-based harvesting, predation rate, disease/parasite outbreaks, or large-scale disturbance events. these strategies table 1. hypothesized drivers of population change by moose management unit (mmu) in the boreal plain ecozone of saskatchewan and western manitoba, 1985–2015. winter population abundance and density (in brackets) are modelled estimates projected from the available survey data. mmu winter population (moose/km²) hypothesized driver(s)4 31 yr ave. (1985–2015) current (2015) meadow1 2554 (0.208) 2407 (0.201) cc, hc (fire suppression) bronson1 1719 (0.424) 1174 (0.241) cc, uhh, hc (increased access, oil and gas disturbance, forest cattle grazing, fire suppression) divide1 4627 (0.462) 3238 (0.282) cc, uhh, hc (increased access, fire suppression) sled-panp1 2142 (0.142) 1385 (0.091) cc, uhh, hc (increased access, fire suppression) candle-cub1 2834 (0.273) 1811 (0.157) cc, uhh, hc (increased access, fire suppression) cumberland delta1 3678 (0.401) 2553 (0.223) cc, uhh, hc (hydroelectric development altering delta ecology and allowing increased human and predator access, and vegetation succession) pasquia1 4555 (0.603) 3500 (0.411) cc, uhh, hc (increased access, anthropogenic disturbance, fire suppression) porcupine2 6160 (0.658) 4705 (0.496) cc, uhh, hc (increased access, anthropogenic disturbance, fire suppression) duck mtn.2 2643 (0.452) 1491 (0.248) cc, uhh, hc (anthropogenic disturbance, fire suppression) swan-pelican3 1515 (0.280) 152 (0.030) cc, uhh, hc (increased access) riding mtn.3, 4 3105 (1.009) 3,054 (0.995) cc, hc (fire suppression) the pas3 340 (0.196) 234 (0.130) cc, uhh, hc (fire suppression) tom lamb3 585 (0.206) 178 (0.057) cc, uhh, hc (fire suppression) red deer bog3 490 (0.107) 195 (0.041) cc, uhh 1 = saskatchewan, 2 = inter-provincial population, 3 = manitoba, 4 = largely not hunted (hc = habitat change, cc = climate change, uhh = unsustainable hunter harvest). alces vol. 55, 2019 moose population demography – arsenault et al. 53 generally allocated more licenses to harvest adult bulls, particularly since the mid-2000s with the removal of calves from non-draw hunting seasons, and presumably created a skewed adult sex ratio in favor of females. licensed calf harvest was restricted to drawonly seasons of limited allocation since the mid-2000s, and for cows in the entire study period. we found a significant density dependent relationship between the adult sex ratio and population trend (fig. 5b). most of the populations with density above the regression line were increasing (λ > 1.0) and those below decreasing (λ < 1.0), suggesting that population growth was impaired when the adult sex ratio fell below a density dependent threshold. eberhardt (2002) also observed a sequence of changes in vital rates and demographic measures in relation to population abundance and trend. the relative rate of increase in a moose population is greater when the population is skewed towards females, and when most adults are in the “prime” age classes, whereby fecundity and survival are maximized (van ballenberghe 1983). in a moose population with a given sex ratio, harvest allocations to maintain density above the regression line (fig. 5b) while maintaining a stable bull:cow ratio are likely to result in stable or increasing populations. population density varies across moose range (0.1 to 1.1 moose/km²) but is generally <0.5 moose/km² in the boreal forest of north america (crête 1987, messier 1994, timmermann and buss 1998, arsenault 2000, maier et al. 2005). populations that are widely distributed at low winter density require a higher bull:cow ratio to ensure adequate reproduction (schwartz 1998). timmermann (1992) recommended an adult bull:cow ratio > 0.5 at a density of 0.30 moose/km², and messier (1996) a bull:cow ratio of 0.4 to 0.5 to maximize sustainable harvest of 0.025 moose/km² (under a selective harvest regime) at 0.28 to 0.35 moose/km²; our assessment of the boreal plain ecozone populations support these recommendations. for example, fig. 5b illustrates adult sex ratios across a range of population density relative to λ that could be used to set population objectives and evaluate performance within a management-by-objective framework based on management units. unbalanced adult sex ratios can result in several negative consequences depending on the degree of the imbalance and density of the moose population. a protracted breeding season resulting from a skewed sex ratio favoring females can shift neonate sex ratios in favor of males which can reduce population growth rate (ballard et al. 1991, boer 1992). low bull:cow ratios can impair breeding effectiveness (crête et al. 1981, schwartz 1998, laurian et al. 2000) because of the inability of bulls to locate and breed estrous cows in low-density populations (page 1983, sæther et al. 2003) and protract contraception over 2–3 estrous cycles that cause reduced twinning (aiken and childs 1993) and late-born calves more susceptible to winter mortality (bubenik and timmermann 1982, sæther et al. 2003). lower body mass of males can be associated with a low proportion of adult males in the population (solberg and sæther 1994), and can influence reproductive success in polygamous cervids (sæther et al. 2003). collectively, the relationships depicted in fig. 5 suggest a negative feedback system; specifically, that stable to growing populations at higher winter densities are characterized by a lower adult sex ratio and proportionately lower calf recruitment rate than depressed populations at lower density with declining λ. from a population performance perspective, this suggests that abundant populations approaching their upper limit in size and density (i.e., carrying moose population demography – arsenault et al. alces vol. 55, 2019 54 capacity) have proportionately lower calf production and winter calf recruitment rates than populations below their mean abundance state. it is important to note that we assessed calf recruitment with mid winter calf:cow ratios which likely overestimate true recruitment because predation of ungulate calves occurs continuously in natural systems (musante et al. 2010, environment canada 2012, hurley 2016, jones et al. 2017). recruitment is the most variable of demographic metrics for ungulate populations, but it reflects fecundity and survival of offspring and strongly influences inter-annual variation in population growth (gaillard et al. 2000, monteith et al. 2015). to effectively manage a species, its population dynamics must be thoroughly understood. therefore, consistent methods to survey moose population demographics and long-term data sets are essential to understand population performance. management-by-objective provides a foundation of actionable science that drives management decisions and informs ongoing survey needs from which management prescriptions are evaluated and adjusted based on population performance through an adaptive management process (artelle et al. 2018). it is essential to link population objectives and performance within an appropriate scale of management and to implement population monitoring programs that provide information with direct relevance and use to evaluate population status and management prescriptions. monitoring populations with consistent methods over the long-term will provide insights about population performance metrics used to develop population objectives and guide management under a management-by-objective framework (arsenault 2000, lyons et al. 2008). in the absence of reliable harvest data at the management unit scale, this framework requires relevant data about population demography and performance over time to monitor the consequences of management actions imposed on a population and/or its habitat, and to monitor outcomes relative to the desired objective (arsenault 2000, lyons et al. 2008, sauer et al. 2013). assessing population performance relative to numerical objectives is a crucial step towards establishing population performance metrics for informed and structured decision-making within the managementby-objective framework (strickland 1985, lyons et al. 2008, sauer and knutson 2008). high male harvest under a selective harvest strategy that results in considerable over-kill of males is neither optimal nor viable over the long-term for effective breeding of receptive females (messier 1996, sylvén 2003). saskatchewan came to this conclusion and subsequently altered its harvest strategy to include population performance measures within a management-byobjective framework (arsenault 2000), but failed to fully implement the framework by linking sustainable harvest to long-term demographic objectives for each management unit. manitoba has not established moose population objectives or management planning at any scale, potentially subjecting the population to chronic and unsustainable hunting mortality, and contributing to population decline and implementation of hunting moratoriums in some ghas to induce population recovery. in addition, there is no coordinated moose management between manitoba and saskatchewan for shared populations or biologically sustainable population or harvest objectives. our study suggests moose populations in both provinces would benefit from the development of these objectives and full implementation of a managementby-objective framework. alces vol. 55, 2019 moose population demography – arsenault et al. 55 acknowledgements this manuscript significantly benefitted from comments and contributions provided by saskatchewan ministry of environment (e. h. kowal (retired) and r. tether), manitoba conservation and water stewardship (v. f. j. crichton (retired), k. rebizant, h. hristenko (retired), f. m. van beast (aarhus university), m. purcell (trent university), d. sleep (national council for air and stream improvement (ncasi), inc.), r. k. brook (university of saskatchewan), and j. wiens (manitoba hydro). we appreciate the efforts of associate editor steve windels and 2 anonymous reviewers in providing suggestions that greatly improved our paper. s. frey (parks canada) provided annual moose population survey data for riding mountain national park. k. m. brookes (wood canada limited) provided gis services. references aitken, d. a., and k. n. childs. 1993. relationships between in utero productivity of moose and population sex ratios: an exploratory analysis. alces 28: 175–187. arsenault, a. a. 2000. status and management of moose (alces alces) in saskatchewan. fish and wildlife branch technical report 2000–01. saskatchewan environment and resource management, saskatoon, saskatchewan, canada. artelle, k. a., j. d. reynolds, a. treves, j. c. walsh, p. c. paquet, and c. t. darimont. 2018. hallmarks of science missing from north american wildlife management. science advances 4: eaao0167. doi: 10.1126/sciadv.aao0167 ballard, w. b., j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildlife monographs 114: 1–59. bjorneraas, k., i. herfindal, e. j. solberg, b.-e. saether, b. van moorter, and c. m. rolandsen. 2012. habitat quality influences population distribution, individual space use and functional responses in habitat selection by a large herbivore. oecologia 168: 231–243. doi: 10.1007/s00442-011-2072-3 _____, e. j. solberg, i. herfindal, b. c. van moorter, m. rolandsen, j.-p. tremblay, c. skarpe, b.-e. saether, r. eriksen, and r. astrup. 2011. moose alces alces habitat use at multiple temporal scales in a human-altered landscape. wildlife biology 17: 44–54. doi: 10.2981/10-073 boer, a. h. 1992. fecundity of north american moose (alces alces): a review. alces (supplement 1): 1–10. broders, h. g. 1998. population genetic structure and the effect of founder events on the genetic variability of moose (alces alces) in canada. m.sc. thesis, memorial university of newfoundland. 83 pp. _____, a. b. coombs, and j. r. mccarron. 2012. ecothermic responses of moose (alces alces) to thermoregulatory stress on mainland nova scotia. alces 48: 53–61. broms, k., j. r. skalski, j. j. millspaugh, c. a. hagen, and j. h. schultz. 2010. using statistical population reconstruction to estimate demographic trends in small game populations. journal of wildlife management 74: 310–317. doi: 10.2193/2008-469 brown, g. s. 2011. patterns and causes of demographic variation in a harvested moose population: evidence for the effects of climate and density-dependent drivers. journal of animal ecology 80: 1288–1298. doi: 10.1111/j.1365-2656. 2011.01875.x bubenik, a. b., and h. r. timmermann. 1982. spermatogenesis in the taiga moose of north central ontario – a pilot study. alces 18: 54–93. crête, m. 1987. the impact of sport hunting on north american moose. swedish moose population demography – arsenault et al. alces vol. 55, 2019 56 wildlife research supplement 1: 553–563. _____, r. j. taylor, and p. a. jordan 1981. optimization of moose harvest in southwestern quebec. journal of wildlife management 45: 598–611. crichton, v. f. j. 1992. management of moose populations: which parameters are used? alces (supplement 1): 11–15. darimont, c. t., p. c. paquet, t. e., reimchen, and v. crichton. 2005. range expansion by moose into coastal temperate rainforests of british columbia. diversity and distributions 11: 235–239. doi: 10.1111/j.1366-9516.2005.00135.x delgiudice, d. g. 2013. 2013 aerial moose survey final results. http://files.dnr. state.mn.us/recreation/hunting/moose/ moose_survey_2013.pdf (accessed june 2017). doak, d. f., and w. f. morris. 2010. demographic compensation and tipping points in climate-induced range shifts. nature 467: 959–962. doi: 10.1038/ nature09439 dussault, c., j.-p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioral responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. doi: 10.1080/ 11956860.2004.11682839 eberhardt, l. l. 2002. a paradigm for population analysis of long-lived vertebrates. ecology 83: 2841–2854. doi: 10.1890/0012-9658(2002)083[2841: apfpao]2.0.co;2 ecological stratification working group. 1995. a national ecological framework for canada. agriculture and agri-food canada. research branch, center for land and biological resources research and environment canada, state of the environment directorate, ecozone analysis branch, ottawa/hull. report and national map at 1:7,500,000 scale. http://sis.agr.gc.ca/cansis/publications/ ecostrat/cad_report.pdf (accessed june 2017). environment canada. 2012. recovery strategy for the woodland caribou (rangifer tarandus caribou), boreal population, in canada. species at risk act recovery strategy series. environment canada, ottawa, ontario, canada. fahrig, l., and t. rytwinski. 2009. effects of roads on animal abundance: an empirical review and synthesis. ecology and society 14(1): 21. doi: 10.5751/es-02815 140121 faison, e. k., g. motzkin, d. r. foster, and j. e. mcdonald. 2010. moose foraging in the temperate forests of southern new england. northeastern naturalist 17: 1–18. doi: 10.1656/045.017.0101 foster, d. r., g. motzkin, d. bernardos, and j. cardoza. 2002. wildlife dynamics in the changing new england landscape. journal of biogeography 29: 1337– 1357. doi: 10.1046/j.1365-2699.2002. 00759.x funk, w. c., j. k. mckay, p. a. hohenlohe, and f. w. allendorf. 2012. harnessing genomics for delineating conservation units. trends in ecology and evolution 27: 489–496. doi: 10.1016/j.tree.2012. 05.012 gaillard, j. m., m. festa-bianchet, n. g. yoccoz, a. loison, and c. toigo. 2000. temporal variation in fitness components and population dynamics of large herbivores. annual review of ecology and systematics 31: 367–393. doi: 10.1146/annurev.ecolsys.31.1.367 gasaway, w. c., and s. d. dubois. 1987. estimating moose population parameters. swedish wildlife research (supplement) 1: 603–617. _____, _____, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska number 22. institute of arctic biology, fairbanks, alaska, usa. government of manitoba. 2014. wildlife: five-year report (reporting period 2007–2012). https://www.gov.mb.ca/sd/ http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2013.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2013.pdf http://files.dnr.state.mn.us/recreation/hunting/moose/moose_survey_2013.pdf http://sis.agr.gc.ca/cansis/publications/ecostrat/cad_report.pdf http://sis.agr.gc.ca/cansis/publications/ecostrat/cad_report.pdf https://www.gov.mb.ca/sd/wildlife/pdf/fiveyear_report2007to​2012.pdf alces vol. 55, 2019 moose population demography – arsenault et al. 57 wildlife/pdf/fiveyear_report2007to 2012.pdf (accessed june 2017). haufler, j. b., r. k. baydack, h. campa iii, b. j. kernohan, c. millar, l. j. o’neil, and l. waits. 2002. performance measures for ecosystem management and ecological sustainability. wildlife society technical review 02-1. the wildlife society, bethesda, maryland, usa. hebblewhite, m. 2008. a literature review of the effects of energy development on ungulates: implications for central and eastern montana. report prepared for montana fish, wildlife and parks, miles city, montana, usa. herfindal, i., j.-p. tremblay, b. b. hansen, e. j. solberg, m. heim, and b.-e. sæther. 2009. scale dependency and functional response in moose habitat selection. ecography 32: 849–859. doi: 10.1111/j.1600-0587.2009.05783.x hurley, m. a. 2016. mule deer population dynamics in space and time: ecological modelling tools for managing ungulates. ph. d. dissertation. university of montana, missoula, montana, usa. jones, h., p. j. pekins, l. e. kantar, m. o’neil, and d. ellingwood. 2017. fecundity and summer calf survival of moose during 3 successive years of winter tick epizootics. alces 53: 85–98. karns, p. d. 1998. population distribution, density and trends. pages 125–139 in a. w. franzmann, and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, dc, usa. knudsen, b. 2007. manitoba aerial survey data management manual, version 1.1. manitoba wildlife and ecosystem protection branch, winnipeg, manitoba, canada. kuzyk, g. w. 2016. provincial population and harvest estimates of moose in british columbia. alces 52: 1–11. laforge, m. p., n. l. michel, a. l. wheeler, and r. k. brook. 2016. habitat selection by female moose in the canadian prairie ecozone. journal of wildlife management 80: 1–10. doi: 10.1002/ jwmg.21095 laliberte, a. s., and w. j. ripple. 2004. range contractions of north american carnivores and ungulates. bioscience 54: 123–138. doi: 10.1641/00063 5 6 8 ( 2 0 0 4 ) 0 5 4 [ 0 1 2 3 : r c o n a c ] 2.0.co;2 lande, r., s. engen, and b. e. sæther. 2003. stochastic population dynamics in ecology and conservation. oxford university press, new york, new york, usa. laurian, c., j. p. ouellet, r. courtois, l. breton, and s. st-onge. 2000. effects on intensive harvesting on moose reproduction. journal of applied ecology 37: 515–531. doi: 10.1046/j.1365-2664. 2000.00520.x _____, c. dussault, j.-p. ouellet, r. courtois, m. poulin, and l. breton. 2008. behaviour of moose relative to a road network. journal of wildlife management 72: 1550–1557. doi: 10.2193/2008-063 lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013–1023. doi: 10.2193/2009-493 _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–510. doi: 10.2193/ 2008-265 lowe, s. j., b. r. patterson, and j. a. schaefer. 2010. lack of behavioral responses of moose (alces alces) to high ambient temperatures near the southern periphery of their range. canadian journal of zoology 88: 1032–1041. doi: 10.1139/z10-071 lyons, j. e., m. c. runge, h. p. laskowski, and w. l. kendall. 2008. monitoring https://www.gov.mb.ca/sd/wildlife/pdf/fiveyear_report2007to​2012.pdf https://www.gov.mb.ca/sd/wildlife/pdf/fiveyear_report2007to​2012.pdf moose population demography – arsenault et al. alces vol. 55, 2019 58 in the context of structured decision-making and adaptive management. journal of wildlife management 72: 1683–1692. doi: 10.2193/2008-141 lynch, g. m., and g. e. schumaker. 1995. gps and gis assisted moose surveys. alces 31: 145–151. maier, j. a. k., j. ver hoef, a. d. mcguire, r. t. bowyer, l. saperstien, and h. a. maier. 2005. distribution and density of moose in relation to landscape characteristics: effects of scale. canadian journal of forest research 35: 2233–2243. doi: 10.1139/ x05-123 marshall, i. b., p. h. schut, and m. ballard. 1999. a national ecological framework for canada: attribute data. agriculture and agri-food canada, research branch, centre for land and biological resources research, and environment canada, state of the environment directorate, ecozone analysis branch. ottawa/hull, ontario, canada. http://sis.agr.gc.ca/cansis/nsdb/ e c o s t r a t / 1 9 9 9 r e p o r t / i n d e x . h t m l (accessed june 2017). mccann, n. p., r. a. moen, and t. r. harris. 2013. warm season heat stress in moose (alces alces). canadian journal of zoology 91: 893–898. doi: 10.1139/ cjz-2013-0175 messier, f. 1994. ungulate population models with predation: a case study with the north american moose. ecology 75: 478–488. doi: 10.2307/1939551 _____. 1996. moose co-management in the trilateral agreement territory: principles and recommendations based on scientific knowledge and aboriginal rights. report to the algonquins of berriere lake trilateral secretariat. department of biology, university of saskatchewan, saskatoon, saskatchewan, canada. monteith, k. l., r. w. klaver, k. r. hersey, a. a. holland, t. p. thomas, and m. j. kauffman. 2015. effects of climate and plant phenology on recruitment of moose at the southern extent of their range. oecologia 178: 1137–1148. doi: 10.1007/s00442-015 3296-4 murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. doi: 10.2193/0084-0173(2006)166 [1:pndaci]2.0.co;2 _____, k. f. hussey, l. a. finnigan, s. j. lowe, g. n. price, j. benson, k. m. loveless, k. r. middel, k. mills, d. potter, a. silver, m.-j. fortin, b. r. patterson, and p. j. wilson. 2012. assessment of the status and viability of a population of moose (alces alces) at its southern range limit in ontario. canadian journal of zoology 90: 422– 434. doi: 10.1139/z2012-002 musante, a. r., p. j. pekins, and d. l. scarpitti. 2010. characteristics and dynamics of a regional moose (alces alces) population in the northwestern united states. wildlife biology 16: 185–204. doi: 10.2981/09-014 padbury, g. a., d. f. acton, and c. t. stuchnoff. 1998. ecoregions of saskatchewan. saskatchewan environment and resource management and canadian plains research center/ university of regina, regina, saskatchewan, canada. page, r. e. 1983. managing moose with the knowledge of population dynamics theory. alces 19: 83–97. popescu, v. d., k. a. artelle, m. i. pop, s. manolache, and l. rozylowicz. 2016. assessing biological realism of wildlife population estimates in datapoor systems. journal of applied ecology. 53: 1248–1259. doi: 10.1111/1365 2664.12660 raedeke, k. j., j. j. millpaugh, and p. e. clark. 2002. population characteristics. pages 449–513 in d. e. toweill, http://sis.agr.gc.ca/cansis/nsdb/ecostrat/1999report/index.html http://sis.agr.gc.ca/cansis/nsdb/ecostrat/1999report/index.html alces vol. 55, 2019 moose population demography – arsenault et al. 59 and j. w. thomas, editors. north american elk: ecology and management. smithsonian institute press, washington, dc, usa. raithel, j. d., m. j. kauffman, and d. h. pletscher. 2007. impact of spatial and temporal variation in calf survival on the growth of elk populations. journal of wildlife management 71: 795–803. doi: 10.2193/2005-608 rayl, n. d., t. k. fuller, j. f. organ, j. e. mcdonald jr., r. d. otto, g. bastill-rousseau, c. e. soulliere, and s. p. mahoney. 2015. spatiotemporal variation in the distribution of potential predators of a resource pulse: black bears and caribou calves in newfoundland. journal of wildlife management 79: 1041–1050. doi: 10.1002/jwmg.936 renecker, l. a., and r. j. hudson. 1990. behavioral and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspen-dominated forests. alces 26: 66–72. sæther, b. -e., e. j. solberg, and m. heim. 2003. effects of altering sex ratio structure on the demography of an isolated moose population. journal of wildlife management 67: 455–466. sauer, j. r., p. j. blank, e. f. zipkin, j. e. fallon, and f. w. fallon. 2013. using multi-species occupancy models in structured decision making on managed lands. journal of wildlife management 77: 117–127. doi: 10.1002/ jwmg.442 _____, and m. g. knutson. 2008. objectives and metrics for wildlife monitoring. journal of wildlife management 72: 1663–1664. doi: 10.2193/2008-278 schwab, f. e., and m. d. pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071–3077. doi: 10.1139/ z91-431 schwartz, c. c. 1998. reproduction, natality and growth. pages 141–172 in a. w. franzman, and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, dc, usa. skalski, j. r., k. e. ryding, and j. j. millpaugh. 2005. wildlife demography: analysis of sex, age and count data. elsevier academic press, amsterdam, netherlands. smith, r. e., h. veldhuis, g. f. mills, r. g. eilers, w. r. fraser, and g. w. lelyk. 1998. terrestrial ecozones, ecoregions, and ecodistricts. an ecological stratification of manitoba’s landscapes. technical bulletin 98-9e. land resource unit, brandon research center, research branch, agriculture and agri-food canada, winnipeg, manitoba, canada. snaith, t. v., and k. f. beazley. 2002. application of population viability theory to moose in mainland nova scotia. alces 38: 193–204. solberg, e. j., and b. e. sæther. 1994. male traits as life history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammalogy 75: 1069–1079. doi: 10.2307/1382491 stewart, r. r. 1983. the stratified random block aerial survey technique. wildlife population management information base 83-wpm-12. saskatchewan parks and renewable resources, wildlife branch, regina, saskatachewan, canada. _____, e. h. kowal, r. beaulieau, and t. w. rock. 1985. the impact of black bear removal on moose calf survival in east-central saskatchewan. alces 21: 403–418. street, g. m., a. r. rodgers, and j. m. fryxell. 2015. mid-day temperature variation influences seasonal habitat selection by moose. journal of wildlife management 79: 505–512. moose population demography – arsenault et al. alces vol. 55, 2019 60 strickland, d. 1985. managing wildlife resources by objective. transactions of the north american wildlife and natural resources conference 50: 296–301. sylvén, s. 2003. management and regulated harvest of moose (alces alces) in sweden. p.h. d. dissertation. department of conservation biology, swedish university of agricultural science, uppsala, sweden. taber, r. d., and k. j. raedeke. 1979. population dynamics. pages 98–106 in r. d. teague, and e. decker, editors. wildlife conservation principles and practices. the wildlife society, bethesda, maryland, usa. tarleton, p. 1992. cervid monitoring and status report, riding mountain national park, 1990–1991. unpublished canadian parks service report. onanole, manitoba, canada. thiele, d. 2007. chapter 2: mule deer (odecoileus hemionus). pages 2–1 to 2–27 in s. a. tessmann, and j. bohne, editors. handbook of biological techniques: 3rd edition. wyoming game and fish department, cheyenne, wyoming, usa. timmermann, h. r. 1992. moose sociobiology and implications for harvest. alces 28: 59–77. _____, and m. e. buss. 1998. population and harvest management. pages 559–615 in a. w. franzman, and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institute press, washington, dc, usa. _____, and a. r. rodgers. 2017. the status and management of moose in north america – circa 2015. alces 53: 1–22. van ballenberghe, v. 1983. the rate of increase in moose populations. alces 19: 98–117. van beest, f. m., b. van moorter, and j. m. milner. 2012. temperaturemediated habitat use and selection by a heat-sensitive northern ungulate. animal behavior 84: 723–735. doi: 10.1016/j. anbehav.2012.06.032 ________, and j. m. milner. 2013. behavioural responses to thermal conditions affect seasonal mass change in a heat-sensitive northern ungulate. plos one 8(6): e65972. doi: 10.1371/journal. pone.00065972 wasser, s. k., j. l. keim, m. l. taper, and s. r. lele. 2011. the influences of wolf predation, habitat loss, and human activity on caribou and moose in the alberta oil sands. frontiers in ecology and environment 9: 546–551. doi: 10.1890/100071 white, g. c. 2000. modeling population dynamics. pages 84–107 in s. demarais, and p. r. krausman, editors. ecology and management of large mammals in north america. prentice-hall, upper saddle river, nj. https://pdfs.semanticscholar. org/265c/e706871c297be46b88d7db118081aeb1b7db.pdf (accessed june 2017). wilton, m. l. 1992. implications of harvesting moose during pre–rut and rut activity. alces 28: 31–34. https://pdfs.semanticscholar.org/265c/e706871c297be46b88d7db118081aeb1b7db.pdf https://pdfs.semanticscholar.org/265c/e706871c297be46b88d7db118081aeb1b7db.pdf https://pdfs.semanticscholar.org/265c/e706871c297be46b88d7db118081aeb1b7db.pdf alces 31_45.pdf integrating thermal constraints into habitat suitability for moose in the adirondack state park, new york catherine g. haase1 and h. brian underwood2 1state university of new york college of environmental science and forestry, department of environmental and forest biology, 212 illick hall, 1 forestry drive, syracuse, new york 13210; 2usgs patuxent wildlife research center, state university of new york college of environmental science and forestry, department of environmental and forest biology, 426 illick hall, 1 forestry drive, syracuse, new york 13210, usa. abstract: moose (alces alces) survive cold winter temperatures due to their large body size, thick skin, and dense, dark pelage. these same characteristics impede heat dissipation under thermal conditions often encountered in spring-fall. while thermal cover has long been recognized as an important component of moose habitat suitability, it has not been explicitly incorporated into published models. we integrated the biophysical construct of operative temperature, te, into an existing habitat suitability index (hsi) model for moose in the adirondack state park (asp) of new york. te is a thermal index that incorporates the effects of radiative and convective heat transfer on air temperature. we modeled air temperature with respect to elevation and calculated solar radiation transmitted through the canopy as a function of topography, location, forest cover-type, and time of year. we classified 1028, 25 km2 evaluation units for thermal suitability based on a modified upper critical threshold for te derived from published studies. compared to a published model for asp, our hsi better classified moose observations in low, moderate, and high suitability categories, especially during april. we discuss the complexities of modeling thermal suitability for moose. alces vol. 49: 49–64 (2013) key words: adirondacks, alces alces, habitat suitability, heat stress, moose, operative temperature, thermal cover. wildlife biologists and forest managers focus on the relationship between an animal species and its preferred habitat for developing appropriate management practices (guisan and zimmerman 2000). this relationship is important to evaluate an area relative to the animal's survival and reproduction (puttock et al. 1995). there are various tools for habitat analyses, including resource selection functions, occupancy models, and habitat suitability index (hsi) models. hsi models are graphical constructs that quantify habitat quality in response to food and cover requirements of a species (koitzsch 2002). these models rate suitability on a scale of 0.0 (unsuitable habitat) to 1.0 (optimal habitat) with different compartments or areas scored in relation to life requisites (romito et al. 1999), and allow for comparisons among managed areas as well as focusing on increasing suitability scores by managing for resources that are limiting (koitzsch 2002, dussault et al. 2006). it is assumed that a species will be present and more abundant in areas with higher suitability; wildlife managers can therefore validate hsi models by measuring habitat characteristics and correlating the calculated scores with population data (dettki et al. 2003, dussault et al. 2006). hsi models have been used in moose (alces alces) management since the u.s. fish and wildlife service (usfws) first published procedures in the early 1980s when allen et al. 49 (1987) described the first 2 usfws models: model i required detailed vegetation measurements and habitat assessment, whereas model ii focused on remotely-sensed data in its classification of suitable habitat. remote sensing has allowed both models to be applied to large tracts of land without the time and spatial constraints of field-based methods (koitzsch 2002, hickey 2008). although originally created for the lake superior region, model ii was used with remotely-sensed data in geographic information system (gis) to assess habitat suitability relative to regenerating forests and non-forested wetlands in vermont (koitzsch (2002). hickey (2008) used a similar approach and speculated about future growth of the recolonizing moose population in the adirondack state park (asp) in new york. in addition to requiring large tracts of land with diverse vegetation types (dussault et al. 2006), moose require 40–50% of an area to be comprised of suitable habitat with regenerating woody stems essential for latesummer and winter forage, 5–10% in nonforested, macrophyte-rich wetlands necessary for summer forage, and 5–55% comprising dense forest stands critical for thermal cover in late winter and summer months (renecker and hudson 1986, schwab and pitt 1991, puttock et al. 1995, koitzsch 2002). while the gis-based models of koitzsch (2002) and hickey (2008) acknowledged the importance of these criteria, neither measured nor included thermal cover explicitly. renecker and hudson (1986) first noted that although moose are adapted to live in cold environments, they exhibit heat stress at temperatures as low as �5 °c in the late winter and 14 °c in the summer. chronic heat stress may lead to increased susceptibility to parasitism and disease, reduced productivity, and starvation (renecker and hudson 1990, lenarz et al. 2008). moose respond behaviorally to heat stress by seeking cover under dense coniferous forest canopies, by prostrating themselves on cool substrates (e.g., soil or snow), by immersing themselves in water (dussault et al. 2004), and by reducing voluntary food intake (belovsky 1981). the gradual 40-year decline in moose populations in isle royale national park in michigan and in the agassiz national wildlife refuge in northern minnesota are correlated with increasing temperatures associated with climate change, and speculation exists about the role of increased parasitism and heat stress (murray et al. 2006, lenarz et al. 2008). projections of climate-warming scenarios indicate a future asp forest with fewer coniferous trees, warmer and shorter winters with decreased snowpack and longer ice-free periods, and hotter, longer summers (jenkins 2004). we attempted to incorporate an index of thermal suitability into a model ii approach for assessing moose habitat in the asp. we evaluated critical thermal environments (moen 1968, parker and gillingham 1990) for moose by computing operative temperature (te), an index that integrates the combined effects of ambient temperature, total absorbed radiation, and wind velocity on the thermal environment experienced by an animal (bakken 1992). te considers the effects of pelage on heat loss and heat absorption and incorporates seasonal variation in surface albedo due to accumulated snow. air temperature was incremented above ambience to indicate the temperature of a space that would feel the same as the heat load in the sun (campbell and norman 1997). integrating thermal cover as a habitat variable not only allows assessment of current suitability, but also facilitates prediction of future suitability under climate-warming scenarios (koitzsch 2002). by mapping te across the asp, our goal was to stimulate discussion and additional research regarding critical thermal environments as components 50 thermal cover and moose – haase and underwood alces vol. 49, 2013 of habitat suitability for moose. our specific objectives were to 1) develop an hsi model for moose that incorporates thermal suitability in a spatially explicit manner, and 2) compare the hsi performance against a published model (hickey 2008) lacking a specific thermal component. study area the asp is located in the northeastern part of new york state (latitude 44e 00" n, longitude 74e 13" w; fig. 1). it is comprised of >6 million acres of both privately owned land (3.3 million acres) and state protected forest preserve (2.7 million acres; apa 2001) and contains the entire range of the adirondack mountains. the asp is mostly forested with a combination of northern hardwood and softwood stands scattered among numerous lakes and wetlands. the high peaks area is mountainous with >40 peaks from 1,200 to >1,500 m elevation (jenkins 2004). this area contains most of the original, old growth forests in the asp as little harvesting or forest management occurred prior to the establishment of the asp in 1984 (dinunzio 1984). the unique geographical characteristics and location with respect to lake ontario control most of the weather patterns, making it one of the coldest regions in the continental united states. the interior is usually 3 °c cooler than the bordering counties of upstate new york and vermont; monthly winter temperatures range from �12 to �6 °c and summer temperatures range from 20–26 °c. the colder climate results in a short growing season (∼180 days) with a mean of 85.8 mm of monthly precipitation (garner 1989, jenkins 2004). topography and meteorological patterns influence the vegetation; southern species are more common along the periphery with cold-hardy species more plentiful in the boreal center fig. 1. map of new york state, the adirondack state park, and huntington wildlife forest. alces vol. 49, 2013 haase and underwood – thermal cover and moose 51 (jenkins 2004). forestland in the asp consists mostly of northern hardwoods including beech (fagus grandifolia), red maple (acer rubrum), sugar maple (a. saccharum), striped maple (a. pensylvanicum), and yellow birch (betula alleghaniensis), and softwoods such as balsam fir (abies balsamea), eastern hemlock (tsuga canadensis), red spruce (picea rubens), white cedar (thuja occidentalis), and white pine (pinus strobus). shrubby vegetation, important for moose forage, includes witch hobble (viburnum alnifolium), wild raisin (v. cassinoides), and other viburnum species (garner 1989). methods we used the united states geological survey's national land cover datasets (nlcd 2001; resolution 30 m) obtained from the adirondack park agency (apa) to quantify the necessary habitat characteristics of model ii in arcgis™ 9.3 (table 1). all layers were clipped to the asp boundary as it was our central focus for this study. using hawth's tools (beyer 2004) we overlaid a sampling grid on the asp to designate 1028 evaluation units that reflected the approximate annual home range size of a moose, or about 25 km2 (allen et al. 1987, koitzsch 2002, hickey 2008). regenerating habitat (v1) and wetland habitat (v2) were quantified using the methods of hickey (2008) and evaluated against suitability figures developed for model ii approaches by allen et al. (1987). softwood (v3) and old and mixed hardwoods (v4) stands were designated as winter and summer cover, respectively. because forest stand type can range in thermal cover characteristics, we condensed v3 and v4 into a single “thermal” cover variable (v3) based on modeled operative temperature, and evaluated it against 2 suitability figures developed for the previous variables (allen et al. 1987). thermal cover (habitat variable v3) we separated the habitat analysis for thermal cover into 2 months (april and july) when moose are presumably susceptible to heat stress. in april moose are confronted with increasing daytime temperatures and insolation along with high surface albedo from accumulated snow, and in combination with molting winter pelage and increased metabolism, heat gain can be substantial (renecker and hudson 1986, 1990). long days near maximal ambient temperature can also result in heat stress during july (dussault et al. 2004). to incorporate spatial variation associated with elevation, slope, and aspect, we used arcgis™ (esri 2010) to calculate mean monthly te for hardwood and softwood stands for each evaluation unit with the following equation (campbell and norman 1997): te ¼ ta þ re rabs � esrt4a � � qcp ð1þ where ta is the ambient temperature (°c) evaluated at each unit, re is the animal's parallel resistance to convective and radiative heat transfer (s m−1), rabs is the total amount table 1. suitability classes, habitat variables and their optimal specifications (developed by allen et al. 1987) for a new habitat suitability index (hsi) model for moose in the adirondack state park, new york. suitability class habitat variables optimal specification browse habitat v1: hardwoods/mixed < 20 years 40% ≤ area ≤ 50% aquatic habitat v2: wetlands/open water 5% ≤ area ≤ 10% thermal cover v3: operative temperature < ucte 1 5% ≤ area ≤ 55% 1upper critical operative temperature (see text for details). 52 thermal cover and moose – haase and underwood alces vol. 49, 2013 of solar and thermal radiation absorbed by the animal (w m−2), esrt 4 a is the thermal emittance (w m−2) of the surface of the animal (through tissue, skin, and pelage) at ta (k), and ρcp is the volumetric specific heat of air (1200 j m−3 per k; table 2). table 2. inputs to a model of operative temperature (te) for moose in the adirondack state park, new york. constant, parameter or variable source reference air temperature (ta; c) modeled in arcgis ncdc, castnet resistance to heat flow (re; sm −1) rha·rr/rha+rr beaver et al. 1996 resistance to convective transfer (rha) 307 ffiffiffiffiffiffiffiffi d=u p � tf campbell and norman 1997 characteristic dimension (d) 0.70 m3 mitchell 1976, haase 2010 average wind velocity (u) 1 ms−1 parker and gillingham 1990 turbulence factor (tf) 0.7 campbell 1981 resistance to long-wave transfer (rr) qcp=4esrt 3 a parker and gillingham 1990 emissivity of surface (ɛs) 0.97 belovsky 1981 stephan-boltzmann constant (σ) 5.67·10−8 wm−2 per k4 monteith 1973 total radiation absorbed (rabs;/wm −2) sw + lw monteith 1973 short-wave radiation as(ap/a · sp + 0.5sd + 0.5swgr) monteith 1973 absorptivity to radiation (as) 0.74 summer, 0.89 winter belovsky 1981 view factor, or projected shadow area on a surface perpendicular to the beam (ap/a) as a function of θ 0.29−0.01(θ/90)−0.31 (θ/90)2+0.12(θ/90)3 kubaha et al. 2004; haase 2010 direct radiation corrected for the angle of incidence (sp) sb/sinθ parker and gillingham 1990 diffuse short-wave radiation (sd) st * difffrac boland et al. 2001 reflected short-wave radiation from the ground (swgr) albedo · st parker and gillingham 1990 beam radiation incident on a horizontal surface (sb) st * beamfrac boland et al. 2001 total global radiation (st) modeled in arcgis™ esri 2010 solar elevation angle (θ) dependent on time of day monteith 1973 albedo 0.8 for snow, 0.2 for grass monteith and unsworth 1990 long-wave radiation (lw) al (0:5eskyrt 4 a þ 0:5egrrt4a þ monteith 1973 absorptivity to lw (al) 1.0 for caribou monteith 1973 emissivity of sky (ɛsky) 0.67 + 0.007*ta in c gates 1980 emissivity of the ground (ɛgr) 0.97 parker unpublished thermal emittance of the surface of the animal at ta in k esrt 4 a parker and gillingham 1990 volumetric specific heat of air ρcp monteith 1973 alces vol. 49, 2013 haase and underwood – thermal cover and moose 53 ambient temperature (ta) — we obtained air temperature data for april and july 2009 from 12 weather stations from the environmental protection agency's clean air status and trend network (castnet; http://www.epa.gove/castnet) and the national climatic data center (ncdc; http://www.ncdc.noaa.gov) to model ambient temperature across the set of evaluation units. we used ordinary co-kriging of air temperature against a digital elevation model in arcgis™ to create a raster map with air temperature as a function of station elevation and separation distance. geostatistical analyst™ uses a take-one-out, cross-validation scheme for assessing goodness-of-fit of kriged surfaces. we used a three-step diagnostic process (johnston et al. 2001, pp. 190–191) to validate modeled ambient temperatures, and then corrected them to 1 m above ground level using micro-meteorological stations established in the 2 forest canopy types (haase 2010). resistance to total heat flow (re) — we calculated the thermal resistance to heat flow as a combination of the resistance to longwave radiative heat transfer (rr) and convective heat transfer (rha). rr was computed from the volumetric specific heat of air (ρcp), the emissivity of the surface of the pelage (ɛs), the stephan-boltzman constant (σ), and ambient air temperature (ta); rha was calculated from the characteristic dimension (d) of a moose, the average wind velocity (u), and the turbulence factor (tf). all constants, parameters, and variables were initialized from the published literature except for characteristic dimension, wind velocity, and air temperature (table 2). because wind velocities below the forest canopy are usually low (demarchi 1991) and difficult to model, we used a constant wind velocity of 1m s−1 across all evaluation units (parker and gillingham 1990). characteristic dimension was approximated as the volume of a sphere raised to the one-third power (campbell and norman 1998). we used data from cameron et al. (1999: 96) and hundertmark et al. (1997) to estimate mean volume to calculate the ratio of ingesta-free body mass to the density of water, muscle, fat, and bone of a 400 kg female moose (0.345 m3). shortand long-wave radiation (rabs) — thermal radiation emitted from the surface of the animal (esrt 4 a ) was calculated from the stephan-boltzmann constant in relation to air temperature and the emissivity of moose pelage (table 2; monteith 1973, belovsky 1981). total radiation (rabs) was calculated as the sum of longand shortwave radiation (parker and gillingham 1990). long-wave radiation (lw) was calculated from emissivities of the ground and sky (dependent on air temperature) and absorptivities to long-wave heat transfer (table 2; beaver et al. 1996). total short-wave radiation (sw) was calculated from modeled solar radiation in reference to seasonal resistances of the moose pelage to short-wave heat transfer (table 2; belovsky 1981, parker and gillingham 1990, beaver et al. 1996). the amount of solar radiation an animal is exposed to on the ground is a function of the amount of global radiation (direct and diffuse radiation) that is transmitted through the atmosphere, cloud cover, and forest canopy. hourly global radiation was collected from a castnet weather station within the asp on the huntington wildlife forest (hwf187) and extraterrestrial radiation (i.e., above the atmosphere) was calculated based on the earth's distance from the sun in a model developed for the solar energy research institute (bird and hulstrom 1991). we calculated monthly averaged, hourly clearness indices (kt, also referred to as cloudiness index), which is the ratio of global to extraterrestrial radiation for april and july (haase 2010). we computed the mean monthly diffuse and beam fractions (boland et al. 2001) with these 54 thermal cover and moose – haase and underwood alces vol. 49, 2013 http://www.epa.gove/castnet http://www.ncdc.noaa.gov values by extrapolating average solar radiation at each intersection of a 100-m2 grid using the solar analyst™ tool in esri's spatial analyst™. finally, we kriged the point data to create raster maps representing monthly mean global, direct, and diffuse solar radiation at the top of the forest canopy for april and july. we calculated the monthly fraction of solar radiation transmitted through the canopy associated with the 2 forest covertypes as: g go ¼ beamfrac 1 � sofð þ þ difffrac � c�o ð2þ where beamfrac is the fraction of beam (i.e., direct) radiation through the atmosphere, difffrac is the fraction of diffuse radiation through the atmosphere, sof is the sky obscuration factor of the canopy as a function of solar elevation zenith angle, and c � o is the total canopy opening factor for the canopy type (lindroth and perttu 1981). we used hemispherical photography to calculate sof and gap light analyzer (gla) software (version 2.0, frazer et al. 1999) to calculate c � o (hardy et al. 2004). we used a nikon n2000 camera (integralmotor multi-mode 35 mm single-lens reflex) to take true color hemispherical photographs with a sigma 8 mm f4 hemispherical lens (22.5 filter size) in april and july at 3 sites in each forest canopy cover-type (i.e., hardwood and softwood; 6 sites total) on huntington wildlife forest. each photograph was taken with the camera facing skyward, placed on a level tripod 1 m above the ground. the camera was positioned using a compass so the bottom faced south, allowing consistent registration of every photo. in gla each photo was separated into “sky” or “non-sky” pixels using the image classification tools and overlaid with a sky map of grid cells to calculate canopy openness in each grid. average transmission fractions were calculated for each forest covertype and month (lindroth and perttu 1981). the nlcd 2001 raster was reclassified using the calculated transmission fractions for hardwood and softwood forest covertypes for each month. we averaged transmission fractions from both forest cover-types to represent the mixed forest cover-type; all other land-cover classes exhibited 100% radiation transmission (lindroth and perttu 1981). we multiplied the direct, diffuse, and global radiation raster maps by the canopy transmission fraction layer to produce new solar radiation layers with values adjusted for transmission through the forest canopy. classifying thermal cover — schwab and pitt (1991) adjusted the upper critical temperatures (uct) measured by renecker and hudson (1986) onto te values of 0 °c in late winter and 20 °c in summer; likewise, they adjusted changes in respiration rate that resulted in panting for both late winter (ucte = 8 °c) and summer (ucte = 30 °c). because thresholds were derived from moose acclimated to local conditions of alberta, canada, we re-scaled them proportionally to the prevailing difference in mean temperature regimes for the same period in the adirondacks (chaffee and roberts 1971). the result was a respiration threshold increasing by 37.5% (i.e., ucte = 11 °c) for april and no change for july (i.e., ucte = 30 °c; data obtained by weather underground, www. wunderground.com). we computed te across all the evaluation units for both april and july and then re-classified the raster values as “thermally” suitable on the basis of te 0.67 were considered highly suitable (koitzsch 2002, hickey 2008), and those in between were considered moderately suitable. we computed the percentage of evaluation units in each suitability class to compare our classification to a published hsi model for the asp that did not incorporate thermal cover (hickey 2008). in order to test that moose select suitable habitat designated by our model, we obtained moose observations (i.e., both visual and telemetry) from the new york state department of environmental conservation (c. dente, nysdec, pers. comm.) and the adirondack program of the wildlife conservation society (m. glennon, wildlife conservation society, pers. comm.). we sorted the data by month and projected moose observations for april and july onto the appropriate seasonal suitability map. we performed a x2 goodness-of-fit test on moose locations relative to the numbers expected in each suitability class based on area and compared standardized residuals among the 3 models (i.e., april, july, and hickey 2008). no significant difference between expected and observed proportions indicated a correct classification relative to moose habitat use. results mean monthly air temperatures derived from ordinary co-kriging ranged from 4.7– 10.3 °c in april and 16.2–20.5 °c in july; prediction standard errors were relatively small (fig. 2). standardized root mean square errors (april: 0.88, july: 0.80) indicated that the model modestly overestimated variability in predicted temperatures. forest canopy radiation transmission fractions varied by month and forest cover-type (fig. 3), with the greatest decrease between april and july for hardwood stands (66.7 to 19.7% transmitted; p <0.0002). softwood canopy transmission also declined (from 42.8 to 36.4%; p = 0.355), but not as dramatically. average kt values differed between months with april (kt = 0.13) cloudier on average than july (kt = 0.25; p-value = 0.031). below-canopy solar radiation in april ranged from 69.0–212.2 wm−2 and in july from 42.5–280.5 wm−2. te ranged from 5.6–15.3 °c during april (fig. 4a), and from 19.0–29.1 °c during july (fig. 4b). thermal suitability increased from 59.9% of evaluation units below ucte in april, to 85.7% of evaluation units below ucte in july. the lowest te values corresponded closely to the highest elevations. in april approximately 36.3% of evaluation units were characterized as low, 23.8% as moderate, and 39.7% as high suitability (fig. 5a). because average snow depth never exceeded 0.9 m, it was not included in our hsi (schwab and pitt 1991). during july the model classified 8.7% of the asp as low, 16.6% as moderate, and 64.5% as high suitability (fig. 5b). highest suitability in both april and july occurred in a broad, crescent-shaped swathe extending from the north-central asp westward and then south and eastward to the south-central asp. moderate to low suitability predominated in the eastern half of asp. there were 88 observations of moose in april in which 36.4% (n = 32) were located 56 thermal cover and moose – haase and underwood alces vol. 49, 2013 fig. 2. average air temperatures (c) for april and july and respective standard prediction error maps. fig. 3. hemispherical photographs of (a) hardwood (hwd) and (b) softwood (swd) stands in (1) april and (2) july on huntington wildlife forest, new york and percent radiation transmission through respective forest canopy types (error bars represent one standard error). alces vol. 49, 2013 haase and underwood – thermal cover and moose 57 fig. 4. operative temperature (c) maps of the adirondack state park, new york for april (left) and july (right). fig. 5. april (left) and july (right) moose observations (white circles) on a modified habitat suitability map for the adirondack state park, new york. 58 thermal cover and moose – haase and underwood alces vol. 49, 2013 in unsuitable, 23.9% (n = 21) in moderately suitable, and 39.8% (n = 35) in highly suitable habitat (table 3, fig. 5a). there were 150 moose observations in july, with 8.7% (n = 13) in unsuitable, 16.7% (n = 25) in moderately suitable, and 74.7% (n = 112) in highly suitable habitat (table 3, fig. 5b). the april model placed moose in suitability classes in proportion to expectation (x2 = 4.5, p = 0.104, df = 2), whereas the july and hickey (2008) models showed independence from the expected values (p <0.001, df = 2). discussion habitat suitability modeling has been a mainstay of wildlife habitat management for over 30 years (allen et al. 1987), and modern gis technology has changed the use and application of hsi models. the use of remotely-sensed data has grown rapidly, while on-site, intensive habitat evaluation is becoming less common (koitzsch 2002). in addition, complex mathematical operations on entire map layers can now be performed with relative ease (fig. 4), and constructs like esri's modelbuilder™ allows for a largely automated processing of data. modeling represents the only practical way to explore spatial patterns of solar radiation, temperature, and physical elements affecting sunlight transmission through the forest canopy (demarchi and bunnell 1995). our approach combined modeling (i.e., parsing of the te equation), geospatial analysis (i.e., kriging of broad-scale spatial data), and field assessments of sub-canopy temperature and light regimes (i.e., hemispherical photography and micro-meteorology). the statistical and computational methods for modeling geospatial data are welldocumented (johnston et al. 2001); for example, the digital elevation model was key to modeling solar radiation over complex topography (fu and rich 2002), and to co-kriging ta to generate map layers as input to modelbuilder™. even with 12 stations, prediction standard errors for ta were under 2 °c (fig. 2), which we consider adequate for coarse evaluation of suitability. we acknowledge the limited scope of data used to estimate an average kt for april and july, but because the majority of our weather results from synoptic-scale (≥1000 km) atmospheric disturbances (bluestein 1992), we deemed the application of a measured kt value at a single location as reasonable. hemispherical photography is widely used to characterize forest canopy structure (frazer et al. 1999, beaudet and messier table 3. number of observed (o) and expected (e) moose observations (1980–2000) in each suitability class by month, and the related x2 statistic for a new hsi model developed for the adirondack state park, new york. low suitability is an hsi score <0.31, moderate suitability is between 0.32 and 0.66, and high suitability is >0.67. total number of evaluation units is 1028 and expected number of moose is proportional to area of each suitability class. april july hickey (2008) suitability class o e (o-e)2/e std. resid o e (o-e)2/e std. resid o e (o-e)2/e std. resid low 32 36 0.38 −0.62 13 30 9.56 −3.09 182 272 29.81 −5.46 medium 21 26 1.00 −1.00 25 32 1.35 −1.16 1313 832 277.53 16.66 high 35 26 2.96 1.72 112 89 6.20 2.49 204 594 256.64 −16.02 x2-statistic 4.5 17.1 564.0 asl1 0.104 <0.0001 <0.0001 1attained significance level, df = 2. alces vol. 49, 2013 haase and underwood – thermal cover and moose 59 2002, fu and rich 2002, hardy et al. 2004) and provides detailed characterization of the size and distribution of openings in the canopy, which we used to estimate radiation transmission fraction under forest covertypes (fig. 3). while we replicated canopy measurements (n = 3 sites for each forest cover-type), we certainly did not capture sufficient range in variation of sub-canopy radiation transmission due to obscuration and stand management history. a much larger sampling effort would have been required; however, our transmission fractions fall within the range published for the northern hardwood forest (domke et al. 2007; fig. 3). hardy et al. (2004) criticized the method we used to modify the abovecanopy total irradiance, which neglects the roles of path lengths, sunflecks, tree geometry, and micro-topography on sub-canopy irradiance. we do not discount the criticism, but we believe the effects of those factors are less critical over the broad spatial scales we modeled. the formulation of te we used is based on a model for mule deer (odocoileus hemionus; parker and gillingham 1990) and domestic livestock (beaver et al. 1996), and we made several modifications to fit our purpose (table 2). first we modeled the diffuse fraction of solar radiation as a function of kt and solar elevation angle (boland et al. 2001). we also required a variable form for the view factor (kubaha et al. 2004) to permit its calculation for any time of day (haase 2010). in addition, rather than use a linear measurement as a surrogate for characteristic dimension, we computed it directly from volume by estimating carcass composition, body mass, and specific density of tissues (mitchell 1976, haase 2010). the more difficult challenge was mapping the respiratory threshold of renecker and hudson (1986) onto the te scale. assuming that organisms become locally acclimated, it seemed reasonable to adjust their thresholds (schwab and pitt 1991) proportionally to the difference in average regional temperatures collected over the same time span, until thermoregulatory responses by moose in the northeast are observed directly. developing a te model in arcgis™ allows for mapping of the thermal environment across the landscape. despite the complexity of the calculations, crudeness of several data layers, and approximations to key determinants of te, our hsi model generates a more reasonable classification of moose habitat suitability than heretofore available (table 3). because we combined all 3 forest cover-types (i.e., hardwood, softwood, mixed) that were thermally suitable (i.e., te < ucte), our model classified fewer evaluation units in the high suitability class; therefore, the amount of thermal cover available for moose decreased as a consequence of the conflation. assuming that few hardwood stands are thermally suitable in april, our hsi more accurately classified evaluation units as moderate and low suitability. though the july model did a poorer job of correctly classifying moose observations, it improved upon the hickey (2008) model that over-classified evaluation units into low and high suitability classes leaving the moderate habitat substantially underclassified (table 3). due to the presence of industrial forest lands, it is generally accepted that the western half of the asp has higher forage value for moose (garner 1989). it also contained the largest area of highly suitable habitat based on our hsi, despite exhibiting mostly moderate thermal suitability (fig. 5), indicating the complexity of evaluating the relative importance of forage and/or thermal cover to moose. the dramatic reduction in radiation transmitted through the forest canopy and the inclusion of open water boosted the suitability of eastern asp during july; however, the 2 monthly maps differ mostly in extent, rather than location of highly suitable habitat. 60 thermal cover and moose – haase and underwood alces vol. 49, 2013 classifying habitat suitability by season is important, because aspects of the thermal environment and forage availability necessary for moose survival change throughout the year, particularly in the asp (fig. 5; koitzsch 2002, dussault et al.2006). moose cope with thermal and nutritionally stressful environments and seasons through physiological and behavioral adaptations (schwab and pitt 1991, dussault et al. 2004). moose, in the short-term, cope with stressful thermal conditions by trading off time for space in favorable microhabitats (bakken 1992, parker and gillingham 1990, sargeant et al. 1994, mysterud and ostbye 1999). but, moose must maintain homeostasis in the long-term or face potentially deleterious individual and population consequences (belovsky 1981, dussault et al. 2004). our development of a moose hsi that incorporated thermal suitability agreed in general with known locations and suitable habitat of moose in the asp. along their southern range boundary moose are declining in certain areas (e.g., minnesota, lenarz et al. 2008) and appear to be thriving in others (e.g., quebec, ontario; dussault et al. 2004, lowe et al. 2010) despite ∼5 ° difference in latitude. habitat quality, forage abundance, effects of disease and parasites, density of white-tailed deer (see lankester 2010), and combined effects of stress associated with warmer temperatures (see murray et al. 2006, lenarz et al. 2008) are possible explanations of regional and local differences in moose density and population response. further elucidating the interrelationships of these factors, including the interaction of forage and thermal cover, is warranted to address a warming climate, differential population responses, and potential range shifts of moose. acknowledgements we thank the members of the quantitative studies laboratory and the adirondack ecological center at the state university of new york college of environmental science and forestry and dr. m. glennon of the wildlife conservation society for field and technical support. we thank 2 anonymous reviewers for substantially improving an earlier draft of this manuscript. funding was provided by the edna b. sussman foundation and the sigma xi research society. references adirondack park agency (apa). 2001. state of new york adirondack park state master plan. ray brook, new york, usa. allen, a.w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. united states fish and wildlife service biological report 82 (10.155), fort collins, colorado, usa. bakken, g. s. 1992. measurement and application of operative and standard operative temperatures in ecology. american zoology 32: 194–216. beaudet, m., and c. messier. 2002. variation in canopy openness and light transmission following selection cutting in northern hardwood stands: an assessment based on hemispherical photographs. agricultural and forest meteorology 110: 217–228. beaver, j. m., b. e. olson, and j. m. wraith. 1996. a simple index of standard operative temperature for mule deer and cattle in winter. journal of thermal biology 21: 345–352. belovsky, g. e. 1981. optimal activity times and habitat choices of moose. oecologica 48: 22–30. beyer, h. l. 2004. hawth's analysis tools for arcgis. (accessed 2010). bird, r. e., and r. l hulstrom. 1991. a simplified clear sky model for direct and diffuse insolation on horizontal s urfaces. seri technical report seri/ alces vol. 49, 2013 haase and underwood – thermal cover and moose 61 http://www.spatialecology.com/htools http://www.spatialecology.com/htools tr-642–761. energy research institute, golden, colorado, usa. bluestein, h. b. 1992. synoptic-dynamic meteorology in midlatitudes: principles of kinematics and dynamics. vol. 1. oxford university press, oxford, england. boland, j., l. scott, and m. luther. 2001. modeling the diffuse fraction of global solar radiation on a horizontal surface. environmetrics 12: 103–116. cameron, j. r., j. g. skofronick, and r. m. grant. 1999. physics of the body. second edition. medical physics publishing, madison, wisconsin, usa. campbell, g. s. 1981. fundamentals of radiation and temperature relations. pages 11–40 in o. l. lange, p. s. nobel, c. b. osmond, and h. ziegler, editors. encyclopedia of plant pathology. vol. 12a, physiological plant ecology i. springer-verlag, berlin, germany. ———, and j. m. norman. 1997. an introduction to environmental biophysics, 2nd edition. springer, new york, new york, usa. chaffee, r. r., and j. c. roberts. 1971. temperature acclimation in birds and mammals. annual review of physiology 33: 155–202. demarchi, m. w. 1991. influence of the thermal environment on forest cover selection and activity of moose in summer. m. s. thesis, university of british columbia, vancouver, british columbia, canada. ———, and f. l. bunnell. 1995. forest cover selection and activity of cow moose in summer. acta theriologica 40: 23–36. dettki, h., r. löfstrand, and l. edenius. 2003. modeling habitat suitability for moose in coastal northern sweden: empirical vs. process-oriented approaches. ambio 32: 549–56. dinunzio, m. g. 1984. adirondack wildguide. adirondack conservancy committee and adirondack council, elizabethtown, new york, usa. domke, g. m., j. p. caspersen, and t. a. jones. 2007. light attenuation following selection harvesting in northern hardwood forests. forest ecology and management 239: 182–190. dussault, c., r. courtois, and j. j. -p. ouellet. 2006. a habitat suitability index model to assess moose habitat selection at multiple spatial scales. canadian journal of forest resources 36: 1097–1107. ———, j. -p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioral responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321–328. frazer, g. w., c. d. canham, and k. p. lertzman. 1999. gap light analyzer (gla), version 2.0: imaging software to extract canopy structure and gap light transmission indices from true-colour fisheye photographs, user's manual and program documentation. simon fraser university, burnbay, british columbia, canada and the institute of ecosystem studies, millbrook, new york, usa. fu, p., and p. m. rich. 2002. a geometric solar radiation model with applications in agriculture and forestry. computers and electronics in agriculture 37: 25–35. garner, d. l. 1989. ecology of the moose and the feasibility for translocation into the greater adirondack ecosystem. m. s. thesis, suny college of environmental science and forestry, syracuse, new york, usa. gates, d. m. 1980. biophysical ecology. springer-verlag, new york, new york, usa.. guisan, a., and n. e. zimmermann. 2000. predictive habitat distribution models in ecology. ecological modeling 135: 147–186. haase, c. g. 2010. characterizing critical thermal environments for moose (alces alces) in the adirondack mountains of new york. m. s. thesis, suny college 62 thermal cover and moose – haase and underwood alces vol. 49, 2013 of environmental science & forestry, syracuse, new york, usa. hardy, j. p., p. melloh, g. koenin, d. marks, a. winstral, j. w. pomeroy, and t. link. 2004. solar radiation transmission through conifer canopies. agricultural and forest meteorology 126: 257–270. hickey, l. 2008. assessing the recolonization of moose in new york with hsi models. alces 44: 117–126. hundertmark, k. j., c. c. schwartz, and t. r. stephenson. 1997. estimation of body composition in moose. federal aid in wildlife restoration final report. alaska department of fish and game, division of wildlife conservation, juneau, alaska, usa. jenkins, j. 2004. the adirondack atlas: a geographic portrait of the adirondack park. wildlife conservation society, bronx, new york, usa. johnston, k., j. m. ver hoef, k. krivoruchko, and n. lucas. 2001. using arcgis™ geostatistical analyst™. esri, redlands, california, usa (accessed 2010). koitzsch, k. b. 2002. application of a moose habitat suitability index model to vermont wildlife. alces 38: 89–107. kubaha, k., d. fiala, j. toftum, and a. h. taki. 2004. human projected area factor for detailed direct and diffuse solar radiation analysis. international journal of biometeorology 49: 113–129. lankester, m. w. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53–70. lenarz, m., m. e. nelson, m. w. schrage, and a. j. edwards. 2008. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503–511. lindroth, a., and k. perttu. 1981. simple calculation of extinction coefficient of forest stands. agricultural meteorology 25: 97–110. lowe, s. j., b. r. patterson, and j. a. schaefer. 2010. lack of behavioral responses of moose (alces alces) to high ambient temperatures near the southern periphery of their range. canadian journal of zoology 88: 1032–1041. mitchell, j. w. 1976. heat transfer from spheres and other animals. biophysical journal 16: 561–569. moen, a. n. 1968. critical thermal environment: a new look at an old concept. bioscience 18: 1041–1043. monteith, j. l. 1973. principles of environmental physics. american elsevier, new york, new york, usa. ———, and m. h. unsworth. 1990. principles of environmental physics. second edition. edward arnold, new york, new york, usa. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1–30. mysterud, a., and e. ostbye. 1999. cover as a habitat element for temperate ungulates: effects on habitat selection and demography. wildlife society bulletin 27: 385–394. parker, k. l., and m. p. gillingham. 1990. estimates of critical thermal environments for mule deer. journal of range management 43: 73–81. puttock, g. d., p. shakotko, and j. g. rasaputra. 1995. an empirical habitat model for moose, alces alces, in algonquin park, ontario. forest ecology and management 81: 169–178. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory response of moose. canadian journal of zoology 64: 322–27. ———, and ———. 1990. behavioral and thermoregulatory responses of moose to high ambient temperatures and insect alces vol. 49, 2013 haase and underwood – thermal cover and moose 63 http://www.esri.com http://www.esri.com harassment in aspen dominated forests. alces 26: 66–72. romito, t., k. smith, b. beck, j. beck, m. todd, r. bonar, and r. quinlan. 1999. moose winter habitat suitability index model. version 5. foothills model forest, alberta, canada. sargeant, g. a., l. e. eberhardt, and j. m. peek. 1994. thermoregulation by mule deer (odocoileus hemionus) in arid rangelands of southcentral washington. journal of mammalogy 75: 536–544. schwab, f. e., and m. d. pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071–3077. 64 thermal cover and moose – haase and underwood alces vol. 49, 2013 integrating thermal constraints into habitat suitability for moose in the adirondack state park, new york study area methods thermal cover (habitat variable v3) hsi model ii results discussion acknowledgements references alces34(1)_59.pdf alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 135 the impact of moose (alces alces andersoni) on forest regeneration following a severe spruce budworm outbreak in the cape breton highlands, nova scotia, canada craig smith, karen beazley, peter duinker, and karen a. harper school for resource and environmental studies, dalhousie university, 6100 university avenue, halifax, nova scotia b3h 3j5, canada. abstract: two interacting disturbances such as stand-level defoliation by spruce budworm (choristoneura fumiferana) and subsequent herbivory by moose (alces alces) may affect landscapes differently than if they occurred in isolation. we studied moose (a. a. andersoni) browsing on sites disturbed approximately 25 years ago by a severe spruce budworm outbreak in a region historically dominated by balsam fir (abies balsamea) forest on northern cape breton island, nova scotia, canada. our objectives were to 1) describe the impact of a large resident moose population on post-budworm regeneration of balsam fir and white birch (betula papyrifera), and 2) to examine the interplay between moose abundance, site conditions, and variation in post-budworm forest regeneration. fifty-eight randomly located sites were sampled for composition and structural characteristics, moose browse severity, moose pellet group density, and site conditions. we used univariate general linear modelling (glm) and multivariate redundancy analysis (rda) to examine relationships between moose abundance as indicated by pellet-groups, site conditions, and post-budworm regeneration. approximately 65% of all balsam fir and white birch saplings tallied were severely browsed by moose, exhibiting stunted, abnormal growth forms. both the glm and the rda indicated that moose abundance was the best predictor of variation in the density of post-budworm regeneration of balsam fir and white birch. site conditions were less useful predictors of variation in regeneration. the relationship between moose abundance and regeneration of balsam fir and white birch was positive, suggesting that moose may be more abundant in areas where regeneration is more dense. sustained, severe browsing in areas regenerating after spruce budworm outbreak may significantly inhibit future forest development and alter the well documented spruce budworm-balsam fir cyclic successional system of northern cape breton island. alces vol. 46: 135-150 (2010) key words: alces alces andersoni, balsam fir, browse, defoliation, forest succession, herbivory, moose, natural disturbance, regeneration, spruce budworm. natural disturbances are drivers of change in the composition and structure of forest ecosystems. while classic succession theory predicts cyclic patterns, many recent studies offer an opposing view, suggesting that postdisturbance development may take alternate successional pathways or trajectories in a wide variety of ecosystems (see dublin et al. 1990, augustine and frelich 1998, frelich and reich 1999, payette et al. 2000, jasinski and payette 2005). interactions between disturbances and the resultant impact on landscapes remain poorly studied despite their relative prevalence (radeloff et al. 2000). paine et al. (1998) argued that interacting or compounding disturbances carry greater potential ramifications for the long-term transformation of a community than that of large, infrequent disturbances. the spruce budworm (choristoneura fumiferana), a native insect defoliator in the spruce-fir forests of eastern canada, is prone to population explosions that can result in stand-replacing disturbance in balsam fir moose impacts on forest regeneration smith et al. alces vol. 46, 2010 136 (abies balsamea) dominated forests (maclean 1984). the ability of balsam fir forests to succeed themselves through advanced regeneration has contributed to claims that spruce budworm and balsam fir interact in a cyclic successional system (e.g., baskerville 1975, maclean 1984, 1988). the impacts of moose (alces spp.) browsing on forest ecosystems have been well studied in north america and scandinavia. these studies have shown that browsing by moose can alter the development of individual trees (bergerud and manual 1968, snyder and janke 1976, brandner et al. 1990), alter the compositional and structural make-up of a forest (risenhoover and maass 1987, thompson et al. 1992, connor et al. 2000), and serve as an important function in the cycling of nutrients within an ecosystem (pastor and naiman 1992, pastor et al. 1993, pastor and kjell 2003). interactions between an outbreak of spruce budworm and browsing by a significant population of moose may have created a previously undocumented successional pattern in the forest ecosystem of northern cape breton island. radeloff et al. (2000) studied the effects of disturbance by an outbreak of the jackpine budworm (choristenuera pinus pinus) and subsequent salvage logging and found that these 2 interacting disturbances had together affected the landscape differently than if they had occurred in isolation of each other. citing examples of interplay among forestry practices, agricultural development, and forest fires in the southern boreal forest, and between climatic conditions and exotic invasive species in san francisco bay, among others, paine et al. (1998) found that the impacts of these “compounded perturbations” were more significant, long-term, and fundamentally different than the impacts of even “large infrequent disturbances” on the host ecosystem. in this paper we examine the impacts of moose (alces alces andersoni) on regeneration of balsam fir and white birch (betula papyrifera) following an outbreak of spruce budworm on northern cape breton island, nova scotia. a large portion of the area affected by spruce budworm outbreaks is not presently following the expected cyclic successional trajectory (smith 2007). we believe a second, unanticipated disturbance – herbivory by moose – is inhibiting post-budworm successional development. our research objectives were to 1) describe the impact of moose on post-budworm regeneration of balsam fir and white birch, and 2) examine the interplay between moose abundance, site conditions, and variation in post-budworm forest regeneration. the response of sites disturbed by spruce budworm followed by moose herbivory is of increasing interest to land and resource management agencies in the atlantic provinces. monitoring their succession, or alternatively their regression, will yield insight into interacting population irruptions of spruce budworm and moose and their resultant impact on the boreal forest of northern cape breton, with potential relevance for similar interacting disturbances in forests elsewhere. methods study area the study area was between 60º 56´ and 60º 94´ latitude and 46º 61´ and 46º 90´ longitude and encompassed 2 protected areas: 1) cape breton highlands national park (hereafter referred to as the national park) a 948 km2 protected area spanning the island from the west coast to east, and 2) the provincially designated pollet’s cove-aspy fault wilderness area (hereafter referred to as the wilderness area), which abuts the northern border of the national park and extends almost to the northernmost tip of the island (fig. 1). specifically, the study sites were located on the cape breton plateau in areas of the national park and the wilderness area affected by the spruce budworm outbreak of 1974-84 (see smith 2007). alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 137 the close proximity of the atlantic ocean influences the climate of northern cape breton in all seasons. the sharp rise of the plateau from the ocean produces orographic precipitation with a notable gradient from west to east. the plateau region of the highlands is cooler than lower-lying coastal areas. snowfall accumulations are greater and generally persist well into spring (bridgland and millette 1995). the maritime influenced boreal forest, unique in nova scotia, is the dominant forest type on the plateau, is close to the southern limit of its range, and has historically been dominated by balsam fir (fernow 1912, collins 1951). black spruce (picea mariana) is also abundant, while white spruce (p. glauca), mountain ash (sorbus americana) and red maple (acer rubrum) are less common. a landcover type characterized by barrens, stunted black spruce trees, and lichen-heath communities, locally referred to as taiga, is also at elevations >400 m (a.s.l.). spruce budworm in northern cape breton -s pruce budworm is an important agent of disturbance in balsam fir-dominated forests. the most recent outbreak (1974-1984) on northern cape breton island was intense, resulting in 87% average mortality in affected stands (ostaff and maclean 1989). balsam fir mortality levels were evenly distributed among diameter classes and were not correlated with site characteristics (maclean and ostaff 1989). despite the severity of the cape breton outbreak, it was widely believed that advanced regeneration would facilitate a return to a canopy largely dominated by balsam fir. this expected return would follow stages of secondary succession characterized by a mix of deciduous (primarily white birch) and coniferous (primarily balsam fir with some fig. 1. study area and site locations in polletts cove-aspy fault wildnerness area and cape breton highlands national park, nova scotia, canada. moose impacts on forest regeneration smith et al. alces vol. 46, 2010 138 spruce [picea spp.]) stands (maclean 1988, pardy 1997, smith 1998). moose in northern cape breton - moose are native to all of nova scotia but were extirpated from cape breton island in the early twentieth century, likely due to over-hunting and habitat alteration (cameron 1958). in 1947 and 1948, 18 moose were transported from elk island national park in alberta and released in northern cape breton (pulsifer and nette 1995). these moose were of a different subspecies (a. a. andersoni) than those endemic (a. a. americana) to nova scotia. while initial population growth after re-introduction was slow, the moose population has increased drastically since the 1970s. mid-winter surveys in recent years suggest that as many as 5000 (± 1000) moose inhabit northern cape breton, with local mid-winter densities as high as 20 animals/km2 in some areas (parks canada, unpublished data); however, the most recently (2008) compiled estimates suggest the population may be in decline (d. quann, cape breton highlands national park, unpublished data). moose in northern cape breton forage heavily on white birch (basquill and thompson 1996) during summer, whereas balsam fir is a more important forage of moose in eastern canada than in other parts of north america (bergerud and manuel 1968, risenhoover and maass 1987, mcinnes et al. 1992, basquill and thompson 1996). sample site selection we imposed a 10-ha minimum patch size requirement on potential sites, enabling us to minimize potential edge effects when orienting sampling units. this requirement restricted the observed sites to the western portion of the national park and the central portion of the wilderness area (fig. 1). potential sample sites were stratified across a gradient of moose abundance within the study area. abundance values derived from 4 years of aerial surveys (parks canada, unpublished data) were summed for each survey block (approximately 5 km2), and classified as either low or high (moose abundance) using an arbitrary cut-off of an average abundance rating of 9 moose per survey block (≤ 9 moose/survey block = low abundance, >9 moose/survey block = high abundance). abundance ratings were derived from sightings, tracks, and other moose sign. a subjectively ranked sightability factor was included in the calculation of abundance scores to account for differences in tree cover and other variables affecting overall ability to see moose or moose sign. we used gridded random tessellation sampling to randomly select 15 sample sites from the varied number of possibilities for 4 different site types. site types were determined from spectral signatures derived from a supervised classification of a spot 5 ms 10 satellite scene of the post-budworm landscape. the spectral signatures were identified as predominant from photographic analysis of sites associated with each spectral signature and refined using field verification and statistical analysis. the final classified and filtered image resulted in 4 classes identified by their seemingly predominant characteristics: grassland (grass), snag (snag), downed wood (blowdown), and browsed deciduous (browsed deciduous). access issues and misclassifications altered the intended sample size so that final sample sizes ranged from 13-18 sites per class, for a total of 58 sites (see smith 2007 for further description). field measurements we sampled trees (≥2.0 m in height), saplings (<2.0 m and ≥0.5 m in height) and seedlings (<0.5 m in height) using a series of 4 parallel, rectangular plots placed at 5 m intervals. plots were 4 m x 50 m for trees and snags (4 x 200 m2 = 800 m2 total), and 2 m x 50 m for saplings and seedlings (4 x 100 m2 = 400 m2 total). start-points and orientation of plots were pre-determined and designed to maximize randomization and minimize alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 139 potential edge effects. the density of live and dead woody species was recorded. all live individuals were classified subjectively according to browsing severity: unbrowsed, lightly browsed, moderately browsed, severely browsed, and dead. unbrowsed individuals exhibited no evidence of moose browsing. dead individuals had no living parts; we assumed that dead saplings exhibiting signs of severe browsing suffered mortality as a direct result of that browsing. while no dead saplings devoid of evidence of severe browsing were encountered, it is possible that some individuals perished as a result of several cumulative factors. however, in the absence of disease, drought, interspecific competition, or other known stressors, and given the abundance of moose in the region, we concluded that it was a justified assumption. the browse severity classification was based on growth form, the proportion of stems browsed, and the presence/absence of live foliage (table 1). while severity classifications have been criticized elsewhere (e.g., mclaren et al. 2004) for being subjective and potentially variable among species, we believe that ours was an efficient method to obtain a relative measure of the proportion of individuals browsed by moose versus those that were unbrowsed or dead. while the conditional approach taken is admittedly subjective, the simplicity of the method, the low number (3) of species classified, and the ubiquity of severely altered growth forms afforded us confidence in our measure. trees were assigned to 1 of 2 height classes (2-4 m, >4 m) and measured for diameter at breast height (dbh). ground vegetation was sampled using 1-m2 quadrats placed every 10 m within each of the 4 plots composing one sample site, for a total of 20 per site (20 m2 total). herbaceous plants were identified to species, with the exception of aster spp., which were identified to genus. grasses and sedges were identified to family, with the exception of calamagrostis spp. lichens and mosses were simply identified as such, with the exception of sphagnum spp., which were identified to genus. nomenclature followed zinck (1998). at each site we measured the continuous variables elevation (elev) and slope (slope), and the categorical variable aspect (aspect, 8 categories). additional categorical environmental variables included topography (topo, 4 categories), drainage (drain, 2 categories), and soil texture that were obtained from the nova scotia department of natural resources (nsdnr) ecological land classification (neily et al. 2003). soil texture was later dropped from the analysis question unbrowsed lightly browsed moderately browsed severely browsed dead live foliage? yes yes yes yes no evidence of browse? no yes yes yes n/a growth form altered? no no yes some evidence yes drastic alteration n/a multiple dead stems? no no no yes n/a(white birch, mountain ash) cylindrical and dense? no no no yes n/a(balsam fir) proportion of stems browsed? none < 1/3 > 1/3, < 2/3 >2/3 n/a table 1. criteria and decision making used in the browse severity classification used to assess impact of moose browsing on forest regeneration, cape breton highlands, nova scotia. n/a = not applicable moose impacts on forest regeneration smith et al. alces vol. 46, 2010 140 because it was homogeneous across all sample sites. moose feces, deposited as pellets in winter and non-pelletized clumps in summer, were recorded within the 2 x 50 m portion of the plot. non-pelletized fecal deposits and pellet groups consisting of >30 pellets were counted (neff 1968). the predictor variable (pellet), used here as a surrogate for moose abundance, represents the density of moose fecal deposits found within the total area of each sample site (800 m2). data analyses we used both univariate and multivariate modeling approaches to examine the relationship between environmental variables and post-budworm regeneration of balsam fir and white birch. for the dependent variable in the univariate analysis, we developed a regeneration index an additive measure of post-budworm regeneration success (white birch (wb) and balsam fir (bf)) on each site. our regeneration index is the sum of the densities of trees (t), saplings (s), and seedlings (sd) weighted by their growth form: regeneration index = 3wb(t) +3bf(t) + 2wb(s) + 2bf(s) + wb(sd) + bf(sd). we then used the proc genmod function in sas version 9.1 (littell et al. 1996, sas institute 2003) to fit a generalized linear model (glm) to the response variable regeneration index. goodness-of-fit statistics indicated that a model fit to the gamma distribution of the ‘regeneration index’ using the log-link function was most appropriate. we confirmed this with a visual inspection of a histogram. the gamma distribution is a continuous probability distribution arising from the poisson process and may take a variety of shapes. the link function is a mathematical transformation embedded in the model used to normalize the distribution of errors. in the case of the gamma distribution, the log-link is recommended (mccullagh and nelder 1989). to fit the model, we tested the significance of individual model terms using chi-square probability values from likelihood ratio tests (type iii analyses; α = 0.05) that are independent of the order in which variables are entered into the model (johnstone n.d.). significant predictor variables were used as a base model to which additional model terms were added individually to test whether they improved the model fit. models were ranked using akaike’s information criterion (aic), a part of the information-theoretic approach to model selection (burnham and anderson 1998). the aic formula is: -2(log likelihood) + 2(p) where p is the number of parameters in the model. differences between models (∆i) of >10 aic units or more from the model with the lowest aic value indicate strong support that the model is not the best. differences of <2 aic units are considered small and suggest that identification of a best model between the two is not pragmatic (burnham and anderson 1998). to examine the multivariate relationship between environmental variables and postbudworm regeneration of balsam fir and white birch, we conducted constrained ordinations for each of white birch and balsam fir using redundancy analysis (rda; rao 1964). the multivariate species variables were the densities of white birch and balsam fir in various stages of growth, including seedling density and the densities of saplings within 4 browse severity classes (unbrowsed, lightly browsed, moderately browsed, severely browsed) excluding dead individuals. we also included trees in the 2-4 m height class as post-budworm regeneration, believing them to be young enough to have germinated and grown in years following the spruce budworm outbreak; trees >4 m in height were identified as potential budworm survivors and were not regarded as post-budworm regeneration. alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 141 redundancy analysis is a technique akin to a multiple regression for all dependent variables simultaneously, combining elements of ordination and regression to detect patterns in response data that can be best explained by the subset of environmental or predictor variables used (ter braak 1995). the final set of environmental variables was selected in a preliminary rda using forward selection (α = 0.10) and only significant variables from the preliminary analysis were used in the final iteration, as recommended by lepš and šmilauer (2003). significance testing was by monte carlo permutation. the redundancy analysis was performed in canoco for windows ver. 4.54 (ter braak and šmilauer 2002). non-parametric spearman correlations were calculated to examine associations between pellet and the browse severity classes. significance testing was performed in spss 11.5 (spss inc. 2005). results moose browse severity evidence of moose browsing was ubiquitous across our study sites. the classification of browse severity showed that few saplings had been browsed by moose, and the vast majority had been severely browsed or were dead as a result of browsing. greater than two-thirds of all white birch and balsam fir saplings encountered were severely browsed (fig. 2). the browse classifications for white birch were dominated by the severely browsed and dead categories; more variation existed for balsam fir that had more stems classified as unbrowsed, lightly browsed, and moderately browsed than white birch (fig. 2). mountain ash (sorbus spp.) was recorded far more infrequently; 84% were browsed severely (not shown) but no dead individuals were encountered at any site. predicting variation in post-budworm regeneration pellet was the most important variable in the fit of the glm (tables 2, 3). adding subsequent model terms resulted in a slightly improved fit but higher aic value in most cases (table 2), with the exception of the pellet + slope model. because the difference between the pellet model and the pellet + slope model is <2 aic units, it is not possible to declare either of them best. therefore, we present additional information for only the pellet + slope model that had the lowest aic value (table 3). pellet and slope had similar effect sizes in relation to the density of post-budworm regeneration; the coefficients indicate that for approximately every 1-unit increase in pellet and slope (1 pellet-group, 1° slope), regeneration increased by 1 unit (1 stem). the subset of predictors fit the gamma distribution of the dependent variable regeneration index well (scaled deviance = 1.00) despite known cases of clustering and outliers. the scaled deviance is the measure of the fit of the residuals to the predicted distribution, in fig. 2. mean stem density of balsam fir and white birch and the proportional breakdown among the 5 browse severity classes, cape breton highlands, nova scotia. moose impacts on forest regeneration smith et al. alces vol. 46, 2010 142 this case, the gamma. one (1.00) is a centre point that indicates a good fit; the farther the deviance is from 1.00 in either direction, the poorer the fit. results of the multivariate rda examining the relationship of the environmental variables and pellet coincided with the glm indicating that pellet was the best predictor of variation of post-budworm regeneration. redundancy analyses explained 27% of variation in balsam fir regeneration and 22% of the variation in white birch regeneration, indicating that factors other than pellet and the interaction between slope and aspect contribute to the variation in post-budworm regeneration. in the rda, the first axis accounted for >98% of the explained regenerationenvironment relations for both balsam fir and white birch regeneration (table 4). the variable pellet and the interaction term slope*aspect were significantly related to variation in post-budworm regeneration of both species. the first axis of each rda analysis explained 27% and 22% of the overall variation (not shown) for balsam fir and white birch, respectively. for balsam fir, interset correlations showed that axis 1 was strongly positively correlated with pellet, and negatively correlated with all other variables. interset correlations in the ordination for white birch showed that the first axis was positively correlated with all variables, most strongly with pellet and the interaction term slope*aspect. axis 2 explained < 1% of the variation in regeneration for both species (not shown). the spearman correlations examining the association between pellet and the 5 sapling browse categories for each of white birch and balsam fir indicated that the relationship between pellet and the regeneration index is likely a function of significant positive associations between moose pellet abundance and the density of severely browsed balsam fir and white birch saplings (table 5). discussion moose and post-budworm regeneration evidence of sustained, heavy moose model term n l-likelihood aic δi pellet 58 -355.41 712.82 0.9 pellet + slope 58 -353.96 711.92 0 pellet + elev 58 -355.4 714.8 2.88 pellet + topo 58 -355.74 715.48 3.56 pellet + drain 58 -355.08 714.16 2.24 pellet + aspect 58 -355.76 715.52 3.6 pellet + slope*aspect + slope + aspect 58 -355.31 718.7 6.78 table 2. generalized linear model results examining the influence of predictor variables on post-budworm regeneration (models were fit to the gamma distribution of errors using the log link function), cape breton highlands, nova scotia. l-likelihood = log likelihood for the model, aic = akaike’s information criterion, and δi = difference between a given model and the model with the lowest aic in units of aic. variables are described in methods. model term coefficient s.e. p value intercept 126.23 1.19 – pellet 1.08 1.03 0 slope 1.01 1.03 0.33 scaled deviance = 1.00 table 3. details for the pellet + slope model used to assess impact of moose browsing on forest regeneration, cape breton highlands, nova scotia. intercept coefficient represents mean regeneration holding other model terms constant. coefficient = effect size, s.e. = standard error, and p value is for chi-square goodness of fit test. the scaled deviance is obtained by df/deviance; a value closer to 1 indicates a good fit. alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 143 browsing in the form of severely stunted, atypical growth forms was noted on >65% of all white birch and balsam fir saplings. low runners or shoots emerging from the root collar were often the only live vegetation noted on saplings. mortality as a result of browsing was much more common for white birch than balsam fir, suggesting a lower tolerance to browsing or possibly reflecting moose preference for white birch. however, as mclaren et al. (2009) concluded, foraging preference is a useful, but not holistic predictor of vegetation responses in the short term, and shade tolerance and the ability to invest in below ground resources will also dictate a species’ response to the onset of, or release from, severe browsing by moose. strong survivorship under intense browsing can also be inferred from our field data for mountain ash which was relatively common on the landscape in a severely browsed condition, yet never recorded as dead. our hypothesis regarding the impact of moose on post-budworm regeneration was that the density of post-budworm regeneration would exhibit a negative relationship with moose abundance. we also anticipated more significant mortality levels. however, we found that where regeneration was most dense, moose were more abundant, and while browsing pressure was extreme, mortality balsam fir white birch axis 1 axis 2 axis 1 axis 2 eigenvalue 0.268 0.003 0.223 0.002 species-environment correlation 0.576 0.239 0.491 0.305 cumulative % variation explained 26.8 27.2 22.3 22.5 cumulative % species-environment 98.4 99.5 98.6 99.6 interset correlations pellet 0.48 0.12 0.398 0.142 slope*aspect -0.096 0.052 0.267 -0.146 aspect -0.313 0.131 0.075 -0.226 slope -0.105 0.075 0.244 0.012 table 4. results of constrained ordination (redundancy analysis) analyzing the relationship between predictor variables and variation in post-budworm regeneration of balsam fir and white birch, cape breton highlands, nova scotia. only variables determined to be significant (α = 0.10) using a preliminary stepwise selection were included. eigenvalues for the axes represent the percentage of variation explained by each. the species-environment correlation is a measure of the relationship between each axis and the species data. also given are the cumulative percentage of the variation explained by the axes and the cumulative percentage of the total species-environment relations. pearson interset correlations represent the association of each variable with each axis; slope*aspect is the interaction between slope and aspect. browse class correlation coefficient p value balsam fir unbrowsed 0.065 0.63 lightly browsed -0.034 0.8 moderately browsed 0.124 0.35 severely browsed 0.493 0 dead 0.013 0.92 white birch unbrowsed -0.188 0.15 lightly browsed -0.067 0.61 moderately browsed 0.171 0.19 severely browsed 0.554 0 dead 0.152 0.25 table 5. non-parametric spearman correlations between pellet and the sapling browse severity classes, cape breton highlands, nova scotia. moose impacts on forest regeneration smith et al. alces vol. 46, 2010 144 was not. the positive relationship between moose abundance (pellet) and the density of post-budworm regeneration is likely a function of: 1) the significant correlation with the density of severely browsed saplings (table 5), 2) growth form adaptations in response to browsing, and 3) rapid degradation of moose feces by insects and precipitation (as found by timmerman and buss 1997). the significant correlations between pellet and the density of severely browsed white birch and balsam fir indicate that severely browsed saplings are significantly associated with increased moose abundance while all other browse classification categories are not. both balsam fir (brandner et al. 1990, mclaren 1996) and white birch (mclaren et al. 2009) adapt their growth form under browsing pressure. balsam fir has been shown to adopt exaggerated apically or laterally oriented forms, depending upon the nature of the browsing and access to light (brandner et al. 1990, mclaren 1996), whereas white birch (among other hardwoods) has been shown to invest more in below-ground-growth (mclaren et al. 2009). while detailed structural data were not collected for saplings, we observed many altered growth forms similar to those described elsewhere (e.g., brandner et al. 1990, mclaren 1996). we are not aware of any data regarding rates of decomposition for moose feces in northern cape breton, though neff (1968) reported that the time span required for ungulate pellet decomposition can be as little as several months. comparing the short amount of time required for decomposition of moose feces in conjunction with the presumably longer period of time required for mortality induced by repeated browsing indicates that mortality occurring beyond the decomposition period would not be reflected in our snapshot of the current situation. we believe the nonsignificant correlations between pellet and dead saplings are representative of this. the cumulative result of these 3 factors is that sites exhibiting the ultimate impact of sustained, severe moose browsing – a reduction in the densities of preferred browse species as a result of mortality are not adequately represented by our regeneration index. however, the positive relationship between moose abundance and post-budworm regeneration (particularly the ubiquity of severely browsed saplings) describes the relationship between the ecosystem response to the stand initiating disturbance caused by spruce budworm, and the large resident moose population of the northern cape breton highlands. site factors as predictors of post-budworm regeneration with the exceptions of slope in the glm and the interaction term slope*aspect in the rdas, environmental variables were not important in explaining variation in post-budworm regeneration. an increase in slope was associated with an increase in regeneration; however, only 3 sites had a slope of >10 degrees (none >14), making it difficult to interpret any real significance from this association. while it is possible that site factors are not contributing to variation in regeneration, it is also possible that the landscape-level data we used, derived from an ecological land classification of the entire province of nova scotia, was too coarse to detect subtle site conditions that may be contributing to the emergence of varied regeneration. particularly, site-specific information on soil moisture and site productivity during the growing season may improve understanding of the presence and potential expansion of calamagrostis spp. capacity for recovery our field data suggest that with sustained high levels of moose browsing, the sparsely treed conditions prevailing in most budwormaffected areas of the national park and the wilderness area are likely to persist. we derive this conclusion by considering several points. alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 145 first, the current condition and proportion of both balsam fir and white birch saplings recorded as severely browsed indicates that these individuals will either perish or, at a minimum, not grow to become canopy trees. in addition, a lack of trees having escaped the moose browse zone (0-3 m in height) indicates a diminished and potentially inadequate seed source from continued tree regeneration. an extreme example of this has been noted in terra nova national park, newfoundland, canada, where annual seed rain was consistently ~0.0014 seeds/m2 over 8 years of collection (l. hermanutz, memorial university, personal communication). data on cone production and seed rain in regenerating areas and adjacent sites would be useful in understanding the potential for a return to forested conditions if browsing were reduced. additionally, the recent establishment of moose exclosures will facilitate an understanding of the effects of the removal of moose browsing pressure, and could provide insight into the capacity for restoration of a balsam fir-dominated canopy. experimental work allowing for reduced, but not absent, browsing pressure would be of greater interest in understanding current potential resilience within this system, and would allow managers to simulate the effect of a moose population reduction. the ubiquity of calamagrostis in some sites (smith 2007) may inhibit future forest development by affecting seedling germination and success. hogg and lieffers (1991) reported notable changes in soil thermal regimes under calamagrostis sod, noting midsummer soil temperatures fully 4.0º c warmer than in open sites. calamagrostis reduces root growth and water and nutrient uptake in spruce seedlings (grossnickle 1988) and can potentially cause drought stress (macdonald 1986). further, cater and chapin (2000) found that white birch seedlings were also suppressed under the dense sod of calamagrostis. in addition, calamagrostis expands rapidly, requiring just a few years to fully colonize sites after canopy removal (eis 1981). coomes et al. (2003) noted expanded niche occupation by non-palatable species as a potential cause of a lack of regeneration following deer population reduction in new zealand. if calamagrostis is contributing to seedling mortality thereby furthering conditions conducive to its expansion, in essence a positive feedback switch (wilson and agnew 1992), it may substantially alter future forest development even with reduced moose browsing. further understanding of the dynamics of sites dominated by calamagrostis would provide insight into the future successional trajectory of the affected areas of the cape breton plateau. the most salient consideration in understanding the future state of these sites is the growth and condition of the moose population on northern cape breton island. in addition to the absence of wolves (canis lupus) and low levels of predation by black bear (ursus americanus), disease and parasites present among moose populations in other regions, including parelaphostrongylus tenuis (brainworm) and dermacentor albipictus (winter tick), appear to exist at low levels in the cape breton population. this is likely because there is no significant range overlap with deer, and a function of moose preference for the higher elevations where they exist (telfer 1967, 1968); it may also reflect lack of formal monitoring (t. nette, nsdnr, personal communication). the most recent population estimate (2008) for the study area suggests that the moose population of northern cape breton island is declining (d. quann, cape breton highlands national park, unpublished data). population estimates from 2009 and 2010 will assist in confirming whether this decline is short-term or trending. an alternate stable state? the spruce budworm outbreak acting alone would not have resulted in the landmoose impacts on forest regeneration smith et al. alces vol. 46, 2010 146 scape described herein. while smith (2007) reported moderate densities of regeneration at the sapling stage, the severe levels of browsing and the low densities of trees indicate that the compounding effects of intense herbivory by moose may be giving rise to what paine et al. (1998) called an “ecological surprise.” while not studied to date, it seems likely that the moose population and its herbivory following the spruce budworm outbreak have been above the ecological carrying capacity for this region. given lower levels of herbivory by moose following the spruce budworm disturbance, we assume the forest would have progressed slowly through the reorganization of nutrients and energy to a mature balsam fir forest, coinciding with the concept of the budworm-fir cyclic succession model (baskerville 1975, maclean 1984, 1988) and model predictions for our study area (smith 1998). as browsing by moose has become more severe, it is arguable that the system has been, or will be, pushed beyond a critical threshold or a “catastrophic bifurcation” as described by scheffer and carpenter (2003), causing dramatic change in the successional trajectory of budworm affected area. asselin et al. (2006) suggested only 2 simple criteria for assessing whether an alternate stable state exists: 1) occurrence of the states under the same environmental conditions (sutherland 1974), and 2) persistence over time (connell and sousa 1983, jasinski and asselin 2004). two drastically different states are currently occurring side by side in northern cape breton in previously forested areas: forested and nonforested areas. factors such as a paucity of established trees on many sites and a potential expansion of calamagrostis may perpetuate the observed lack of forest development. management implications this study offers some insight into the current condition of budworm-affected areas in northern cape breton, and will assist managers in guiding future research activities; however, determining whether current conditions threaten ecological integrity as it was understood or simply imply new baseline conditions will be paramount. understanding the ecology of the disturbed sites is critical to informing decision-making regarding provincially endangered species such as the american marten (martes americana) and the canada lynx (lynx canadensis), both of which persist currently in small numbers in the area. by monitoring regeneration on sites with high levels of moose browsing and high abundance of balsam fir and white birch, an understanding of stand-level responses to browsing pressure over time may be gained. in particular, permanent sample plots would provide data to monitor the evolution of these sites over time. measures of moose population density and abundance with more longevity will be required to understand trends in browse-induced mortality as they relate to landscape pattern. given the amount of damage by moose to post-budworm regeneration, it is questionable whether even a drastic reduction in the moose population would prevent further regression of this landscape. the hardy, resilient nature of species such as white birch and balsam fir may be conducive to recovery, however, and as such, active management should not be precluded. we caution strongly that sites sampled in this study are almost certainly in the areas most affected by spruce budworm and being affected most severely by moose. we advocate for further research investigating the impact of moose on sites that were not affected by the spruce budworm during the last outbreak. our findings indicate northern cape breton island is currently an ecosystem where studying ecological responses to severe, compounding disturbances over time may yield increased knowledge of the dynamics of the spruce-fir forests of eastern canada. acknowledgements we acknowledge the valuable contribualces vol. 46, 2010 smith et al. moose impacts on forest regeneration 147 tions of geordon harvey and dawn allen who provided gis support, dave mealiea who assisted in the fieldwork, and james bridgland, cape breton highlands national park. funding for this project was provided by parks canada. references asselin, h., a. belleau, and y. bergeron. 2006. factors responsible for the cooccurrence of forested and unforested rock outcrops in the boreal forest. landscape ecology 21: 271-280. augustine, d., and l. e. frelich. 1998. white-tail deer impacts on populations of an understory forb in fragmented deciduous forests. conservation biology 12: 995-1004. baskerville, g. l. 1975. spruce budworm: super silviculturalist. forestry chronicle 51: 138-140. basquill, s., and r. g. thompson. 1996. moose (alces alces) browse availability and utilization in cape breton highlands national park. parks canada technical report, ecological science, 1200-3298; x. bergerud, a. t., and f. manual. 1968. moose damage to balsam fir-white birch forests in central newfoundland. journal of wildlife management 4: 729-746. brandner, t. a., r. o. peterson, and k. l. risenhoover. 1990. balsam fir on isle royale: effects of moose herbivory and population density. ecology 71: 155164. bridgland, j., and f. millette. 1995. resource description and analysis. cape breton highlands national park. ingonish beach, nova scotia, canada. burnham, k. p., and d. r. anderson. 1998. model selection and multi-model inference: a practical information-theoretic approach. springer-verlag inc., new york, new york, usa. cameron, a. w. 1958. mammals of the islands in the gulf of st. lawrence. national museum of canada bulletin 154. queens printer, ottawa, ontario, canada. cater, t. c., and f. s. chapin, iii. 2000. differential effects of competition or microenvironment on boreal tree seedling establishment after fire. ecology 81: 1086-1099. collins, e. h. 1951. a study of the boreal forest formation in northern cape breton island. b.a. thesis, acadia university, wolfville, nova scotia, canada. connell, j. h., and w. p. sousa. 1983. on the evidence needed to judge ecological stability or persistence. american naturalist 121: 789-824. connor, k., b. w. allard, t. dilworth, s. mahoney, and d. anions. 2000. changes in structure of a boreal forest community following intense herbivory by moose. alces 22: 111-131. coomes, d. a., r.b. allen, d. m. forsyth, and w. g. lee. 2003. factors preventing recovery of new zealand forests following control of invasive deer. conservation biology 17: 450-459. dublin, h. t., a. r. e. sinclair, and j. mcglade. 1990. elephants and fire as causes of multiple stable states in the serengeti-mara woodlands. journal of animal ecology 59: 1147-1164. eis, s. 1981. effects of vegetation competition on revegetation of white spruce. canadian journal of forest research 11: 1-8. fernow, b. e. 1912. forest conditions of nova scotia. commission of conservation canada: ottawa, ontario, canada. frelich, l., and p. b. reich. 1999. neighbourhood effects, disturbance severity, and community stability in forests. ecosystems 2: 151-166. grossnickle, s. c. 1988. conifer seedling establishment on boreal reforestation sites: environmental influences and ecophysiological responses. pages 71-78 in c.r. smith and r.j. reffle, editors. taking moose impacts on forest regeneration smith et al. alces vol. 46, 2010 148 stock: the role of nursery practice in forest renewal. ontario forestry research committee proceedings, 1417 sept. 1987, kirkland lake, ontario. cojfrc symposim o-p-16. hogg, e. h., and v. j. lieffers. 1991. the impact of calamagrostis canadensis on soil thermal regimes after logging in northern alberta. canadian journal of forest research 21: 387-394. jasinski, j. p., and h. asselin. 2004. alternate views on alternate stable states. forest ecology and environment 2: 10-11. _____, and s. payette. 2005. the creation of alternate stable states in the southern boreal forest. ecological monographs 75: 561-583. johnstone, g. no date. sas software to fit the generalized linear model. sas institute inc., cary, north carolina, usa. lepš, j., and p. šmilauer. 2003. multivariate analysis of ecological data using canoco. cambridge university press, cambridge, england. littell, r. c., g. a. milliken, w. w. stroup, and r. d. rolfinger. 1996. sas system for mixed models. sas institute inc., cary, north carolina, usa. macdonald, p. m. 1986. grasses in young conifer plantations – hindrance and help. northwest science 60: 271-277. maclean, d. a. 1984. effects of spruce budworm outbreaks on the productivity and stability of balsam fir forests. forestry chronicle 60: 273-279. _____. 1988. effects of spruce budworm outbreaks on vegetation, structure and succession of balsam fir forests on cape breton island, canada. pages 253-261 in m.a.werger, p. j. m. van der aart, h. j. during, and j. t. a. verhoeven, editors. plant form and vegetation structure. spb academic publishing, the hague, netherlands. _____, and d. p. ostaff. 1989. patterns of balsam fir mortality caused by an uncontrolled spruce budworm outbreak. canadian journal of forest research 19: 1087-1095. mccullagh, p., and j. a. nelder. 1989. monographs on statistics and applied probability. generalized linear models, 2nd edition. chapman and hall, london, england. mcinnes, p., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73: 2059-2075. mclaren, b. 1996. plant specific response to herbivory: simulated browsing of suppressed balsam fir on isle royale. ecology 77: 228-235. _____, b. a. roberts, n. djan-chékar, and k. p. lewis. 2004. effects of overabundant moose on the newfoundland landscape. alces 40: 44-59. _____, l. hermanutz, j. gosse, b. collet, and c. kasimos. 2009. broadleaf competition interferes with balsam fir regeneration following experimental removal of moose. forest ecology and management 257: 1395-1404. neff, d. j. 1968. the pellet-group census technique for big game trend, census and distribution: a review. journal of wildlife management 32: 597-613. neily, p. d., e. quiqley, l. benjamin, b. stewart, and t. duke. 2003. ecological land classification for nova scotia. vol. 1 – mapping nova scotia’s terrestrial ecosystems. nova scotia department of natural resources, renewable resources branch, shubenacadie, nova scotia, canada. ostaff, d. p., and d. a. maclean. 1989. spruce budworm populations, defoliation, and changes in stand condition during an uncontrolled spruce budworm outbreak on cape breton island, nova scotia. canadian journal of forest research 19: 1077-1086. alces vol. 46, 2010 smith et al. moose impacts on forest regeneration 149 paine, r. t., m. j. tegner, and e. a. johnson. 1998. compounded perturbations yield ecological surprises. ecosystems 1: 535-545. pardy, a. b. 1997. forest succession following a severe spruce budworm outbreak at cape breton highlands national park. msc thesis, university of new brunswick, fredericton, new brunswick, canada. pastor, j., b. dewey, r. naiman, p. mcinness, and y. cohen. 1993. moose browsing and soil fertility in the boreal forests of isle royale national park. ecology 74: 467-480. _____, and d. kjell. 2003. moose-vegetationsoil interactions. alces 39: 177-192. _____, and r. naiman. 1992. selective foraging and ecosystem processes in boreal forests. american naturalist 139: 690-705. payette, s., n. bhiry, and m.simard. 2000. origin of the lichen woodland at the southern range of its limit. canadian journal of forest research 30: 288-306. pulsifer, m. d., and t. n. nette. 1995. history, status and present distribution of moose in nova scotia. alces 31: 209-219. radeloff, v. c., d. j. mladenoff, and m. s. boyce. 2000. effects of interacting disturbances on landscape patterns: budworm defoliation and landscape pattern. ecological applications 10: 233-247. rao, c. r. 1964. the use and interpretation of principal component analysis in applied research. sankhya a 26: 329-358. risenhoover, k. l., and s. a. maass. 1987. the influence of moose on the composition and structure of isle royale forests. canadian journal of forest research 17: 375-364. sas institute. 2003. sas version 9.1. sas institute, cary, north carolina, usa. scheffer, m., and s. r. carpenter. 2003. catastrophic regime shifts in ecosystems: linking theory to observation. trials in ecology and evolution 18: 648-656. smith, c. 2007. the impact of moose on forest regeneration following disturbance by spruce budworm in the cape breton highlands, nova scotia, canada. mes thesis, dalhousie university, halifax, nova scotia, canada. smith, m. j. 1998. an examination of forest succession in the cape breton highlands of nova scotia. msc. thesis, university of new brunswick, fredericton, new brunswick, canada. snyder, j. d., and r. a. janke. 1976. impact of moose browsing on boreal-type forests of isle royale. american midland naturalist 95: 79-92. spss inc. 2005. spss base 14.0 for windows. spss inc. chicago, illinois, usa. sutherland, j. p. 1974. multiple stable points in natural populations. american naturalist 108: 859-873. telfer, e. s. 1967. comparison of moose and deer winter range in nova scotia. journal of wildlife management 31: 418-425. _____. 1968. distribution and association of moose and deer in central new brunswick. transcripts of the new england wildlife society 25: 41-70. ter braak. c. j. f. 1995. ordination. pages 91-173 in r. h. g. jongman, c. j. e ter braak, and 0. f. r. van tongeren, editors. data analysis in community and landscape ecology. new edition. cambridge university press, cambridge, england. _____, and p. šmilauer. 2002. canoco reference manual and canodraw for window’s use guide. software for canonical community ordination (version 4.54), microcomputer power, new york, new york, usa. thompson, i. d., w. j. curran, j. a. hancock, and c. e. butler. 1992. influence of moose browsing on successional forest growth on black spruce sites in newfoundland. forest ecology and management moose impacts on forest regeneration smith et al. alces vol. 46, 2010 150 47: 29-37. timmerman, h. r., and m. e. buss. 1997. population harvest and management. pages 559-616 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington dc, usa. wilson, j. b., and a. d. q. agnew. 1992. positive feedback switches in plant communities. advances in ecological research 23: 263-336. zinck, m. 1998. roland’s flora of nova scotia. nimbus publishing and nova scotia museum, halifax, nova scotia, canada. alces 31_87.pdf alces32_1.pdf alces 31_209.pdf alces 31_139.pdf alces34(1)_117.pdf alces 31_255.pdf alces35_165.pdf alces30_21.pdf alces30_65.pdf alces30_45.pdf alces34(1)_165.pdf alces30_159.pdf alces30_109.pdf alces35_135.pdf alces36_53.pdf alces34(2)_347.pdf alces29_1.pdf alces35_213.pdf alces36_147.pdf alces34(1)_21.pdf alces29_99.pdf alces36_85.pdf alces37(2)_497.pdf alces29_149.pdf alces34(2)_459.pdf alces34(2)_409.pdf alces32_101.pdf alces32_149.pdf alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 135 compositional analysis of moose habitat in northeastern minnesota mark s. lenarz1, robert g. wright2, michael w. schrage3, and andrew j. edwards4 1minnesota department of natural resources, forest wildlife populations and research group, 1201 east highway 2, grand rapids, mn 55744; 2minnesota department of natural resources, 5463-c west broadway, forest lake, mn 55025; 3fond du lac resource management division, 1720 big lake road, cloquet, mn 55720; 41854 treaty authority, 4428 haines road, duluth, mn 55811. abstract: it is well accepted that moose (alces alces) often use early successional habitats in the boreal forest. it is not clear, however, whether use of disturbed habitats represents a preference or simply that moose are more detectable. previous research based on visual observations assumed that moose were equally detectable in all cover types. we evaluated habitat selection of moose in northeastern minnesota using telemetry locations and land-use land-cover (lulc) type information. we calculated measures of habitat selection within their home range (third-order) and selection of habitats to create a home range (second-order) using a technique called composition analysis. the analyses indicated that the cutover cover type ranked highest in summer and winter in both secondand thirdorder selection and its rank generally was significantly higher than most other cover types. selection for aquatic habitats during the summer was not evident in our analysis. cover types that could provide lower operative temperatures from shade ranked higher than aquatic cover types. inferences from these analyses should be treated with caution because of inherent weaknesses of use-availability analyses and biases associated with vhf telemetry locations. alces vol. 47: 135-149 (2011) key words: alces, home range, habitat preference, composition analysis, land-use land-cover cutover, aquatic habitat. habitat management for moose (alces alces) in the boreal forest is predicated on the inference that early successional habitats created by disturbance (e.g., forest fire, logging, and wind or insect damage) are preferred habitats and that they result in higher moose numbers (aldous and krefting 1946, spencer and chatelain 1953, krefting 1974, peek et al. 1976). while it is clear that moose often use disturbed habitat, it is not clear whether this use represents a preference or simply that moose are more detectable in this cover type. early successional habitats tend to be more open which can increase the visibility of moose and present a bias. it is likely that the high quality forage available in early successional habitats should benefit moose if the population is food limited. however, even with good hypothetical explanations of why moose would benefit from this association, until this preference can be demonstrated empirically, continued repetition has the potential of leading to dogmatic thinking (romesburg 1981). to more clearly evaluate the importance of disturbance to moose habitat, analyses require data in which animals have an equal probability of being detected in all habitats. aquatic habitat has also been noted as preferred habitat in much of the boreal forest because moose are commonly observed feeding in ponds and lakes from late may to midaugust (murie 1934, peterson 1955, devos 1956). while not ubiquitous in the boreal forest, most authors suggest that these habitats are preferred (peek 1997). as with early successional habitats, aquatic habitats tend to be more open resulting in a higher detection rate for moose. it has been reasoned that moose use compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 136 these habitats for the high quality forage, the sodium content of this forage, insect relief, and possibly for relief from warmer temperatures (ritcey and verbeek 1969, peek et al. 1976, jordan 1987, peek 1997). cover for shade may also be important to moose in summer. based on metabolic research, renecker and hudson (1986) indicated that moose are intolerant of heat but superbly adapted to cold and that summer temperatures may well define their southerly distribution. recent research has suggested that warming temperatures associated with climate change may be linked to the increased mortality observed in moose populations found along the southern edge of their distribution (murray et al. 2006; lenarz et al. 2009, 2010). microclimate can vary substantially in different cover types (ackerman 1987, schwab and pitt 1991, demarchi and bunnell 1993, dussault et al. 2004, mcgraw 2011) and moose shift to more shaded habitats as temperatures rise above some metabolic threshold (ackerman 1987). habitat selection can take place at several different scales. the geographic range of a species, for example, represents first-order selection (johnson 1980), and second-order selection determines the home range of an individual or social group within that range. the use of various habitat components within the home range represents third-order. based on the preceding interpretations, moose should select summer home ranges that contain early successional habitats and aquatic habitat for food and shaded habitat to prevent overheating. during the winter, moose should select primarily for early successional habitats as a source of food. a clear understanding of how moose utilize their habitat is needed for management. this information is difficult to determine entirely from direct observations which are subjected to visibility biases. even during winter under ideal snow and light conditions, visual obstruction due to forest cover results in many moose being missed (e.g., roughly 50% of moose were missed during helicopter surveys in minnesota; giudice et al. 2011, lenarz, unpubl. data). unbiased determination of habitat utilization requires the use of vhf or gps telemetry. in the following we report from results of a vhf telemetry study in northeastern minnesota. although the study was designed to document annual adult moose mortality, the telemetry locations of radioed individuals should provide insights into moose habitat use in northeastern minnesota. the primary objective of this analysis was to determine whether this regional moose population displayed third-order seasonal selection for specific cover types. we also examined whether moose exhibited second-order seasonal selection relative to the cover types available in northern minnesota. methods study area the 3,953 km2 study area was defined by telemetry locations for moose in northeastern minnesota (47°30’n, 91º21’w; fig. 1). the forests were transitional between canadian boreal forests and northern hardwood forests common further south (pastor and mladenoff 1992). wetlands, including bogs, swamps, small to medium-sized lakes, and small streams are interspersed throughout. the study area is on a low plateau that rises abruptly from lake superior and reaches about 700 m above sea level (heinselman 1996). a northeastsouthwest continental divide runs down the middle of the plateau with water flowing southeast into lake superior or northwest into hudson bay. the study area was primarily a mosaic of mixed wood (39%), conifer (23%), and bog (11%, table 1) classified as the northern superior upland section (minnesota department of natural resources [mndnr] 2007). the main conifer species were northern white cedar (thuja occidentalis), black spruce (picea alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 137 mariana), and tamarack (larix laricina) on the lowlands and balsam fir (abies balsamea), white spruce (picea glauca), and jack (pinus banksiana), white (p. strobus), and red pine (p. resinosa) on the uplands. deciduous species, primarily quaking aspen (populus tremuloides) and white birch (betula papyrifera), occurred on the uplands in hardwood stands or were intermixed with the conifers. the majority of land within the study area (74%) fell within the superior national forest with the balance in state, county, or private ownership. the area was sparsely inhabited and communities within the study area contained <100 permanent residents. hunting is restricted by permit and <2% of the estimated population is harvested annually (mndnr 2011). july is the warmest month in the study area with an average high temperature of 26° c and january the coldest with an average high temperature of -10° c (noaa 20012010). total annual precipitation averaged 71 cm with 55% occurring between june and september. precipitation usually occurred as snow between late october and mid-april and snow sometimes accumulated >100 cm (noaa 2001-2010). study animals and telemetry we captured and handled moose according to methods described by lenarz et al. (2009). between 2002 and 2008, we captured adult male and female moose (≥1.7 yr old) in early february or march by helicopter net-gunning (2002, wildlife capture services, marysvale, fig.1. study area used in composition analysis of moose in northeastern minnesota, usa, 2002-2010. the large polygon represents available habitat for second-order analysis. small open polygons represent 95% mcp home ranges during the winter; cross-hatched polygons represent 95% mcp home ranges during the summer. compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 138 utah) or darting (2003-2005 and 2008, quicksilver air, inc., fairbanks, alaska). we fitted each moose with a vhf radiocollar (advanced telemetry systems, isanti, minnesota). from february 2002-may 2009 we located each moose approximately once weekly (x = 7.7 days, se = 0.177, n = 346) from fixed-wing aircraft and recorded a gps location. moose were observed visually on ~28% of the locations. animal capture and handling protocols met the guidelines recommended by the american society of mammalogists (gannon et al. 2007). the use of point data that ignores locational error increases the risk of drawing erroneous conclusions about relative preference (retti and mclouglin 1999). to identify error, we conducted blind tests of locational error by positioning test collars in trees throughout the study area. the error distances were right-skewed, with a few large values (n = 50, median = 254 m, g1 = 3.6) best represented by a log normal distribution. we estimated parameters of this distribution using maximum likelihood with built in non-linear optimizers in program r (r development core team 2006). because mean patch size (table 1) was generally quite small relative to the telemetry error, there was an increased probability of habitat misclassification at the telemetry location. as recommended by samuel and kenow (1992), we used a subsampling approach in which the error distribution (log normal) was used to simulate a subsample of all possible points surrounding each telemetry location. this procedure reduces the bias associated with misclassification (samuel and kenow 1992). for each telemetry location we simulated 100 points using a random angle between 0 and 360° and a random distance chosen from the lulc cover data cover type proportion mean patch size (ha) final cover type proportion mixed-wood forest mixed forest 38.60% 6.1 mixed 38.60% conifer forest conifer forest 22.50% 2.7 conifer 22.50% wetlands bogs bog 11.20% 6.8 bog 11.20% regeneration/young forest cutover 6.90% 3.3 cutover 6.90% open water water 6.10% 4.4 aquatic 9.50% deciduous forest deciduous forest 6.00% 6.9 deciduous 9.00% marshes and fens marsh 3.40% 1.8 aquatic shrubby grassland shrub 0.20% 6.3 other 2.30% grassland shrub other gravel pits and open mines other 0.30% 0.5 other bare rock other other cultivated land other other farmsteads other other other rural development other other urban/industrial other other hardwood regen.* 3.00% 0.5 deciduous blowdown* 1.30% 1.2 other conifer regen.* 0.50% 0.2 other table 1. land use land cover (lulc) data, cover types used in the analyses, and the proportion of each cover type within the study area in northeastern minnesota. *data from p. t.wolters (unpublished data). alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 139 fitted log normal distribution. because of small patch size (table 1) and potential gps error (moose occasionally moved in response to the aircraft), we used this approach even when moose were observed visually. habitat availability and use we classified telemetry locations according to season with locations dated 1 october-31 march assigned as “winter” and the remainder of the year as “summer.” while these dates were not ideal from a life history standpoint, they did maximize the number of radioed moose that could be included in the seasonal analyses. we censored moose which dispersed out of the study area or displayed distinct migratory behavior from further analysis. we calculated 95% minimum convex polygon (mcp) seasonal home ranges based on telemetry points (not including simulated points) for each moose using home range tools for arcgis (rodgers et al. 2007). this procedure was used to identify outlier points (1 to 3 per seasonal mcp) by calculating the mean of all x and y coordinates for an individual moose and then selecting 95% of the points closest to this mean (rodgers et al. 2007). we then censored moose with fewer than 30 seasonal telemetry points. we tested for differences in mcp size using a linear fixed effects model using program r (r development core team 2006). for the third-order analyses, we used the cover type at the telemetry and simulated points to calculate proportional habitat “use” for each moose in each season. we calculated the seasonal mcp home range for each moose based on telemetry locations and simulated points. the proportion of each cover type within an mcp represented “availability” for an individual moose. for the second-order analyses, we used the proportion of each cover type within a seasonal home range as “use” (availability in the third-order analyses). we defined “availability” as the proportion of each cover type within a polygon calculated as the mcp from the pooled points (both seasons, telemetry, and simulated) of all moose included in the analysis (fig. 1). hereafter, this polygon that represents “availability” for both seasonal second-order analyses (fig. 1) is referred to as the study area. to classify habitat we created a vegetative cover layer using a land use-land cover (lulc) raster layer provided by mndnr (1998a, b). this source layer was derived from landsat 30-meter thematic satellite imagery dated summers 1991-1996, divided into 16 lulc classes based on imagery dated 1995 and 1996, then further processed to replace transportation class values with surrounding values (mndnr 1998a, b). overall classification accuracy was assessed at >95% (mndnr 1998b). in arcgis 9.2 (environmental systems research institute, redlands, california, usa) we used the reclassify command to combine these 15 classes into 9 classes (table 1). forest disturbance data (unpublished) obtained from p. t. wolters (university of minnesota, duluth) were comprised of 3 classes of regenerating forests: deciduous and coniferous logging activities (1975-2000) and blowdown that occurred during a 1999 windstorm. incorporating these data using the mosaic command with the last option allowed us to further refine the vegetative cover layer. finally, we used the arcgis spatial analyst extract values to points command to identify the cover type at each telemetry location and simulated point. the study area was composed primarily of the mixed, conifer, and bog cover types and they made up >72% of the total area (table 1). the mixed cover type had a mature canopy that was composed of approximately equal amounts of hardwood and conifer species. forest with a canopy ≥67% conifer species was classified as conifer cover type and was primarily upland conifers including balsam fir and jack, white, and red pine. bog was characterized as peat lands with a high water table compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 140 and varying amounts of tree cover, typically white cedar, black spruce, or tamarack. the deciduous forest included areas with at least 67% canopy cover of woody deciduous species, primarily white birch and quaking aspen. cutover represented areas where commercial timber was removed in 1980-1995. the marsh cover type was composed of grassy, wet areas with standing or slowly moving water. water covered about 6% of the area and represented permanent water bodies such as lakes, rivers, and ponds. the other cover type represented an amalgam of covers such as bare rock, gravel pits, and rural development and represented only 0.3% of the study area. habitat selection we used compositional analysis (aebischer et al. 1993) to determine seasonal secondand third-order selection of 12 cover types by moose. this technique uses the individual animal as the sampling unit and compares proportional habitat use with proportional habitat availability. hypothesis testing of the nonstandard multivariate data is done using manova/mancova-type linear models (aitchison 1986). aebischer et al. (1993) indicated that a minimum of 6 animals were necessary for statistical inferences; we combined sexes because we had data on 26 females but only 3 males (table 2). missing data occurred in 2 situations: 1) a particular habitat was available but not used, and 2) a particular habitat was not available for use by an individual. in the first case, aebischer et al. (1993) recommended that zero use should be replaced with a small positive value, less than the smallest recorded nonzero proportion. in the first case, bingham and brennan (2004) and bingham et al. (2007) found that replacing zero use with a small positive value increased the potential for type i error and could lead to a systematic bias. the authors recommended reclassifying habitat categories so that no habitat categories with 0% use are included in the analysis. in our analysis, we combined cover types with no use to the most similar type with use. hardwood regeneration was added to deciduous (hereafter termed deciduous), conifer regeneration, shrub, and blowdown were added to other, and water was added to marsh to form the aquatic category (table 1). in the second case, aebischer et al. (1993) indicated that one approach was to delete the animal but cautioned the resultant loss of information could induce bias. because specific cover types were not available for 3 moose (1 moose had no cutover and 3 moose had no other cover type available), we dropped these individuals from the analyses. analyses were conducted in sas using the program bycomp.sas (ott and hovey 1997). results home range size as measured with the 95% mcp (not including simulated points) varied according to season and sex with the largest home ranges occurring among males and during winter (table 2). only 4% of the 90 radioed moose (for which we had ≥30 telemetry locations) in northeastern minnesota were classified as migratory (distinctly separated seasonal ranges); an additional 2% exhibited dispersal (permanent one-way shifts) and 18% displayed exploratory movements (returned to original home range) of up to 77 km. in female moose, the winter home range averaged 30.4 km2 (se = 6.0, n = 12) and during the summer 13.6 km2 (se = 1.9, n = 23). average annual home range for females was 29.9 km2 (se = 3.3, n = 26). in male moose, the winter home range averaged 44.2 km2 (se = 13.3, n = 3), during the summer 29.4 km2 (se = 1.6, n = 3) and annual home range averaged 58.0 km2 (se = 15.1, n = 3). home range size differed between seasons (t = 4.036, p = 0.002) and between sexes (t = -2.417, p = 0.002). third-order selection seasonal cover type selection was not random. during winter male and female alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 141 moose used proportionately more cutover and mixed cover types than was available in their winter home ranges (fig. 2). proportional use of bog, conifer, and other cover types was close to the proportion available. proportional use of deciduous and aquatic cover types was substantially less than the proportion available. during summer proportional use of the cutover cover type was substantially higher than available in the summer home range, and use of conifer was slightly higher than availability (fig. 2). use of mixed, deciduous, aquatic, summer (n = 26) winter (n = 15) annual (n = 29) moose id sex # locations 95% mcp (km2) # locations 95% mcp (km2) # locations 95% mcp (km2) 41430 f 43 18.9 47 72.2 86 69.9 41530 f 53 26.2 47 31.0 95 55.4 41620 m 39 32.2 42 70.8 77 87.8 41720 f 37 7.5 62 42.2 42030 f 43 13.5 66 42.7 43640 f 31 17.5 52 19.6 43840 f 30 36.1 61 64.3 44620 f 60 17.5 52 13.3 107 21.9 45020 f 46 8.3 47 12.3 89 13.0 45530 m 35 26.6 30 31.9 62 46.6 45830 f 35 9.3 59 36.1 46530 f 31 8.7 49 13.5 46740 f 52 23.7 50 70.1 97 47.0 47140 m 42 29.4 47 29.9 85 39.5 47522 f 36 21.9 30 13.1 61 17.1 48310 f 37 12.8 64 17.9 48430 f 32 9.3 48 16.4 49120 f 49 16.7 41 30.7 86 35.2 49430 f 32 6.0 36 23.5 65 18.3 49640 f 30 13.8 52 18.1 49820 f 34 8.0 51 8.9 50550 f 31 9.8 49 18.2 51120 f 31 9.7 50 20.1 51830 f 38 11.1 62 19.8 52030 f 58 15.8 52 21.8 105 26.0 52420 f 37 9.3 57 22.6 52630 f 33 19.8 58 52.4 52933 f 34 12.6 55 32.5 53440 f 32 26.4 62 27.2 mean 40 15.5 41 33.1 68 32.8 table 2. annual and seasonal home range size (95% mcp) of moose in northeastern minnesota, usa, 2002-2009. compositional analyses were conducted only on the seasonal data. compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 142 and bog cover types was slightly lower than the proportion available. proportional use of the other cover type was substantially lower than what was available. compositional analysis indicated that during winter cutover had the highest rank among the 7 cover types, and the aquatic cover type had the lowest rank (table 3). there was no detectable difference between proportional use of cutover and all of the remaining cover types except deciduous and aquatic, implying that the order of their assigned ranks (among cutover, mixed, bog, conifer, and other) means little. the proportional use of the aquatic cover type was significantly less than any other cover types except other. during summer cutover cover was used significantly more than all other cover types, and use of other was significantly less than all other cover types. second-order selection analyses indicated the home ranges selected by northeastern moose were not random. in winter the average home range contained a substantially higher proportion of cutover and a lower proportion of conifer and other cover types than was available in the study area (fig. 3). the other cover types were present in about the same proportion as their availability. during summer the average home range contained more cutover and bog than available, and both mixed and deciduous cover types were found in proportions similar to their availability (fig. 3). there was less conifer, aquatic, and other cover types in the average summer home range than was found in the summer study area. compositional analysis indicated that cutover again had the highest rank in winter home ranges, and was significantly higher than all other cover types (table 4). in contrast, other had the lowest rank and its presence was significantly less than all other cover types. a similar pattern existed in summer; cutover had the highest rank and was significantly higher than all other cover types except bog, whereas other had the lowest rank and was significantly lower than cutover, bog, mixed, and conifer cover types. discussion the area of mcp seasonal and annual home ranges for moose in northeastern minnesota were comparable to those reported for non-migratory moose found elsewhere in north america (hundertmark 1997), but larger than those of moose in northwestern minnesota fig. 2. third-order selection (i.e., comparing cover type composition of simulated locations to the cover type composition of the individual home range) by moose in northeastern minnesota, usa, 2002-2010. habitats are arranged from most to least preferred for winter and summer. gray bars indicate mean proportional utilization and white bars represent mean proportional habitat available. alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 143 (phillips et al. 1973). moose habitat in northwestern minnesota was highly fragmented by agricultural development and smaller home ranges might result if moose are restricted to smaller patches of habitat. phillips et al. (1973) also found that approximately 20% of northwestern minnesota moose migrated between seasonal ranges which contrasts with the 4% we identified in this study in northeastern minnesota. the inference that early successional habitats are preferred by moose was reinforced by our analyses. in both the secondand thirdorder analyses, cutover was ranked highest among the 7 cover types in both seasons (tables 3 and 4). in the third-order analysis the mean proportional use of cutover exceeded its availability and in the second-order analysis, home ranges contained a higher proportion of cutover than was available in the study area (fig. 2 and 3). while early successional habitats may be preferred, these results do not necessarily imply that the availability of cutover results in higher moose numbers. moose populations limited by predation, hunting, and or disease likely will not benefit from an abundance of early successional forest. selection for aquatic habitats during summer was not evident in our analysis. in both the secondand third-order analyses, the mean proportional use of the aquatic cover type was less than its availability. in the thirdorder analysis, the rank of the aquatic cover type was significantly lower than cutover and conifer (table 3). in the second-order analysis, the rank of the aquatic cover type was significantly lower than that for cutover, bog, and mixed (table 4). open water accounted for 6.1% of the area (table 1) and included a few larger lakes up to 95 km2. if the limnetic zone (open water beyond most plant growth) were excluded from both the secondand third-order analyses, the availability of the aquatic cover type would decline and its rank would likely increase. the width of the littoral zone (open water containing plant growth) varies widely in the study area and was dependent in part on lake size, depth, and winter cutover mixed bog conifer deciduous other aquatic rank cutover * 0.745 0.770 0.184 0.004 0.060 0.001 1 mixed 0.745 * 0.991 0.184 0.550 0.050 0.001 2 bog 0.770 0.991 * 0.242 0.111 0.046 0.001 3 conifer 0.184 0.184 0.242 * 0.207 0.132 0.001 4 deciduous 0.004 0.550 0.111 0.207 * 0.699 0.029 5 other 0.060 0.050 0.046 0.132 0.699 * 0.053 6 aquatic 0.001 0.001 0.001 0.001 0.029 0.053 * 7 summer cutover conifer mixed deciduous aquatic bog other rank cutover * 0.001 0.001 0.001 0.001 0.001 0.001 1 conifer 0.001 * 0.246 0.274 0.007 0.043 0.001 2 mixed 0.001 0.246 * 0.618 0.095 0.184 0.001 3 deciduous 0.001 0.274 0.618 * 0.480 0.434 0.001 4 aquatic 0.001 0.007 0.095 0.480 * 0.796 0.001 5 bog 0.001 0.043 0.184 0.434 0.796 * 0.001 6 other 0.001 0.001 0.001 0.001 0.001 0.001 * 7 table 3. simplified ranks and randomized p values of third-order selection of moose in northeastern minnesota, usa, 2002-2010. gray cells represent p < 0.05. the order of cover types represents rank with cutover having the highest rank compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 144 substrate; hence, it would be very difficult to accurately parameterize this sub-type for reanalysis. most authors (peterson 1955, devos 1958, ritcey and verbeek 1969, peek et al. 1976) document that use of aquatic habitat is concentrated from june until august. because our analysis treated summer as a much longer period (1 april-30 september), selection for the aquatic cover type may have been eclipsed by selection for other cover types during the remainder of this season. traditionally, shaded habitats have been viewed as thermal habitat during summer (parker and gillingham 1990, demarchi and bunnell 1993, cook et al. 1998, 2004) because shade will reduce solar radiation and the resultant operative temperature (campbell and norman 1998). the mature component of the conifer, mixed, and deciduous cover types could produce such shade, and aquatic and bog cover types could also act as a thermal refuge because bedding in a wet bog or immersion in water could reduce body temperature through conduction and evaporation. in the third-order analysis (summer) the mean proportional use of conifer exceeded availability while use of mixed, deciduous, aquatic, and bog cover types was slightly less than their availability (fig. 2). conifer was ranked significantly higher than either aquatic or bog, however, there were no significant differences among the ranks for mixed, deciduous, aquatic, and bog cover types. assuming that moose behaviorally reduce their thermal load through habitat selection, it appears that moose prefer conifer habitat over other cover types. presumably, if we could have restricted analysis to days when temperature was above the thresholds identified by renecker and hudson (1986), there might have been a more refined preference of one cover type over another. moreover, if use of aquatic and bog habitats is limited to early morning and evening, our analysis would obscure the importance of these cover types because most of the telemetry locations occurred during mid-day. the second-order analysis suggests that moose selection of home ranges is not random. during summer moose selected home ranges containing a higher proportion of cutover and bog than their availability; conversely, conifer, aquatic, and other cover types had significantly lower preference ranks. during winter cutover again had a significantly higher rank and the other cover type had a fig. 3. second-order selection (i.e., comparing cover type composition of individual home ranges to the cover type composition of the study area) by moose in northeastern minnesota, usa, 2002-2010. habitats are arranged from most to least preferred for winter and summer. gray bars indicate mean proportional utilization and white bars represent mean proportional habitat available. alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 145 significantly lower rank. the cutover cover type identified in the lulc layer represented regenerating forest that was 7-22 years old at the beginning of our research (2002). research has suggested that moose benefit most from regenerating forest when it is 11-30 years old (kelsall et al. 1977, schwartz and franzmann 1989). we attempted to refine our analysis of cover type selection by including areas of hardwood regeneration, conifer regeneration, and blow down from a gis layer provided by wolters (unpubl. data). although these data covered a wider span of time than the lulc data (19752000 vs. 1980-1995), they represented only a small subset of the overall cutover cover type and represented a small proportion of the home ranges and study area. ultimately we needed to combine them with other cover types to prevent type i error (bingham and brennan 2004). problems exist in most analyses of habitat selection because of biases associated with the measurement of habitat use and availability and the interpretation of results (garshelis 2000). in our third-order analysis, we attempted to improve the accuracy of our measures of habitat use by incorporating telemetry locations as well as simulated points to account for telemetry error. nonetheless, recent research indicated that techniques used to incorporate telemetry error were inherently inaccurate with patch sizes < 20 ha (cf. table1, montgomery et al. 2010). unless patches consistently encapsulate the telemetry polygon, it is questionable whether resource use studies such as third-order selection can be accurate (saltz 1994, rettie and mcloughlin 1999). analyses of second-order selection, however, are not dependent on the accuracy of telemetry locations but rather on the characterization of the home ranges used by individuals and the availability of important habitats. garshelis (2000:123) argued that while “animals may be able to choose borders that encompass the best mix of habitats from what exists on the landscape; they cannot alter the mix to suit their needs.” the 7 cover types winter cutover bog mixed conifer aquatic deciduous other rank cutover * 0.001 0.001 0.002 0.001 0.002 0.001 1 bog 0.001 * 0.777 0.271 0.254 0.296 0.001 2 mixed 0.001 0.777 * 0.453 0.230 0.107 0.001 3 conifer 0.002 0.271 0.453 * 0.471 0.645 0.001 4 aquatic 0.001 0.254 0.230 0.471 * 0.927 0.002 5 deciduous 0.002 0.296 0.107 0.645 0.927 * 0.001 6 other 0.001 0.001 0.001 0.001 0.002 0.001 * 7 summer cutover bog mixed conifer deciduous aquatic other rank cutover * 0.405 0.023 0.001 0.012 0.002 0.001 1 bog 0.405 * 0.325 0.005 0.122 0.002 0.001 2 mixed 0.023 0.325 * 0.019 0.029 0.001 0.001 3 conifer 0.001 0.005 0.019 * 0.794 0.097 0.001 4 deciduous 0.012 0.122 0.029 0.794 * 0.393 0.058 5 aquatic 0.002 0.002 0.001 0.097 0.393 * 0.125 6 other 0.001 0.001 0.001 0.001 0.058 0.125 * 7 table 4. simplified ranks and randomized p values of second-order selection of moose in northeastern minnesota, usa, 2010. gray cells represent p < 0.05. the order of cover types represents rank with cutover having the highest rank. compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 146 used in this analysis were fairly ubiquitous and patches within the study area were typically quite small (table 1). that only 3 moose were eliminated from the summer analysis because specific cover types were not available implies that virtually every moose had access to all cover types and was able to create a home range containing the most important habitat. that these 3 moose existed in home ranges not containing all 7 cover types further implies that none of the cover types is critical to moose survival. management implications the primary objective of this analysis was to determine whether moose in northeastern minnesota display third-order selection for specific cover types previously identified as important – cutovers, aquatic habitat, and summer shade. while our results may indicate that the cutover cover type is important, it is dangerous to make precise management prescriptions based on use-availability information because of inherent biases. the preferences identified here tend to represent fairly gross patterns and our results do not indicate anything about the relative size of cutover areas, the juxtaposition of other cover types, or specific resources within the cover type (e.g., forage) important to moose. moreover, osko et al. (2004) demonstrated that preferences are not fixed but change as the relative abundance of available habitat changes. finally, any demonstration of preference identified in these results does not imply that any of these habitats are critical or that they are relevant to population productivity (balsom et al. 1996, osko et al. 2004). it is important to remember that this study was not designed to evaluate habitat preferences in moose of northeastern minnesota. rather, it was designed to determine annual levels of adult mortality. as a result, the sample of vhf telemetry locations for each moose was generally small and the timing was generally limited to mid-day. clearly, a study that incorporates gps locations with much larger sample sizes would have been preferable. nonetheless, the general preferences observed tend to support selection for (not a requirement for) early successional habitat. acknowledgments the mndnr, the fond du lac band of lake superior chippewa, and the 1854 treaty authority provided funding and field support for this research. the u. s. geological survey, northern prairie wildlife research center provided in-kind support. the u. s. fish and wildlife service’s tribal wildlife grants program provided additional funding. we thank usgs research biologist m. nelson for countless hours locating moose from the air. we also thank mndnr pilots a. buchert, j. heineman, and d. murray for logistical support and j. fieberg for analysis of telemetry error and statistical advice. we thank gordon easton and an anonymous reviewer for their comments. references ackerman, t. n. 1987. moose response to summer heat on isle royale. m.s. thesis. michigan technological university, houghton, michigan, usa. aebischer, n. j., p. a robertson, and r. e. kenwood. 1993. compositional analysis of habitat use from animal radio-tracking data. ecology 74: 1313-1325. aitchison, j. 1986. the statistical analysis of compositional data. chapman and hall, london, england. aldous, s. e., and l. w. krefting. 1946. the present status of moose on isle royale. transactions of the north american wildlife conference 11: 296-308. balsom, s., w. b. ballard, and h. a. whitlaw. 1996. mature coniferous forest as critical moose habitat. alces 32: 131140. bingham, r. l., l. a. brennan, and b. m. ballard. 2007. misclassified resource alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 147 selection: compositional analysis and unused habitat. journal of wildlife management 71: 1369-1374. _____, and _____. 2004. comparison of type i error rates for statistical analyses of resource selection. journal of wildlife management. 68: 206-212. campbell, g. s., and j. m. norman. 1998. an introduction to environmental biophysics. springer-verlag. new york, new york, usa. cook, j. g., l. l. irwin, l. d. bryant, r. a. riggs, and j. w. thomas. 1998. relations of forest cover and condition of elk: a test of the thermal cover hypothesis in summer and winter. wildlife monographs 141. _____, _____, _____, _____, and _____. 2004. thermal cover needs of large unuglates: a review of hypothesis tests. transactions of the 69th north american wildlife and natural resources conference 69: 708-726. demarchi, m. w., and f. l. bunnell. 1993. estimating forest canopy effects on summer thermal cover for cervidae (deer family). canadian journal of forestry research 23: 2419-2426 de vos, a. 1956. summer studies of moose in ontario. transactions of the north american wildlife conference 21: 510-525. dussault, c., j. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioural responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321-328. gannon, w. l., r. s. sikes, and the animal care and use committee of the american society of mammalogists. 2007. guidelines of the american society of mammalogists for the use of wild mammals in research. journal of mammalogy 88: 809-823. garshelis, d. l. 2000. delusions in habitat evaluation: measuring use, selection, and importance. pages 111-164 in l. boitani and t. k. fuller, editors. research techniques in animal ecology: controversies and consequences. columbia university, new york, new york, usa. giudice, j. h., j. r. fieberg, and m. s. lenarz. 2011. spending degrees of freedom in a poor economy: a case study of building a sightability model for moose in northeastern minnesota. journal of wildlife management 75: in press. heinselman, m. 1996. the boundary waters wilderness ecosystem. university of minnesota press, minneapolis, minnesota, usa. hundertmark, k. j. 1997. home range, dispersal, and migration. pages 303-335 in a. w. franzman and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., usa. johnson, d. h. 1980. the comparison of usage and availability measurements for evaluating resource preference. ecology 61: 65-71. jordan, p. a. 1987. aquatic foraging and the sodium ecology of moose: a review. swedish wildlife research (supplement) 1: 119-137. kelsall, j. p., e. s. telfer, and t. d. wright. 1977. the effects of fire on the ecology of the boreal forest, with particular reference to the canadian north: a review and selected bibliography. occasional paper 323. canadian wildlife service, ottawa, ontario, canada. krefting, l. w. 1974. moose distribution and habitat selection in north america. le naturalist canadien 101: 81-100. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management. 74: 1013-1023. _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management compositional analysis of moose habitat lenarz et al. alces vol. 47, 2011 148 73: 503-510. mcgraw, a. m. 2011. characteristics of post-parturition areas of moose and effective temperature of cover types in moose home ranges in northeastern minnesota. m.s. thesis. university of minnesota, duluth, minnesota, usa. minnesota department of natural resources (mndnr). 1998a. landsat-based land use-land cover (raster). minnesota department of natural resources, st. paul, minnesota, usa. < http://deli.dnr.state.mn.us/ metadata.html?id=l250000120604> (accessed april 2011). _____. 1998b. landsat-based land use-land cover (vector). minnesota department of natural resources, st. paul, minnesota, usa. (accessed april 2011). _____. 2007. ecological classification system. minnesota department of natural resources, st. paul, minnesota, usa. < http://www.dnr.state.mn.us/ecs/index. html> (accessed april 2011). _____. 2011. minnesota moose research and management plan. minnesota department of natural resources, st. paul, minnesota, usa (in press). montgomery, r. a., g. j. roloff, j. m. ver hoef, and j. j. millspaugh. 2 010. can we accurately characterize wildlife resource use when telemetry data are imprecise. journal of wildlife management 74: 1917-1925. murie, a. 1934. the moose of isle royale. miscellaneous publication number 25. university of michigan, museum of zoology, ann arbor, michigan, usa. murray, d. l., e. w. cox, w. b. ballard, h. a.whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166. national oceanic and atmospheric administration (noaa). 2001-2010. climatological data for ely, minnesota. national climatic data center, ashville, north carolina, usa. osko, t. j., m. n. hiltz, r. j. hudson, and s. m. wasel. 2004. moose habitat preferences in response to changing availability. journal of wildlife management 68: 576-584. ott, p., and f. hovey. 1997. programs: bycomp.sas and macomp.sas. research branch british columbia forest service, victoria, british columbia, canada. (accessed april 2011). parker, k. l., and m. p. gillingham. 1990. estimates of critical thermal environments for mule deer. journal of range management 43: 73-81. pastor, j., and d. j. mladenoff. 1992. the southern boreal northern hardwood forest border. pages 216-240 in h. h. shugart, r. leemans, and g. b. bonan, editors. a systems analysis of the global boreal forest. cambridge university press, cambridge, england. peek, j. m. 1997. habitat relationships. pages 351-375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d. c., usa. _____, d. l. urich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48. peterson, r. l. 1955. north american moose. university of toronto press. toronto, ontario, canada. phillips, r. l., w. e. berg, and d. b. siniff. 1973. moose movement patterns and range use in northwestern minnesota. journal of wildlife management 37: 266-278. r development core team. 2006. r: a alces vol. 47, 2011 lenarz et al. compositional analysis of moose habitat 149 language and environment for statistical computing. r foundation for statistical computing, vienna, austria. isbn 3-900051-00-3, (accessed december 2006). renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditure and thermoregulatory response of moose. canadian journal of zoology 64: 322-327. retti, j. w., and p. d. mcloughlin. 1999. overcoming radiotelemetry bias in habitat selection studies. canadian journal of zoology 77: 1175-1184. ritcey, r. w., and n. a. m. verbeek. 1969. observations of moose feeding on aquatics in bowron lake park, british columbia. canadian field-naturalist 83: 339-343. rodgers, a. r., a. p. carr, h. l. beyer, l. smith, and j. g. kie. 2007. hrt: home range tools for arcgis. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. (accessed april 2011). romesburg, h. c. 1981. wildlife science: gaining reliable knowledge. journal of wildlife management 45: 293-313. saltz, d. 1994. reporting error measures in radio location triangulation: a review. journal of wildlife management 58: 181-184. samuel, m. d., and k. p. kenow. 1992. evaluating habitat selection with radiotelemetry triangulation error. journal of wildlife management 56: 725-734. schwab, f. e., and m. d. pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071-3077. schwartz, c. c., and a. w. franzmann. 1989. bears, wolves, moose, and forest succession, some management considerations on the kenai peninsula, alaska. alces 25: 1-10. spencer, d. l., and e. f. chatelain. 1953. progress in the management of the moose in south central alaska. transactions of the north american wildlife conference 18: 539-552. alces 31_15.pdf alces vol. 47, 2010 young and boertje – antlerless moose harvests 91 prudent and imprudent use of antlerless moose harvests in interior alaska donald d. young jr. and rodney d. boertje alaska department of fish and game, 1300 college road, fairbanks, ak 99701-1551, usa abstract: liberal antlerless moose (alces alces) hunts which allow the take of substantial numbers of largely female moose have been controversial and divisive since the alaska department of fish and game instituted ill-timed, liberal antlerless hunts in the early 1970s that contributed to a precipitous population decline. thus, we initially found the governing, citizen (non-agency) advisory committees largely skeptical of implementing liberal antlerless harvests in the early 2000s in game management unit 20a (unit 20a). to help justify the hunts, we focused on presenting information about the notably low nutritional status of the current moose population relative to moose populations worldwide. however, to gain broader credibility and trust, we needed to directly address public perceptions regarding former “mismanagement” of antlerless hunts, including admitting past mistakes that contributed to long-term poor hunting opportunities. we subsequently presented major differences between recent antlerless hunts and those in the 1970s. specifically, we contrasted relevant circumstances between the 2 time periods, including moose population trajectories, harvest rates of males and females, survey techniques and related technology, winter severity and frequency, and reproductive rates. illustrating the major, time-period differences in these parameters was key to assuring the public that harvest of female moose could be prudent. by directly addressing public anxieties, we were successful in gaining and maintaining public support for liberal antlerless hunts in unit 20a. subsequently, our success in unit 20a has helped ease recent expansion of antlerless hunts into adjacent areas. alces vol. 47: 91-100 (2011) key words: alaska, alces alces, antlerless, harvest rates, management, moose, overharvest, population trajectory, productivity, winter severity. we provide an example where decadesold, and at times imprudent moose (alces alces) management had an overwhelming influence on the implementation of current management strategy. the moose population in game management unit 20a (unit 20a) increased to an estimated 23,000 moose in the early 1960s following large-scale wildfires in the early 1940s, federal predator control in the 1950s, and low bull-only harvests (rausch et al. 1974, gasaway et al. 1983). a dramatic population decline to an estimated 2,800 moose occurred by early winter 1975. causes for the decline included at least 5 harsh winters between 1961 and 1975, accompanying high rates of predation, and liberal harvest of female moose in 1972-1974, simultaneous with increased numbers of hunters, improved access, and increased use of snow machines and airplanes (gasaway et al. 1983). in this system where black bears (ursus americanus), brown bears (ursus arctos), and wolves (canis lupus) are all significant predators on moose, boertje et al. (2007) defined liberal antlerless harvest as harvests ≥2.0% of the prehunt moose population. in retrospect, managers underestimated the effects of predation especially during harsh winters, and therefore the severity of the decline during the mid-1960s to mid-1970s, and mistakenly promoted liberal harvests of female moose, in part, to improve productivity. in response to intense public pressure concerning mismanagement that contributed to the depressed population and reduced hunting opportunity, the alaska legislature in 1975 granted veto authority for antlerless moose harvests – young and boertje alces vol. 47, 2011 92 antlerless hunts to the majority of the locally affected citizen (non-agency) advisory committees (young et al. 2006). following the population decline, a period of growth ensued from 1976-2003. causes for the increase included state wolf control (1976-1982, 1993-1994), public harvest of predators, predominantly conservative bullonly harvests, and nearly 3 decades of mostly mild winters (boertje et al. 1996, 2009). by november 2004, unit 20a had the highest moose density (~1.3 moose/km2) in alaska for any equivalent-sized area. given the high and increasing moose density and related poor nutritional status (boertje et al. 2007, 2009), the relevant primary goals of the alaska department of fish and game (adfg) were to 1) protect the moose population’s health and habitat, and 2) fulfill an intensive management (im) mandate for achieving high levels of harvest (alaska statutes 2009). in order to meet these objectives, it was necessary to reduce the population through harvest of substantial numbers of cows, because bull:cow ratios were already at or below the objective of 30 bulls:100 cows (young and boertje 2004, young et al. 2006, young and boertje 2008, boertje et al. 2009). to gain public support for these antlerless moose hunts, adfg needed to convince local citizen advisory committees that 1) the hunts were required to achieve the department’s goals, and 2) management “mistakes” that occurred in the 1970s would not be repeated. study area the study area, unit 20a, is in interior alaska immediately south of fairbanks (alaska, usa) across the tanana river, and is centered on 64°10′n latitude and 147°45′w longitude (fig. 1). unit 20a encompasses 17,601 km2, but only 13,044 km2 contains topography and vegetation characteristic of moose habitat; the study area was described in detail by gasaway et al. (1983). the northern portion consists of the northern lowlands (tanana flats) with elevations ranging from 110-300 m. the southern portion consists of the northern foothills and mountains of the alaska range with elevations varying up to 4000 m. lowland vegetation is a mosaic of shrub and young forest dominated seres, climax bogs, and mature black spruce (picea mariana) forest. vegetation in the hills, foothills, and mountains grades from taiga at lower elevations into shrub dominated communities with alpine tundra at higher elevations. the climate is typical of interior alaska where temperatures frequently reach 25° c in summer and -10° to -40° c in winter. snow depths are generally >80 cm. boertje et al. (1996) and keech et al. (2000) described the physiography, habitat, climate, and factors limiting moose in 1963-1997. young and boertje (2004) described hunter access, moose seasons, and bag limits from the 1960s through the early 2000s, moose population status from 1997-2003, and the use of calf hunts to increase yield. young et al. (2006) detailed the regulatory and biological history of moose from the 1960s through the early 2000s, and described impediments and achievements of managing moose for elevated yield in unit 20a. moose in unit 20a (1997-2005) exhibited the lowest nutritional status documented for noninsular, wild moose in north america (boertje et al. 2007). boertje et al. (2009) described how predation and reproduction affected the harvest of moose in 1996-2007. young and boertje (2008) described the use of selective harvest strategies to recover low bull:cow ratios in unit 20a. methods we defined harvest rate as (estimated harvest)/(prehunt population estimate). estimated harvest was calculated as reported harvest × 1.15 to lend consistency to past studies (gasaway et al. 1983, boertje et al. 1996). the prehunt population estimate was calculated as the estimated november population size + estialces vol. 47, 2010 young and boertje – antlerless moose harvests 93 mated harvest. we used estimated harvest and prehunt moose population estimates reported by gasaway et al. (1983) for years 1963-1978, boertje et al. (1996) for years 1979-1994, and unpublished data for years 1995-2009. gender-specific harvest rates were calculated as (estimated harvest per gender)/(prehunt moose population estimate). we monitored the annual moose harvest and gender of the harvest using a mandatory harvest report card system with reminder letters (schwartz et al. 1992, boertje et al. 1996). we used the november moose population estimates reported by gasaway et al. (1983; years 1963-1978), boertje et al. (1996; years 1979-1994), and boertje et al. (2007; years 1996-2006). in 2008 and 2009 we flew 158 and 116 of the 987 sample units available using methods described by boertje et al. (2007). we did not conduct moose population estimation surveys in 1995, 2002, or 2007, but used interpolations from adjacent years. since 1999 we used spatial statistics to estimate moose abundance (delong 2006, kellie and delong 2006, ver hoef 2008) and applied a sightability correction factor of 1.21 (boertje et al. 2009). to estimate the finite annual population growth rate (λ), we fitted population estimates during 1996-2004 and 2003-2009 with a trend line using mixed effects models (ver hoef 1996, delong and taras 2009). we estimated finite annual population growth rates for the overall population and, in order to reduce variability and improve precision, the adult female (≥1 year-old) segment of the population. we calculated twinning rate as the number of adult females observed with ≥2 newborns divided by the number of adult females observed with ≥1 newborn (boertje et al. 2007). staff flew late may or early june surveys in fig. 1. location of game management unit (gmu) 20a in interior alaska. antlerless moose harvests – young and boertje alces vol. 47, 2011 94 the central tanana flats during 43 years from 1960-2009 to estimate moose twinning rates. staff flew transect surveys during 1-day to 4-day periods in bellanca scout or piper pa-18 aircraft with both an observer and pilot searching for newborns; staff circled to determine if twins were present. to evaluate which winters were severe (i.e., winters defined as those with ≥80 cm accumulated snow depth, coady 1974) we used snow data reported by gasaway et al. (1983; winters 1959-1960 through 19781979), boertje et al. (1996; winters 19791980 through 1993-1994) and the national weather service at fairbanks, alaska and archived by the alaska climate research center, geophysical institute, university of alaska (alaska climate research center 2010; winters 1994-1995 through 2009-2010). results to address concerns of committees with veto authority over antlerless hunts, we documented 5 major differences (table 1) between moose management in the early 1970s versus the early 2000s: 1. the moose population was clearly declining prior to the 1970s liberal antlerless harvests (fig. 2). in contrast, prior to initiation of the 2004 liberal antlerless hunts, we estimated that the moose population increased (λ = 1.053, se = 0.013, 1996-2004) from approximately 11,500 to 17,800 and cows (λ = 1.04, se = 0.015, 1996-2003) from approximately 7,700 to 11,000 (fig. 3 and 4). 2. during 1996-2003 the population grew with overall harvest rates averaging 5.1% (3.5-6.5%) and female harvest rates averaging 0.6% (0-1.1%; fig. 3 and 4). during 2004–2007 the population declined with overall harvest rates of 7.0% (6.3-7.5%) and female harvest rates of 3.5% (3.1-4.2%; fig. 5). during the liberal antlerless hunts in 1972-1974 harvest rates averaged 14.2% (10.4-18.5%) overall and 6.7% (4.3-9.5%) for females (fig. 6). 3. moose managers in the early 1970s had inadequate survey techniques to estimate moose numbers, and therefore could not discern appropriate harvest rates. in contrast, since 1978 wildlife managers have had statistically defensible aerial survey techniques for estimating moose population parameters (gasaway et al. 1983, gasaway et al. 1986, kellie and delong 2006). 4. dmanagers in the 1970s did not appropriately account for the successive years of severe winters (fig. 7) that contributed to a sharp decline in moose (fig. 2). in contrast, during winters 1993-1994 through 2007-2008, maximum accumulated snow depth never reached the critical threshold affecting calf moose survival (coady 1974). moreover, except for winters 1989-1990, 1990-1991, and 1992-1993, the unit 20a moose population experienced a 37 year period when maximum accumulated snow depth was below the critical threshold affecting adult table 1. major differences in management components relative to imprudent (1970s) and prudent (2000s) moose management, game management unit 20a, interior alaska, usa. component 1970s 2000s moose population trajectory decreasing increasing harvest rates (% of females harvested in prehunt population) higher (4.3–9.5%) lower (3.1–4.2%) survey techniques to measure moose density and related technology unavailable available and proven frequency and severity of harsh winters series of harsh winters decades-long mild winters nutritional status higher (x = 15% twinning rates in central area 6 years prior to liberal antlerless harvest, 1966– 1971) lower (x = 7% twinning rates in central area 6 years prior to liberal antlerless harvest, 1998– 2003) alces vol. 47, 2010 young and boertje – antlerless moose harvests 95 moose survival. 5. the poor nutritional status perceived in the early 1970s as a rationale for liberal harvests was not realized until after 1996. during the years prior to initiation of the 1970s liberal antlerless hunts (1966-1971), twinning rates averaged 15% (boertje et al. 2007; fig. 8). in contrast, during the 6 year period (1998-2003) prior to the initiation of recent liberal antlerless hunts, twinning rates averaged 7%. discussion as initial justification for liberal antlerless harvest, adfg provided convincing information to the public on the moose population’s low nutritional status (boertje et al. 2007). during 1997-2005, moose in unit 20a exhibited the lowest nutritional status reported at the time for wild, noninsular, north american populations, shown by: 1) delayed reproduction until ≥36 months of age, 2) low parturition rate among 36 month old moose (29%, n = 147), 3) low average multi-year twinning rates (7%), 4) delayed twinning until moose reached 60 months of age, 5) low average mass of female short-yearlings in alaska (x = 155 ± 1.6 se kg), and 6) high removal (42%) of current annual browse biomass (boertje et al. 2007). when considering similarly high moose density in unit 20a and a study area in sweden (cederlund and sand 1991), recent studies (1997-2007) indicate that unit 20a produced only 75 calves:100 cows ≥36 months of age versus 117 calves in sweden (boertje et al. 2009). we concluded that low nutritional status in unit 20a resulted from the cumulative effects of having periodically high moose density in unit 20a (fig. 2) and a lower carrying capacity than the study area in sweden. in order to improve productivity and increase overall harvest yields, adfg reasoned that it would be necessary to lower the moose population by elevating the harvest rate of cow moose. to gain broader credibility and trust among fig. 2. moose population trends in game management unit 20a, interior alaska 1955-2009. large circles were back-calculated based on an index of abundance (moose/hour) linked to the 1978 population estimate (gasaway et al. 1983:6). error bars = 90% confidence limits. fig. 3. moose population trend using parametric bayes methods, game management unit 20a, interior alaska, 1996-2004. error bars = 90% confidence limits; sightability correction factor (scf) = 1.21; λ = 1.053 (se = 0.013). fig. 4. cow moose population trend using parametric bayes methods, game management unit 20a, interior alaska, 1996-2003. error bars = 90% confidence limits; sightability correction factor (scf) = 1.21; λ = 1.04 (se = 0.15). antlerless moose harvests – young and boertje alces vol. 47, 2011 96 those citizens who remembered the results of the 1970s antlerless harvests, we needed to admit past mistakes that contributed to longterm poor hunting opportunities. we also needed to specify why past mistakes would not be repeated by identifying major differences between the current antlerless hunts and those conducted in the 1970s. thus, we contrasted moose population status in the early 1970s with recent data and described the interim advancement in our ability to assess moose populations and habitat. we deemed the initiation of liberal antlerless harvests prudent only when 1) moose numbers were increasing, and 2) specific long-term, low nutritional indices were reached as density increased and without the effects of severe winters (boertje et al. 2007, 2010). it was imperative that population size and trajectory be monitored diligently and credibly. to assist with credibility issues, we solicited a representative private citizen with a seat on a local advisory committee to participate in and become knowledgeable with the population surveys, and to help describe the surveys at key committee meetings. a lay person’s involvefemale harvest rate male harvest rate fig. 6. gender-specific harvest rates of moose, game management unit 20a, interior alaska, 19632009. fig. 5. cow moose population trend using parametric bayes methods, game management unit 20a, interior alaska, 2003-2009. error bars = 90% confidence limits; sightability correction factor (scf) = 1.21; λ = 0.96 (se = 0.021). alces vol. 47, 2010 young and boertje – antlerless moose harvests 97 ment and perspective reduced the adversarial nature of the meetings and improved credibility. we also described the historical advancement in a manager’s ability to track population trend with confidence. in the 1970s moose were surveyed using moose/h counts in a few, small moose-concentration areas which left managers with no reliable way to confidently assess large-scale population trajectories or to estimate harvest rates. only in the late 1970s, after simultaneous moose/h counts and largescale surveys were conducted, did gasaway et al. (1983) use extrapolation to back-calculate the 1970s population trajectories. in contrast, in the early 2000s we sampled randomly selected, gps-defined, 5.8 mi2 survey units throughout the study area, and sample units were selected from both lowand high-density strata to improve confidence in the combined total estimate (delong 2006). we also combined estimates from several years to improve our confidence in detecting trend (ver hoef 2008). these population estimation techniques greatly improved the scientific basis for moose management in the 2000s. in addition, radio-telemetry was not generally available to wildlife managers in the early 1970s, but was a common tool for monitoring moose by the 2000s. prior to beginning liberal antlerless hunts in 2004, we had 8 years of age-specific productivity and survival data for radio-collared yearling and older moose, and 2 years of calf mortality studies in unit 20a. we incorporated those data into a simple quantitative model to illustrate why the population was increasing (boertje et al. 2007, 2009). moreover, during the antlerless hunts from 2004-2009, we continued using radiocollared moose to monitor age-specific productivity, survival, causes of mortality, unreported harvest, and wounding loss (boertje et al. 2009). lacking moose population estimates, managers in the early 1970s could not estimate harvest rates, thus had no practical experience with determining prudent harvest rates. instead fig. 7. maximum accumulated snow depth (cm) during winters 1965-1966 through 2009-2010. fairbanks, alaska. hashed vertical bars represent years with liberal antlerless harvests. horizontal bar represent critical snow depth thresholds for calf (80 cm) and adult (90 cm) moose (coady 1974). antlerless moose harvests – young and boertje alces vol. 47, 2011 98 managers relied on a prevailing management philosophy from scandinavia where harvests were being elevated to cope with increasing moose numbers and corresponding reduced nutritional status. however, the scandinavian systems were largely free of large predators, had abundant moose forage due to widespread clearcutting, and were not experiencing severe winters (lavsund et al. 2003). recent comparisons showed the moose population in unit 20a could sustain a 5% harvest of the prehunt population in 1996-2004, whereas the moose population in sweden could sustain a 33% harvest of the prehunt population, with calves constituting 48% of the harvest (cederlund and sand 1991, lavsund et al. 2003, boertje et al. 2009). these differences were not incorporated in alaska’s 1970s moose management, so managers erred in advocating liberal antlerless harvests, particularly immediately after a series of severe winters (fig. 6 and 7). lastly, we benefitted from 40 years of comparative data in unit 20a and elsewhere which allowed us to develop a convincing strategy for prudent liberal harvest of female moose, based largely on relative nutritional status (boertje et al. 2007, 2009, 2010). at this time, prudent management in unit 20a included preventing a reoccurrence of the extremely high moose densities of the mid-1960s that compromised nutritional status. thus, it is rewarding that the 2004-2006 liberal antlerless harvests led to a slight decline in the moose population, as desired (fig. 2 and 5). in addition, this history led to recommendations for future managers; specifically, for any study area in inland alaska where moose numbers gradually increase and twinning rates gradually decline to a 2 year average of <20%, female moose should be harvested in increasing numbers to stabilize population size. as population size is stabilized by harvest, we envision maintaining a 2 year average twinning rate between ~10-17% (boertje et al. 2010). if the 2 year average twinning rate declines to ≤10%, harvest should be slightly increased to reduce population size. this strategy appears to both manage moose responsibly below long-term carrying capacity and provide for elevated yield. conclusions modeling indicated recent liberal antlerless hunts were vital to keeping the moose population from growing to the unsustainable high levels observed in the mid-1960s. the key to achieving these liberal harvests and decreasing moose numbers was overcoming fig. 8. moose twinning rates, game management unit 20a, interior alaska, 1960-2009. alces vol. 47, 2010 young and boertje – antlerless moose harvests 99 skepticism related to the ability of adfg to responsibly manage harvest of female moose. overcoming skepticism entailed admitting to mistakes of the 1970s to show adfg was cognizant of past management mistakes and had incorporated what was learned into presentday moose management strategies. given the opportunity to provide the historical rationale for the overharvest of female moose also allowed us the unique opportunity to explain how severe winters and accompanying high predation played a more significant role in the 1960s and 1970s decline (rausch et al. 1974, gasaway et al. 1983, boertje et al. 1996). this history was crucial to convincing the public that moose could not be “stockpiled” and that elevated antlerless harvest of an increasing, high-density moose population experiencing nutritional effects was prudent. equally important, this unit 20a history eased the expansion of antlerless harvests into adjacent urban and agricultural areas (boertje et al. 2010). increased antlerless hunting opportunities in the 2000s were consistent with the mandate to manage for elevated yields and to meet the fiduciary responsibility of adfg to protect the health and habitat of the moose population over the long term. acknowledgements funding for this work was provided by adfg and federal aid in wildlife restoration. we wish to thank laura mccarthy for her help preparing the manuscript for publication. references alaska climate research center. 2010. snow depth data for fairbanks area, alaska, usa. (accessed july 2010). alaska statutes. 2009. section 16.05.255, regulations of the board of game; management requirements. pages 27-28 in alaska fish and game laws and regulations annotated, 2009-2010 edition. lexisnexis, charlottesville, virginia, usa. (accessed may 2010). boertje, r. d., m. a. keech, and t. f. paragi. 2010. science and values influencing predator control for alaska moose management. journal of wildlife management 74: 917-928. _____, _____, d. d. young, k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314-327. _____, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494-1506. _____, p. valkenburg, and m. mcnay. 1996. increases in moose, caribou, and wolves following wolf control in alaska. journal of wildlife management 60: 474-489. cederlund, g. n., and h. k. g. sand. 1991. population dynamics and yield of a moose population without predators. alces 27: 31-40. coady, j. w. 1974. influences of snow on behavior of moose. naturaliste canadien 101: 417-436. delong, r. a. 2006. geospatial population estimator software user’s guide. alaska department of fish and game, fairbanks, alaska, usa. _____, and b. d. taras. 2009. moose trend analysis user’s guide. alaska department of fish and game, fairbanks, alaska, usa. gasaway, w. c., r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. ———, s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological paper 22. university of alaska antlerless moose harvests – young and boertje alces vol. 47, 2011 100 fairbanks, fairbanks, alaska, usa. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, and t. r. stephenson. 2000. life-history consequences of maternal condition in alaskan moose. journal of wildlife management 64: 450-462. kellie, k. a., and r. a. delong. 2006. geospatial survey operations manual. alaska department of fish and game, fairbanks, alaska, usa. (accessed may 2010). lavsund, s., t. nygren, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. rausch, r. a., r. j. somerville, and r. h. bishop. 1974. moose management in alaska. naturaliste canadien 101: 705712. schwartz, c. c., k. j. hundertmark, and t. h. spraker. 1992. an evaluation of selective bull moose harvest on the kenai peninsula. alces 28: 1-13. ver hoef, j. m. 1996. parametric empirical bayes methods for ecological applications. ecological applications 6: 10471055. _____. 2008. spatial methods for plotbased sampling of wildlife populations. environmental ecological statistics 15: 3-13. young, d. d., jr., and r. d. boertje. 2004. initial use of moose calf hunts to increase yield, alaska. alces 40: 1-6. _____, and _____. 2008. recovery of low bull:cow ratios of moose in interior alaska. alces 44: 65-71. _____, _____, c. t. seaton, and k. a. kellie. 2006. intensive management of moose at high density: impediments, achievements, and recommendations. alces 42: 41-48. alces34(1)_75.pdf alces37(2)_447.pdf alces32_9.pdf alces 31_53.pdf alces36_205.pdf alces vol. 46, 2010 becker et al. – moose condition in wyoming 151 nutritonal condition of adult female shiras moose in northwest wyoming scott a. becker 1,3, matthew j. kauffman2, and stanley h. anderson 2,4 1wyoming cooperative fish and wildlife research unit, university of wyoming, department 3166, 1000 east university avenue, laramie, wy 82071, usa; 2u.s. geological survey, wyoming cooperative fish and wildlife research unit, university of wyoming, department 3166, 1000 east university avenue, laramie, wy 82071, usa. abstract: t�� ������� ��������� �������� ������� ���� ������� �� ������ �� � ������� �� ��� ���t�� ������� ��������� �������� ������� ���� ������� �� ������ �� � ������� �� ��� ��� vironment, it likely reflects the quality of its environment. although this concept has been applied to assess population condition and habitat quality for alaskan moose (alces alces gigas), to our knowledge this is the first time it has been used to assess the nutritional status of a shiras moose (a.a. shirasi) population. we investigated the physical condition and nutritional status of adult (≥ 2 years) female shiras moose captured in northwest wyoming during the winters of 2005-2007. rump fat depth was measured via ultrasonography and biological samples were collected and analyzed for hematology, serum chemistry, microand macronutrients, endoand ectoparasites, and bacterial and viral serology. five blood parameters believed to be important predictors of moose condition (packed cell volume, total serum protein, hemoglobin [hb], calcium [ca], and phosphorous [p]) were compared to data from alaskan moose considered to be in average-above average condition. microand macronutrient values were evaluated based on published deficiency levels for domestic herbivores. we conducted a correlation analysis to determine if a significant relationship existed between hematological and serum chemical parameters and rump fat depth. mean rump fat depth did not differ among years and was greater than reported values for alaskan moose. however, a high proportion of sampled moose had hb, ca, and p values lower than alaskan moose that were considered to be in average condition. hair and serum microand macronutrient analyses indicated a high proportion of moose were potentially deficient in copper, zinc, manganese, and p. we observed a marginally significant relationship between depth of rump fat and two serum chemical parameters (aspartate amimotransferase and lactate dehydrogenase). the results are suggestive of a shiras moose population in marginal physical condition that is probably related to less than optimal habitat quality. these findings should assist managers in evaluating the health of shiras moose populations throughout their range. alces vol. 46: 151-166 (2010) key words: alces alces shirasi, condition, disease, hair, hematology, moose, nutrients, nutrition, parasites, rump fat, serum chemistry, ultrasound, wyoming. t�� ������� ��������� �������� �� ���� to provide managers with a relative index �� ���������� ������ w��� ������� �� ������� carrying capacity (franzmann 1985). this �������� ������� ���� �� ������ �� � ������� of its environment and therefore will reflect the quality of its environment. early work focused on hematological and serum chemi� ��� ���������� �� ������ ����������� �� ������� quality among populations of pronghorn ante� lope (antilocapra americana; seal and hosk� inson 1978), white-tailed deer (odocoileus virginianus; seal et al. 1978), and elk (cervus elaphus; weber et al. 1984). franzmann and leresche (1978) expanded this concept by evaluating blood parameters in relation to in� 3present address: wyoming game and fish department, 2820 state highway 120, cody, wy 82414, usa. 4d�������. moose condition in wyoming – becker et al. alces vol. 46, 2010 152 ����� �� ��y����� ��������� ��� a���k�� ����� (alces alces gigas). this provided managers w��� �������� ���� ���� ����� �� ���� �� ������ ���������� ���������, ��������� ����������v� performance, and ultimately, habitat quality (franzmann and schwartz 1985, stephenson 2003). packed cell volume (pcv) was the single best predictor of body condition in moose, followed by hemoglobin (hb), total serum protein (tsp), calcium (ca), and phos� phorous (p; franzmann and leresche 1978). although the value of using tsp, ca, and p has been questioned (keech et al. 1998), these 5 blood parameters were effective in identifying populations at the extremes (i.e., very good or v��y ���� ���������), ��� w��� ���� �������v� w��� ���� �� ������� ����������� �� ������ ate condition (franzmann et al. 1987). more �������y, ���������� ������������ �� ���� ��� depth have been used to successfully quantify ����� ��������� ��� ����������v� ������� (stephenson et al. 1998, testa and adams 1998, keech et al. 2000). ev�� �� �� ������ ������� �� �� �� ����� tively good physical condition, nutritional deficiencies can create physiological imbal� ����� ���� ��y ������ ���������� ����������� (combs 1987, gogan et al. 1989). due to high variability in forage mineral concentra� tions among sites and seasons, free-ranging herbivores rarely acquire sufficient quantities of particular nutrients (mcdowell 2003). the nutritional quality of moose browse is most limited during winter (kubota et al. 1970, oldemeyer et al. 1977, ohlson and staaland 2001) ��� ������� �������������� �� ����� hair show similar temporal trends (franzmann et al. 1974, flynn et al. 1977, stewart and flynn 1978, flynn and franzmann 1987). mineral deficiencies can lead to reduced sur� vival, especially among calves and yearlings, ��� ������� ����������v� ������ �� �������� herbivores (wallisdevries 1998). although clinical deficiencies are difficult to diagnose in wild ungulate populations, deficiencies of trace elements, specifically cu, have been suggested as a contributing factor to moose population declines in alaska (flynn et al. 1977, o’hara et al. 2001), minnesota (custer et al. 2004), and sweden (frank et al. 1994). i������ �� ���� ����������� ��� ���������� density suggest a downward trend in shiras moose (a.a. shirasi) ������� �� �����w��� wyoming (brimeyer and thomas 2004, becker 2008). several factors have been hypothesized as contributing to this decline (brimeyer and thomas 2004), but no sys� ������� �������� ��� ���� ����������� �� �v������ ����� �������. t� ������� ��� ������ of habitat quality, disease, and parasites, we ���� ��� ������ ��������� ������� �� �������� ��� ��y����� ��������� ��� ����������� ������ of adult (≥ 2 years) female shiras moose via a suite of physiological parameters. although houston (1969) and kreeger et al. (2005) ���v�����y �������� ����� ��������� v����� for shiras moose in wyoming, to our knowl� edge, this study is the first to use the animal ��������� ������� �� ������ ��� ��������� �� � s����� ����� ����������. t��������, ���� work provides data that will aid managers in ������ �v�������� �� s����� ����� ����������� throughout their range. our research objec� tives were to: 1) compare hematological and ����� �������� ���������� �� �������� ���� ���� a���k�� �����, 2) �v������ ��� ������v� ��������� �� ��� ����� ���� ���� ���������� ���� ��� ������������, 3) ���� ��� � ��������� ship between rump fat depth and hematological and serum chemical parameters, 4) examine ������ ��� ������������� ������� �� ����� ����� ��� ���� ��� ������� �� �������� ��� ficiency values for domestic ruminants, 5) �v������ ��� �������� �� ���������� ��������, ��� 6) ������ ����� ��� ������������ �����. study area t�� ����y ���� w�� �������� �� ��� b������ valley (43° 42’ n, 110° 22’ w) approximately 50 km north of the town of jackson, wyoming, usa. it encompassed nearly 6,400 km2 �� ������������y ������ ���� �� �����w��� wy�� alces vol. 46, 2010 becker et al. – moose condition in wyoming 153 ming and included portions of grand teton national park, yellowstone national park, and the bridger-teton national forest where eleva� tions ranged from 1,866-4,197 m. the climate was characterized by short, cool summers and cold winters (houston 1968). in general, sage� brush (artemisia ���.) ��������� ��� v����y floors while coniferous forests and open forest parks were the most abundant vegetation types at moderate elevations (knight 1994); alpine tundra occurred at the highest elevations. riparian areas were characterized by willow (salix ���.) ������������ w��� �����w���� cottonwood (populus angustifolia) �� ��w�� elevations and on more mesic sites at higher ���v������. m���� �� ��� ����y ���������� w������� �� ��w����v�����, ������������������ habitats along the snake river and its primary tributaries (becker 2008). during summer, migratory moose traveled to more dispersed, mid-elevation ranges (becker 2008), whereas non-migratory individuals remained on low elevation ranges (houston 1968). methods a���� ������ ����� w��� �������� �� w��� ter range in january-march, 2005-2007. they were darted from the ground or helicopter and immobilized with 10-mg thiafentanil (a-3080, wildlife pharmaceuticals inc., fort collins, colorado, usa; mcjames et al. 1994, arnemo et al. 2003, kreeger et al. 2005) in 2005 and 2006, and 10-mg carfentanil (wildnil, wildlife pharmaceuticals inc., fort collins, colorado, usa; kreeger 2000) in 2007. samples were collected and moose were fitted with global po� sitioning system (model tgw-3700, telonics, mesa, arizona, usa) or very high frequency radio transmitters (model m2710, advanced telemetry systems, isanti, minnesota, usa). once handling was completed, thiafentanil and carfentanil were antagonized with an intramuscular injection of 300-mg naltrexone (trexonil, wildlife pharmaceuticals, fort collins, colorado, usa; kreeger et al. 2005). c������� w��� ��������� �� ���������� w��� approved university of wyoming animal care and use committee protocols (approved 2005, 2006, 2007). we collected approximately 50-ml of blood from each moose via jugular venipunc� ture for hematological analyses, serum chemi� ��� ����y���, ����� ����� ������� ������, ��� bacterial and viral serology. hematological ����y��� �������� w���� ����� �������������� of pcv, hb, mean corpuscular hemoglobin content (mchc), red blood cells (rbcs), total white blood cells (wbcs), composition �� w���� ����� �����, ��� ���������. s���� �������� ����y��� �������� �������������� of albumin (alb), alkaline phosphate (alp), aspartate aminotransferase (ast), blood urea nitrogen (bun), creatine kinase (ck), gammaglutanyl transferase (ggt), globulins (glob), glucose (gluc), lactate dehydrogenase (ldh), tsp, and the macronutrients ca, magnesium (mg), and p. levels of 5 micronutrients were analyzed w��� ����� ����� ������� ������� ��� �������� cu, iron (fe), manganese (mn), molybdenum (mb), and zinc (zn). blood was analyzed for the presence of antigens against leptospira, ���������� ��v��� ��������������� v����, ��v��� viral diarrhea virus, parainfluenza-3 virus, and ��v��� ����������y �y��y���� v���� �� 2005; ����y��� w�� ��������� ��� brucella abortus in 2005, 2006, and 2007. hair samples were ��������� ���� ��� ������ ������� ���w��� ��� shoulders and analyzed for concentrations of arsenic (as), barium (ba), cadmium (cd), chromium (cr), cobalt (co), cu, fe, lead (pb), mn, mercury (hg), mb, nickel (ni), selenium (se), thallium (tl), vanadium (v), tin (sn), and zn. f���� ������� ��� ��� �w��� w��� ���� ������ �� �v������ ����� ��� ������������ loads. although encapsulation would have resulted in few, if any, fluke eggs transported through the feces, fecal examinations were ���� �� ������ ����� ��������, �����������y ��� giant liver fluke (fascioloides magna) w���� is undocumented in wyoming and the com� moose condition in wyoming – becker et al. alces vol. 46, 2010 154 mon liver fluke (fasciola hepatica) w���� �� ���������� ����. a 30������� ���k ����� w�� performed along the dorsal midline posterior �� ��� ���k �� ���� ����� �� �������� ��� ��� verity of winter tick (dermacantor albipictus) infestations. all diagnostic analyses were performed at the wyoming state veterinary laboratory (laramie, wyoming, usa). body condition was subjectively evalu� ated and a score from 0-10 was assigned to each moose (franzmann 1977). depth of rump ��� w�� �������� w��� ���������� �������� �� the nearest 0.1 cm using an omega i portable ultrasound unit (e.i. medical, loveland, colorado, usa) in 2005 and a bantam xls portable ultrasound unit (e.i. medical, love� land, colorado, usa) in 2006 and 2007. we measured to the midpoint between the coxal tuber (hip bone) and the ischial tuber (pin bone), then located maximum rump fat depth from that point. maximum rump fat depth was closer to the ischial tuber than the coxal tuber in all cases; however, since our starting point differed slightly from that described by stephenson et al. (1993, 1998), the measure� ���� w�� ������� �w�y ���� ��� ����� ��� �� was unknown how this might affect subsequent ����������� w��� ����� ����. blood parameter (hematological and ����� ��������) v����� ��� ������� ������� trations (serum and hair microand macro� ���������) ��� ��� ����� w��� ������ w����� years. annual means for pcv, hb, tsp, ca, and p were compared to baseline data for a���k�� ����� ���� w��� ���������� �� �� in average-above average condition (fran� zmann and leresche 1978); we report the ���������� �� ��� ������� ���������� ����w ����� �������� v�����. m����� ��� �������� trient requirements for moose have not been �����������, �� ��� ���������� �� ��� ������� population that was deficient was estimated based on published deficiency thresholds for domestic ruminants (puls 1994, mcdowell 2003). t�� ��������� ��������� v����� ��� c� (8.0 mg/dl) and mg (1.8 mg/dl) are not true deficiency thresholds and only represent the ��w�� ������ ����� ��� �������� ��������� (puls 1994, mcdowell 2003). w� ����� ������y��� �� ��� ������ ���� because we expected among-year variation in female reproductive status (i.e., cost of lacta� tion) and environmental conditions (i.e., winter ��v����y, ������ ��������v��y) �� ��v� � ����� nant influence on individual condition. since ���� ��� ������������ ��� ������������v� �� the variation that can be expected in adult female moose condition among years (testa and adams 1998, keech et al. 2000, boertje et al. 2007), we plotted rump fat depth for moose sampled in 2 (n = 5) or 3 (n = 2) years against ����� ������� �� ���y ��� y���. t��� ����w�� �� �� �v������ �� �������� �������� ���� ��� ���� ����� �v�� �������� y���� ������ �� cluster (suggesting a lack of independence) or were variable among years (schwartz et ��. 2010). v����� ���������� �� ����� ���w�� ���� ���� ��� �� ����� ������� �� �������� y���� w�� �� v������� �� ����� ������� ���y once, suggesting that moose-year was an ap� propriate sampling unit. we used a one-way analysis of variance and a tukey’s honestly significant difference (hsd) test to examine among year differences (α = 0.05) in rump fat depth, body condition scores (bcs), all blood ����������, ��� ������ ��� �������������� ���� w��� ���v� ��� ������� ��������� ����� (mdl) in order to quantify between-year v���������. w� ��������� � s������� ���k correlation analysis (α = 0.05) with bonfer� roni corrections to determine if a significant relationship existed between hematological (α = 0.001) and serum chemical parameters (α = 0.004) ��� ����� �� ���� ���. a�� ����������� analyses were performed with statistix 8.0 software (analytical software, tallahassee, florida, usa). results moose capture, rump fat, disease, and parasites forty-eight adult female moose were cap� alces vol. 46, 2010 becker et al. – moose condition in wyoming 155 tured 61 times during the course of this study. most captures occurred in february (n = 54) from a helicopter (n = 53). n����y ��� ������ ����� ���� ��� ������������ w��� �������� in february (n = 41) with the exception of 5 �� ����y �� ����m����. w� ��� ��� ������� �� distinguish rump fat depth between cows with ��� w������ ���v���������� �� w����� ������� of inconsistency in reporting presence of a calf during capture. mean rump fat depth was not different among years (f(2,43) = 0.9, p = 0.399; table 1). there were no differ� ences between rump fat depth for pregnant cows observed with (x = 24.1 ��, se = 2.4, n = 8) or without calves (x = 27.4 mm, se = 1.3, n = 31) in the spring following capture (t = 1.18, df = 37, p = 0.246). d���������� were observed in bcs among years (f(2,51) = 4.8, p = 0.012) ��� post hoc ����y��� ��������� that bcs in 2005 were significantly higher than in either 2006 or 2007 (table 1). moose (n = 59) were negative for anti� gens against b. abortus �� ��� y���� ��� ��� leptospira, ���������� ��v��� ��������������� virus, bovine viral diarrhea virus, parainflu� enza-3 virus, and bovine respiratory syncytial virus in 2005 (n = 20). w����� ���k ����� w��� relatively low and averaged 2.8 ticks/moose with 55 of 59 moose hosting <10 ticks. no moose (n = 56) ��� �v������ �� ��� ����� ��� fluke eggs were not observed in any sample (n = 43). fecal examinations (n = 44) ��������� a low infection (≤12 eggs/g) of nematodirus spp. �� 13 ����� ��� trichostrongylus spp. �� 2 �����. hematological, serum chemical, and macroelement analyses there were no among-year differences in hb (p = 0.053) or platelets (p = 0.104), ��� differences were found for pcv (f(2,48) = 9.5, p <0.005), mchc (f(2,48) = 8.3, p <0.005), rbc (f(2,48) = 6.9, p = 0.002), and wbc (f(2,48) = 4.7, p = 0.013; table 2). no consistent increasing or decreasing patterns were observed for pcv, mchc, or rbc, but wbc exhibited a gener� ally increasing trend with 2005 significantly lower than 2007. the percent composition of wbc did not differ among years for lympho� cytes (p = 0.089), eosinophils (p = 0.353), �� monocytes (p = 0.168), but differences were observed for neutrophils (f(2,48) = 4.7, p = 0.014), ��� post hoc ����y��� ��������� ���� 2007 was significantly lower than 2005 and 2006 (table 2). f�� ����� �������� ����y���, ����� w��� no among-year differences for alp (p = 0.149) and ggt (p = 0.339), but differences were found for alb (f(2,54) = 19.0, p < 0.005), ast (f(2,54) = 10.3, p < 0.005), bun (f(2,54) = 4.7, p < 0.005), ck (f(2,53) = 6.5, p = 0.003), globulins (f(2,54) = 23.6, p <0.005), glucose (f(2,54) = 12.5, p <0.005), ldh (f(2,54) = 47.1, p <0.005), and tsp (f(2,54) = 48.3, p < 0.005; table 3). no consistent increasing or decreasing patterns were observed for alb, ast, bun, globulins, glucose, ldh, and tsp. however, ck values exhibited a generally increasing y��� parameter n x ± se 95% ci1 2005 rump fat (mm) 13 27.6 ± 3.5 19.9 � 35.3 bcs 17 7.5 ± 0.3 6.9 � 8.2 2006 rump fat (mm) 18 26.4 ± 1.3 23.7 � 29.0 bcs 19 6.6 ± 0.2 6.1 � 7.0 2007 rump fat (mm) 15 23.6 ± 1.3 20.8 � 26.5 bcs 18 6.6 ± 0.2 6.1 � 7.0 table 1. count (n), mean (x) ± standard error (se), and 95% confidence intervals (ci) for rump fat depth and body condition scores (bcs) by year for adult female moose captured in northwest wyoming during winter 2005-2007. 1upper and lower confidence interval. moose condition in wyoming – becker et al. alces vol. 46, 2010 156 pattern with 2007 significantly higher than 2005 (table 3). a���y��� �� ����� �������� ���������� indicated among-year differences for all 3 macronutrients (ca: f(2,54) = 35.7, p <0.005; mg: f(2,53) = 16.1, p <0.005; p: f(2,54) = 4.93, p = 0.011; table 3), but no consistent increasing or decreasing trend was evident. annual means of serum ca exceeded the lower normal limit threshold for domestic ruminants (8.0 mg/dl) in 2006 and 2007, but were slightly below this level in 2005 (table 3). when moose were compared individually, 18% (11 of 58) had ca levels <8.0 mg/dl threshold, and 57% (33 of 58) were below the domestic ruminant deficiency threshold (4.5 mg/dl) for serum p; parameter1 (units) 2005 (n = 19) 2006 (n = 16) 2007 (n = 16) pcv (%) 54.7 ± 7.9 45.6 ± 4.5 49.7 ± 5.2 hb (g/dl) 16.5 ± 2.1 15.6 ± 1.6 17.2 ± 1.7 mchc (g/dl) 30.6 ± 4.8 34.2 ± 1.9 34.7 ± 2.0 rbc (x 106/μl) 7.9 ± 1.3 6.8 ± 0.6 7.3 ± 0.7 total wbc (/μl) 5296.8 ± 1581.2 5967.5 ± 1466.2 6952.5 ± 1706.7 lymphocytes (%) 56.1 ± 9.6 56.4 ± 9.9 63.7 ± 13.2 neutrophils (%) 36.2 ± 8.7 37.3 ± 8.9 27.9 ± 11.2 eosinophils (%) 4.4 ± 3.2 3.7 ± 2.9 5.4 ± 4.1 monocytes (%) 3.3 ± 1.8 2.7 ± 1.1 2.4 ± 1.0 platelets (x 103/μl) 189.4 ± 53.0 148.4 ± 58.5 177.8 ± 58.7 table 2. mean ± standard deviation for hematological analyses of adult female moose captured in northwest wyoming during winter 2005-2007. 1pcv = packed cell volume; hb = hemoglobin; mchc = mean corpuscular hemoglobin concentration; rbc = ��� ����� ����; wbc = w���� ����� ����. parameter1 (units) 2005 (n = 20) 20062 (n = 18) 2007 (n = 17) albumin (g/dl) 2.9 ± 0.5 3.8 ± 0.5 3.4 ± 0.4 alp (u/l) 255.9 ± 99.1 338.1 ± 151.1 297.3 ± 125.5 ast (u/l) 62.4 ± 17.5 87.1 ± 18.6 103.7 ± 42.5 bun (mg/dl) 3.4 ± 1.0 5.0 ± 2.4 3.4 ± 1.9 ca (mg/dl) 7.9 ± 1.2 10.2 ± 0.4 10.5 ± 0.9 ck (u/l) 111.8 ± 76.6 238.8 ± 175.6 328.9 ± 267.1 ggt (u/l) 10.2 ± 5.7 16.2 ± 6.2 15.5 ± 22.3 globulins (g/dl) 3.3 ± 0.8 4.6 ± 1.0 5.1 ± 0.7 glucose (mg/dl) 102.6 ± 20.3 79.7 ± 20.6 72.0 ± 18.8 ldh (u/l) 161.5 ± 37.8 275.2 ± 58.2 310.5 ± 53.3 mg (mg/dl) 2.0 ± 0.1 2.4 ± 0.1 2.4 ± 0.1 p (mg/dl) 3.7 ± 0.2 4.7 ± 0.3 4.3 ± 0.2 tsp (g/dl) 6.1 ± 1.1 8.4 ± 0.8 8.5 ± 0.6 t���� 3. m��� ± �������� ��v������ ��� ����� �������� ����y��� �� ����� ������ ����� �������� �� northwest wyoming during winter 2005-2007. 1alp = alkaline phosphate; ast = aspartate aminotransferase; bun = blood urea nitrogen; ca = calcium; ck = creatine kinase; ggt = gamma-glutanyl transferase; ldh = lactate dehydrogenase; mg = magnesium; p = phosphorous; tsp = total serum protein. 2alp, ck, and mg (n = 17). alces vol. 46, 2010 becker et al. – moose condition in wyoming 157 ������ ����� w��� ����w ���� ��v�� �� 2005 and 2007 (table 3). the annual means of serum mg exceeded the lower normal limit threshold for domestic ruminants (1.8 mg/ dl) in all years (table 3); 12% (7 of 57) were ����w ���� ��v��. t���� w�� v�������� �� ��� ���������� �� moose with pcv, hb, tsp, ca, and p values ����w ����� �������� ��� a���k�� ����� considered to be in average-above average condition (table 4). most moose fell below the average thresholds for hb, ca, and p; approximately 50% and 33% were below average for pcv and tsp, respectively. mean hb concentrations were lower in all years and pcv was lower in 2006 and 2007 (table 2, table 4). serum levels of ca and p were lower in all years and tsp was lower in 2005, but higher than the average threshold in 2006 and 2007 (table 3, table 4). o� ��� 13 ����� �������� ���������� analyzed, 2 exhibited a marginally signifi� cant relationship with rump fat depth (n = 43 moose; α = 0.05). aspartate aminotransferase (r� = -0.339, p = 0.041; fig. 1) and ldh (r� = -0.327, p = 0.049; fig. 2) were both negatively correlated with depth of rump fat. the enzyme ck was partially correlated and negatively related to rump fat (r� = -0.317, p = 0.057); however, when the single ck value >1000 u/l was removed, the direction �� ����������� ��v����� ��� ��� ������������ was insignificant (r� = 0.237, p = 0.130). w��� b��������� ����������� w��� �������, ��� ������������ �����v�� ���w��� ���� ���, ast, and ldh were insignificant. no sig� nificant relationship was observed between rump fat and any hematological parameter (n = 38 moose). serum and hair trace mineral analyses serum cu, fe, and zn were detected in all moose (table 5), whereas mn and mb had ��v��� ����w ��� mdl ��� w��� ����������. there were no among-year differences in cu (p = 0.329), but differences were found for fe (f(2,47) = 3.79, p = 0.030) and zn (f(2,47) = 25.1, p <0.005). no consistent increasing parameter1 (units) n range r��������2 proportion below reference pcv/hct (%) 51 35.1 � 50.0 50 0.51 hb (g/dl) 51 12.1 � 18.6 18.6 0.88 tsp (g/dl) 58 3.6 � 7.5 7.5 0.33 ca (mg/dl) 58 5.2 � 10.4 10.4 0.81 p (mg/dl) 58 2.1 � 5.2 5.2 0.78 table 4. total adult female moose sampled (n), range, and the proportion of the sample that was below the reference value for alaskan moose considered to be in average-above average condition for 5 ����� ���������� ���� ��� ��������� ����������. 1pcv = packed cell volume; hb = hemoglobin; tsp = total serum protein; ca = calcium; p = phos� �������. 2values for alaskan moose in average-above average condition (franzmann and leresche 1978). y = -0.095x + 34.322 r2 = 0.099, p = 0.041 12 18 24 30 36 42 48 54 35 45 55 65 75 85 95 105 115 125 135 145 155 aspartate aminotransferase (u/l) r um p fa t ( m m ) fig. 1. scatterplot describing the relationship between rump fat depth (mm) and aspartate ami� notransferase (u/l) concentrations of captured adult female moose in northwest wyoming, winter 2005-2007 (n = 43). t�� ������������ was significant at the α = 0.05 level, but became insignificant when bonferroni corrections were applied (α = 0.004). moose condition in wyoming – becker et al. alces vol. 46, 2010 158 or decreasing patterns were observed for the annual means of fe and zn. when compared �� �������� ���������, ������� ����� w��� deficient in cu during all years and deficient in zn in 2005 and 2007 (table 5). when examined individually, a high proportion of moose were deficient in cu and zn (table 5); only in 2006 were moose (n = 15) above the zn deficiency threshold. annual means of serum fe exceeded the 1.1 ppm threshold during all years; only 2% (1 of 50) of individual moose were below this level (table 5). hair concentrations of as, cd, co, hg, m�, n�, s�, t�, v, ��� s� w��� �����������y ����w mdl, w������ ��� ������� ��� ������� able levels of ba, cr, cu, fe, mn, pb, and zn (table 5). there were no among-year differ� ences in cu (p = 0.279), mn (p = 0.429), and pb (p = 0.080); differences were found for ba (f(2,56) = 3.34, p = 0.043), cr (f(2,56) = 4.80, p = 0.012), fe (f(2,56) = 4.52, p = 0.015), and zn (f(2,56) = 11.80, p < 0.005). no consistent in� creasing or decreasing patterns were observed in concentrations of ba, cr, and zn, but fe t��� element (units) s����� �y�� 2005 2006 2007 published deficiency levels (ppm) proportion ����w deficiency ��v�� copper (ppm) s���� 0.51 ± 0.09 0.46 ± 0.14 0.45 ± 0.10 < 0.61,2 0.84 hair 4.76 ± 0.72 4.63 ± 0.56 4.43 ± 0.65 < 6.72 1.00 iron (ppm) s���� 2.78 ± 0.38 2.33 ± 0.74 2.32 ± 0.46 < 1.11 0.02 hair 26.35 ± 16.70 19.12 ± 8.32 15.37 ± 6.54 ≤ 402 0.95 zinc (ppm) s���� 0.58 ± 0.13 1.42 ± 0.60 0.71 ± 0.09 < 1.01 0.70 hair 82.64 ± 7.74 89.49 ± 3.52 89.86 ± 2.98 < 1001 1.00 manganese (ppm) hair 1.09 ± 0.16 0.79 ± 0.08 1.00 ± 0.25 < 5.01 1.00 barium (ppm) hair 1.29 ± 0.74 1.79 ± 0.63 1.73 ± 0.64 c������� (ppm) hair 1.73 ± 0.55 1.39 ± 0.20 1.57 ± 0.29 lead (ppm) hair 0.17 ± 0.09 0.11 ± 0.06 0.26 ± 0.35 table 5. annual mean ± standard deviation, published deficiency levels, and the proportion of sampled adult female moose that were deficient in microand macronutrients analyzed in serum and hair from northwest wyoming during winter 2005-2007. no published deficiency levels were reported ��� ������, ��������, ��� ����. 1deficiency level for cattle and sheep; mn levels are indicative of slight deficiency (mcdowell 2003). 2deficiency level for cattle (puls 1994). y = -0.027x + 33.218 r2 = 0.083, p = 0.049 12 18 24 30 36 42 48 54 50 150 250 350 450 lactate dehydrogenase (u/l) r um p fa t ( m m ) fig. 2. scatterplot describing the relationship between rump fat depth (mm) and lactate de� hydrogenase (u/l) concentrations of captured adult female moose in northwest wyoming, winter 2005-2007 (n = 43). t�� ������������ was significant at the α = 0.05 level, but was insignificant when bonferroni corrections were applied (α = 0.004). alces vol. 46, 2010 becker et al. – moose condition in wyoming 159 concentrations showed a generally decreasing pattern; 2007 means were significantly lower than in 2005. annual means for cu, fe, zn, and mn were below the deficiency thresholds for domestic ruminants in all years (table 5). when examined individually, all moose were deficient in cu, zn, and mn, and all but 3 moose were below the deficiency threshold for fe (table 5). discussion blood parameters and rump fat although pcv, hb, tsp, ca, and p have been used to evaluate habitat quality and the nutritional status of alaskan moose (fran� zmann and leresche 1978), we, like keech et al. (1998) with alaskan moose, found none of ����� ���������� w��� ���������� w��� s����� moose rump fat depth. our results suggest that the serum enzymes ast and ldh may be good predictors of shiras moose condition as indexed by ultrasonic rump fat measurements. even though neither variable was significant w��� ��� b��������� ���������� w�� �������, they were significant at the α = 0.05 level and the negative relationship between ast and ���� ��� �� ���������� w��� ���v���� w��k w��� alaskan moose . keech et al. (1998) suggested ���� ������� ��v��� �� ast w��� ��������v� �� ����� ���� w��� �� ������ ��y����� ��������� w���� ��k��y ������� ����� �������������y �� disease. although this may be true, ast and ldh are indicators of muscle or organ damage generally associated with exertional myopathy (em; williams and thorne 1996). levels of ast ��� s����� ����� w��� ��� ��������v� �� em ��� w��� w��� ����w v����� �������� for bighorn sheep that were stressed or sub� sequently developed em (kock et al. 1987). a����������y, ��v��� �� ast w��� w��� ����w normal values reported for moose (haigh et al. 1977) which suggests that em had little influence on these relationships. the negative relationships that we ob� served between ast, ldh, and rump fat are consistent with increased utilization of body proteins from muscle and organ tissues as lipid �����v�� ������� �� ���� �������. c����� �� ��. (1992) observed a similar trend in which lean rats utilized greater amounts of muscle protein during phase ii fasting (i.e., protein sparing) ���� ��� ����� ����. w���� w� �����v�� � marginally significant relationship between two serum enzymes and rump fat depth, as with caribou (rangifer tarandus; m������ �� al. 1987), elk (cook et al. 2001), and moose (keech et al. 1998), we cannot identify a set �� ����� ���������� �� s����� ����� ���� ��� curately reflects nutritional status as an index of rump fat. because not all managers have ������ �� �� ���������� �� �v������ ����� ���������, ������� �v�������� �� ����� ��������� ����� ������� w�������� ����� ����� ������� ��� ���� �����y ��������. although rump fat depth should be in� ��������� w��� ������� ����� w� �������� in a slightly different location than previous studies, our field measurements indicated that ����� �� ��� ����y ���� w��� �� ������v��y good physical condition. furthermore, the �������k� ���� w��� ���� �� ������ �������� ����� �������� �� ��� ���� �������� �� ���� moose were oftentimes difficult to locate, suggesting that most study animals carried high amounts of subcutaneous fat. when �������� �� ���� ��� �� ����� �������� �� early to mid-march in alaska (keech et al. 1998, bertram and vivion 2002, boertje et al. 2007), this population displayed nearly 2x more rump fat. although we were unable to ������� ���� ��� ��� ����� w��� ��� w������ calves-at-side during capture, other studies found that cow moose with greater amounts of rump fat were not tending calves (testa and adams 1998, keech et al. 2000). since calf recruitment has declined for approximately 20 years (becker 2008), the high rump fat values may reflect fewer cows with calves. w���� ���� ��� ��y �� � ������ ��������� �� ����������v� ������� w����� ����� ������� tions, it appears to be an insensitive index of fitness when compared across populations moose condition in wyoming – becker et al. alces vol. 46, 2010 160 (boertje et al. 2007). additionally, heard et al. (1997) suggested that moose populations living in relatively harsh environments, or in areas with low forage quality or quantity, may have a higher fat-fertility threshold than moose populations living in milder climates with quality forage. thus, our high rump fat values may indicate a population needing to maintain high fat levels to realize their optimal reproductive potential. nonetheless, a larger sample size collected across multiple locations may provide researchers and managers with ���� �������� ��������� �� �������� ���� ��� ��v�� ��� ����������v� ����������� ��� moose in wyoming. as with evaluations of elk condition (cook et al. 2001), the thick� ness of specific muscles measured via ultra� sonography could provide an additional index ���� w��� ���� ��� ����� �� ���v��� � ���� �������� ���������� �� ��� ��y����� ��������� (i.e., protein versus fat catabolism) of shiras ����� �����������. although we are confident that most ���������� ������������ w��� �������� ���������y, ������� �������� w��� ������ �y other professionals sent images for their ��������������. w� w��� ���� �� v������� ��� ������������ �� ��� ����� ������ ����� ���� was euthanized during capture and another that died within a month of capture; both had high ������� �� ������������ ���. n����������, �� �� �������� ���� ��������������� �������� because we had insufficient training and may have measured the wrong tissue layer for some �����, ���������y ����� w��� ������ ��� �����v�� (cook et al. 2007). similarly, the difference �� bcs ���w��� 2005 ��� ����� y���� ��k��y resulted from inexperience in the application of this subjective method, as well as multiple individuals scoring moose in 2005. for con� �������y, ��� ������ ���� �������� w��� ��� scoring method palpated moose and provided the bcs scores in 2006 and 2007; this approach ��k��y ������� v�������� �� ����� y����. m���� ���� ��� ����y ���� �������� �� �� in marginal physical condition based on the 5 blood parameters (pcv, hb, tsp, ca, and p) ���������� �� ���������� �� ����������� ������ of moose (franzmann and leresche 1978). this suggests that habitat conditions may be slightly suboptimal, but it is clear that condi� tions are not extreme. when compared to alaskan moose considered in good-excellent condition (franzmann and leresche 1978), ���� ����� ������ ����� w��� ����w ��� reference values for pcv, hb, ca, and p and above the reference value for tsp. when these ����� ���������� w��� ������� �������� �� �� expanding, highly productive population and ��� ���� w�� �� ���� ��������� ���� a���k� (i.e., populations on the extremes; franzmann et al. 1987), moose from the study area fell in ��� ������. b������ ����� ���������� v��y across winters of differing severity (ballard et al. 1996), are best used to identify popula� tions at nutritional extremes (franzmann at el. 1987), and are not always representative of other indices of physical condition (keech et al. 1998), their efficacy in assessing condition �� ���y ����������� �� ��k��y �������. microand macronutrients adult female moose in the study area ex� ������� ������ v�������� �� �����y ��� ������ ��� macronutrients. these results suggest that the nutritional quality of moose browse exhibits ������� ������ v��������. i�����, ����������� in alaska and sweden have reported high an� ���� v�������� �� ��� ������� ������� �� ����� browse (oldemeyer et al. 1977, ohlson and staaland 2001). it has been suggested that a ��v�����y �� ���w�� ������� ��� ������ ���� the nutritional requirements of moose than a single, highly abundant species (oldemeyer et al. 1977, miquelle and jordan 1979, ohl� ��� ��� s������� 2001). m���� �� ��� ����y area utilized low-elevation, riparian habitats dominated by large communities of willow intermixed with small stands of conifers and aspen during winter (becker 2008), suggesting that willow composed a high proportion of the winter diet. if willows are deficient in certain alces vol. 46, 2010 becker et al. – moose condition in wyoming 161 nutrients, moose that consume high quantities of willow may also be deficient in these ele� ments. direct analysis of forage quality is a more precise indicator of deficiency in most cases (mcdowell 2003), thus future investiga� tions may explore potential links between diet diversity and nutritional deficiencies on winter and summer ranges of shiras moose. since moose acquire nutrients directly from the plants they consume (mcdowell 2003), ��w �������������� �� ���� �������� in serum and hair suggest nutritional limita� ����� ���������� w��� ����� ������� �� ��� ����y ����. o�� ������� �������� ���� w����� forage may have been limited in cu, zn, mn, and p. deficiencies in any nutrient are most likely to occur during winter when the avail� ability and mineral content of forage is most limited (kubota et al. 1970, oldemeyer et al. 1977, ohlson and staaland 2001). increased intraand interspecific competition for lim� ited winter forage (o’hara et al. 2001) may exacerbate existing nutritional deficiencies due to overutilization of resources (barboza et al. 2003). for example, overabundant freeranging elk would remove preferred vegetation and reduce forage quality earlier in winter in willow-dominated, riparian range in the b������ v����y. moose may be highly susceptible to nu� tritional deficiencies (murray et al. 2006), and although cu, zn, mn, and p deficiencies are extremely difficult to diagnose in wild popula� tions, the physiological imbalances that they ��y ������ ����� ��v� ������������ ������ �� ��� ����������� �� ��� ����������, �����������y the developing fetus and calf. while we cannot conclude that low or marginal cu has been the ������y ����� �� ��� ������ ����� �������, �� remains a possible contributing factor because �������������� �� ����� ��� ���� ��������� � potential deficiency among moose in the study ����. m��� c� �� ������ �� ��� ��v��, ��� w��� levels are <20 μg/g, serum and hair become sensitive indicators of cu deficiency among domestic ruminants (combs 1987, blakley et al. 1992, mcdowell 2003). copper is an essential nutrient for the developing fetus, and fetal demand of cu greatly increases during the final trimester of pregnancy (puls 1994, mcardle 1995, rombach et al. 2003); the ��k������� �� ����������v� ������� ���������y increases if maternal cu is deficient (hidiro� glou and knipfel 1981, mcdowell 2003). s���� c� ��v��� �� ����� ���� ��� ����y ���������� w��� ������� �� ��v��� �������� in cu deficient elk that experienced reduced adult survival and poor recruitment (gogan et al. 1989). although we did not observe faulty hoof keratinization associated with cu deficiency, ����� ��� ������� ����������v� ������ (becker 2008) similar to populations from the kenai peninsula, alaska (flynn et al. 1977), the north slope of alaska (o’hara et al. 2001), and minnesota (custer et al. 2004). however, low c� ��v��� ����� w��� ��� ����������� ��� ��� reduced reproductive success among pregnant moose in northwest wyoming (becker 2008). i� ��y �� ���� ��� ��������v� ������� �� ������ sors (i.e., low quality forage, moderate physical ���������, ��v���������� ����������) ���� ��� third trimester of pregnancy combined with potential deficiencies in several other nutrients (i.e., mn, zn, p) created physiological imbal� ances (frank et al. 1994) that compromised ����������v� �����������. concentrations of mn in hair and zn in serum and hair indicated a potential deficiency in the study area. all hair samples suggested a deficiency in these nutrients while approxi� mately two-thirds of moose were serum zn deficient. all moose that were above serum zn deficiency thresholds were sampled in 2006; however, these higher levels were likely � ������ �� ������ ������������� ���� ������ rect collection procedures (puls 1994). in domestic ruminants, clinical signs of mn and zn deficiencies include reduced reproductive performance and calf survival (hidiroglou 1979, hidiroglou and knipfel 1981, mcdowell 2003). to our knowledge, clinical signs of mn moose condition in wyoming – becker et al. alces vol. 46, 2010 162 and zn deficiencies have not been observed �� w��� ����� �����������. t�� ����������y �� using serum and hair to assess dietary intake of mn and zn is relatively low (smart et al. 1981, combs 1987, mcdowell 2003), but the possibility remains that deficiencies occurred �� ��� ����y ����������. the low serum p observed in the sample population in 2005 and 2007 may have been partially due to the effect of capture. although franzmann and leresche (1978) did not observe changes in serum p concentrations during their study, karns and crichton (1978) observed a decrease in p in caribou from ��� ���� �� ������� �� �������. o�� ������� techniques may have delayed sample collec� tion in some moose causing a decrease in p concentration. nonetheless, mcdowell (2003) noted that p has to be consistently below the deficiency threshold to consider a population deficient. since moose were not deficient all 3 years of the study, further investigation ap� ����� w��������. parasites and disease insignificant loads of endoparasites in fecal samples and tick counts suggested a ������v��y ��w ����������� �� w����� ���k� �� ���� �� ��� ����y ����. t�� ��w ���k ������ may have been due partially to inexperience in identifying the nymph stage which is com� mon during february. however, patterns �� ���� ������������� ��� ���� �� m���� ��� april (lankester and samuel 1997, samuel 2004) also suggest relatively low tick loads �� ����� ���� ��� �������� ���� �� ��� ����y area, whereas moose occupying winter ranges further south appear to carry higher tick loads. f���� �����v������ ��������� ���� ���w ��v�� remained longer into spring on the northern winter ranges, but disappeared rapidly to the south. snow cover during april adversely affects tick reproductive success (drew and samuel 1986) whereas warm, dry spring ���������� ��y ������� ���k ��������� ��� following autumn (samuel 2004). the 6 disease antigens did not appear to play a large role in the dynamics of moose in northwest wyoming. brucellosis seropreva� lence in elk was 12.5% in the buffalo valley (barbknecht 2008), thus there was potential for transmission on winter range. however, experimental studies of brucellosis in moose �������� ���� ���y ��y �� � �������� ���� ��� ��� ������� ������� ��������� ����� �� ����� mortality (forbes et al. 1996). deaths as� �������� w��� ����������� ��������� ��v� ��� been observed among moose in the greater yellowstone ecosystem (cook and rhyan 2003), but due to the rapid progression of the �������, �������� �y������ �� ��������� ��y ��� �� �����v�� ����� �� �����. acknowledgements funding was provided by teton county conservation district, wyoming animal damage management board, wyoming de� partment of transportation, wyoming game and fish department, and wyoming gover� nor’s big game license coalition/wildlife heritage foundation of wyoming. we thank bridger-teton national forest, grand teton national park, wyoming game and fish department, and yellowstone national park for logistical support and our pilots g. lust (mountain air research [retired]), d. savage (savage air services), and d. stinson (sky aviation) for their expertise in the air. we wish to acknowledge the efforts of numerous personnel who assisted with field, office, and logistical support, especially c. anderson, d. brimeyer, s. dewey, t. fuchs, h. harlow, w. hubert, s. kilpatrick, t. kreeger, f. lindzey, w. long, s. smith, and t. thurow. we also ����k w. e�w����, t. c������, m. r������k, r. siemion, and others from the wyoming state veterinary laboratory for diagnostic analyses and guidance. t. schuff provided assistance with ultrasonography and c. c. schwartz was always available for thoughtprovoking discussions about animal nutrition. a. e. b���k�����, e. j. w���, ��� ������ ��v� alces vol. 46, 2010 becker et al. – moose condition in wyoming 163 improved this manuscript through their con� �������v� ��v��w�. references arnemo, j. m., t. j. kreeger, ��� t. soveri. 2003. chemical immobilization of freeranging moose. alces 39: 243-253. ballard, w. b., p. j. macquarrie, a. w. franzmann, ��� p. r. krausman. 1996. e������ �� w������ �� ��y����� ��������� �� ����� �� ������������� a���k�. a���� 32: 51-59. barbknecht, a. e. 2008. ecology of elk parturition across winter feeding oppor� �������� �� ��� ����������� ������� ���� of wyoming. m.s. thesis, iowa state university, ames, iowa, usa. barboza, p. s., e. p. rombach, j. e. blake, ��� j. a. nagy. 2003. c����� ������ of muskoxen: a comparison of wild and �����v� �����������. j������ �� w������� diseases 39: 610-619. becker, s. a. 2008. habitat selection, ���������, ��� ���v�v�� �� s����� ����� in northwest wyoming. m.s. thesis, university of wyoming, laramie, wyo� ming, usa. bertram, m. r., ��� m. t. vivion. 2002. m���� ��������y �� ������� �������� alaska. journal of wildlife management 66: 747-756. blakley, b. r., j. c. haigh, ��� w. d. mccarthy. 1992. concentrations of copper �� ������� �� w����� ������ �� s��k����� �w��. t�� c������� v��������y j������ 33: 549-550. boertje, r. d., kellie, k. a., c. t. seaton, m. a. keech, d. d. young, b. w. dale, l. g. adams, ��� a. r. aderman. 2007. ranking alaska moose nutrition: sig� nals to begin liberal antlerless harvests. journal of wildlife management 71: 1494-1506. brimeyer, d. g., ��� t. p. thomas. 2004. his� tory of moose management in wyoming and recent trends in jackson hole. alces 40: 133�143. cherel, y., j. p. robin, a. heitz, c. calgari, ��� y. l. maho. 1992. relationships ���w��� ����� �v���������y ��� ������� utilization during prolonged fasting. journal of comparative physiology b 162: 305�313. combs, d. k. 1987. hair analysis as an indica� ��� �� ������� ������ �� ��v�����k. j������ of animal science 65: 1753-1758. cook, r. c., j. g. cook, d. l. murray, p. zager, b. k. johnson, ��� m. w. gratson. 2001. d�v�������� �� ��������v� ������ �� ����������� ��������� ��� r��ky m���� tain elk. journal of wildlife management 65: 973-987. _____, t. r. stephenson, w. l. meyers, j. g. cook, ��� l. a. shipley. 2007. validating ��������v� ������ �� ����������� ������ ���� ��� ���� ����. j������ �� w������� management 71: 1934-1943. cook, w., ��� j. rhyan. 2003. b���������� vaccines and non-target species. pages 6165 in t. j. kreeger, editor. brucellosis in e�k ��� b���� �� ��� g������ y����w����� area. greater yellowstone interagency brucellosis committee, 17-18 september 2002, jackson, wyoming, usa. custer, t. w., e. cox, ��� b. gray. 2004. trace elements in moose (alces alces) ����� ���� �� �����w������ m��������, usa. the science of the total environ� ment 330: 81-87. drew, m. l., ��� w. m. samuel. 1986. r����������� �� ��� w����� ���k, dermacentor albipictus, ����� ����� ���������� �� a������, c�����. c������� j������ �� zoology 64: 714-721. flynn, a., ��� a. w. franzmann. 1987. m������ ������� ������� �� n���� a����� ��� �����. sw����� w������� r������� (supplement) 1: 289-299. _____, _____, p. d. arneson, ��� j. l. oldemeyer. 1977. indications of copper ���������y �� � ������������� �� a���k�� moose. journal of nutrition 107: 1182moose condition in wyoming – becker et al. alces vol. 46, 2010 164 1189. forbes, l. b., s. v. tessaro, ��� w. lees. 1996. experimental studies on brucella abortus in moose (alces alces). j������ of wildlife diseases 32: 94-104. frank, a., v. galgon, ��� l. r. petersson. 1994. secondary copper deficiency, �������� ���������y ��� ����� ������� imbalance in the moose (alces alces l.): effect of anthropogenic activity. ambio 23: 315-317. franzmann, a. w. 1977. condition assess� ment of alaskan moose. proceedings of ��� n���� a������� m���� c��������� and workshop 13: 119-127. _____. 1985. assessment of nutritional status. pages 239-259 in r. j. hudson and r. g. white, editors. bioenergetics of wild herbivores. crc press, boca raton, florida, usa. _____, a. flynn, p. d. arneson, ��� j. l. oldemeyer. 1974. monitoring moose mineral ���������� v�� ���� ������� ����y���. proceedings of the north american moose c��������� ��� w��k���� 10: 1�21. _____, ��� r. e. leresche. 1978. alaskan ����� ����� ������� w��� �������� �� ��������� �v��������. j������ �� w������� management 42: 334-351. _____, ��� c. c. schwartz. 1985. moose twinning rates: a possible population ��������� ����������. j������ �� w������� management 49: 394-396. _____, _____, ��� d. c. johnson. 1987. moni� toring status (condition, nutrition, health) �� ����� v�� �����. sw����� w������� research (supplement) 1: 281-287. gogan, p. j. p., d. a. jessup, ��� m. akeson. 1989. copper deficiency in tule elk at point reyes, california. journal of range management 42: 233-238. haigh, j. c., r. r. stewart, g. wobeser, ��� p. s. macwilliams. 1977. capture �y�����y �� � �����. j������ �� ��� a������� v��������y m������ a���������� 171: 924-926. heard, d., s. barry, g. watts, ��� k. child. 1997. fertility of female moose (alces alces) in relation to age and body com� position. alces 33: 165-176. hidiroglou, m. 1979. trace element de� ���������� ��� ��������y �� ���������: � ��v��w. j������ �� d���y s������ 62: 1195-1206. _____, ��� j. e. knipfel. 1981. maternal-fetal relationships of copper, manganese, and ������ �� ���������. a ��v��w. j������ �� dairy science 64: 1637-1647. houston, d. b. 1968. the shiras moose in jackson hole, wyoming. technical bul� letin no. 1. grand teton natural history association, moose, wyoming, usa. _____. 1969. a note on the blood chemistry of shiras moose. journal of mammalogy 50: 826. karns, p. d., ��� v. f. j. crichton. 1978. effects of handling and physical restraint �� ����� ���������� �� w������� �������. journal of wildlife management 42: 904-908. keech, m. a., r. t. bowyer, j. m. ver hoef, r. d. boertje, b. w. dale, ��� t. r. stephenson. 2000. l��� ������y ������ quences of maternal condition in alaskan moose. journal of wildlife management 64: 450�462. _____, t. r. stephenson, r. t. bowyer, v. van ballenberghe, ��� j. m. ver hoef. 1998. relationships between blood-serum v�������� ��� ����� �� ���� ��� �� a���k�� moose. alces 34: 173-179. knight, d. l. 1994. mountains and plains: the ecology of wyoming landscapes. yale university press, new haven, con� necticut, usa. kock, m. d., r. k. clark, c. e. franti, d. a. jessup, ��� j. d. wehausen. 1987. effects of capture on biological parameters in freeranging bighorn sheep (ovis canadensis): �v�������� �� ������, �������� ��� ���� �����y �������� ��� ������������� �� ����������� ���v�v��. j������ �� w������� alces vol. 46, 2010 becker et al. – moose condition in wyoming 165 d������� 23: 652�662. kreeger, t. j. 2000. xylazine-induced aspira� ���� ��������� �� s����� �����. w������� society bulletin 28: 751-753. _____, w. h. edwards, e. j. wald, s. a. becker, d. brimeyer, g. fralick, ��� j. berger. 2005. health assessment of shiras moose immobilized with thiafen� tanil. alces 41: 121-128. kubota, j., s. rieger, ��� v. a. lazar. 1970. mineral composition of herbage browsed �y ����� �� a���k�. j������ �� w������� management 34: 565-569. lankester, m. w., ��� w. m. samuel. 1997. pests, parasites, and diseases. pages 479-517 in a. w. franzmann and c. c. schwartz, editors. ecology and management of north american moose. smithsonian institute press, washington, d.c., usa. mcardle, h. j. 1995. the metabolism of copper during pregnancy – a review. food chemistry 54: 79-84. mcdowell, l. r. 2003. m������� �� a����� and human nutrition. second edition. e���v��� s������ b.v., a��������, n����������. mcjames, s. w., j. f. kimball, ��� t. h. stanley. 1994. immobilization of moose with a-3080 and reversal with nalmefene hcl or naltrexone hcl. alces 30: 21-24. messier, f., j. huot, f. goudreault, ��� a. v. tremblay. 1987. reliability of blood ���������� �� ������ ��� ����������� ������ of caribou. canadian journal of zoology 65: 2413�2416. miquelle, d. g., ��� p. a. jordan. 1979. t�� ���������� �� ��v�����y �� ��� ���� �� moose. proceedings of the north ameri� ��� m���� c��������� ��� w��k���� 15: 54-79. murray, d. l., e. w. cox, w. b. ballard, m. s. lenarz, t. w. custer, t. barnett, ��� t. k. fuller. 2006. pathogens, ����������� ���������y, ��� ������� ������ ences on a declining moose population. wildlife monographs 166. o’hara, t. m., g. carroll, p. barboza, k. mueller, j. blake, v. woshner, ��� c. willetto. 2001. m������ ��� ���vy ����� ������ �� ������� �� � ��������y �v��� ��� ���� ����������� �� � ����� ���������� �� a���k�. j������ �� w������� d������� 37: 509-522. ohlson, m., ��� h. staaland. 2001. m������ ��v�����y �� w��� ������: �������� ��� ���� for moose. oikos 94: 442-454. oldemeyer, j. l., a. w. franzmann, a. l. brundage, p. d. arneson, ��� a. flynn. 1977. browse quality and the kenai ����� ����������. j������ �� w������� management 41: 533-542. puls, r. 1994. mineral levels in animal health: diagnostic data. sherpa inter� ��������, c��������k, b������ c�������, c�����. rombach, e. p., p. s. barboza, ��� j. e. blake. 2003. costs of gestation in an arctic ruminant: copper reserves in muskoxen. comparative biochemistry and physiol� ogy part c 134: 57-168. samuel, w. m. 2004. w���� a� � g����: w����� t��k� ��� m����. f��������� �� a������ n����������. e�������, a������, c�����. schwartz, c. c., s. l. cain, s. podruzny, s. cherry, ��� l. frattaroli. 2010. c��� trasting activity patterns of sympatric and allopatric black and grizzly bears. journal of wildlife management 74: in press. seal, u. s., ��� r. l. hoskinson. 1978. m�������� ���������� �� ������� ��������� and capture stress in pronghorns. journal of wildlife management 42: 755-763. _____, m. e. nelson, l. d. mech, ��� r. l. hoskinson. 1978. metabolic indicators �� ������� ����������� �� ���� m�������� ���� �����������. j������ �� w������� management 42: 746-754. smart, m. e., j. gudmundson, ��� d. a. christiensen. 1981. trace mineral defi� �������� �� ������: � ��v��w. t�� c������� moose condition in wyoming – becker et al. alces vol. 46, 2010 166 veterinary journal 22: 372-376. stephenson, t. r. 2003. physiological ecol� ogy of moose: nutritional requirements ��� ������������ w��� ������� �� ���y ��������� ����������. f������ a�� �� wildlife restoration research final per� formance report. project 1.52. alaska d��������� �� f��� ��� g���, j�����, alaska, usa. _____, k. j. hundertmark, c. c. schwartz, ��� v. van ballenberghe. 1993. ultra� sonic fat measurements of captive yearling bull moose. alces 29: 115-123. _____, _____, _____, ��� _____. 1998. pre� dicting body fat and body mass in moose with ultrasonography. canadian journal of zoology 76: 717-722. stewart, r. r., ��� a. flynn. 1978. mineral ������� ��v��� �� s��k�����w�� ����� hair. proceedings of the north american m���� c��������� ��� w��k���� 14: 141�156. testa, j. w., ��� g. p. adams. 1998. body condition and adjustments to reproduc� tive effort in female moose (alces alces). journal of mammalogy 79: 1345-1354. wallisdevries, m. f. 1998. habitat quality and the performance of large herbivores. pages 275-320 in m. f. w�����d�v����, j. p. bakker, and s. e. van wieren, editors. grazing and conservation management. kluwer academic publishers, dordrecht, n����������. weber, b. j., m. l. wolfe, g. c. white, ��� m. m. rowland. 1984. physiologic re� sponse of elk to differences in winter range quality. journal of wildlife management 48: 248-253. williams, e. s., ��� e. t. thorne. 1996. exertional myopathy (capture myopathy). pages 181-193 in a. f����������, l. n. locke, and g. l. hoff, editors. nonin� �������� d������� �� w�������. s����� edition. iowa state university press, ames, iowa, usa. alces 31_93.pdf alces 31_145.pdf alces 31_221.pdf alces30_1.pdf alces34(1)_125.pdf alces30_71.pdf alces30_173.pdf alces30_117.pdf alces36_61.pdf alces30_51.pdf alces34(1)_173.pdf alces vol. 46, 2010 gardner reducing moose capture in wolf snares 167 reducing non-target moose capture in wolf snares craig l. gardner alaska department of fish and game, 1300 college road, fairbanks, ak 99701-1599, usa. abstract: i investigated the characteristics of moose (alces alces) bycatch in kill snares set for wolves (canis lupus) in interior and south-central alaska, usa. my objective was to design a kill snare that would reduce moose vulnerability and injury if captured without reducing its effectiveness for capturing wolves. i documented at close range (<30 m) snare encounters by captive moose in natural habitat at the kenai moose research center (mrc) in south-central alaska. moose contacted 153 cm or 183 cm snares (n = 184) with their chest–shoulder area (59.8%), neck-head region (34.2%), upper legs (3.8%), and along the ribs (2.2%). i documented the fate of moose following 225 snare contacts; 13.8% were captured by the nose (5.8%), leg (4.9%), or unknown (3.1%) with the remainder either knock-downs (65.3%) or push-asides (21.0%). moose did not attempt to avoid snares. of the 147 knock-downs, 86.4% formed a loop 15-38 cm in diameter that laid near the snow surface continuing to present a potential trap for moose. i also evaluated capture rates by loop size for wild moose in 3 study areas in interior alaska. capture rate and type were not influenced by snare loop size or snow depth in the wild or the mrc. capture vulnerability of wild and captive moose was higher in snares that were knock-downs by other moose or wind. i subsequently developed a snare that incorporated an additional wire (diverter) placed at a height that allowed moose or any ungulate taller than the set height of a wolf snare to contact and push the snare away prior to contact. this design reduced the vulnerability of moose but not wolves to capture. i also placed a cinch stop at 24.1-26.7 cm from the end stop of the snare loop to reduce injury to moose and act as a breakaway system without reducing the snare’s effectiveness for capturing wolves. results of this study are applicable to areas where wolf or coyote (canis latrans) snaring occurs in the presence of moose and other large hoofed mammals. alces vol. 46: 167-182 (2010) key words: accidental capture, alaska, alces alces, breakaway snares, canis lupus, moose vulnerability, snare effectiveness, snare efficiency, trapping, wolf snares, wolves. kill snares are an effective trap to catch wolves (canis lupus), lynx (lynx canadensis), fox (vulpes vulpes), and coyotes (canis latrans) (phillips 1996, roy et al. 2005, blejwas 2006), and are used throughout alaska (usa), canada, and russia. although snares were found to be 10 times more selective than foothold traps for coyotes and lynx (guthery and beasom 1978), incidental captures occur (proulx et al. 1994). furthermore, wolf snares can be even less selective than snares set for smaller furbearers because cable diameter and loop circumference are larger, set height is higher, and the size and strength of a wolf require that minimum breaking forces must be high. historically, the problem of snares not being selective has been a concern for wildlife managers and trappers (phillips 1996), resulting in areas closed to snaring throughout north america (shivik and gruver 2002) due to concerns that indiscriminate capture could negatively impact other wildlife populations. also, public pressure exists to improve snare selectivity (traps, trapping, and furbearer management, the wildlife society technical review 90-1, 1990) and this is an issue addressed by the international program best management practices (bmp) for regulated trapping conducted by the international association of fish and wildlife agencies. moose (alces alces), caribou (rangifer tarandus), and sitka black-tailed deer (odoreducing moose capture in wolf snares gardner alces vol. 46, 2010 168 coileus hemionus sitkensis) are caught in wolf snares every year in alaska (gardner 2007). in separate 5-year studies using radiocollared moose (75-125 active radios/yr), boertje et al. (2009) and m. a. keech (alaska department of fish and game (adfg), unpublished data, fairbanks) documented 0-3 moose killed/yr in wolf snares (0.5%/yr). wolf trapping was common in both study areas with snaring the preferred capture method. based on my 15 years of experience releasing nearly 40 moose from snares and discussions with other alaskan biologists, i concluded that most moose restrained in wolf snares die either at the capture site or from frozen limbs or nose subsequent to release. for example, steve dubois (adfg, personal communication) radio-collared and released 4 moose caught in snares that were without obvious injury, yet died 2 days later. although necropsies were not performed, the timing of deaths indicates that death was probably due to complications associated with restraint in the wolf snare. previous studies found that accidental ungulate catch in coyote snares could be reduced through trapper education and use of snares with improved selectivity (phillips et al. 1990, phillips 1996, roy et al. 2005). in alaska, development and testing of wolf snares designed to release moose and caribou, but restrain wolves, has been ongoing since 1993 by adfg and private trappers. one difficulty in designing a breakaway wolf snare is the tradeoff between achieving desired selectivity and maintaining acceptable efficiency for wolves, because wolves and moose exert powerful forces on the snare when captured. two prototypes, the thompson split lock (thompson snares 2009) used with 0.28 cm diameter cable and the camlock soft pin breakaway designs (fig. 1), showed promise in the laboratory and were used as part of a wolf control program by adfg in 1993-1994. during the program 30 wolves, 9 moose, and 5 caribou were caught in snares with the thompson split lock breakaway mechanism. of these, 29 wolves (96.7%), 6 moose (66.7%), and 3 caribou (60.0%) did not escape. i evaluated these data using fisher’s exact tests (fet) and found that the release rate was higher for moose (p = 0.03) and caribou (p = 0.047) than wolves; however, the restraining rate of moose and caribou remained unacceptably high. three wolves were caught by the camlock soft pin design and 1 escaped due to the mechanism release; no moose or caribou were caught by this design. alaska trappers continued to improve ungulate release from wolf snares with a fig. 1. common breakaway mechanisms used by alaskan trappers on wolf snares: (a) the thompson split lock, (b) camlock soft pin, and (c) s-hook. a b c alces vol. 46, 2010 gardner reducing moose capture in wolf snares 169 variety of breakaway mechanisms, most commonly a thompson split lock used on a smaller diameter cable (0.24 cm) or s-hooks with varying breakaway strengths (fig. 1). a trapper survey conducted by adfg (blejwas 2006) suggested that these breakaway systems worked for leg-caught moose, unless the moose had entangled the snare wire around flexible brush and could not generate enough force to break the release mechanism; moose caught by the nose or neck rarely broke free. moose that remain restrained were vulnerable to injury and death due to freezing limbs at the snare attachment point. these deficiencies illustrated the need for a wolf snare design that minimized moose capture, particularly by the nose, and reduced the chance of injury when the breakaway mechanism failed. these findings were consistent with results from studies that evaluated breakaway snare performance for capturing coyotes and releasing deer (odocoileus hemionus and odocoileus virginianus; phillips et al. 1990, phillips 1996, roy et al. 2005). roy et al. (2005) documented 74-88% release rates of deer using snares with the national 813 s-hook as the breakaway device. deer that remained restrained were mostly fawns and all were caught by the neck. phillips et al. (1990) found that coyotes and deer fawns generated similar force on a snare and concluded it would be difficult to design a system that released all deer yet restrained coyotes. previous efforts to reduce the accidental restraint of moose in wolf snares and other ungulates in coyote snares were to design breakaway systems that allow these ungulates to escape. although completely eliminating moose capture by wolf snares is improbable, snares could be made more selective and humane if differences in behavior or physical stature of moose related directly to modifications that reduced their capture vulnerability. accounting for behavioral differences proved beneficial in reducing incidental capture by other snare types (proulx et al. 1994). my primary objective was to design a wolf snare that would be less accessible to moose and contain a breakaway system that would minimize injury without reducing its effectiveness for catching wolves. snare effectiveness for any new design needs to be consistent with current designs to be accepted by trappers (naylor and novak 1994). i took an innovative approach by directly observing hundreds of moose–snare encounters at close range (<30 m) in natural habitat to develop and test snare designs. my original hypotheses were: 1) wolf snare loop-size affects moose vulnerability to capture, 2) moose were equally vulnerable to being caught by the nose or leg in wolf snares, and 3) moose became more vulnerable to wolf snares as snow depth increases. the primary contributions of this study to wildlife research and management are: 1) demonstrating that repeated direct observations of ungulate–snare encounters are invaluable for designing effective snares that minimize the chance for bycatch of ungulates, 2) the importance of reducing vulnerability to capture and incorporating an effective breakaway mechanism, and 3) the development of a wolf snare that will likely protect moose and other ungulates from being captured without significantly reducing effectiveness for wolf capture. results from this study will benefit the ongoing bmp process and be directly relevant to areas throughout the world that have wolves, large ungulates, and wolf trapping with kill snares. study area i field tested various designs of wolf snares on captive moose at the kenai moose research center (mrc) in south-central alaska and wild moose on the tanana river flats in game management unit (gmu) 20a in interior alaska (fig. 2). the mrc allowed me to observe 100s of moose–snare encounters in a relatively short period of time, while in gmu 20a i evaluated snares in habitat and circumstances directly comparable to wolf reducing moose capture in wolf snares gardner alces vol. 46, 2010 170 trapping in interior alaska. the primary overstory–shrub vegetation types at the mrc were paper birch (betula paperifera), alder (alnus crispa), willow (salix spp.), and spruce (picea mariana and p. glauca). snow depths were 10-15 cm in february 2005 and 40-50 cm in january 2007. trappers commonly set snares in these vegetative types and snow conditions in south-central alaska. i tested snares in 3 areas within central gmu 20a that supported high moose densities (>800/1,000 km2; boertje et al. 2007) and were adjacent to areas trapped commonly for wolves. the primary vegetative types in the dry creek area were dwarf birch (betula nana), willow, alder, and paper birch; spruce, paper birch, willow, dwarf birch, and alder were the common overstory-shrub species in the clear and mcdonald creek areas. these areas were representative of habitats and climates commonly trapped in interior alaska (gasaway et al. 1983, 1992). snow depth was reasonably similar in the 3 areas during snare testing, ranging 28 (december 2005)-56 cm (march 2006). methods moose vulnerability to wolf snares on 1-4 february 2005 and 6-9 january 2007, i observed moose-wolf snare encounters at the mrc by setting 153 cm and 183 cm wolf snares in areas that maximized the chance that moose would encounter the snare (areas of highest moose use), but in a manner that mimicked typical snare sets for wolves. i used these loop sizes because they are the most commonly used in alaska, are effective in catching wolves by the neck, and are the most readily available from commercial snare dealers. i set the snares following methods used by successful wolf trappers including dying and boiling the snares and setting them in a manner that they blended with the surrounding vegetation. each set included 1-24 snares, closely divided between 153-cm and 183-cm loop sizes; 3-10 moose were monitored daily. when a group of moose moved beyond observation, i pulled the snares and reset them in another area. i simulated the standard method of alaskan wolf trappers (alaska trappers association 2007) by setting both 153 and 183 cm circumference loop snares at 46 cm above the supportive surface of the snow. this height has proved effective in promoting neck catches by causing wolves to contact the bottom of the loop with their chest. i recorded the initial contact point of a moose encountering a snare and described the characteristics of the encounter including snare loop size, snow depth, fate, and moose reaction. i categorized the fate of a moosesnare encounter as knock-down, push-aside, or caught. a knock down occurred when a moose contacted the snare and caused it to drop from its original height and form a smaller loop, pushed aside was when the moose contacted the snare but it returned to its original position and retained its loop size. to prevent restraining or injuring of moose, i modified fig. 2. study areas were located at the kenai moose research center, ca. 30 km northeast of soldotna, alaska and in game management unit 20a south of fairbanks, alaska, usa. alces vol. 46, 2010 gardner reducing moose capture in wolf snares 171 each snare by removing the nut or stop behind the lock (fig. 3). this modification allowed the test snare to cinch normally but the lock would slide off the cable quickly (<10 sec) freeing any captured animal with minimal discomfort. this approach also minimized learned behavior effects. i compared moose capture rates (moose caught/encountered snare) and capture types (nose, neck, and leg) between wolf snares with 153-cm and 183-cm loop sizes at 2 different snow depths (46 cm and 10 cm). for each catch, i recorded the snare loop size and capture type. initially i would reset the snare attempting to increase encounters and captures. however, there were incidences when a different moose would encounter a previously knocked-down snare and become caught by the leg. to examine the capture rate in previously knocked-down snares (another moose or wind), i recorded the circumference and position of the resulting loop following 18 knock-downs and evaluated the vulnerability of subsequent moose contacting the fallen snare. from 30 december 2005-31 march 2006, i set and monitored 34 153-cm and 30 183cm circumference loop snares divided among the 3 study sites (8-12 of each type/site) in gmu 20a. i purposely set individual snares along natural trails (simulated trail set) or in a gang set with 6-11 snares blocking off most of the natural trails in a 30 m radius (simulated bait-kill set). both snare sizes were placed together, but not always in equal numbers, to evaluate moose capture rates by snare loop size. to be consistent with check times followed by most alaskan trappers in areas without a defined check period, i waited at least 7 days and as long as 21 days due to periods of severe bad weather. using tracks in the snow and position of the snare and lock in relation to the original set, i determined if a snare was encountered by a moose and was either knock-down, push-aside, or had caught the moose. snare modifications to reduce moose capture by the nose i used the results from the moose-snare encounter tests conducted at the mrc to design a wolf snare that reduced moose vulnerability to capture. i attached a 2.30 mm diameter “diverter wire” to standard 153-cm wolf snares so that it extended 70 cm perpendicular to the plane defined by the snare loop, at an angle 10-20o from the horizontal plane tangent to the top of the snare (fig. 4). the intent was for a moose to contact the wire with its nose or chest, and push the snare away before its nose entered the noose. length of the diverter wire was based on measuring the distance from tip of nose to chest on 3 taxidermy-mounted adult male moose (≥6 yr). i used the longest measurement (70 cm) to ensure that a moose would contact the diverter before the snare. a b fig. 3. test snare without a cable end stop (a) that allows the lock to slide off the cable if an animal is caught to prevent injury, and a snare that includes an end stop (b). reducing moose capture in wolf snares gardner alces vol. 46, 2010 172 i compared capture rates and types between the diverter test snare design and 153-cm and 183-cm loop standard snares by setting diverter snares alongside these snares. since the snares in gmu 20a were not checked for 7-21 days, the number of days that a snare was a knock-down and could potentially capture moose was unknown. i tested if diverter snares would be more prone to being knock-downs by wind or snow due to the additional wire because increased knock-down rates would reduce snare efficiency for wolves and possibly increase vulnerability of moose to leg capture. i compared the knock-down rate between diverter snares and standard 153-cm and 183-cm snares due to wind in the clear creek and mcdonald creek study areas in gmu 20a. data from dry creek were not included in my analysis because the periods of observations were not aligned with those of the other 2 areas. in the clear creek area, 11-12 diverters and 10-11 standard snares (153-cm or 183-cm) were monitored for 6 periods of 8-29 days (99 total days and 2,291 trap nights). in the mcdonald creek area, 8 diverters and 36 standard snares (153-cm or 183-cm) were monitored for 6 periods of 7-29 days (95 total days and 4,224 trap nights). period length varied due to periodic cold snaps (<-40o c for 6-13 days) that precluded safe travel. i categorized a snare as a knock-down from wind if it had dropped from its original set position if animal tracks, measurable snowfall, and high wind (snow off trees, drifting) were not evident. i timed my visits after high wind events but before subsequent snowfall. i censored the data in only 2 instances because i could not discern if wind or animals caused the knock-down. to compare selectivity and effectiveness of diverter snares in the 2006-2007 trapping season, i contracted 2 trappers in gmu 20a to use 100 diverter snares in their normal trapping activity. they were trained in data collection protocol and provided with data forms; they recorded the number of diverters set at each site, how each snare was anchored (flexible or solid anchor), species caught, and fate of captured wildlife. they also interpreted fig. 4. modified wolf snare showing the diverter wires that extend ca. 70 cm perpendicular to the snare loop at a 10-20o angle from the top of the snare. the positioning of the diverter wire allows wolves to travel underneath without contact and moose or large ungulate to contact the wires causing the snare to be pushed away from the nose. the ends are recurved to minimize chance of injury when encountered by a moose. alces vol. 46, 2010 gardner reducing moose capture in wolf snares 173 tracks to document if wolves avoided the set. location, snare anchor point, and the number of snares were not random; each trapper made decisions from site-specific wolf sign and available vegetation to anchor the snare. snare modifications to reduce injury and death for leg-caught moose to reduce injury to leg-caught moose and other ungulates, i investigated the possibility of incorporating a cinch stop that would prevent the snare from cinching tight on a moose leg but not reduce the effectiveness in killing neck-caught wolves. i selected the placement of the cinch stop by comparing loop sizes that killed trapper-caught wolves by the neck (n = 62) with the circumferences of loops cinched on hunter-killed moose legs (n = 9). i also asked trappers to record sex of wolves and if practical, provide the carcass or front leg to age wolves (pup, adult) using the epiphyseal closure on the radius and ulna (rausch 1967). trappers in gmu 20a also caught known-aged wolves marked in another study (gardner and beckmen 2008). to determine if the cinched down loop size differed due to snare cable size or sex and age of the wolf, i compared final loop circumferences of 0.24, 0.28, and 0.32 cm diameter snare from wolf kills. wolves were classified as pups (5-11 months), subadult (17-22 months), or adult. my rationale for these analyses was if a certain size cable cinched tighter, or if the circumference of cinched loop size on certain age or gender of wolves is comparable to a moose leg, the position of the cinch stop may need to vary by cable size or not be a viable option. to determine the minimum loop size for leg-caught moose, i attached a snare cable to the front and rear legs of hunter-killed 5 month calf (n = 1), adult female (n = 4), and adult male (n = 4) moose at the most common catch point on the leg, cinched it snug but not so tight to cause injury, and measured the final loop circumference. i then tested the cinch stop snare in the laboratory by cinching the snare down on legs of a 5 month calf, an adult female moose, an adult male moose, and a simulated wolf neck (phillips et al. 1990, roy et al. 2005). i observed that if the lock contacted the cinch stop, the lock deformed (flattened out) as force was added; this led me to investigate whether this contact force would be sufficient for the cinch stop to also function as a breakaway mechanism. i hypothesized that the breaking force would be less when the lock came into contact with the cinch stop, which would occur when cinched down on a leg of a moose or a smaller ungulate, thus increasing the chance of release. i constructed the breakaway component by cutting the snare within the loop at either 24.1 or 26.7 cm from the cable end stop (circumference range of largest moose leg and smallest wolf neck) and inserting a 0.24 cm double ferrule on 0.24 cm snare cable, or 0.32 cm double ferrule on 0.28 and 0.32 cm snare cables. the ferrule was attached by swaging each end using a 0.24or 0.32-cm swage tool. each ferrule was inspected to ensure that inconsistent manufacturing was not a factor in breaking strength. i initially evaluated breaking strengths of the cinch stop breakaway (csb) mechanism in the laboratory by measuring the breaking force by cinching down csb snares until the mechanism released on a front leg collected from a female moose (circumference = 22.7 cm) and a simulated wolf neck (i.e., 27.9 cm circumference steel pipe wrapped with cotton; phillips et al. 1990, roy et al. 2005). the simulated wolf neck was 32.6 cm in circumference matching the mean neck size from 62 wolves collected from trappers; cotton was added to allow the snare cable to embed and absorb energy to better mimic when a wolf is snared by the neck, and to make it more similar to a moose leg. i measured the breaking force necessary to break the csb using a dynalink dynamometer strain gauge (model 7200; measurement system international, seattle, wa, usa) attached to a hydraulic tee reducing moose capture in wolf snares gardner alces vol. 46, 2010 174 cylinder (model sae-9012; prince manufacturing corporation, sioux city, ia, usa). i tested the csb system on 1×19 twist 0.24 cm, 0.28 cm, and 0.32 cm snare cable. each snare type was tested 20 times each on the simulated wolf neck and a moose leg. i compared the breaking strength for the csb for 0.24, 0.28, and 0.32 cm diameter cable sizes to the 0.28 cm diameter thompson split lock design field and laboratory tested during the wolf control program by adfg in 1993-1994 (adfg, unpublished data). the breaking force of the thompson split lock was determined by different researchers at adfg with the same methodology and equipment as described above. the measured breaking forces for all the tested breakaway types do not necessarily replicate the actual force that captured moose or wolves exert on a snare, but indicated possible differences that were field-tested. i first tested the efficiency of the csb mechanism for moose at the mrc in 2005 by catching 2 male moose by the leg in a natural setting. the csb was attached on a 0.28 cm 1×19 snare. i documented how moose were caught, their behavior while caught, and the elapsed time to release. the efficiency of the csb snare was further tested in the 2005-2006 trapping season by the 2 contract trappers. they set these snares under the same conditions explained for the diverter snares. to maximize the number of encounters and catches of moose and wolves, only csb snares without diverter wires were set by these trappers recognizing the possibility of nose catches. moose capture using the test snare without the end stop complied with acceptable methods for field studies adopted by the american society of mammalogists (animal care and use committee 1998, adfg protocol #0604). field testing by trappers of the diverter and breakaway snare designs as kill snares (end stop attached) followed state trapping regulations but was not included under the protocol. data analysis i used the software r® (r development core team 2008) to perform statistical analyses. i used chi-square tests (cochran 1977), or fet if any expected cell count was <10 in 2×2 contingency tables, (single degree of freedom) to identify difference in capture rate and capture type by snare type, snow depth, captive and wild moose, and to distinguish how moose initially contacted different snare types. i employed both chi-square tests and fets when expected cell counts were low as a check against the potential for the exact tests to be overly conservative (d’agostino et al. 1988). lack of balance in the experimental design precluded using generalized linear models to test for interactions due to snow depth when examining capture rates and types. to test for differences in capture type, i followed the method specified by scott and seber (1983) that accounts for the covariance associated with sampling a multinomial distribution. i used t-tests to compare breaking forces for the different breakaway mechanisms. i used generalized linear models to assess the effect of a diverter on the binary response, knock-down by wind, or not. i used quasi-aic (qaic) (lebreton et al. 1992) and likelihood ratio tests to compare these models and present the goodness-of-fit metric, ĉ. results moose vulnerability to wolf snares i documented 304 moose–snare encounters at mrc through direct observation or from tracks in the snow and found no evidence that moose modified their behavior due to the presence of snares; moose did not shy away or abruptly change course when encountering a snare. i observed 184 encounters between moose and standard wolf snares; the impact points were the chest-shoulder area (59.8%; se = 3.6%), neck-head (34.2%, se = 3.5%), legs (3.8%, se = 1.4%), or ribs (2.2%; se = 1.1%) (table 1). i documented the fate through observation and by tracks alces vol. 46, 2010 gardner reducing moose capture in wolf snares 175 of 225 moose-snare encounters; 65.3% (se = 3.2%) were knock-downs, 20.9% (se = 2.7%) were push-asides, and 13.8% (se = 2.3%) were caught moose (table 2). snare impact points were not related to snare loop size. for 183-cm snares, moose initially contacted their neck-head area 37.5% (se = 5.7%) of the time, similar to the initial contact rate of 32.1% (se = 4.4%; table 1) for 153-cm snares (χ2 = 0.56, p = 0.46). capture rate was not affected by snare loop size (χ2 = 1.31, p = 0.25; table 2); capture rates of the 153and 183-cm loop snares were 17.3% (n = 84, se = 4.2%) and 11.8% (n = 147, se = 2.7%), respectively. snow depth did not influence capture rate (p = 0.83, fet; table 3). capture rates of wild moose by 183-cm loop snares (15 of 35, 42.9%, se = 8.5%) were not different (χ2 = 0.99, p = 0.32) from that with 153-cm loop snares (12 of 38, 31.6%, se = 7.6%). capture rate of wild moose (27 of 73, 37.0%, se = 5.7%) was higher than that of captive moose (31 of 225, 13.8%, se = 2.3%; χ2 = 18.9, p <0.001). i was able to determine capture type in snares encountered at the original set height for 24 of 31 (77.4%) moose caught at the mrc; 54% (se = 10.4%) were caught by the nose and 46% (se = 10.4%) by the leg (table 2). unobserved captures occurred due to the short time necessary to escape the test snare, as well as attempting to observe multiple moose simultaneously. all nose catches occurred in snares encountered at impact point contacts chest–shoulder neck–head ribs legs snare type n n % n % n % n % 153-cm loop 112 67 59.8 36 32.1 4 3.6 5 4.5 183-cm loop 72 43 59.7 27 37.5 0 0 2 2.8 subtotal 184 110 59.8 63 34.2 4 2.2 7 3.8 diverter 23 17 73.9 6 26.1 0 0 0 0 table 1. observed impact points where captive moose initially contacted 153-cm loop, 183-cm loop, and diverter wolf snares (n = 207). this phase of the study was conducted at the kenai moose research center, alaska, february 2005 and january 2007. fate capture typea snares encountered knockeddown pushed-aside caught nose leg snare type n % n % n % # %b # %b 153-cm loop 144 104 72.2 23 16 17 11.8 7 5 5 3.6 183-cm loop 81 43 53.1 24 29.6 14 17.3 6 7.6 6 7.6 subtotal 225 147 65.3 47 20.9 31 13.8 13 6 11 5 knock-down (153 and 183-cm snares)c 18 n/a n/a 6 33.3 0 0 6 0 diverter 42 40 95.2 2 4.8 0 0 0 0 0 0 diverter knock-downc 19 n/a n/a 0 0 0 0 0 0 table 2. capture rate and type in 153-cm, 183-cm, and diverter wolf snares measured by observing captive moose at the kenai moose research center, alaska, february 2005 and january 2007. a the sample size of capture type is less than # caught because all captures were not observed. b percent capture determined without including unknown capture types. c snares that were previously knocked down but left until another moose encounter occurred. reducing moose capture in wolf snares gardner alces vol. 46, 2010 176 original height; all leg catches occurred in a knock-down when a moose stepped in with its front (n = 3) or hind foot (n = 8). the proportion of nose and leg catches did not differ (pleg – pnose = –0.08; 95% ci = –0.48, 0.32; n = 24). capture type did not depend on snare loop size (p = 1, fet; table 2) or snow depth (p = 0.38, fet; table 3). moose were caught more frequently by knock-downs from another moose or wind (6 of 18, se = 11%; table 2) than snares encountered at original set height (31 of 225, se = 2.3%; p = 0.04, fet); leg captures occurred only in previous knock-downs. at the mrc, 86.4% (102 of 118, se = 3.2%) of knock-downs by moose formed loops 15-38 cm in circumference, remaining in the trail at snow level and available for leg captures. there was no difference in the number of knock-downs of 153-cm (6 of 74, se = 3.2%) and 183-cm snares (0 of 36; p = 0.17, fet) forming loops <15 cm. snare modification to reduce moose capture i observed 23 moose-diverter snare encounters at the mrc and the impact points were either at the chest-shoulder (73.9%) or neck-head area (26.1%; table 1). based on observations and tracks, moose contacting a diverter wire caused knock-downs in 40 of 42 cases (95.2%) with 2 push-asides (4.8%; table 2). no moose contacting a diverter snare (n = 42) was caught, and the capture rate was less than that for standard snares (p = 0.007, fet; table 2). assuming the next encounter with a diverter snare would result in a capture, the capture rate for the diverter snares would have remained lower than that for standard snares (p = 0.04, fet). moose knocked down diverter snares more frequently than standard snares (p < 0.001, fet), and once knocked down, 85.0% formed 15-38 cm circumference loops on the snow. due to the high knock-down rate, i hypothesized that moose would be more vulnerable to leg catches in diverter snares; however, no moose at the mrc was caught in a knock-down from diverter snares (n = 19) compared to 6 of 18 caught in knock-downs from standard snares (p = 0.008, fet). i observed 6 knock-downs from diverter snares contacted by moose, and in all cases the diverter wire was still contacted first causing the snare loop to move away. encounters of 1-2 additional contacts caused no damage to the diverter wire. the capture rate of wild moose in diverter snares (without a cable end stop) was 12.1% (7 of 58) in gmu 20a. as snares were unattended, i was not able to determine capture types and the frequency of encounter for knockdowns of diverter snares. diverter wires on the 7 snares that caught moose were bent and no longer functional, but i could not confirm if this damage was preor post-capture. moose were only caught in diverter snares unchecked 12-21 days; no moose were caught in snares unchecked 7-11 days. standard test snares set in gmu 20a caught moose more frequently (27 of 73) than diverter modified snares (p = 0.002, fet). the 2 contracted trappers caught and killed 9 wolves by the neck after setting 96 diverter snares in gmu 20a in december 2005-march 2006. no moose encountering snow type encounters catch rate (%) nose catch (%) neck catch (%) leg catch (%) unknown deep snowa 218 12.8 9 (4.3) 0 12 (5.7) 7 shallow snowb 62 11.3 4 (6.6) 0 2 (3.3) 1 table 3. catch rate and catch type of captive moose in standard wolf snares at 2 snow depths at the kenai moose research center, alaska, february 2005 and january 2007. a snow depth ca. 46 cm. b snow depth ca. 10 cm. alces vol. 46, 2010 gardner reducing moose capture in wolf snares 177 a diverter snare was captured (n = 9); no wolf or moose approached any other snare. based on binomial probabilities (95% confidence level), the diverter snares would catch at least 71% of wolves and prevent capture of ≥71% moose (proulx et al. 1994) diverter snares were not knocked down more by wind than standard 153and 183-cm snares. the global generalized linear model, with qaic = 95.1, indicated that area and period effects were significant or marginal (area: x1 2 = 4.5, p = 0.03; period: x4 2 = 8.9, p = 0.06), while the diverter effect and the area:period interaction were not significant (diverter: x1 2 = 1.4, p = 0.23; area:period: x4 2 = 2.8, p = 0.58). a comprehensive comparison of realistic models indicated that the best fit model included area as the only covariate (qaic = 84.4 and weight of evidence = 48.2%). the goodness of fit statistic (ĉ) was 1.7 for the best model indicating reasonable fit. snare modifications to reduce injury to moose loop circumference of cinched snares on moose legs was 23.5-24.1 cm for 3 adult males, 22.5 cm for 1 yearling male, 20.922.7 cm for 4 adult females, and 19.7 cm for 1 calf; average cinch size was 22.4 cm (sd = 0.32). the average loop circumference of neck-caught wolves (n = 62) was 32.6 cm (sd = 2.48, range = 26.7-38.7); the smallest was on a 5 month old female (22.7 kg). the cinch stop could be placed 22.7-26.7 cm from the cable end stop based on the age (subadult/ad) and sex of 31 of these wolves; therefore, i placed the cinch stop at either 24.1 cm or 26.7 cm for testing. the breakaway force required to release the csb mechanism depends upon snare cable size, circumference of the cinched loop, and proximity of the lock to the csb mechanism (table 4, fig. 5). on a moose leg, the cinched loop stopped at the cinch stop as the lock contacted the mechanism. on the simulated wolf neck, the cinched loop size was 32.6 cm and the lock stopped 5.9-8.5 cm from the csb. the breakaway force was higher on the simulated wolf neck than the moose leg, increased with cable size, and decreased when the csb mechanism was placed further from the cable end stop (p ≤ 0.01; table 4). the breaking force for csb equipped snares was less than the breaking force of the 0.28-cm snares with a thompson split-lock (325.4 kg; se = 8.2, p < 0.001), regardless of cable size and csb placement (table 4). during the initial field test a 12 year and a 3 year old male moose were caught at the mrc in a csb snare with the mechanism placed at 24.1 cm and attached with solid anchor. the 3 year old male was caught by the hind foot and broke free in <2 sec; the 12 year old male was caught by the front leg and broke free in 2 min and 21 sec. upon capture, the 12 year old male tangled the snare wire around surrounding flexible shrubs preventing it from pulling directly against the solid anchor; the lock was tight against the breakaway mechanism but the snare loop rotated around the foot. after inspecting the leg and verifying that the restraining loop caused no injury, i determined that the design was adequate for further testing by the 2 contract trappers. the contract trappers set 24.1 cm (n = 150 175 200 225 250 275 300 325 350 375 22 24 26 28 30 32 34 36 final cinch circumference (cm) b re ak in g st re ng th ( kg ) csb at 26.7 cm 0.28 cm cable csb at 24.1 cm 0.28 cm cable csb at 26.7 cm 0.24 cm cable csb at 24.1 cm 0.24 cm cable stop contact moose leg wolf neck fig. 5. comparisons of breaking strengths for a cinch stop breakaway mechanism by placement and cable size. reducing moose capture in wolf snares gardner alces vol. 46, 2010 178 212) and 26.7 cm (n = 80) csb snares without diverter wires during the course of their normal wolf trapping in 2005-2006. they neck-caught and killed 20 wolves with the 24.1 cm csb snare (16 flexible and 4 solid anchors), and 9 wolves (0 escaped) with the 26.7 cm csb snare (6 flexible and 3 solid anchors). five of 6 moose (2 calves and 4 adult) caught in the 24.1 cm csb escaped, and all 3 adults escaped the 26.7 cm csb snare; captures occurred with 5 flexible and 4 solid anchors. the single moose (yearling female) not escaping was neck-caught (flexible anchor). i assumed that escaped moose were those caught by the leg because the csb mechanism was not designed to release neck or nose-caught moose. i combined results to test efficiency and selectivity because no wolves, but all leg-caught moose, escaped from both csb snare types. the csb breakaway system restrained and killed all 29 wolves and allowed the release of all 8 leg-caught moose; no wolves or moose approached any other available snares. based on the binomial probabilities (95% confidence level), this breakaway system should kill ≥90% of wolves captured and allow escape of at least 68% of leg-caught moose (proulx et al. 1994). discussion my data indicate that moose are vulnerable to wolf snares because 1) moose are largely unaware of wolf snares and do not try to avoid them even if detected, 2) the top of the loop of wolf snares is set at a height that corresponds closely to the height at which moose carry their head while walking or sometimes feeding, and 3) even knock-downs mostly retain loop sizes large enough to catch a moose by the leg. reducing vulnerability to wolf snares and developing an effective breakaway mechanism is difficult because moose are caught in different manners; most are caught in wolf snares by the nose or leg (tables 2 and 3). capture type and rate depend on whether the snare is encountered at its original set height or is a knock-down lying on the trail. i found no difference in catch type or rate due to snare loop size or snow depth. both nose and leg catches occur at the same proportion if the snare is encountered at original height, but leg-caught moose have to cause a knock-down and step into the loop; i only observed leg catches in knock-downs. moose are more vulnerable to knock-downs caused by other moose or wind due to the loop size and position on the trail. not surprisingly, managers and trappers have type/location cable size breaking strength (kg) moose se wolf se csb/24.1a 2.4 192.6 3.53 240.4 5.97 2.8 246.6 6.44 314 7.43 csb/26.7b 2.4a 166.4 3.62 201.1 3.86 2.8b 228.4 3.46 246.5 6.21 3.2c 276 6.89 312.9 8.2 split lockc 2.4 264.2 3.6 2.8 325.4 8.2 s-hookd 2.4 198.5 12.2 table 4. breaking strength (kg) of breakaway snares used on simulated wolf necks and actual moose legs in fairbanks, alaska, 2004–2006. each snare cable diameter combination was tested 20 times. acinch stop breakaway (csb) located 24.1 cm from the cable end stop. bcinch stop breakaway located 26.7 cm from the cable end stop. cthompson split lock. ds-hook attached to a thompson lock. alces vol. 46, 2010 gardner reducing moose capture in wolf snares 179 concentrated on designing snare types more effective in releasing leg-caught ungulates than improving capture selectivity. i found that moose vulnerability to wolf snares can be reduced by adding a diverter wire that extends from the snare about 70 cm at a 10-20o angle from the horizontal plane tangent to the top of the snare (fig. 4). the placement and length of this wire ensures that moose will initially contact it instead of the snare, thereby pushing the snare aside or creating a knock-down, and minimizing the chance of a nose/neck-caught moose. unfortunately, there is no efficient breakaway mechanism that will allow escape of a neck/nose-caught large ungulate. i believe that diverter snares will also minimize neck/nose-caught caribou and other non-target species taller than wolves because the diverter wires would be struck prior to contact with the snare. importantly, the efficiency of wolf captures was not affected by adding the diverter wire. diverter wires did not increase the frequency of knock-downs by wind, but did cause more knock-downs by moose than occurred with standard snares. however, there was no related increased capture of moose suggesting that the diverter snare continued to be effective. my 23 observations of moose contacting diverter snares indicate that the snare usually falls to the trail after contact forming a 15-38 cm loop with the diverter wire maintaining its original orientation. therefore, subsequent moose on the trail should still contact the diverter wire prior to stepping into the loop. the most likely situations when moose are caught in diverters occur when moose do not follow the trail and bypass the diverter wires, or when diverter wires are damaged. the diverter wires in this study were not damaged after 1-2 knock-downs. all moose caught in diverters were in snares unchecked ≥12 days indicating that the efficacy of diverters may be reduced from repeat contacts with the diverter wire or a moose eventually did not follow the trail. these failures illustrate the need to incorporate a breakaway system to allow leg-caught moose to escape. i found only one reference evaluating breakaway efficiency for wolf snares (thompson snares). most information describing the efficiency of breakaway snares has come from trappers who report good success with several breakaway mechanisms, particularly the thompson split lock on 0.24 cm diameter cable and s-hooks (blejwas 2006). however, there are no reports of trappers or researchers incorporating a cinch stop with any of the breakaway mechanisms on wolf snares. due to extreme cold temperatures in most of alaska, moose that do not break free from snares often sustain mortal injuries due to freezing. therefore, a cinch stop would be a remedial measure for leg-caught moose especially if the snare was anchored to a flexible anchor and more time was required for the moose to break free. the ideal wolf snare would incorporate a breakaway system that released all legcaught moose but no neck-caught wolves. the breaking force necessary to cause release of the csb mechanism placed either at 24.1 or 26.7 cm tested during this study was low enough for all leg-caught moose to break free regardless of the anchor type, but was sufficient to hold all neck-caught wolves. the advantage of the csb mechanism over other breakaway mechanisms is that it breaks easiest when the lock comes in contact and pushes against the ferrule. thus the breaking force necessary for release of a leg-caught moose, where the lock contacts the ferrule, will be less than that for a neck-caught wolf where contact is not achieved. the breaking force increases the further the cinch down point is from the csb mechanism because the force is no longer concentrated on the release, but spread around the entire loop. this is not the case for breakaway mechanisms that are dependent on the lock separating or s-hooks pulling apart; the breaking force is similar for moose and wolves, or possibly less for wolves reducing moose capture in wolf snares gardner alces vol. 46, 2010 180 as loop size is larger (roy et al. 2005). not using a cinch stop can be problematic if the breakaway mechanism does not release the moose because the chance of injury and even death is high due to freezing limbs. to minimize the chance of injury, a cinch stop should be included when s-hooks are the primary breakaway mechanism. unfortunately, a cinch stop does not work with the split lock on any size cable because a split lock releases when the cable is pulled through the cut. if a cinch stop is incorporated, it would also have to be pulled through the cut. i recommend that trappers use the csb or s-hooks incorporated with a cinch stop as their primary breakaway mechanisms on wolf snares. an apparent disadvantage of the csb was that breaking forces decreased with smaller diameter cable because of less contact surface (less friction) between the cable and ferrule, increasing the possibility that wolves could escape. some trappers may be reluctant to use the csb mechanism on 0.24 cm cable using the attachment methods described herein. to alleviate that concern, higher breaking forces can be achieved by increasing the contact surface between the ferrule and cable by increasing the number of times the ferrule is swaged or by using a longer ferrule. placement of the csb on the snare loop is an important consideration because breaking force declines with greater spacing between the csb and the end stop. i recommend that the csb be placed at 26.7 cm to minimize the breaking force for moose or other smaller ungulates yet ensure adequate loop size and holding strength to kill wolves. my analysis of loop size relative to cable diameter indicated that this would be adequate for wolf snare cable set 0.24-0.32 cm. for snares using s-hooks as the breakaway mechanism, i recommend placing the cinch stop at 26.7 cm. trappers, other researchers (phillips 1996, roy et al. 2005), and i have found effective release mechanisms to release ungulates from snares. none of these breaking mechanisms, including the csb, are efficient in releasing nose-caught moose from wolf snares; the diverter wire is presumably the only mechanism that reduces nose catches. management implications snares are an effective method to catch wolves and are a preferred trapping method in alaska. however, the associated accidental capture of moose is problematic. based on the characteristics of how moose encounter a wolf snare, i found that incorporating 2 modifications (diverter wire and cinch stop) to the snare resulted in fewer caught and injured moose, and higher escape rate. these changes did not affect the snare's effectiveness to catch wolves as i found no instance where wolves either escaped or evaded capture because the breakaway mechanism released, or by actively avoiding the snare. both modifications can be easily done by trappers and commercial suppliers of wolf snares on snare cable with 0.24-0.32 cm diameter. although results are particularly pertinent to wolves and moose, these results are likely applicable in other areas where wolf or coyote snaring occurs in the presence of other large hoofed mammals. importantly, these modifications will improve selectivity without reducing efficiency of wolf snares. in areas of high moose density where wolves are trapped intensively, i recommend that a cinch stop be required, and possibly a diverter wire, to reduce the chance of accidentally catching and restraining moose. furthermore, a maximum 7-day snare check should be considered because knock-downs make moose more vulnerable to capture, albeit, recognizing that trapping in rural alaska and canada often requires long trap-lines and severe weather conditions that may require special consideration. using captive moose to evaluate vulnerability to snares and test snare modifications proved to be an opportunistic and valuable approach. if possible, further study to improve alces vol. 46, 2010 gardner reducing moose capture in wolf snares 181 selectivity and efficiency of snares should be conducted with tractable moose to realize optimal sample sizes and testing design. specifically, i recommend evaluating the influence of snare loop size by investigating loop sizes <153 cm. i documented no reduced capture rate in 153-cm snare loops as compared to 183-cm loops, despite the top of the tear-dropped shaped loop of 183-cm snares being at least 7.9 cm higher. the ideal loop size would be >153 cm and reduce the chance of caught moose, yet maintain high efficiency in wolf capture. acknowledgements the federal aid in wildlife restoration project w-33-3 provided financial support. i thank j. crouse, s. jenkins, l. lewis, t. lohuis, j. selinger, and t. mcdonough for their assistance in the field and t. hollis and j. caikoski for helping construct snares. t. hollis, r. perkins, and j. whitman helped test snares. i thank b. taras, s. brainerd, r. boertje, j. burns, l. mccarthy, and s. szepanski for their editorial comments to this manuscript. n. pamperin and m. ross assisted constructing the figures. i particularly want to thank m. keech and t. lohuis for many discussions on possible snare designs and b. taras for his help with data analysis. references alaska trappers association. 2007. alaska wolf trapping manual. fairbanks, alaska, usa. (accessed july 2009). animal care and use committee. 1998. guidelines for the capture, handling, and care of mammals as approved by the american society of mammalogists. journal of mammalogy 79: 1416-1431. blejwas, k. 2006. trapper questionnaire statewide annual report 1 july 2004 through 30 june 2005. alaska department of fish and game, division of wildlife conservation. juneau, alaska, usa. (accessed september 2009). boertje, r. d., m. a. keech, d. d. young, jr., k. a. kellie, and c. t. seaton. 2009. managing for elevated yield of moose in interior alaska. journal of wildlife management 73: 314-327. ———, k. a. kellie, c. t. seaton, m. a. keech, d. d. young, jr., b. d. dale, l. g. adams, and a. r. aderman. 2007. ranking alaska moose nutrition: signals to begin liberal antlerless harvests. journal of wildlife management 71: 1494-1506. cochran, w. g. 1977. sampling techniques. john wiley and sons, new york, new york, usa. d’agostino, r. b., w. chase, and a. belanger. 1988. the appropriateness of some common procedures for testing the equality of two independent binomial proportions. the american statistician 42: 198-202. gardner, c. l. 2007. development and testing of breakaway snares. alaska department of fish and game. federal aid in wildlife restoration. final research performance report. grants w-33-2 through w-33-5. project 15.12. juneau, alaska, usa. (accessed october 2009). ———, and k. b. beckmen. 2008. evaluating methods to control an infestation by the dog louse in gray wolves. alaska department of fish and game. federal aid in wildlife restoration. research progress report. grant w-33-5. project 14.25. juneau, alaska, usa. gasaway, w. c., r. d. boertje, d. v. grangaard, d. g. kelleyhouse, r. o. stephenson, and d. g. larsen. 1992. the role of predation in limiting moose at low densities in alaska and yukon and implications for conservation. wildlife reducing moose capture in wolf snares gardner alces vol. 46, 2010 182 monographs 120. ———, r. o. stephenson, j. l. davis, p. e. k. shepherd, and o. e. burris. 1983. interrelationships of wolves, prey, and man in interior alaska. wildlife monographs 84. guthery, f. s., and s. l. beasom. 1978. effectiveness and selectivity of neck snares in predator control. journal of wildlife management 42: 457-459. lebreton, j.-d., k. p. burnham, j. clobert, and d. r. anderson. 1992. modeling survival and testing biological hypotheses using marked animals: a unified approach with case studies. ecological monographs 62: 67-118. naylor, b. j., and m. novak. 1994. catch efficiency and selectivity of various traps and sets used for capturing american martens. wildlife society bulletin 22: 489-496. phillips, r. l. 1996. evaluation of three types of snares for capturing coyotes. wildlife society bulletin 24: 107-110. ———, f. s. blom, and r. e. johnson. 1990. evaluation of breakaway snares for use in coyote control. pages 255-259 in l. r. davis and r. e. marsh, editors. proceedings 14th vertebrate conference, university of california davis, california, usa. proulx, g., a. j. lolenosky, m. j. badry, p. j. cole, and r. k. derscher. 1994. a snowshoe hare snare system to minimize capture of marten. wildlife society bulletin 22: 639-643. r development core team. 2008. r: a language and environment for statistical computing. r foundation for statistical computing.vienna, austria. (accessed february 2009). rausch, r. a. 1967. some aspects of the population ecology of wolves, alaska. american zoology 7: 253-265. roy, l. d., c. twitchell, and m. hiltz. 2005. factors influencing the effectiveness of breakaway snares to capture coyotes and release deer in alberta. alberta research council, vegreville alberta, canada. (accessed february 2009). scott, a. j., and g. a. f. seber. 1983. differences of proportions from the same survey. the american statistician 37: 319-320. shivik, j. a., and k. s. gruver. 2002. animal attendance at coyote trap sites in texas. wildlife society bulletin 30: 502-507. thompson snares. 2009. thompson snares homepage. (accessed february 2009). alces35_177.pdf alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 53 status and management of moose in the northeastern united states david w. wattles1 and stephen destefano2 1massachusetts cooperative fish and wildlife research unit, department of environmental conservation, university of massachusetts, amherst, ma 01003 usa, 2u. s. geological survey, massachusetts cooperative fish and wildlife research unit, university of massachusetts, box 34220, amherst, ma 01003 usa abstract: moose (alces alces) populations have recolonized much of their historic range in the northeastern united states in the past 30 years, with their southern range edge extending to southern new england and northern new york. this southerly expansion occurred when certain other populations in the united states were in decline along the southern range edge, with climate change often cited as a probable cause. the areas that moose have recently occupied in the northeastern united states are some of the most densely human populated in moose range, which has raised concern about human safety and moose-vehicle collisions (mvc). we conducted a literature search about moose in the northeastern united states, and distributed a questionnaire and conducted phone interviews with regional biologists responsible for moose management to determine the status of moose, management activity, and research deficiencies and needs. moose numbers appear stable throughout much of the region, with slow population growth in northern new york. management activity ranges from regulated harvest of moose in maine, new hampshire, and vermont, to no active management in southern new england and new york. the combined annual harvest in maine, new hampshire, and vermont is >3,000. mvcs are a widespread regional concern with >1,000 occurring annually involving several human fatalities. research should address impacts of parasitism by winter tick (dermacentor albipictus) and brain-worm (parelaphostrongylus tenuis) on productivity and mortality of moose, influence of climate change on population dynamics and range, and conflicts in areas with high human population density. alces vol. 47: 53-68 (2011) key words: alces alces, management, moose, new england, new york, status. although exact records of historic moose (alces alces) distribution and numbers are difficult to document, goodwin (1936) claimed through anecdotal evidence that moose once ranged as far south as the alleghany mountains of pennsylvania in eastern north america. by 1870 moose had likely been eliminated throughout the southern portion of their range by unregulated and commercial hunting, and forest clearing for agriculture. allen (1870) claimed that moose were extinct in massachusetts, southern vermont, southern new hampshire, and southern maine, but inhabited northern portions of maine and were probably still in northern new hampshire, vermont, and the adirondack mountains of new york. the eventual recovery and expansion of moose populations in the northeastern united states (northeast) likely resulted from a number of factors, the 2 most important being regulation of moose hunting and reforestation of abandoned farmland. the 1936 closure of moose hunting in maine provided protection of a core population of moose in the northeast, and as farms were abandoned across the region, reforestation and subsequent logging that created patchy younger forest amid evenaged stands increased and improved habitat for moose (alexander 1993, bontaites and guftason 1993). other contributing factors to the population increase were the reintroduction and spread of beaver (castor canadensis) and status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 54 corresponding increase in wetland habitat, and the decline of white-tailed deer (odocoileus virginianus) populations and their associated parasite parelaphostrongylus tenuis (alexander 1993, bontaites and guftason 1993). by the 1970s moose had increased in sufficient number in maine to disperse to and augment the small population in adjacent new hampshire in which there were ~500 moose by 1977. exploiting unoccupied habitat, moose in new hampshire quickly increased to ~1,600 in 1982 and 5,000 by 1993 (bontaites and guftason 1993); the same pattern followed in vermont, with the population increasing from 200 in 1980 to 1,500 in 1993 when a hunting season was reinstated (alexander 1993). despite the historical presence of moose in southern portions of the northeast, many biologists considered the region to have marginal habitat and thought it unlikely that moose would establish viable populations (karns 1997; w. woytek, massachusetts division of fisheries and wildlife [mdfw], pers. comm.), particularly given the humandominated landscape and high potential for human conflict (vecellio et al. 1993, peek and morris 1998). other factors that could impede their reestablishment were the highly fragmented mid-late stage mixed deciduous forest, relatively limited early successional habitat, and lack of key browse species found in the boreal forest such as balsam fir (abies balsamea), willow (salix spp.), mountain ash (sorbus aucuparia), and trembling aspen (populus tremuloides). the regional northern forest type where moose are common is dominated by spruce (picea spp.), balsam fir, beech (fagus grandifolia), birch (betula spp.), hemlock (tsuga canadensis), and maple (acer spp.); transitional hardwood forests occur more southerly and are increasingly dominated by oak (quercus spp.) and white pine (pinus strobus) where little is known about moose habitat use and requirements. exceptions are in northern new york and the berkshire mountains in western massachusetts with forests similar to that in northern new england. presumably, higher temperatures in southern portions would increase the likelihood of negative impacts due to thermal stress (renecker and hudson 1986, murray et al. 2006, lenarz et al. 2009). moose must also cohabit with higher deer densities from north to south; however, the effects of p. tenuis on moose populations may be less severe than previously believed (whitlaw and lankester 1994). despite these presumed barriers to regional expansion, moose were sighted occasionally in the 1960s (presumably from vermont and new hampshire) in massachusetts where few public reports occurred prior to 1966 (vecellio et al. 1993). regular occupation in new york began in 1980, initially in the border regions near quebec, ontario, and vermont, and spread quickly into the adirondack mountains (hicks 1986). by the late 1980s-early 1990s moose appeared in connecticut, and by 1998 there was evidence of a breeding population (kilpatrick et al. 2003). moose are now considered established in new york, massachusetts, and connecticut. to best describe moose status and management in the northeast, we defined 2 regions relative to the time of establishment and size of population. southern new england and new york included massachusetts, connecticut, and upstate new york where moose are more recently established and management policies are forming. northern new england included maine, new hampshire, and vermont where moose populations are well established and been actively managed for several decades. our objectives were to report on the current status of moose populations and management policies, identify differences and similarities between the 2 regions, and identify research and management strategies to aid management of the regional population. study area the states of maine, new hampshire, vermont, massachusetts, connecticut, and alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 55 new york currently have resident moose populations, and rhode island, new jersey, and pennsylvania are states where moose occurred historically (goodwin 1936; l. gibson, rhode island division of fish & wildlife [ridfw], pers. comm.; c. condolf, new jersey division of fish and wildlife [njdfw], pers. comm.). new york, connecticut, and massachusetts represent the southern edge of current moose range in eastern north america. the population is between 66° 57’w longitude in eastern maine and 76° 10’ w longitude on the western side of the adirondack mountains in new york, and between 47° 28’ n latitude in northern maine and 41° 38’ n in central connecticut (ed reed, new york department of environmental conservation, bureau of wildlife [nydec], pers. comm.; h. kilpatrick, connecticut department of environmental protection [cdep], pers. comm.; l. kantar, maine department of inland fisheries and wildlife [mdifw], pers. comm.) (fig. 1). new jersey, rhode island, massachusetts, and connecticut are the most densely populated states in the united states (u.s. census bureau n.d.) where human development and road networks make forest habitat patchy and highly fragmented; southeastern new hampshire and coastal southeastern maine also have high levels of human development. in general, human density decreases to the north and west as forested area and available moose habitat increases. the region is heavily forested with extensive streams, rivers, lakes, ponds, and wetlands; elevation ranges from sea-level to 1,916 m in the white mountains of new hampshire. degraaf and yamasaki (2001) identified 5 forest regions, each with characteristic tree species and specific physiographic and climatic conditions. the spruce-balsam fir forest occurs in the coldest areas of the northeast, >150 m in maine and at higher elevations in new hampshire, vermont, and new york. the northern hardwoods-spruce forest is at lower elevations in maine and at <850 m in mountains of new hampshire, vermont, and northern new york; small pockets are found in the mountains of western massachusetts. the northern hardwood forest is at 150-790 m in maine, new hampshire, vermont, new york, and western massachusetts. the transitional hardwoods-white pine forest is at lower elevations in the uplands of northern new england, and is the dominant forest in massachusetts and northeast connecticut. the central hardwoods-eastern hemlock-white pine forest is found throughout connecticut, southern and eastern massachusetts, and extreme southeastern new hampshire and maine (degraaf and yamasaki 2001). methods we conducted an electronic mail (e-mail) survey of state agency deer and moose biologists managing established moose populations in maine, new hampshire, vermont, massachusetts, new york, and connecticut. we asked about the abundance, distribution, status, population goals, management practices including hunting and habitat management, issues and concerns, and experience with public perception of moose. we asked follow-up questions via telephone and e-mail when addifig. 1. range of moose in the northeastern united states (dashed line represents southern edge of moose range). status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 56 tional information or clarification was needed. we also conducted telephone interviews with the deer biologists of rhode island, new jersey, and pennsylvania (adjacent to states with moose populations and where moose were believed to be historically) and asked about sightings and anecdotal information about moose in their state. we also gathered, reviewed, and summarized literature on the status and management of moose populations in the region. results southern new england and new york massachusetts moose numbers increased rapidly in massachusetts in the 1990s after re-colonizing the state in the 1960s. the mdfw estimated a population of 850-950 in 2010 (s. christensen, mdfw, pers. comm.) based on a regression model developed in new hampshire that uses moose sighting surveys by deer hunters and available suitable habitat to estimate moose abundance (bontaites et al. 2000). the population has stabilized since 2001 with the possible exception of slight increase in the berkshires hills in the western part of the state. the mdfw prefers to maintain the population at the current level, with the overall goal “to maintain and sustain a resident breeding moose population in the state in areas of suitable habitat throughout its historic range at levels which support ecological and cultural values while minimizing human-moose conflicts” (s. christensen, pers. comm.). moose habitat in massachusetts is found primarily in the central and western portions of the state, west of the city of worcester. however, moose are frequently reported farther east in an area that constitutes the greater boston metropolitan region where patches of suitable habitat are smaller and more fragmented; high human population density makes it likely that moose are considered problem animals in this area. in western massachusetts the 2 main forest regions are separated by the connecticut river valley and the interstate 91 highway corridor, both of which run n-s. these regions are fragmented by state highways and towns, but enough forest habitat remains to support a stable moose population. as the number of moose increased in the late 1980s and early 1990s, vecellio et al. (1993) and mcdonald (2003) questioned whether a state as densely populated as massachusetts could support a larger moose population; massachusetts has the third highest density of people in the united states, averaging about 313 people/km2 (u.s. census bureau n. d.). they speculated that the cultural carrying capacity would be exceeded and the issue would become untenable and conflict would be inevitable without proactive management. as predicted, the moose population increased rapidly after 1993, as did the number of moose-vehicle collisions (mvc) that peaked at 52 in 2004 (fig. 2; vecellio et al. 1993; mdfw, unpublished data); 2 human fatalities occurred in 2003 and 2007 (table 1). despite the increase in costly and dangerous mvcs, moose have apparently not exceeded cultural carrying capacity as public perception is almost universally positive, based on mdfw’s and our interactions with the public. moose density is relatively low and a moose sighting is still somewhat of a novelty as most residents have never seen a moose in massachusetts. the mdfw regards the return of moose as a conservation success; however, as of 2010 it does not have authority to initiate a regulated hunt because moose hunting is specifically prohibited by state statute. legislation was first introduced in 2002 to give management authority to the mdfw, but the bill has not progressed beyond the legislative committee stage which is influenced by both those desiring a moose hunt and a large animal rights and anti-hunting population in massachusetts. management activities include monitoring mvcs, continued analysis of deer hunter surveys, and response to problem animals. alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 57 massachusetts has developed a large animal response team (lart) composed of mdfw and environmental police personnel who respond to problem moose and other large mammals. such situations occur when a moose is a threat to their own or public safety by wandering into towns or onto busy roadways. the current policy has 3 stages: 1) the animal is hazed or herded back to suitable habitat, 2) if hazing fails and immediate public safety is not an issue, the animal is immobilized and relocated to a wildlife management area, state forest, or other suitable area away from development, or 3) the animal is euthanized if an immediate threat and hazing and immobilization are unfeasible. the lart has performed 1-9 relocations and 0-5 euthanasias annually in the past 10 years; the number of problem animal responses of all types has declined in the last 5 years. in southern new england, as elsewhere, regenerating forests are an important source of browse. in any given season, moose preferentially use (50-65%) early successional habitat created by logging (usgs massachusetts cooperative fish and wildlife research unit [mcru], unpublished data). large tracts of managed state land tend to support higher moose density and appear to have greater browsing impacts. commercial foresters and large logging companies are increasingly concerned with the long-term impacts of browsing on the species composition and structure of massachusetts forests. state year maine new hampshire vermont massachusetts connecticutt new york total 1998 5 0 2 0 0 0 7 1999 1 1 1 0 0 0 3 2000 3 2 0 0 0 0 5 2001 1 1 0 0 0 0 2 2002 2 1 1 0 0 0 4 2003 3 1 1 1 0 0 6 2004 4 2 0 0 0 0 6 2005 1 0 1 0 0 0 2 2006 2 0 2 0 0 0 4 2007 5 0 1 1 1 0 8 totals 27 8 9 2 1 0 47 table 1. human fatalities resulting from moose-vehicle collisions in the northeastern united states, 1998-2007. fig. 2. reported moose vehicle collisions in southern new england, 1989-2007. status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 58 research includes studying how moose use the landscape, respond to dense human populations, cope with high temperature, and interact and influence the deciduous forest. the mcru and mdfw began studying movements and habitat use of moose equipped with global positioning system (gps) collars (n = 35) in 2006. faison et al. (2010) studied moose browsing in the deciduous forest of massachusetts; additional studies will integrate vegetation and gps data to further evaluate moose browsing. also, several sets of 20 x 20 m fenced exclosures have been built with paired control plots to estimate the effect of moose browsing on species composition and rate of forest regeneration (compton and destefano 2009, faison et al. 2010). new york re-colonization began around 1980 from moose dispersing from vermont and canada. by 1990 the population was estimated as ~20 animals, with a bull to cow ratio of 3:1 typical of a colonizing population (garner and porter 1990). the nydec considered augmenting the population with relocations in the early 1990s, but did not due to lack of public support, concern of increased moose-human conflict, and a desire for moose to repopulate naturally (hicks and mcgowan 1992, lauber and knuth 1997). the population has grown steadily as moose exploit unoccupied habitat; nydec estimated the population at 300-500 animals in 2008 with most in and around the 25,000 km2 adirondack park and reserve in the northern third of the state. public sightings indicate that moose are also present and increasing in the taconic highlands on the vermont and massachusetts borders. moose density was highest on private land along the northern edge of the adirondack park where forest management is more active. conversely, the adirondack park represents the majority of moose habitat in the state, but logging is not permitted. presumably the lower density and population growth rate in new york, as compared to that in vermont and new hampshire, reflects less forest harvesting and early successional habitat. the few reports of moose south of the interstate 90 corridor are usually young bulls presumably dispersing. the nydec does not expect moose to establish in southern new york because of the high level of development and warmer climate. state biologists predicted the population to exceed 1,000 animals by 2010; however, population growth was lower than expected and the 2010 population was estimated at 500-800 (c. dente and e. reed, nydec, pers. comm.). the goal of nydec is to increase the population in northern new york, and it monitors the population through reports of mortalities, reproduction, and public sightings. aerial surveys have been conducted to document range distribution but not to estimate the population. the population was monitored with a survey of successful deer hunters conducted by the wildlife conservation society that documented moose sightings and sign; response to these surveys was low (10%). the nydec has conducted surveys since 2008 and the response rate has increased to 30%. due to the overlap of moose and deer in new york, there is concern about the effect of p. tenuis and several cases of brainworm have been documented (c. dente and e. reed, pers. comm.). due to concern about increasing mvcs (fig. 2), the nydec and the new york department of transportation (nydot) have increased signage and public information about mvcs; nydot is investigating moose road crossings and mvcs in the state. in densely populated southern new york (north of new york city) concern about increasing mvcs and the need for a hunt have been raised in the popular press; however, it is believed that moose involved in local mvcs dispersed from adjacent connecticut and massachusetts, not northern new york (c. dente, pers. comm.). moose-human conflicts are increasing and the state has developed a plan to coordinate alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 59 and standardize response actions, including relocation of moose considered a threat to public safety. most response actions result from moose wandering into developed areas in the greater albany region and the interstate 90 corridor. management options are limited because moose are fully protected by state law; therefore, nydec is evaluating the possibility of securing regulatory change to allow broader management actions (e.g., hunting season). public opinion about moose recovery in new york remains positive despite increasing moose-human conflicts (e. reed, pers. comm.). connecticut moose began to re-colonize connecticut in the late 1980s with young bulls dispersing from massachusetts (kilpatrick et al. 2003). sightings of females occurred by 1990, and evidence of a resident breeding population was confirmed with the first cowcalf sighting in 2000; cow-calf sightings are consistent since. the increase in cow-calf sightings corresponded with increased public sightings in the late 1990s, from <5 in the early 1990s to 32 in 2002 (kilpatrick et al. 2003). similarly, there were no mvcs in connecticut before 1995; the frequency of mvcs has increased to 1-4 annually since 2003 (fig. 2; kilpatrick et al. 2003; h. kilpatrick, cdep, pers. comm.) and the first human fatality occurred in 2007 (table 1; a. labonte, cdep, pers. comm.). most moose are located in the more rural and forested northern third of the state, with higher density in the northwest than northeast (h. kilpatrick, pers. comm.). these 2 areas are largely separated from each other by the heavily developed portion of the connecticut river valley between springfield, massachusetts and hartford, connecticut; however, each is contiguous with moose habitat in massachusetts. in 2008 the population was estimated at >100 animals and increasing (h. kilpatrick, pers. comm.). continued growth was expected despite the belief that high summer temperatures, range overlap with white-tailed deer and threat of brainworm, and marginal habitat should all limit population growth. with a more conservative method, the population was estimated at ~75 animals in 2010 (a. labonte, pers. comm.). these estimates were based largely on observation rates and public reports, but the latter may be decreasing as moose are less of a novelty. this general stabilizing trend after a slight population reduction matches what seemingly occurred in massachusetts and southwest new hampshire (s. christensen, pers. comm.; k. rines, new hampshire fish and game department [nhfg], pers. comm.). connecticut is the fourth most densely populated state in the nation with 271 people/ km2 (u.s. census bureau n.d.). this high human population, dense road network, and high traffic volume make a large moose population potentially dangerous to human safety. several moose entering highly developed or high traffic areas are relocated or euthanized annually, and cdep is concerned about continued growth of the moose population as it desires to minimize mvcs. as a result, connecticut is conducting public and hunter opinion surveys, preparing a moose management plan, and considering the possibility of initiating a moose hunt to limit population growth. any initiation of a hunt will likely be met by opposition from anti-hunting groups in the state. the cdep has also deployed gps radio-collars on moose to investigate habitat use and movement patterns (h. kilpatrick, pers. comm.). northern new england maine, new hampshire, and vermont have well established moose populations of social, ecological, and economic importance. wildlife viewing, tourism revenue, and hunting permits and related expenditures generate millions of dollars annually in these states. moose populations have been actively managed with hunting seasons since 1980 in maine, 1988 in new hampshire, and 1993 in status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 60 vermont. the respective histories of moose recovery and development of hunting seasons are found in morris (2007), bontaites and guftason (1993), and alexander (1993) as well as in annual state reports. public participation with varied stakeholder groups has played an important role in determining state management goals (c. alexander, vermont fish and wildlife department [vfwd], pers. comm.; l. kantar, mdifw, pers. comm.; k. rines, pers. comm.). certain management issues are similar to those in southern new england, (e.g., mvc; fig. 3) but state-specific management is often related to balancing sometimes conflicting goals of different interest groups. maine – the moose population was estimated at 30,000-60,000 in 2010. highest density occurs in the forested interior with lower density along the coast. density ranges from 0.2-0.6 moose/km2 in the southern and 1.0-1.7 moose/km2 in the northern forested interior (morris 2007; l. kantar, pers. comm.). management goals and objectives were revised by a public working group after the state legislature granted the mdifw full authority for moose management in 2000 (l. kantar, pers. comm.) . the working group created a more comprehensive set of goals compared to the previous goal of maintaining the population at the 1985 level (morris 2007). goals and objectives continue to be developed through a public process involving representative stakeholders, including potentially conflicting groups associated with expanding moose watching and moose hunting, both of economic importance in maine (morris 2007). the goal is to strike a balance between moose viewing, public safety, and recreational opportunities (l. kantar, pers. comm.). the management guidelines developed in 2000 set population objectives specific to each of 29 wildlife management districts (wmd) and fall into 3 main categories: 1) recreation management which seeks to maintain the population at 60% of carrying capacity to maximize hunting and viewing opportunities, 2) road safety which seeks to reduce the population to decrease mvcs, and 3) compromise which seeks to reduce the population by a third to both reduce mvcs and maintain quality recreational opportunities (morris 2007; l. kantar, pers. comm.). in the remote and heavily managed forests in northern and central western maine where human population is minimal, wmds are in the recreation management category (1), wmds along the northeast-eastern and southwest borders of the state are in the compromise category (3), and wmds along the more densely populated southern interior and southeastern coastline of maine are in the road safety category (2; l. kantar, pers. comm.). despite the risk of mvcs, public opinion indicates that the majority do not favor a large reduction of the moose population along the fig. 3. reported moose vehicle collisions in northern new england, 1980-2008. alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 61 coast (morris 2007). new hampshire the moose population was estimated at ~6,000 in 2008 based upon analyses of deer hunter surveys and infrared thermal imagery surveys (bontaites et al. 2000), but was reduced to ~4,500 by 2010 (k. rines, pers. comm.). moose density declines from north to south in new hampshire, ranging from 1.2/km2 in the northern third of the state to <0.01/km2 along the more densely human populated coastline; numbers are considered relatively stable throughout most of the state. the population decline occurred in the northern third of the state due to increased hunting pressure proposed in the 2006-2015 new hampshire big game plan and to mortality related to winter tick parasitism (k. rines, pers. comm.). musante et al. (2010) found that body weight, survival, and reproduction of adult cows were high, but winter ticks caused measurable mortality of calves in years of high infestation and probably affected productivity of yearling cows. as a result, parasitism rather than habitat is believed more limiting to the population growth rate of moose in northern new hampshire. the new hampshire big game plan 2006-2015 (nhfg 2005) states the goal for moose management as: “new hampshire will regionally manage moose populations by balancing and incorporating social, economic, public safety and ecological factors, using the best available science.” management in each of 6 regions seeks balance among multiple and somewhat opposing goals of limiting browsing impacts, maximizing wildlife viewing and hunting opportunities, and limiting mvcs. management goals vary with regional priorities that are determined largely by the public. for instance, limiting mvcs is a priority in the southeastern region of highest human density, whereas balancing maximal recreational opportunity and limiting browsing impacts are priorities in the north where lower human density makes mvcs less of a concern (nhfg 2005). since 1999 the number of annual moose hunting permits has ranged from 482-678 in response to change in observation rates, hunter success, adult sex ratio, fall calf recruitment, and population growth rates; annual harvest has ranged from 333-482 (k. rines, pers. comm.). vermont the 2008 population was estimated at 4,000-5,000 animals with densities of 1/km2 in the northeast to ≤0.2/km2 elsewhere. an estimated statewide population of 3,000-4,000 moose in 2010 reflects the success of the effort to reduce density in the northeast portion of vermont (c. alexander, pers. comm.). a conservative hunt with limited permits was initiated in 1993 to allow increase of moose throughout the state, with the exception of wildlife management unit (wmu) e in the northeastern corner where stabilization/reduction of the population was desired because moose had or were nearly exceeding cultural carrying capacity. other goals were to monitor the population relative to biological and cultural carrying capacity to determine when and if expansion of the hunt was needed, maximize recreational opportunities, minimize moose-human conflicts, and provide funding for vermont’s moose management program (alexander 1993). as the population grew throughout the state, additional wmus were opened to hunting and permits increased; hunting now occurs in most of vermont. by 2003 the number of annual permits was 440 statewide and harvest was 298 animals. despite the increase in permits, the high moose density in northeastern vermont was causing heavy browse damage as body condition and productivity declined. hunting permits in the northeast were increased annually in an attempt to control the population; permits jumped to 833 in 2004 and 1250 in 2008 when 75% of permits and harvest statewide occurred in the 4 northeastern wmus. population reduction was achieved by allocating half the permits as antlerless only status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 62 (vermont moose management team 2008a, b; c. alexander, pers. comm.). management goals in 2008 were to further reduce the moose population in northeastern wmus, stabilize the population in most other wmus, and allow for controlled growth in a few wmus. it was believed that after several years of high permit numbers, permits would decline to ~500 statewide (vermont moose management team 2008b). harvest levels in 2009, hunter sighting rates, and a reduction in mvcs and non-hunting mortality all indicated that population goals in the northeastern wmus were being met. as a result, permits in wmu d2 were reduced from 337 in 2009 to 90 in 2010, and permits in wmus e1 and e2 were reduced about 30% from 600 combined permits in 2009; further reductions are anticipated in 2011 (c. alexander, pers. comm.; darling and alexander 2010; vfwd 2010). rhode island, new jersey, and pennsylvania annual moose sightings in rhode island are relatively rare (1-2 annually with no cow-calf sightings) and are usually north of scituate. the ridfw has not responded to a moose incident since the early 1990s when a moose was removed from inside the highway 295 corridor; the moose died in transit to new hampshire. it is believed that more sightings would occur if resident moose existed (l. gibson, ridfw, pers. comm.). there have been no reports of moose in new jersey, other than one animal observed crossing into the northwest corner (c. condolf, njdfw, pers. comm.). similarly, in the past 5 years there has been only a single sighting in northeastern pennsylvania, a young bull in a pasture with domestic cows (b.wallingford, pennsylvania game commission, pers. comm.). it is presumed that new jersey and pennsylvania are isolated from established moose populations either by distance or dense human development barriers. discussion population trends since 2001 the massachusetts moose population appears stable at 850-950 animals. the frequency of mvcs, responses to problem moose, relocations, and public safety kills in massachusetts peaked in 2004-05, followed by sharp decline (mdfw, unpublished data). a similar temporal trend toward stability occurred in connecticut (75-100 animals) and in southwest new hampshire where numbers increased sharply during range expansion when moose exploited unoccupied habitat. eventually these populations declined and stabilized at a lower level. it is speculated that brainworm might act as a limiting factor given the relatively high deer density in these areas (k. rines, nhfg, pers. comm.). moose in massachusetts and connecticut are at relatively low density and viewing moose is difficult due to their tendency to inhabit contiguous forest blocks. further, the few carcasses found are usually too deteriorated for necropsy or to determine cause of death; most animals afflicted with brainworm or heavy tick loads likely die unobserved. although no cases of brainworm are confirmed in massachusetts, neither have animals been tested, and several cases of brainworm have been confirmed in connecticut; many suspected cases are noted in both states. winter ticks are observed on captured and free-ranging moose in massachusetts, but the infestation level is not considered as severe as in northern new england (mcru, unpublished data; k. rines, pers. comm.; d. scarpitti, mdfw, pers. comm.). comparison of the heat stress index in ely, minnesota, where lenarz et al. (2009) speculated that warm temperatures were a mortality factor of moose, to that calculated in central massachusetts indicates that moose in southern new england are subjected to more prolonged periods of temperatures above the estimated thermoneutral zone of moose (renecker and hudson 1986). data from gps alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 63 radio-collared moose in massachusetts show reduced use of early successional habitats and a corresponding increase in use of conifer stands and wooded wetlands at spring and summer temperatures associated with thermal stress (mcru, unpublished data), an example of thermoregulatory behavior. if such behavior lead to reduced body condition (e.g., through stress and inefficient foraging), productivity and survival could be compromised. however, conflicting evidence of such includes high pregnancy rates and twinning by radio-collared adult cows (mcru, unpublished data). mortality related directly to winter tick (musante et al. 2007) and brainworm (lankester 2010), and mortality correlated with increasing temperatures (murray et al. 2006, lenarz et al. 2009) are associated with moose on the southern edge of their range. these factors and mvcs likely account for most mortality in southern new england where no hunting exists and predators of moose are few. moose may be at or near carrying capacity in deciduous forest habitat in southern new england, but no relative or comparator estimate of population density exists. despite evidence of heavy use of preferred browse species in regenerating sites, damage causing permanent forest change and/or nutritional impacts on moose has not been measured. the initial irruptive phase of population growth appears to have shifted to a slight decline and stabilization phase in massachusetts and connecticut. research needs include improved population estimation and related techniques, and further study of the influence of habitat, parasitism, mvcs, and temperature on this southern-most population in new england. habitat the long-term future of moose in southern new england is debatable. it is unknown how many moose are sustainable in the fragmented, deciduous forests of southern new england, if long-term occupation by moose will affect forest plant communities, and the impact of continued human development on forest resource availability. preliminary data from the gps radio-collar study in massachusetts indicate that moose (as elsewhere) use early successional forests created by logging. eastern hemlock substitutes for balsam fir, which is typical winter browse in northern new england but uncommon in most of western massachusetts and absent in eastern massachusetts and connecticut. forest harvesting has created a shifting mosaic of small patches of early successional habitat in southern new england where small privately-owned wooded parcels predominate (kittredge et al. 2003, mcdonald et al. 2006). further, public perception of logging in southern new england is often negative, and has lead to societal pressure in massachusetts to limit or eliminate logging on state lands. in fact, creation of new forest management plans in 2010 for state lands has greatly reduced the acreage for, and the types and extent of logging. these restrictions will require that private lands provide most early successional habitat in the state, with no guarantee that logging will continue at current levels. eastern hemlock is threatened with decline due to the hemlock woolly adelgid (adelges tsugae) with outbreaks causing widespread mortality of hemlock in connecticut and elsewhere in the appalachian mountains (orwig et al. 2003); it has recently spread into southern maine. the decline of hemlock in southern new england will presumably limit important winter browse and seasonal thermal cover for moose, and the relative impact may be related directly to higher temperatures associated with climate change. habitat in northern new england will likely not be a primary limiting factor of moose. although early successional habitat in maine is less than after the spruce budworm (choristoneura fumiferana) outbreak of the 1970s and 1980s, availability of commercial forestland and forest harvesting that produces early successional habitat is relatively constatus of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 64 stant in the northern areas of maine, new hampshire, and vermont (morris 2007). as in northeastern vermont, cultural carrying capacity influenced by public safety concern about mvcs and regeneration of commercial forests will mean that certain populations are managed at levels below nutritional carrying capacity. a growing human population with increased development is considered a potential limiting factor for the moose population in southern new hampshire (nhfg 2005). continued human development and urban sprawl probably pose the greatest threat to moose in southern new england and coastal new hampshire and maine. direct habitat loss from development of forested lands is occurring rapidly in massachusetts (denormandie and corcoran 2009), and the combination of habitat loss to development and increased forest fragmentation will likely result in more mvcs as moose move among habitat patches throughout the region. management and the public role public opinion and involvement in the management process has and will continue to drive moose management policies and population goals in the northeastern united states. for example, public meetings and public advisory groups composed of members from various stakeholder groups shaped vermont’s management goals and plan when the initial hunting season was under consideration. providing the public a voice in the decisionmaking process and proactive efforts to address the social issue of morality of moose hunting likely helped to minimize the anti-moose hunting sentiment that existed in vermont (alexander 1993; c. alexander, vfwd, pers. comm.); public involvement also plays an important role in moose management in new hampshire and maine. in massachusetts the policy for response to problem moose was influenced by public sentiment (vecellio et al. 1993), and public opposition in new york factored into the decision to not augment the moose population (lauber and knuth 1997). public opinion will obviously play a large role in whether moose hunting occurs in new york, massachusetts, and connecticut. moose vehicle collisions while relatively rare, mvcs are a primary public safety concern throughout the region because of their devastating nature and the possibility of human fatality. increasing moose populations in densely populated massachusetts and connecticut have led to a corresponding increase in mvcs (fig. 2). likewise, mvcs in new york have increased but the distribution of moose in the lightly roaded northern third of the state has presumably limited the relative number. higher moose density in northern new england has lead to more mvcs and related human fatalities, despite lower human population and traffic density (fig. 3). since 1996 there have been >1,000 mvcs annually in the northeastern united states resulting in >50 human fatalities, or about 1 in 250 (table 1; fig. 2, 3). the 600-700 annual mvcs in maine result in estimated damages of $17.5 million (danks 2007). unfortunately, mvc data in massachusetts have become increasingly unreliable. the number of reported collisions has declined in recent years to 24 and 18 in 2007 and 2008, respectively (fig. 2); however, this decline is at least partially due to inconsistent reporting from conflict over ownership of a moose carcass, lack of communication among state agencies, and the simple fact that a mvc in massachusetts is less novel. anecdotal and second-hand reports of mvcs now outnumber official reports, and comparison of division of law enforcement and mdfw records indicates that only a fraction of mvcs are reported to mdfw. the decline in collisions may represent an actual trend in the population, but any use of mvc data as a population index is compromised. by comparison, in 1999 alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 65 the new york state legislature amended the law concerning the disposition of the moose carcass in a mvcs; people whose vehicle has been damaged can obtain a permit from a law enforcement officer to possess the carcass (nydec 2010). connecticut adopted a law in 2008 that allows motorists to claim deer, moose, and bears killed in collisions (h. kilpatrick, cdep, pers. comm.); a similar approach in massachusetts may improve reporting of mvcs. research is continuing to better understand how habitat associations, landscape characteristics, road features, speed limits, moose density, and traffic volume influence mvcs. flexible management policies in northern new england provide for population reduction through hunting to reduce risk of mvc. wildlife managers in southern new england and new york are without this option and attempt to reduce collisions through signage, public education, and response to problem moose. hunting the number of moose hunting permits has been ~2,900 in maine and 400-700 in new hampshire over the past decade. the number fluctuates by management unit relative to evolving management goals and observation rate, hunter success, adult sex ratio, fall calf recruitment, and population growth rate. the number of permits in vermont fluctuates with change in observation rate to meet wmu-specific population goals and has increased statewide in recent years; however, the greatest increase was in the northeastern section to reduce and stabilize the population below ecological carrying capacity to address concern about forest impacts (vermont moose management team 2008b). in response to the population reduction, more conservative permit levels were established in the northeastern section in 2010. connecticut and new york are exploring the option of instituting moose hunts; however, in both cases legislation is required. it is also unlikely that moose hunting will occur in massachusetts in the near future. research regional cooperation among state moose biologists and managers is high. an annual meeting is held to share information and address management issues within northeastern u.s. states and canadian provinces. other meetings and collaborations are used specifically to help produce regional methods to index and estimate moose populations, and construct a uniform system to classify habitat. despite extensive studies throughout moose range, important regional questions remain regarding biology, foraging ecology, habitat use, life history, and population dynamics. in southern new england where most information about moose biology, habitat interactions, and forest interactions is scarce to non-existent or anecdotal and speculative, current research includes use of gps radiocollared moose and forest exclosures (wattles and destefano 2009, unpublished data). other research needs include methods to estimate population size and growth, causes of and factors influencing mortality and population growth, and ecological carrying capacity in this unique moose environment (h. kilpatrick, cdep, pers. comm.). in northern new england there is considerable interest in the role of black bear (ursus americanus) and coyote (canis latrans) predation, and parasitism (winter tick, lung worm, and brainworm) in population dynamics, especially with regard to calf mortality and recruitment. recent studies have focused on methods to monitor and predict infestation level of winter ticks and associated impacts on moose (musante et al. 2007, bergeron 2011). moose-deer interactions are of interest from the standpoint of interspecific competition for browse and the role of deer density on parasitism (brainworm) in moose. the impact of moose on forest regeneration and production is also a status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 66 focal area of management and research (e.g., bergeron et al. 2011). the future the recovery of moose in the northeast is widely heralded as a unique example of successful wildlife management. moose are the largest native mammal of the region and they recovered naturally due to ecological change largely associated with forest management and regulation of hunting. moose appear to have a stable future in the region, with the population well established in northern new england and relatively stable or growing in southern new england. however, the presence and ecological impact of such a large, charismatic mammal in the highly populated northeastern united states creates unique management issues for researchers and managers, particularly because its range has extended well southward in the past 20 years. the greatest concern and challenge for managers will be how to manage such a large animal in an increasingly human-dominated landscape that also has highly productive commercial forestland. moose management in this region will demand unique approaches due to the interconnected ecological, economic, social, and political factors in this diverse region. acknowledgments we would like to thank the following for their great assistance in making this manuscript possible: cedric alexander, vfwd; kristine rines, nhfg; lee kantar, mdifw; ed reed and chuck dente, nydc, bureau of wildlife; howard kilpatrick and andrew labonte, cdep; sonja christensen and dave scarpitti, mdfw; lori gibson, ridfw; carrol condolf, njdfw; and bret wallingford, pennsylvania game commission. the information provided by the individual state managers was invaluable. the mdfw has provided long-term funding for moose research in massachusetts through the federal aid in wildlife restoration program (w-35-r). the massachusetts department of conservation and recreation, u. s. geological survey, u. s. forest service, and university of massachusetts-amherst have also provided funding and logistical support. we would like to thank todd k. fuller, john e. mcdonald, jr., thomas o’shea, and sonja christensen for their careful review and comments on drafts of this manuscript. references alexander, c. e. 1993. the status and management of moose in vermont. alces 29: 187-195. allen, j. a. 1870. the distribution of moose in new england. american naturalist 4: 535-536. bergeron, d. h. 2011. assessing relationships of moose populations, winter ticks, and forest regeneration in northern new hampshire. m. sc. thesis, university of new hampshire, durham, new hampshire, usa. _____, p. j. pekins, h. f. jones, and w. b. leak. 2011. influence of moose population density on forest regeneration in northern new hampshire. alces 47: 39-51. bontatites, k. m., and k. guftason. 1993. the history and status of moose management in new hampshire. alces 29: 163-167. _____, _____, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69-75. compton, j. a., and s. destefano. 2009. experimental exclosures and the regeneration of forest vegetation in response to moose and deer browsing. usgs massachusetts cooperative fish and wildlife research unit, university of massachusetts, amherst, massachusetts, usa. danks, z. d. 2007. spatial, temporal, and landscape characteristics of moose-vehicle collisions in maine. m. sc. thesis. state university of new york, syracuse, alces vol. 47, 2011 wattles and destefano status of moose in northeastern u.s. 67 new york, usa. darling, s., and c. alexander. 2010. 2010 moose season proposal. vermont department of fish and wildlife, st. johnsbury, vermont, usa. degraaf, r. m., and m. yamasaki. 2001. new england wildlife; habitat, natural history, and distribution. university press of new england, hanover, new hampshire, usa. denormandie, j. and c. corcoran. 2009. losing ground: beyond the footprint, patterns of development and their impact on the nature of massachusetts. mass audubon, may 2009. faison, e. k., g. motzkin, d. r. foster, and j. e. mcdonald. 2010. moose foraging in the temperate forests of southern new england. northeast naturalist 17: 1-18. garner, d. l., and w. f. porter. 1990. movement and seasonal home ranges of bull moose in a pioneering adirondack population. alces 26: 80-85. goodwin, g. g. 1936. big game animals of the northeastern united states. journal of mammalogy 17: 48-50. hicks, a. 1986. the history and current status of moose in new york. alces 22: 245-252. _____, and e. mcgowan. 1992. restoration of moose in northern new york state. draft environmental impact statement. new york department of environmental conservervation, albany, new york, usa. karns, p. d. 1997. population distribution, density, and trends. pages 125-139 in a.w. franzmann and c.c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. kilpatrick, h. j., r. riggs, a. labonte, and d. celotto. 2003. history and status of moose in connecticut. connecticut department of environmental protection, hartford, connecticut, usa. kittredge, d. b., jr., a. o. finley, and d. r. foster. 2003. timber harvesting as ongoing disturbance in a landscape of diverse ownership. forest ecology and management 180: 425-442. lankester, m. w. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53-70. lauber, t. b., and b. a. knuth. 1997. fairness in moose management decision-making: the citizens perspective. wildlife society bulletin 25: 776-787. lenarz, m. s., m. e. nelson, m. w.schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. mcdonald, j. e., jr. 2003. bears and moose in massachusetts: the past, the present and the future possibilities. transactions of the north american wildlife and natural resources conference 68: 225-234. mcdonald, r .i., g. motzkin, m. s. bank, d. b. kitteridge, j. burke, and d. l. foster. 2006. forest harvesting and land-use conversion over two decades in massachusetts. forest ecology and management 227: 31-41. morris, k. i. 2007. moose assessment. maine department of inland fisheries and wildlife, augusta, maine, usa. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monograph 166: 1-30. musante, a. r., p. j. pekins, and d. l. scarpitti. 2007. metabolic impacts of winter tick infestations on calf moose. alces 43: 101-110. _____, _____, _____. 2010. characteristics and dynamics of a regional moose alces alces population in the northeastern united states. wildlife biology 16: 185-204. status of moose in northeastern u.s. wattles and destefano alces vol. 47, 2011 68 new hampshire fish and game department (nhfg). 2005. new hampshire big game plan; species managment goals and objectives 2006-2015. new hampshire fish and game department concord, new hampshire, usa. new york state department of environmental conservation (nydec). 2010. moose. (accessed june 2010). orwig, d. a., d. r. foster, and d. l. mausel. 2003. landscape patterns of hemlock decline in new england due to the introduced hemlock woolly adelgid. journal of biogeography 29: 1475-1487. peek, j. m., and k. i. morris. 1998. status of moose in the contiguous united states. alces 34: 423-434. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322-327. samuel, w. m. 2007. factors affecting epizootics of winter tick and mortality of moose. alces 43: 39-48. u.s. census bureau. n.d. census 2000. population housing units, area and density for states: 2000. (accessed june 2010). vecellio, g. m., r. d. deblinger, and j. e. cardoza. 1993. status and management of moose in massachusetts. alces 29: 1-7. vermont fish and wildlife department (vfwd). 2010. 2009 vermont wildlife harvest report; moose. vermont department of fish and wildlife, st. johnsbury, vermont, usa. vermont moose management team. 2008a. 2008 moose season proposal. vermont department of fish and wildlife, st. johnsbury, vermont, usa. _____. 2008b. executive summary, 2008 moose season proposal. vermont department of fish and wildlife, st. johnsbury, vermont, usa. wattles, d., and s. destefano. 2009. movement and landscape pattern use of a colonizing moose population in massachusetts. usgs massachusetts cooperative fish and wildlife research unit, university of massachusetts, amherst, massachusetts, usa. whitlaw, h. a., and m. w. lankester. 1994. a retrospective evaluation of the effects of parelaphostrongylosis on moose populations. canadian journal of zoology 72: 1-7. alces36_155.pdf alces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 163 first nations moose hunt in ontario: a community’s perspectives and reflections joseph w. leblanc1, brian e. mclaren1, christopher pereira1, mark bell2, and sheldon atlookan2 1lakehead university, faculty of natural resources management, 955 oliver rd, thunder bay, on, canada p7b 5e1; 2aroland first nation, p.o. box 10 aroland, on, canada p0t 1b0. abstract: moose (alces alces) hunting and other means of forest food production employed by members of first nations communities are undertaken as part of their treaty rights in ontario, articulated in specific nation-to-nation agreements with the government of canada on behalf of the british crown. aroland first nation in northwestern ontario is party to treaty 9 (1905), which overtly protects the community’s rights to hunt throughout the unoccupied tracts of crown land claimed as “traditional territory.” traditional use supersedes provincial authority and, as such, is not managed by provincial policy or regulation. this jurisdictional divide has presented an interesting history and many challenges for both provincial managers and first nations land users. strained relationships between provincial authorities and first nations, emergent from decades of misunderstandings of jurisdictional authority, have presented difficulty in all aspects of natural resource management. in this paper, we engaged community-based researchers in an exploration of the community’s perspective of the current and historical management regime. in collaboration with community members, we interpret the results, discuss implications, and provide considerations for future managers and policy makers. we also quantified the annual moose harvest by aroland and ginoogaming first nations that is only estimated by provincial managers; our results show provincial calculations may underestimate total harvests by up to 40%. this error could have significant implications for future moose populations, wildlife managers, and both provincial and first nations hunters. the potential for such errors serves to highlight our call for provincial authorities to seek and engage first nations perspectives and participation in moose management for the benefit of the entire community. alces vol. 47: 163-174 (2011) key words: alces alces, community perception, first nations, forest food production, hunting, moose, management, ontario, treaty rights. archaeological evidence and the petroglyphs of our ancestors show that the relationship between people of ontario’s first nations and moose (alces alces) is very old. it involves human use of meat, internal organs, hide, and skeleton (timmermann and rodgers 2005), while moose benefited from human use of fire that increased production of their forage plants. as natcher et al. (2007) further discovered, humans used fire to influence the movement of moose during fall hunts and to ease their own movement during hunting. 1we 1the use of the first person plural allows us to speak from an inclusive perspective. this perspective and our voice are therefore from those find such stories of our past to be intermittent in the scientific literature, and often told from a modern perspective that suggests the relationships are no longer relevant. we are delighted by how watson and huntington (2008) shared their understanding of a moose hunt, and are sympathetic to the lack of understanding by ecologists and wildlife biologists they experienced in the shared story of human and moose together in the boreal forest. we, who met in aroland first nation of the treaty 9 area of ontario, canada (aboriginal affairs people we see today as the community called aroland first nation, but broaden to include its neighbours in certain contexts. ontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 164 and northern development canada 2008), now wish to share with ecologists and wildlife biologists a review of our relationships with moose. we hope to illustrate that the past is part of our present situation and that the direction the future will take us depends on our acknowledging this singular story that is broader than the moose hunt itself. before we start, we can share what we learned about the present and future elsewhere. in nova scotia, canada, the mi’kmaq peoples of cape breton island have recently worked together with the parks canada agency and with provincial officials to maintain treaty rights to moose hunting (bridgland et al. 2007). in the canadian territories, indigenous peoples are intimately involved in co-management and monitoring of moose (larter 2009). in scandinavia, saami community representatives form part of the administrative boards that set moose quotas (bergman and akerberg 2006). we ask why, among these examples of respect, there is such disrespect for our relationship with moose in ontario. we know that wildlife biologists and sport hunters typically view first nations moose harvest with disdain (lynch 2006). kay (1997) even suggested that traditional hunting was unsustainable and that our ancestors kept moose populations from expanding into much of canada, even though his perspective is solely from british columbia. we appreciate the regional variation in the relationship between people and moose; crichton (1981) reviewed the situation in manitoba and concluded the same as kay (1997), while more recent investigation in alberta suggests that what wildlife biologists call “unregulated” harvest actually can have no detrimental effect on a moose population (lynch 2006). feit’s (1987) review is older, but includes 2 key points to which we will return: 1) if management of sport hunting of moose and management of the forest does not acknowledge first nations practices with respect to moose, conflict will escalate, and 2) conflicts develop when resource users do not share a common cultural heritage. our broader purpose in this paper is not to claim that the moose and first nations relationship has always been a good one; rather, it is to convey how people who hold values might be those best equipped to explain their values and plan their future actions. sharing some of our cultural heritage is our first objective. in timmermann and rodgers’ (2005) detailed summary of values embodied by moose, fear and uncertainty are the tone in describing moose management involving first nations peoples, especially in ontario. thus, offering objective considerations on use of the land in ontario for its forest resources, including moose, is our second objective. who is responsible for managing natural resources and who are they managing for? all those for whom the resource is being managed should have a forum for sharing their values, and those responsible for management must be sensitive to, and incorporate those values. our area our perspective focuses on aroland first nation, an anishnabek community in northern ontario. according to the records of nishnawbe aski nation (our treaty organization), there are 300 people living on-reserve and 400 others living off-reserve, but we feel an unaccounted number exist. we have a long history with the surrounding area, and in our traditions maintain a complexity of mutually beneficial relationships with other beings using this land as home. as a result, our community members include all humans and non-humans with whom we are interdependent. in the past, we participated in the fur trade and made a livelihood through local production of foods that came to us naturally or from agriculture (morrison 1986). gradually, as development activities took up land, the opportunities to make a livelihood shifted and we were officially discouraged from participating in food production (waisberg and holzkamm 1993). forestry offered new alces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 165 economic opportunities that offset these losses to our economy (driben 1985), but created a higher demand from external entities for our land’s resources. aroland first nation no. 242 gained reserve status under the indian act on april 15, 1985 (aboriginal affairs and northern development canada 2008). reserve lands encompass 19,599 ha (75.7 square miles) and extend northwards from highway 643 to lands along the western and northern shores of esnagami lake. this land is the extent upon which we have clear authority under the indian act. as a signatory to treaty 9, our community retains rights to access off-reserve resources among those parts of our territory not taken up with development. our territory extends 1000s of square km, but this land is now developed or restricted from us in a number of ways including parks and protected areas, municipalities, mines, and mills interconnected with vast and complex networks of closed roads and private rails. our traditional territory area includes 5 provincial wildlife management units, hunted by people from thunder bay, ontario, and farther away, and 4 provincial forest management units that are licenced to forestry companies, most with ownership in thunder bay or farther away. respective oversight of these management units is under the direction of the ontario ministry of natural resources (mnr) and the ontario ministry of northern development, mines and forestry (mndmf). in all cases, the ministries are headquartered well away from the areas in which they are actively engaged in making management decisions. in addition to the “managed” portions that ontario calls the “area of the undertaking,” our traditional territory extends into ontario’s less developed “far north.” our approach to start this research in december 2009, community-based researchers distributed a detailed questionnaire to potential moose hunters who lived on-reserve at aroland first nation; participants could be any male or female >18 years old. in addition to the questionnaire, consultations with the chief and council and other hunters also occurred as these people offered their time. this second consultation was administered orally with participants and recorded in writing by the interviewer and/or the survey participant. to ensure consistency, potential problems were discussed before allowing participants to continue with the survey. most concerns about the survey stemmed from long-standing trust issues about land use. there have been many instances over the past few decades of external interests seeking data from community members in relation to their land-use practices. often, the information gathered was taken out of the community to be interpreted externally and it is unclear as to how the interpretation is useful to the community. to conclude the data collection process, our survey data was reviewed by the interviewer and, if necessary, conversations were continued to resolve uncertainties or discrepancies; all surveys were kept anonymous. the survey protocol was reviewed and approved by lakehead university’s research ethics board (reb 113 08-09) and by health canada’s research ethics board (reb 2009-0007). participants indicated, on a 5-point scale (0 = none to 4 = all), how many of their meals included moose in each of winter, spring, summer, and fall. they were also asked why they hunt moose and in what season, how they accessed a hunting area, how they hunted, to what extent they relied on hunting for food, and how much moose meat is shared with the immediate family and with the community. thirty-five community members completed the survey (mean age = 44 years, range = 25-78 years). in conjunction with another “health and well-being” survey that included questions on a broader range of harvested, cultivated, and purchased foods, most participants indicated their agreement with the ontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 166 following on a 5-point likert scale: 1) their physical health (1 = poor to 5 = excellent), 2) their life satisfaction (diener et al. 1985), and 3) their connectedness with the land from the “connectedness with nature” scale (mayer and frantz 2004). they were asked to assess their beliefs about food contamination and their health: whether forest herbicides could affect one’s health if they ate moose or other forest foods, whether past mining practices in the area affected the quality of their food, whether eating local foods causes health problems and the degree to which this worried participants, and the nutritional quality of their diet and the amount of physical exercise they maintain. in 2010 in collaboration with the neighbouring community of ginoogaming first nation, we conducted a second smaller survey specifically about moose hunting. participants were asked a series of specific questions related to hunting moose; between the 2 communities, 40 individuals completed the survey. in addition to questions related to how, where, and why they hunted, respondents were asked how many moose they harvest in a year. survey data were entered into microsoft office excel and explored using correlation analysis to identify relationships and similarities among hunters. these relationships and similarities allowed for hypotheses to be formulated in discussion with community members, based on community hunting history and their relevance to non-aboriginal moose harvest in ontario. to supplement the interest of community members in conveying the extent of forest resources development and the use of forest herbicides in their traditional territory, we also accessed records from annual work schedules and reports to the mnr by the companies leasing the adjacent forest management units. these records included paper copies of maps showing roads, logged areas, and associated excel reports of groundbased and aerial spraying of herbicides from 2000-2007. the data on the maps and in the reports, borrowed from the geraldton, nipigon, and thunder bay district offices of the mnr, were transcribed into a geographic information system (gis) in arcgis version 9 at lakehead university. our story pre-contact, before 1800, the present our relationships are founded in our community and defined by our extended families. to survive, we have always used the local environment to generate our livelihoods. products for trade, sale, and local consumption are cultivated and harvested from within our territory. hunted and fished meats, as well as both cultivated and gathered vegetation from the land once represented the staples of our diet. familial territories that provided these staples were designed and cultivated to ensure enough stock for later years (driben et al. 1997). while familial units (nuclear families) often undertook production activities independently, sharing products among extended families and the community at large was commonplace. as with many indigenous communities throughout the world (e.g., kofinas 1993), our activities were undertaken in accordance with time-honoured systems of authority and knowledge. our ancestors passed on this knowledge of the land that grants us the authority to manage the resources that sustain our community. this knowledge and its authority were never given legal status in canada under the rule of law (herbert 2009). it is only the social relationships we hold within our community that honours the knowledge of our ancestors, ensuring it is passed to future generations. as we ethically engage in relations with non-human members (the plants and animals) of our community by hunting, fishing, cultivating, and gathering, we are undertaking activities that sustain the knowledge of our ancestors while meeting our sustenance needs. honourably engaging in conservation activities relating to harvesting food is part of the continuance of our relationship with the past alces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 167 and our ancestors. from an anthropological perspective, the role of moose hunting in the provision of food staples in first nations communities is a point of contention. while some (e.g., winterhalder 1983) rely on the notion that moose populations have consistently fluctuated due to climatic and anthropogenic influences as evidence of the continued occurrence of moose in our diet, others (e.g., rogers and black 1976, hamilton 2002) reference the “fish and hare period” to support the notion that there were times when moose were rare to non-existent and the dietary staples came from other sources, such as walleye (sander vitreus), lake whitefish (coregonus clupeaformis), caribou (rangifer tarandus), ruffed grouse (bonasa umbellus), snowshoe hare (lepus americanus), and beaver (castor canadensis). our interpretation of the lack of moose in diets during the “fish and hare period” is that it resulted from a need to seek continued sustenance while easing demands on some members of our extended community and allowing time for their populations to replenish. regardless of the anthropological interpretation of dietary inputs, moose have forever been an important member of our community. indeed, our crest is anchored by the image of moose antlers. today, moose forms an important part of our diet in fall and, to a lesser extent, in winter. moose meat is eaten at rates (self-estimated, mean ± standard deviation) of 1.87 ± 1.19 (winter), 1.00 ± 0.96 (spring), 1.64 ± 0.84 (summer), and 2.33 ± 1.40 (fall) meals per week. likely the same as for our ancestors, those who consume more moose in spring (the rarest occasion) report feeling better connected to nature (r = 0.69, p = 0.02) with less food insecurity (r = –0.58, p = 0.04). those who consume moose in winter associate themselves with having a better diet (r = 0.59, p = 0.03); those who consume moose in summer associate themselves with overall better self-rated health, (r = 0.59, p = 0.04); those who consume moose in fall feel they maintain better weight (r = 0.57, p = 0.04) and better overall health (r = 0.55, p = 0.05) than the rest of our population. with no other foods, whether country-harvested or purchased, did as many positive correlations occur as for moose. overall, participants from our community who indicated a larger proportion of their diet from local, country-harvested meats also indicated feeling better about their diet (r = 0.86, p = 0.001). as moose and other non-human members of our community have given their lives to sustain and enrich ours, so the knowledge of our ancestors has guided our relationships with each other, helping us ensure that all life exists in perpetuity. slowly, however, these traditional means of governing our relationships exclusively within our own community were being displaced by new laws with foreign ideas and language. post-contact through railway development, 1800-1874 prior to the establishment of canada, developments within our territory by outsiders focused on resource extraction to ship raw materials to europe. a mercantilist dogma drove the quests for gold, furs, and forest products of canada, exploited for wealthy monarchies, eventually in ontario for the king or queen of england. in this pre-treaty era, we held title over our territory, and foreign interests were mostly contained to sporadic trading posts and mines (driben 1985), as well as the odd town settled by european immigrants. increased inflow of settlers followed the construction of the trans-canada railway, which spawned a concentration of activities within its vicinity. increased external interest in wood and minerals in our territory was the stimulus to seek greater control of the land, and for us to articulate more clearly our interests and desire to protect our traditional way of life. with these often conflicting interests in mind, both parties entered into the treaty-making process. ontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 168 cession of lands and articulation of rights, 1905 to present in treaty 9 rest the legal rights to access the same lands by two opposing parties: first nations and the government of canada. on the matter of two distinct sets of rights, treaty 9 reads as follows: “and his majesty the king hereby agrees with the said indian that they shall have the right to pursue their usual vocations of hunting, trapping, and fishing throughout the tract surrendered as heretofore described, subject to such regulations as may from time to time be made by the government of the country, acting under authority of his majesty, and saving and excepting such tracts as may be required or taken up from time to time for settlement, mining, lumbering, trading or other purposes.” our new neighbours began to exercise their rights to take up tracts of land, eventually creating ontario government acts, regulations, policies, and guidelines, such as contained in the municipalities act (2001), the mining act (1990), and the crown forest sustainability act (1994). logging, mining, and protected areas versus traditional activities in a regulatory era following the imposition of external knowledge and management systems by new authorities, many aspects of our own timehonoured systems of authority and knowledge became disrupted. new human actors from outside our community began restructuring our territory without our input or consent. forest management units, parks and protected areas, wildlife management units, mineral claims, and indian reserves were imposed on our territory. along with these new divisions of the land, the dialogue and decision-making on the management of extended members of our community (all plants and animals) increasingly occurred in urban centers a great distance away, often preferentially protecting the rights of sports hunters or big business. forest managers located themselves at district mnr offices, as well as at consultancy, constituency, and corporate offices in thunder bay and farther away. technological advancements in the areas of remote sensing and gis, along with centralization in support of corporate and government efficiency, obligated decision-makers to be away from the land for which they were responsible. those of us most connected with the forest feel we have been disconnected from the decisions most influential to our community. the source of knowledge maintained by the healthy reciprocal relationships of the past that helped sustain this place and all living things within it was largely disrupted. imposed jurisdictions and outside decision-making have displaced local controls. as a result, our ability to exercise traditional practices and implement the knowledge of our ancestors, which are both actions aimed at sustaining our community in perpetuity, has been greatly restricted. currently, our ability to undertake food production activities, even hunting, feels restricted. undertaking many traditional activities has been reduced in stature and in terms of the time with which we are allowed to practice them, reflecting external perceptions of our culture. the time we take for traditional activities has also been reduced considerably by demands on us to compete with the new economy. our food gathering has been now described – and self-identified – more often as undertaking recreational activity than as participating in a traditional economy. purchased foods provide the staples of our diet today, even though they are increasingly less meaningful to our community health and wellbeing than our traditional foods. we feel that traditional products can retain their economic, social, and cultural significance and remain an important diet component. the majority (73%) responded they still rely on moose as a source of meat. nevertheless, we see a number of factors contributing to fewer people participating in traditional activities like moose hunting. these alces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 169 factors include the larger cultural shifts of the past originating with various assimilation attempts (i.e., relocation to reserves, residential school, and child services) and passive acculturation (i.e., mass media, the culture of convenience, and the application of capitalist modes of development). more importantly of late, changes to the land from newly imposed regulations and management activities have forced much farther travel to undertake traditional activities. most of us no longer migrate seasonally to follow our sources of food, nor do we follow our families to traditional territories. permanent relocation of our community to a reserve was a government solution to providing services, but the decision means we now concentrate our hunting activities and deplete the territory immediately around us of animals. as we travel farther for hunting and spend more money so doing, some of us are now less willing to share what we harvest: 31% of respondents said they harvest moose for their use alone. because our perception is that this trend will continue, our community seeks remedies such as the community freezer we recently obtained for food storage to help those in times of need. employment in resource extraction, primarily logging, provided cash for food purchases, or gasoline to travel farther for hunting; for a time, cash alleviated the pressure to produce food by traditional means. but economic downturns in the forest industry and technological advancements that made logging more efficient also drove a reduction in employment, so the total benefit from the forest industry garnered by local peoples was reduced. new access we gained to the forest from the building of logging roads was taken from us for road closures that paid for new roads, and from bridge removals that were likely designed to restrict our road use. silviculture that followed the new roads is now a source of great disturbance to the forest. the sequence of events employed by forest managers as means to regenerate what they allow to be taken by loggers leaves our ecological community disrupted. the complex network of primary, secondary, and tertiary roads – regardless of whether they are closed to us – fragments the forest, even while it opens new areas to recreational hunters visiting us from the outside. the roads of today also allow us to travel faster and farther than in the past, but we see around them that clear-cut logging removes natural forest stands. following logging, soils are often scarified, a process that leaves permanent scars on the landscape. the furrows and trenches left by scarifiers leave an unnatural footprint on the land that managers claim is for new tree plantings; these trees come from seeds sourced outside the community. when they arrive, they are planted in a manner that optimizes the yield at maturity and ease of future harvest; spaced at ~2 m from each other in rows, these new trees experience almost no competition or other forces of natural selection. many planting sites are later sprayed with chemical treatments (herbicides), some aimed at reducing pest populations, but most aimed at reducing competition against the newly planted trees. the competing trees and shrubs that herbicides eliminate are in many instances food for the human and non-human members of our community. we feel that outside decisionmakers are prioritizing efficiency in industrial production over the production of local goods that sustain our community. we see the resulting forest as foreign and unrecognizable and we are concerned that non-human community members experience the same. moose will not use artificially regenerating forests in the same way as naturally regenerating forests; depending on the extent and pattern of logging, the road network, and the hunting pressure, the length of time needed for moose to repopulate an area can be 15 years or more. government scientists (e.g., rempel et al. 1997) tell us our concerns are valid. our perception of change to an area heavily influences how we use it. the extent of herontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 170 bicide spraying activities over our traditional territory in any one year is small relative to its total area. for a typical moose with home range much larger than even the largest blocks treated with herbicide, food supply is probably affected negligibly by herbicide treatments. the moose that experiences herbicides in its home range simply moves away for one or more years (lautenschlager 1992). however, the ecological, social, and economic impacts of one year’s spraying activities are not restricted to that summer. for years to follow, the conditions created by spraying are evident; some plants are removed from sprayed areas almost completely (e.g., raspberry [rubus idaeus]), and others take years to return to pre-treatment levels of production (e.g., blueberries [viburnum angustifolium and v. myrtilloides]). in our continual interactions with the land, we are acutely aware of the new annual disturbances because logging and the associated silvicultural activities (e.g., spraying) are concentrated along roads. moose and our other food sources become farther from roads and more difficult to find; we retain in our memories records of previous years’ silvicultural activities and we avoid harvesting food in disturbed areas. some community members cease to use treated areas entirely, even after ecological and silvicultural processes restore disturbed areas and make them appear natural again. though the reward is great, hunting requires significant time and economic input on the part of the hunter; 68% of responding hunters now travel >2 hours to moose hunt. even as roads are used to access our territory, the concentrated disturbances to the forest, including extensive logging road networks, create an ever growing perception of cumulative negative impacts. people who eat more moose in winter are those most concerned that herbicides affect the food system (r = 0.60, p = 0.04). economically, all losses of food equate to losses of local production opportunities. current forest and moose management guidelines and our hunting rights forest management guidelines require the collection of our “values” in the form of the native values background report prepared by the industrial and/or provincial forest managers. generally, our community is notified of meetings held in the nearest provincial municipality (greenstone, ontario) as they relate to forest management planning; no meaningful consultation takes place in our community. for the past 5 years our community has been informed directly of only a single information session pertaining to forest management plan amendments in a single forest management units imposed upon our territory; few community members travel to these meetings. the bureaucracy is confusing as our hunters could be in 1 of 5 wildlife management units (17, 18a, 18b, 19 or 21a) or in 1 of 4 forest management units (ogoki, lake nipigon, armstrong, or kenogami forest). each of these jurisdictions is managed according to directive given by government policies and guidelines. the managers responsible for these jurisdictional units must address the “recreationally focused” directive of the government of ontario (e.g., heritage hunting and fishing act 2002), as well as our constitutionally protected rights to harvest moose. finding the balance is often politicized and the debate surrounding hunting rights has been disputed for decades among the citizens and governments of canada. we feel we hunt under duress. in formal debates, the majority of canadians agree that aboriginal people should have the right to subsistence hunting. the supreme court has provided clear guidance on the application of these rights, the circumstances by which they can be infringed upon, and a test by which to determine the validity of arguments for infringement. most importantly, the constitution act was amended in 1982 to include section 35, which protects aboriginal and treaty rights. much of the problem seems alces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 171 to lie in an apparent disconnection between informal public opinion and the official guidance for policy directives and management decisions. while there are many stakeholders on the land base, management initiatives seem to favour wealthy, mainly urban, sport hunters. for many in our community, hunting and fishing provides valuable economic input as well as invaluable cultural, spiritual, and recreational opportunity. in hard economic times, moose and other sources of meat from our traditional territory can be crucial to our survival (george et al. 1995). ontario’s new moose management policy states that “moose management will respect aboriginal peoples’ unique perspectives, traditional knowledge and practices related to moose and the exercise of their constitutionally protected aboriginal or treaty rights.” but this guiding principle retains existing jurisdictional constructs, offering respect in lieu of seeking guidance. respecting our values means acknowledgement of our on-going use and attempt to accommodate our perspectives. seeking guidance means acknowledgement of our expertise and adapting practices, past to present. moving toward reconciliation the actions of decision-makers are made possible by complex governance structures. our inherent marginalization in these structures imposed from the outside limits the extent of our participation in decision-making. to those current architects of government policy and programs, our land is one of many jurisdictions to manage in a vast expanse of crown forests. originally, the british royal family’s wealth and security was afforded by a global amalgamation of crown lands throughout the empire, only made possible by the treaties and land surrenders in areas previously occupied and governed by indigenous people. today, the crown still exercises its rights, granted in these treaties, to build structures supporting continued development and management of land, with natural resource management authority afforded to the provinces of canada. ontario’s jurisdictions, held by the ministries, and the policies and guidelines set by various authorities acting on behalf of ontario or the crown are maintained to continue foreign settlement and the extraction of resources to distant corporations. the constitution act (1982) was structured to support greater independence, protecting aboriginal and treaty rights (section 35), a new structure upon which to build a new relationship. but the aim of all management activities remains on facilitating extraction of resources, and sustained extraction includes accommodations for other uses as our uses are marginalized. we prefer to think and act holistically, engaging all those using our shared lands to manage them together. our economy emerged in this place. while the context for traditional use of the land has changed over time, many resilient elements remain. those aspects of the economy carried forward by culture and tradition remain the backbone of our community’s sustainability. our constitutionally protected rights to access our lands and sustain our community through contextually appropriate foods are jeopardized when they do not guide development. practices and guiding principles rooted in this place are most appropriate to our future. the new moose habitat protection afforded by the site and stand guidelines for the crown forest sustainability act includes provisions for consultations sensitive to our traditions. the directive in ontario’s moose management strategy to respect traditional values represents further potential to include our community’s economy within the realm of other values. we are deeply concerned about the future of our community as more development occurs. we hope that readers understand that management of sport hunting of moose and forest management without acknowledging first nations practices will cause conflict to escalate. our community surveys taught us more about not only the economic, social, and ontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 172 cultural traditions we have maintained within our community, but also about the impacts of marginalizing our use. moose managers and forest managers need to balance consumption and conservation of resources for diverse interests. the results of our survey with moose hunters in aroland and ginoogaming first nations showed the respondents were harvesting 87 moose per year. bissett (2002) reported a total of 210 annual moose harvests recorded by the mnr in the wildlife management units located within our traditional territory. as our harvests are not taken into account in the mnr record, we estimate that there is an error of approximately 40% in the moose harvest reported by the mnr in our traditional territory. as this estimate is based on data from 40 hunters in two of at least 5 first nations sharing overlapping traditional territories, claiming 40% error is likely a conservative estimate. the effects of not accounting for our moose harvest could adversely impact the management of moose and the viability of future populations, but are we to blame? by continuing to restrict dialogue, our uses are not accounted for and an underestimation of moose harvested is allowed to continue by the mnr. a review of the mnr moose tag allocation is currently underway and ontario’s moose management strategy indicates that the government is committed to improving the methods used to estimate moose populations and determine harvest allocation. therefore, it is time to incorporate our perspective into moose population estimates and management planning through a meaningful, consistent, and transparent consultation. developing a working relationship with ours and other first nations communities is imperative to effectively manage moose in ontario. but to date, the mnr solicited our knowledge only as an afterthought (reviewing plans and proposed changes to legislation or policy), not as a consultation with knowledge-holders (informing process and contributing to policy development). we agree with the conclusions of watson and huntington (2008) after their moose hunting trip: that the way to proliferate perspectives is not to translate or interpret knowledge, but to change the way that knowledge is represented to make different perspectives explicit when describing everyday life or scientific knowledge. we believe the incorporation of our perspective in a meaningful way will aid wildlife biologists to manage moose populations more effectively in the future. it will also ensure our use will be recognized and sustained for future generations. moving into the future is about weighing costs and benefits of each new step. together we should be able to look at each period of transition in the bridging of two cultures and be ready to admit when corrections were not made, which would have kept benefits outweighing costs for all users of the land. we are aware that the dominating, jurisdictional traditions guiding current forest and wildlife management are deeply entrenched and very difficult to uproot (caza and neave 2000). however, the sustainability of our community is tied to the sustainability of our economy. misrepresentation of this fact in the current management system has encouraged marginalization of our knowledge. can we review the traditions of the past and recognize them as a part of a whole that includes new traditions and new trade possibilities? acknowledgements we wish to acknowledge and thank the many respondents to our survey, who are our friends and family in aroland first nation, and who formed the collective views we share in this paper. we also thank matthew lebron, lakehead university, for assembling forestry records from annual work schedules and reports and putting them into a useable format, including the maps that are part our story. charlotte bourdignon, philip brown, and chris leale, mnr, helped with accessing and interpreting these records. peter raalces vol. 47, 2011 leblanc et al. – ontario first nation community perspectives 173 sevych of ginoogaming first nation helped to conduct surveys with moose hunters in his community. this project was funded by the first nations environmental contaminants program of health canada. connie nelson and mirella stroink of lakehead university’s food security research network, coapplicants to the grant awarded to aroland first nation by health canada, offered sage advice during the drafting of this study. peggy smith, also of lakehead university, helped supervise christopher pereira in his ambition to learn more about first nations and moose in ontario. finally, we thank the chief and council of aroland first nation for granting us permission on behalf of the community to tell its story. references aboriginal affairs and northern development canada. 2008. the james bay treaty treaty no. 9 (made in 1905 and 1905) and adhesions made in 1929 and 1930. (accessed april 2010). bergman, m., and s. åkerberg. 2006. moose hunting, forestry, and wolves in sweden. alces 42: 13-23. bissett, a. r. 2002. 1999 and 2000 moose harvest in ontario. northwest science & information, ontario ministry of natural resources, thunder bay, ontario, canada. bridgland, j., t. nette, c. dennis, and d. quann. 2007. moose on cape breton island, nova scotia: 20th century demographics and emerging issues in the 21st century. alces 43: 111-121. caza, c. l., and d. neave. 2000. new millennium forestry and the fate of wildlife. forestry chronicle 76: 109-115. crichton, v. f. j. 1981. the impact of treaty indian harvest on a manitoba moose herd. alces 17: 56-63. diener, e., r. a. emmons, r. j. larsen, and s. griffin. 1985. the satisfaction with life scale. journal of personality assessment 49: 71-75. driben, p. 1985. aroland is our home: an incomplete victory in applied anthropology. ams press, new york, new york, usa. _____, d. j. auger, a. n. doob, and r. p. auger. 1997. no killing ground: aboriginal law governing the killing of wildlife among the cree and ojibwa of northern ontario. ayaangwaamizin 1: 108. feit, h. a. 1987. north american native hunting and management of moose populations. swedish wildlife research supplement 1: 25-42. george, p., f. berkes, and r. preston. 1995. aboriginal harvesting in the moose river basin: a historical and contemporary analysis. canadian review of sociology and anthropology 32: 69-91. government of canada. 1982. the constitution act, 1982, being schedule b to the canada act 1982 (uk), c 11. hamilton, s. 2002. environmental studies, environmental reconstruction, ethnography. three articles prepared for the encyclopedia of historical archaeology, e. orser, editor. routledge press, london, england. herbert, r. 2009. meaningful aboriginal consultation in canada. a review of the first nation, inuit, and métis right to consultation and accommodation on wildlife resource management and hunting as defined by common law. christian aboriginal infrastructure developments corporation. (accessed april 2010). kay, c. e. 1997. aboriginal overkill and the biogeography of moose in western north america. alces 33: 141-164. kofinas, g. 1993. subsistence hunting in the global economy: contributions of northern wildlife co-management to community ontario first nation community perspectives – leblanc et al. alces vol. 47, 2011 174 economic development. making waves: a newsletter for community economic development practitioners in canada 4 (3). lautenschlager, r. a. 1992. effects of conifer release with herbicides on moose: browse production, habitat use, and residues in meat. alces 28: 215-222. larter, n. c. 2009. a program to monitor moose populations in the dehcho region, northwest territories, canada. alces 45: 89-99. lynch, g. m. 2006. does first nation’s hunting impact moose productivity in alberta? alces 42: 25-31. mayer, f. s., and c. m. frantz. 2004. the connectedness to nature scale: a measure of individuals’ feeling in community with nature. journal of environmental psychology 24: 503-515. morrison, j. 1986. treaty research report, treaty no. 9 (1905-1906). unpublished report, treaties and historical research centre, indian and northern affairs canada, ottawa, canada. natcher, d. c., m. calef, o. huntington, s. trainor, h. p. huntington, l. dewilde, s. rupp, and f. stuart chapin iii. 2007. factors contributing to the cultural and spatial variability of landscape burning by native peoples of interior alaska. ecology and society 12: 7. (accessed april 2010). rempel, r. s., p. c. elkie, a. r. rodgers, and m. j. gluck. 1997. timber management and natural-disturbance effects on moose habitat: landscape evaluation. journal of wildlife management 61: 517-524. rogers, s.,and m. black. 1976. subsistence strategy in the fish and hare period, northern ontario: the weagamow ojibwa, 1880-1920. journal of anthropological research 32: 1-43. timmermann, h. r., and a. r. rodgers. 2005. moose: competing and complementary values. alces 41: 85-120. waisberg, l. g., and t. e. holzkamm. 1993. a tendency to discourage them from cultivating: ojibwa agriculture and indian affairs administration in northwestern ontario. ethnohistory 40: 175-211. watson, a., and o. h. huntington. 2008. they’re here i can feel them: the epistemic spaces of indigenous and western knowledges. social and cultural geography 9: 257-281. winterhalder, b. 1983. history and ecology of the boreal zone in ontario. pages 9-54 in a. t. steegmann, jr., editor. boreal forest adaptations: the algonkians of northern ontario. plenum press, new york, new york, usa. alces vol. 48, 2012 palo et al. – digestive system of moose 7 seasonal variation of phenols, nitrogen, fiber, and in vitro digestibility in swedish moose r thomas palo1, peter a jordan2, åke pehrson3 and hans staaland4† 1department of natural sciences, engineering and mathematics, mid sweden university, se-851 70 sundsvall, sweden; 2university of minnesota, department of fish and wildlife, st. paul, minnesota 55108-1036 usa; 3department of conservation biology, swedish university of agricultural sciences, grimsö wildlife research station, 730 91 riddarhyttan, sweden; 4department of ecology and natural resource management (ina), norwegian university of life sciences. p.o. box 5003, no-1432 ås norway; †prof. h. staaland died in 2009. abstract: understanding how different components of food are processed and digested within the compartments of the digestive tract of large herbivores has important implications in their foraging behaviour, nutritional ecology, and techniques for measuring diet composition and nutritional quality of forage. analysis of contents from different compartments of the digestive tract of moose in central sweden showed that neutral detergent fiber (ndf) and nitrogen (n) content varied throughout the digestive tract and among individual moose (alces alces). total phenols (tp) had an inconsistent pattern throughout the digestive tract, possibly reflecting variation in diet composition and phenol patterns. the study moose were divided into 2 groups; the winter group had low n in the digestive tract and high ndf and dry matter content, and the summer group had high levels of n and low ndf and dry matter content. the phenol platyphyllane, indicative of consumption of dormant silver birch (betula pendula), was detected throughout the contents of the digestive tract in 2 animals in the winter group. the winter moose had higher ndf than summer moose, indicating the seasonal change in diet quality. in vitro organic matter digestibility (ivomd) was not different between the summer and winter diets. the effect of birch phenols on ivomd was concentration-dependent; differences between seasons were apparent at only the highest concentration. the 2 groups had marked differences in digestive content of major nutrients, ndf, and ability to digest forage which were consistent with typical variation in seasonal diet quality. alces vol. 48: 7-15 (2012) key words: alces alces, digestion, fiber, gastrointestinal tract, plant phenol, rumen function. the digestive tract is a dynamic system that changes throughout the year as the diet varies with seasonal change in forage availability (weckerly 1989, cork and foley 1991, pehrson et al. 1997, hume 1999). diet selection of browsing herbivores is complex because diet composition among individuals can be substantial and diets vary both seasonally and geographically (palo and wallin 1996). seasonal variation in the capacity of rumen microorganisms to digest fiber and handle secondary compounds is also expected (palo et al. 1985, pehrson and faber 1994). digestibility is critical in evaluating food utilization by herbivores because it accounts for both passage time of food in the digestive tract and nutritional benefit from forage (clauss et al. 2007). in ruminants, breakdown of ingested plants is facilitated by mastication and microbial degradation that act as sequential processes that reduce particle size enabling passage from the rumen (duncan and poppi 2008). thus, the rate of and maximal digestibility may be limited by plant metabolites that depress microbial growth, hence reduce substrate extraction from food. further, tannins in the diet may reduce maximal protein digestibility by reducing absorption in the lower digestive tract and eventual loss in faeces (robbins et al. 1987). digestive system of moose – palo et al. alces vol. 48, 2012 8 a critical factor affecting food intake rate and digestibility is neutral detergent fiber (ndf) that can reduce intake as its content increases in forage (meyer et al. 2010). phenols in plants may depress digestibility of ndf and organic matter thereby slowing fermentation and supply of nitrogen for microbial growth (lundberg and palo 1993, duncan and poppi 2008, meyer et al. 2010). moose are commonly classified as a concentrate selector (hoffman 1989, illius and gordon 1991, robbins et al. 1995). yearround they consume birch (betula spp.) and willow (salix spp.) as staple foods in northern sweden, but also consume ~40 other plant species (palo and wallin 1996). scots pine (pinus sylvestris) in winter and bilberry (vaccinium myrtillus) in autumn are considered important foods (cederlund et al. 1980, palo and wallin 1996). the in vitro digestibility of these plant species varies seasonally due to changes in chemical composition (palo et al. 1985, pehrson and faber 1994, stolter 2008). for example, birch species vary in phenolic concentrations by altitude, individual trees, tissues within trees, and tree height (palo et al. 1992, santamour and lundgren 1996, hodar and palo 1997, rousi et al. 1997, nordengren et al. 2003). similar seasonal patterns but with other chemical components are found in scots pine, willow, alder (alnus spp.), and aspen (populus spp.) (palo 1984, bryant et al. 1987, sunnerheim and hämäläinen 1992). in birch, a majority of the phenol compounds are composed of glycosides with fairly low molecular weight as compared to tannins (santamour and lundgren 1996, sunnerheim et al. 1988). platyphylloside is the predominant phenol compound in winter twigs of silver birch (b.pendula), and it inhibits in vitro and in vivo digestibility in moose, goats (capra capra), rabbits (oryctolagus cuniculus), and hares (lepus spp.) (palo 1985, 1987, sunnerheim et al. 1988, iason and palo 1991, palo et al. 1997, bratt and sunnerheim 1999). plant species composition in the rumen of moose is not a good predictor of digestibility variation among individuals (pehrson and faber 1994). changes in the species composition of rumen microorganisms by season and food type, and their digestive capacity may also be critical factors in food utilization and digestibility. changes in fiber, nitrogen, and phenols are the components that most drastically change with season and between plant species (palo et al. 1985, risenhoover 1989). thus, understanding their changes in the digestive process is important to further interpret the digestive process and nutritional benefits of forage. the hypothesis of this investigation was that moose show diminishing nitrogen (n) concentrations from rumen to rectum, and conversely, higher proportion of ndf towards the rear of the digestive tract due to digestion and absorption in the intestine. further, it could be expected that the level of n in the rumen is decisive for the ability of rumen microorganisms to digest food and metabolize plant phenols. we expected that moose in winter have higher ndf and lower n concentrations throughout the digestive tract, and less ability to handle phenols in the diet as compared with other seasons. materials and methods we studied 6 free-ranging moose shot in april-november at grimsö wildlife research station in central sweden; live body weight ranged from 115-189 kg and ages were estimated at 6-13 months. strictly regulated hunting outside of the november hunting season precluded a larger sample size and limited us to a general trend analysis. the functional segments of the digestive tracts were separated and their content was isolated by binding with a thread (fig. 1). gut material was sampled/collected in triplicate from the isolated parts of the digestive tract (staaland et al. 1992, pehrson et al. 1997); material was not present (collected) from each compartment from all animals. alces vol. 48, 2012 palo et al. – digestive system of moose 9 the intestinal samples were dried at 70o c for 48 h and stored until later analysis for phenols, n, and fiber. for phenol analyses, the dry content was eluted with etoh, cleaned through a sepac c18 cartridge (merck inc.), and the remaining etoh phase was dried in an evaporator and dissolved in distilled water. the dilution of extracts for the analysis was 660x from the crude extract. this procedure optimizes the analysis of total water soluble phenols by the folin-ciocalteau method (palo 1985, stolter 2008). this method measures the concentrations of low molecular phenol aglycones and quantifies the amount as with more advanced chemical methods (sunnerheim et al. 1988, hodar and palo 1997, bratt and sunnerheim 1999). the folin-ciocalteu reaction was measured in a spectrophotometer after reaction for 2 h at 740 nm and at room temperature as described by palo et al. (1985). the etoh extract of content from different digestive compartments was analysed for the presence of the compounds platyphyllone and platyphyllane by thin layer chromatography (tlc) (merck silica gel hf-254). these compounds are major phenol metabolites of silver birch twigs in the winter and have been found in the rumen of moose (palo 1987, sunnerheim et al. 1988, sunnerheim and bratt 2004). the extract from the gut content was run on tlc using the solvent chloroform:methanol:water (80:15:1 v/v). phenols were detected by spraying the tlc plates with diazotized sulphanilic acid (sigma-aldrich) in 10% sodium carbonate (w/v), followed by 50% sulphuric acid (v/v). platyphyllone has a rf = 0.53 and platyphyllane a rf = 0.68 on tlc with the solvent used. for ndf measurements, 0.5 g of dried gut content from 5 animals was put in glass filter tubes with a pore size of 0.2 mm and incubated with ndf solution according to van soest and wine (1967). after incubation with solvents, the tubes were dried at 105° c, weighed, and the material combusted at 600° c. after cooling, the tubes were weighed again to calculate ash weight; ndf is expressed as the ash free weight. from one animal, only colon and rectum samples were analysed. the gut content was analysed for total n according to the kjeldahl method using a cnanalyser. in vitro dry matter digestibility (ivomd) was measured with fresh rumen content from the collected moose. immediately after killing, rumen contents were removed and transferred to thermos flasks and saturated with co2. within 1 h of sampling, the content was filtered through a cloth and mixed with a mcdoughal buffer stock solution 1:50 (palo 1985). for ivomd measurements, 0.5 g of dried and milled timothy (phleum pratense) was put in glass filter tubes with a pore size of 0.2 mm. the etoh extract from birch (as described above) was added at concentrations 1, 3, and 6 times that naturally found in 2-6 mm birch twigs in winter. the tubes were dried at 70o c over night, filled with stock buf fig. 1. demarcation of the functional compartments of the moose digestive tract used in this study (from pehrson et al. 1997). digestive system of moose – palo et al. alces vol. 48, 2012 10 fer solution, saturated with co2, and incubated for different times up to 96 h in a water bath at 37o c. the tubes were shaken twice daily during the incubation. after digestion, the tubes were washed with distilled water and rinsed with acetone. they were then dried at 105o c, weighed, and the material combusted at 600o c; after cooling the tubes were reweighed to obtain the ash weight. the organic matter disappearance was compared to those of a control that consisted of timothy treated with pure ethanol and dried. the ivomd was calculated as: (ivomdcontrol − ivomdsample) / ivomdcontrol anova and student t-statistics was used for analysis of the data. results based on the mean n concentrations throughout the digestive tract, 2 distinct groups of moose were evident; a summer group with high n concentration (x = 4.55, sd = 1.55, n = 3) and a winter group with low n concentration (x = 2.65, sd = 2.3, n = 3) (student t-test, p = 0.038, df = 19). the summer group consisted of moose shot in june and early november, and the winter group had moose shot in late november and april. the mean concentration of n in the individual compartments of the digestive tract was somewhat stable at ~4 mg/g in the summer group and ~2 mg/g in the winter group, with the exception of the highest levels at ~9 mg/g in the duodenum in both seasons (fig. 2). these differences occurred in all parts of the digestive tract except the duodenum (fig. 2). the phenol concentration varied among compartments but no difference was found between the groups (spearman rank r = 0.71, p <0.1, df = 4; fig. 3). platyphyllane (from platyphylloside found in silver birch twigs) occurred throughout the digestive tract of 2 winter moose only; these animals had similar amounts of total phenols as those without platyphyllane. dry matter content was high in the omasum, declined to the jejunum, and then increased to the rectum (fig. 4), indicating major water absorption in the omasum and lower digestive tract. dry matter in the rectum ranged from 15-30% among individuals (fig.5), and dry matter and n concentration were inversely related between the summer and winter groups. the % ndf varied (5-60%) throughout the digestive tract with similar seasonal patterns; concentration was highest in the front and rear compartments and lowest in the duodenum (fig. 5). however, ndf was lower in summer (~40%) than in winter (~60%) (fig. 5). therefore, the summer group was characterized by high n concentration and low dry matter and ndf; the opposite described the winter group. fig. 2. nitrogen (n) concentration (mg/g) of gut contents in compartments of the moose digestive tract by season in sweden (x and sd). fig. 3. concentration of total phenols of gut contents in compartments of the moose digestive tract by season in sweden; absolute units (740 nm) (x and sd). alces vol. 48, 2012 palo et al. – digestive system of moose 11 digestibility of birch twigs did not differ between seasons (winter (low n) x = 23.49%, sd = 7.16; summer (high n) x = 23.78%, sd = 0.97), nor did digestibility of hay. addition of phenols reduced ivomd for both birch twigs and hay, but differences between seasons were only apparent at the highest concentration (f = 134.4, p <0.001, df = 2; fig. 6). discussion we have shown that the concentrations of n, ndf, phenols, and dry matter vary within the compartments of the digestive tract and with season. moose collected in winter, versus summer, generally had lower concentrations of n and higher concentrations of ndf and dry matter. further, ivomd was reduced by the concentration of phenols in the diet, but the effect was independent of season except at the highest concentration of phenols. these results corroborate with previous research on moose indicating that rumen content varies seasonally in dry matter content, plant species composition, and digestibility (cederlund and nyström 1981, schwartz et al. 1984, palo and wallin 1996, pehrson et al. 1997). in particular, ndf and digestibility reducing compounds (e.g., tannins) influence digestion and passage rate, hence rate of food intake (robbins et al. 1987, clauss et al. 2007, spalinger et al. 2010). the concentration of n was highest in the duodenum reflecting excretion of endogenous enzymes such as cellulase and nuclease, and was consistent with digestive studies indicating that n compounds are rapidly absorbed in the jejunum (leng and nolan 1984). overall, n concentration in rumen content and faeces reflects the dietary intake of n and supports the use of faecal output as an indicator of range quality (renecker and hudson 1985, staaland et al. 1992, massey et al. 1994, leslie et al. 2008, palo and olsson 2009). however, moose collected in november were represented in both the high and low n groups, an apparent contradiction. one possible explanation is that moose are in a seasonal transition of diet fig. 6. in vitro dry matter digestibility (ivomd) of hay with summer and winter moose rumen inocula; birch phenols were added at 1-6 x that occurring naturally in winter birch twigs (x and sd). different letters denote significant differences among groups (p <0.05). fig. 4. dry matter content (%) of gut contents in compartments of the moose digestive tract by season in sweden (x and sd). fig. 5. neutral detergent fiber (ndf, % average of 2 animals) of gut contents in compartments of the moose digestive tract by season in sweden. digestive system of moose – palo et al. alces vol. 48, 2012 12 which could vary individually and locally in november. plant secondary compounds greatly influence food selection and nutritional value of plants because they reduce digestibility or impose a cost for detoxification (palo 1985, robbins et al. 1987, iason and palo 1991, sunnerheim and bratt 2004). moose have few forage options in northern areas during winter and often consume plants low in palatability, high in secondary metabolites, and of low nutritional value. for example, birch is considered of low quality due to low ivomd and high phenol concentration (palo et al. 1992, rousi et al. 1997). we found that silver birch twigs were in the winter diet because of the presence of platyphyllane that is an inhibitory substance only present in winter twigs of silver birch (palo et al. 1992). because birch could be consumed in other seasons, it is only applicable as a chemical marker in winter. total phenolic glycosides and free phenols varied in the compartments of the digestive tract, but faecal analysis revealed 2 groups with high and low phenol content. the high phenol group included moose collected in april and probably reflects high phenol intake associated with consumption of birch, however, this group also included one animal collected in june. it is possible that the method used to measure total phenols is compromised; for example, if an animal is excreting excess free aromatic amino acids such as tyrosine and tryptophane, the folin-ciocalteau method might overestimate phenol content (folin and ciocalteau 1927). no analysis of amino acid composition was performed to control for this possibility. another explanation is that tannins are more common in the diet in june and would bind to salivary or dietary proteins that would be excreted in faeces as indicated by higher phenol concentrations in faeces (hagerman and robbins 1993). diet composition of moose varies more in summer than winter which is reflected in the phenol concentration patterns, and diet composition may also vary widely within season and locally (pehrson and faber 1994, palo and wallin 1996). the analysis of ndf was done on only 5 animals, but 2 distinct groups of high (winter) and low ndf concentration (june and early november) was apparent. presumably, low sample size accounted for the lack of statistical difference between the groups. the range of ndf contents (30-60%) indicates that intake of fibrous foods is high during all seasons and individuals; these values were similar to those measured in feeding trials with several domestic and captive wild ruminants including moose (palo et al. 1985, renecker and hudson 1985, lechner et al. 2010, meyer et al. 2010). high ndf impairs food intake in most herbivores, yet selective feeding on smaller diameter, more digestible twigs is likely an important behaviour to relax this constraint (vivås et al. 1991, palo et al. 1992, hodar and palo 1997). based on the equation of van soest (1994) as modified by meyer et al. (2010) for the relationship between dry matter intake and ndf in cervids, the estimated daily food intake is in the range 52-62 g dm kg-0.75d-1. since the animals in this study had rumen fill that corresponded to 79 g dm kg-0.75, it could be argued that food intake was in the range of 66-78% of rumen fill, with higher intake in summer (pehrson et al. 1997). since pehrson et al. (1997) found no seasonal difference in the wet weight of gut content in moose, ndf would appear to be the single most important factor affecting dry matter intake by moose (meyer et al. 2010). acknowledgements this paper is in honour of my co-authors and research friends; professor emeritus peter a jordan on his 80-years anniversary in 2010, dr. åke pehrson who is retiring in 2011, and in memory of professor hans staaland. we greatly appreciated the service from the hunting team at grimsö research station and permit to hunt moose outside the legal season. ms. emilia fuchs assisted in the laboratory alces vol. 48, 2012 palo et al. – digestive system of moose 13 at the swedish agricultural university at uppsala. comments by dr. m. clauss, s. öhmark, and 2 anonymous reviewers on an earlier draft of this manuscript were greatly appreciated. references bratt, k., and k. sunnerheim. 1999. synthesis and digestibility inhibition of diarylheptanoids: structure-activity relationship. journal of chemical ecology 25: 2703-2713. bryant, j. p., f. s. chapin, p. b. reichard, and t. p. clausen. 1987. response of winter chemical defense in alaska paper birch and green alder to manipulation of plant carbon/nutrient balance. oecologia 72: 510-514. cederlund, g., h. lundqvist, g. markgren, and f. stålfelt. 1980. foods of moose and roe deer at grimsö in central sweden: results of rumen content analyses. swedish wildlife research 11: 167-247. _____, and a. nyström. 1981. seasonal differences between moose and roe deer in ability to digest browse. holarctic ecology 4: 59-65. clauss, m., a. schwarm, s. ortmann, j. w. streich, and j. hummel. 2007. a case of non-scaling in mammalian physiology? body size, digestive capacity, food intake, and ingesta passage in mammalian herbivores. comparative biochemistry and physiology a 148: 249-265. (doi:10.1016/j.cbpa.2007.05.024) (accessed october 2009). cork, s., and b. foley. 1991. digestive and metabolic strategies of arboreal folivores in relation to chemical defences in temperate and tropical forests. pages 133-166 in r. t. palo and c. t. robbins, editors. plant defenses against mammalian herbivory. crc press, boca raton, florida, usa. duncan, a. j., and d. p. poppi. 2008. nutritional ecology of grazing and browsing ruminants. pages 89-116 in i. j. gordon and h. h. t. prins, editors. the ecology of browsing and grazing. springer verlag, berlin, germany. folin, o., and v. ciocalteau. 1927. on tyrosine and tryptophane determinations in proteins. journal of biological chemistry 73: 627-650. hagerman, a. e., and c. t. robbins. 1993. specificity of tannin-binding salivary proteins relative to diet selection by mammals. canadian journal of zoology 71: 628-633. hodar, j. a., and r. t. palo. 1997. feeding by vertebrate herbivores in a chemically heterogenous environment. ecoscience 4: 304-310. hoffman, r. r. 1989. evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. oecologia 78: 443-457. hume, i. 1999. marsupial nutrition. cambridge university press, cambridge, united kingdom. illius, a. w., and i. j. gordon. 1991. prediction of intake and digestion in ruminants by a model of rumen kinetics integrating animal size and plant characteristics. journal of agricultural science 116: 145. iason, g. r., and r. t. palo. 1991. effects of birch phenolics on a grazing and browsing mammal: a comparison of hares. journal of chemical ecology 17: 1733-1743. lechner, i., p. barboza, w. collins, j. fritz, d. gunther, b. hattendorff, j. hummel, k. h. sudekum, and m. clauss. 2010. differential passage of fluids and different sized particles in fistulated oxen (bos primigenius), muskoxen (ovibos moschatus), reindeer (rangifer tarandus) and moose (alces alces): rumen particle size discrimination is independent from content stratification. comparative biochemistry and physiology, part a 155: 211-222. leng, r. a., and r. v. nolan. 1984. nitrogen digestive system of moose – palo et al. alces vol. 48, 2012 14 metabolism in the rumen. journal of dairy science 67: 1072-1089. leslie, d. m., r. t. bowyer, and j. a. jenks. 2008. facts from feces: nitrogen still measures up as a nutritional index for mammalian herbivores. journal of wildlife management 72: 1420-1433 lundberg, p., and r. t. palo. 1993. resource use, plant defenses, and optimal digestion in ruminants. oikos 68: 224-228. massey, b. n., f. w. weckerly, c. e. vaughn, and d. r. mccullough. 1994. correlations between fecal nitrogen and diet composition in free-ranging black-tailed deer. southwestern naturalist 39: 165–170. meyer, k., j. hummel, and m. clauss. 2010. the relationship between forage cell wall content and voluntary food intake in mammalian herbivores. mammal review 40: 221-245. nordengren, c., a. hofgaard, and j. p. ball. 2003. availability and quality of herbivore winter browse in relation to tree height and snow depth. annales zoologici fennici 40: 305-314. palo, r. t. 1984. distribution of birch (betula spp.), willow (salix spp.) and poplar (populus spp.): secondary metabolites and their potential role as chemical defense against herbivores. journal of chemical ecology 10: 499-520. _____. 1985. chemical defence in birch: inhibition of digestibility in ruminants by phenolic extracts. oecologia 68: 10-14. _____. 1987. chemical defense in woody plants and the role of digestive systems of herbivores. pages 103-107 in f. d. provenza, j. t. flinders, and e. d. mcarthur, editors. proceedings of symposium on plant-herbivore interactions. snowbird, utah, 7-9 august, 1985. intermountain research station, united states forest service, ogden, utah, usa. _____, r. bergstrom, and k. danell. 1992. digestibility, distribution of phenols, and fiber at different twig diameters of birch in winter: implications for browsers. oikos 65: 450-454. _____, j. h. gowda, and j. hodar. 1997. consumption of two birch species by mountain hares (lepus timidus) in relation to resin and phenolic content. gibier et faune 14: 385-393. _____, and g. a. olsson. 2009. nitrogen and carbon concentrations in stomach contents of bank voles (myodes glareolus): does food quality determine abundance? open ecology journal 2: 86-90. _____, k. sunnerheim, and o. theander. 1985. seasonal variation of phenols, crude protein and cell wall content of birch (betula pendula roth.) in relation to ruminant in vitro digestibility. oecologia 65: 314-318. _____, and k. wallin. 1996. variability in diet composition and dynamics of radiocaesium in moose. journal of applied ecology 33: 1077-1084. pehrson, a., and w. e. faber. 1994. individual variation of in vitro dry matter digestibility in moose. journal of range management 47: 392–394. _____, r. t. palo, h. staaland, and p. a. jordan. 1997. seasonal variation in weight of functional segments of the gastrointestinal tract and its content in young moose (alces alces). alces 33: 1-10. renecker, l. a., and r. j. hudson. 1985. estimation of dry matter intake of free-ranging moose. journal of wildlife management 49: 785-792. risenhoover, k. a. 1989. composition and quality of moose winter diets in interior alaska. journal of wildlife management 53: 568-577. robbins, c. t., t. a. hanley, a. e. hagerman, o. hjeljord, d. l. baker, c. c. schwartz, and w. w. mautz. 1987. role of tannins in defending plants against ruminants: reduction in protein availability. ecology 68: 98-107. _____, d. spalinger, and w. van hoven. alces vol. 48, 2012 palo et al. – digestive system of moose 15 1995. adaptation of ruminants to browse and grass diets: are anatomical-based browser-grazer interpretations valid? oecologia 103: 208-213. rousi, m., j. tanvanainen, h. henttonen, d. a. herms, and i. uotila. 1997. clonal variation in susceptibility of white birches (betula spp.) to mammalian and insect herbivores. forest science 43: 396-402. santamour, f. s., and l. n. lundgren. 1996. distribution and inheritance of platyphylloside in betula. biochemical systematics and ecology 24: 145-156. schwartz, c. c., w. l. regelin, and a. w. franzmann. 1984. seasonal dynamics of food intake in moose. alces 20: 223-244. spalinger, d. e., w. b. collins, t. a. hanley, n. e. cassara, and a. m. carnaha. 2010. the impact of tannins on protein, dry matter, and energydigestion in moose (alces alces). canadian journal of zoology 88: 977-987. staaland, h., å. pehrson, p. a. jordan, and r. t. palo. 1992. seasonal variation of alimentary mineral and nitrogen pools in the moose. comparative biochemistry and physiology 102a: 163-171. stolter, c. 2008. intra-individual plant response to moose browsing: feedback loops and impacts on multiple consumers. ecological monographs 78: 167-183. sunnerheim, k., and k. bratt. 2004. identification of centrolobol as the platyphylloside metabolite responsible for the observed effect on in vitro digestibility of hay. journal of agricultural and food chemistry 52: 5869-5872. _____, and m. hamalainen. 1992. multivariate study of moose browsing in relation to phenol pattern in pine needles. journal of chemical ecology 18: 659-672. _____, r. t. palo, o. theander, and p. g. knutsson. 1988. chemical defense in birch: platyphylloside, a phenol from betula pendula inhibiting digestibility. journal of chemical ecology 14:549559. van soest, p. j. 1994. nutritional ecology of the ruminant, second edition. cornell university press, ithaca, new york, usa. _____, and r. h. wine. 1967. use of detergents in the analysis of fibrous feeds. journal of association of agricultural chemistry 50: 50-55. vivas, h., b.-e. saether, and r. andersen. 1991. optimal twig size selection of a generalist herbivore, the moose alces alces: implications for plant-herbivore interactions. journal of animal ecology 60: 395-408. weckerly, f. w. 1989. plasticity in the length of hindgut segments of white-tailed deer (odocoileus virginianus). canadian journal of zoology 67: 189-193. alces29_9.pdf alces36_93.pdf alces34(1)_31.pdf alces29_163.pdf alces36_1.pdf alces32_163.pdf alces34(2)_467.pdf alces vol. 46, 2010 rea et al. moose-train interactions 183 youtubetm insights into moose-train interactions roy v. rea1, kenneth n. child2, and daniel a. aitken3 1natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia v2n 4z9, canada; 26372 cornell place, prince george, british columbia, v2n 2n7, canada; 3college of new caledonia, 3302 22nd avenue, prince george, british columbia v2n 1p8, canada. abstract: to gain a better understanding of the behavioral aspects of moose-train encounters, we reviewed videos of ungulate-train interactions available on youtubetm and from train operators. video footage consisted of 21 animal-train encounters including moose (alces alces; 47.4%), cattle (bos taurus; 15.8%), deer (odocoileus spp.; 10.5%), elk (cervus elaphus; 10.5%), camels (camelus dromedarius; 10.5%), and sheep (ovis aries; 5.3%). footage was recorded predominantly in snow-free conditions, but most moose-train interactions were in winter when moose appeared to be trapped by deep snow banks along rail beds. moose, elk, and deer all ran along the rail bed primarily inside of the tracks and nearer the rails than track center. collision mortality generally occurred on straight stretches of track. escapes occurred where a discontinuity in the habitat/setting occurred and/or when train speed was reduced. we suggest that videos can provide a valuable resource for interpreting ungulate reactions to trains and that videos gathered purposefully on railways and posted on open source databases will be useful for studying the dynamics of moose-train collisions for mitigation planning. alces vol. 46: 183-187 (2010) key words: alces alces, behavior, collision, train, linear corridor, open source database, railway, tactility, winter mortality. where railroads bisect moose habitat, moose (alces alces) are killed by trains. efforts aimed at reducing moose-train collisions through alteration of railway corridors have proven partially effective (child 1987, muzzi and bissett 1990, child et al. 1991, gundersen and andreassen 1998, andreassen et al. 2005). little attention appears to have been devoted to describing and interpreting the behavioral responses of moose to trains for collision mitigation planning. unfortunately, safety regulations of many rail corporations generally prohibit non-personnel (e.g., wildlife biologists) on board locomotives (wells et al. 1999). consequently, obtaining observational records is often not possible which makes it difficult for biologists to describe behavioral reactions of moose to the approach and chase by trains. in an effort to understand circumstances surrounding moose strikes by trains, we studied video records of ungulate-train interactions, most of which we found posted on youtubetm. the objectives of this study were to 1) describe the behavioral reactions of ungulates chased by trains, 2) identify conditions of the rail bed that influence outcomes of ungulate-train encounters, and 3) make recommendations toward minimizing moosetrain collisions based on our findings. methods we viewed 21 video records of ungulatetrain interactions downloaded from www. youtube.com, or received from rail personnel. we categorized interactions by species, time of day, season, group size, sex-age class, and habitat type; we further noted whether the animal(s) survived the encounter, speed of the train, and presence and condition of snow. when possible, we classified moose based on presence of vulval patch, antlers, and size of moose-train interactions rea et al. alces vol. 46, 2010 184 dewlap. we recorded whether animals were walking, trotting, or running, the frequency of attempt to exit the rail bed, and time under chase for each animal. final outcomes of each encounter (strike or escape) were recorded relative to train speed (slow ~1-10 km/h; moderate ~10-20 km/h; fast >20 km/h), recording time, and alignment of the track. we also recorded the position of the animal relative to the rails and developed a reference scale bar of 36 cm by dividing the fixed distance between the steel rails (143.5 cm) into 4 equal segments. this scale was used to approximate the position of an animal inside and outside of the rails, lateral to track center. observations were categorized as all ungulate-train interactions (including moose) and as solely moose-train interactions. we viewed each video record 3-5 times detailing interactions between an animal and its conspecifics, speed of train, snow conditions within and along the rail bed, and if possible, action of the train crew to avoid collision. we made several assumptions: 1) all videos were recorded with hand-held recorders (we saw no evidence for mounted, continuously recording cameras) started upon detection of the event, 2) video footage was likely to be recorded equally in all seasons, but 3) more likely to be recorded in daylight, and 4) video footage was just as likely to capture escapes as strikes. we saw no bias towards recording or posting strikes versus escapes; only 7 of 56 animals recorded were struck. results about half (48%) of the video recordings were of moose, and 70% of all moose videos were filmed in winter (fig. 1). most videos of other ungulates were filmed in summer; only 2 encounters (a deer [odocoileus spp.] and a moose cow-calf pair) were recorded at night, both during summer. the total number of animals observed was 56 (22 domestic cows [bos taurus], 16 moose, 12 elk [cervus elaphus], 2 deer, 2 camels [camelus dromedarius], and 2 domestic sheep [ovis aries]). average group size was 2.78 ± 3.46sd for all ungulates and 1.6 ± 0.97sd for moose. most animals were adults: 38 of 56 ungulates observed and 12 of 16 moose. although determining sex-age class of moose was difficult in most recordings, we identified 4 cows, 3 bulls, and a cow-calf pair. moose, deer, and elk were filmed consistently in forested habitats with some footage showing adjacent features such as clear-cuts, rock face cliffs, roads, and fields; camels were filmed in desert environments and cattle and sheep were generally filmed in pasture or chaparral settings, except for a group of free-range cattle filmed in a forested habitat. five animals were killed in the footage (1 moose, 2 cows, 1 camel, and 1 deer). of those killed, 1 was standing (cow), 2 were running (adult moose and domestic calf), 1 was trotting (camel), and 1 was walking (deer). in all cases, the speed of the train was moderate or fast (i.e., at a speed that the animal(s) could not sustain during flight), with brevity of recording being a predictor of strike probability (fig. 2); animals escaped when train speed was accommodating. all collisions occurred during the day except for a single buck deer, and all animals were killed on straight stretches of track. in addition, carcasses of a cow and calf moose were filmed immediately adjacent to the rails on a straight stretch of track. of the winter footage that we examined spring summer fall winter season 0 20 40 60 80 100 p er ce nt ag e of a ni m al t yp e (% ) all ungulates moose fig 1. percentage of video records of ungulate-train and moose-train interactions by season. alces vol. 46, 2010 rea et al. moose-train interactions 185 (moose, elk, and deer), animals trotting and running down the tracks ahead of the locomotive attempted to exit the rail bed 1.5 ± 1.7sd times on average, then returned to the tracks where deep snow prevented escape. also, from our inspections, neither the distance between the train and animal or the duration of the chase seemed to affect how often animals attempted to exit the rail bed. while under chase, animals spent most of their time running between the rails (fig. 3). escape by individual moose occurred mostly where a discontinuity in the habitat/ setting adjacent to the railbed was encountered (e.g., a creek bed, bridge, road crossing, stationary equipment at a rail siding), although one escape occurred at a seemingly random spot along an embankment. on the other hand, when a group of animals encountered an approaching train, the action of one or more members of the group often facilitated a successful escape. we did notice, however, that this social facilitation between members of a group could possibly increase collision risk. for example, in 2 separate encounters, a domestic calf was struck when attempting to reunite with the cow, and 2 members of a group of 11 elk were nearly struck when closely following the herd across the tracks. discussion most of the video records we observed were encounters between trains and moose. woods and munro (1996) and wells et al. (1999) indicated that the largest portion of ungulate mortality on railroad tracks involves elk and moose in northwestern north america. their mortality appears to be related to occupation of winter habitat in valley bottoms where railways are common (heershap 1982, child et al. 1991) and biodiversity and winter range values are highest (woods and munro1996), consequently increasing chance of animaltrain encounters (bertwistle 2001). all moose observed on video were in forested areas with a mosaic of streams and plantations. heerschap (1982) reported that most moose-train encounters in ontario occurred in forest habitat, where forest edges along the rail corridor can act as an ecotonal trap for ungulates attracted to edges where browse and mature tree cover co-occurs (child et al. 1991). although most ungulate-train interactions were recorded in summer, most moose-train interactions were recorded in winter, as also reported in alaska (rausch 1959, becker and grauvogel 1991, modafferi 1991), british columbia (child et al. 1991, wells et al. 1999), and norway (andersen et al. 1991, gundersen and andreassen 1998). based upon our study of video footage and file photographs, discussions with train crews, and by our field inspections (rea, child, struck escaped outcome of chase 0 20 40 60 80 100 120 r ec or de d t im e u nd er c ha se ( se c) fig. 2. the relationship between recording time and the outcome in recorded ungulate-train interactions; 5 animals were struck. 0 36 72 108 144 180 216 252 288 324 360+ lateral distance from track center (cm) 0 10 20 30 40 50 t im e u nd er c ha se ( se c) fig. 3. the average (± 1 se) number of seconds spent by ungulates fleeing trains relative to distance from track center. the distance between steel rails is 143.5 cm; track center is 0 cm and the position of the steel rails is approximately 72 cm either side of center. moose-train interactions rea et al. alces vol. 46, 2010 186 and aitken, unpublished data), animals under prolonged chase travel mostly between the steel rails, sometimes running just outside or directly on the rails. snow conditions closer to the rails are usually more shallow and denser than at track center (rea, child, and aitken, unpublished data) and likely provide the best footing and most energy-efficient locomotion (geist 1999). these trails of compacted snow (approximately 30 cm on both sides of the steel rails) are created by the continued disturbance of snow by the combined effects of the snowplow, drag of the ballast regulator, and the dual wheels of the service trucks (b. easton, managing engineer, canadian national railway, cn north, personal communication). such trails represent the path of least resistance for moose (geist 1999), but are also the portion of the corridor where the risk of strike is highest (and is often referred to as "the kill zone" by railway personnel) and the chance of escape lowest. speed of the train has been implicated in ungulate-train collisions (espmark 1966, gundersen and andreassen 1998). bubenik (1998) suggested that collisions are inevitable because moose do not readily conceptualize moving objects. although moose can reportedly trot at speeds >60 km/h over a distance of 500 m (geist 1999), this speed and distance might well be above what a running moose can sustain on a snow-covered rail bed. in alaska, train speed was reduced from 79 to 40 km/h to mitigate moose strikes (becker and grauvogel 1991), but collision risk to moose was not reduced as expected. anecdotal reports from train crews suggest that escape is quite common when the speed of the train is accommodating. we observed in one video that when crews reduced train speed to 10-15 km/h (voice recorded in video), moose had sufficient time to exit and avoid collision. in another, we observed a cow-calf moose pair escape by negotiating a high snow bank when the speed of the train was reduced (actual speed unknown). conclusions and recommendations our findings suggest that collisions will continue to be difficult to mitigate because reducing train speed is not always possible, natural escape routes in deep snow are few, and moose preferentially run along snowpacked railbeds where mobility is easiest and less energetically costly during winter. however, collecting more and better information about ungulate-train interactions could help to improve our understanding of the ecology of ungulate-train interactions and assist in the development of strategies to reduce collisions. consequently, we recommend the following actions: 1) expand video recordings of animal reactions to trains, 2) continue to expand, integrate, and standardize data collection, and 3) reduce train speed in known collision hotspots when strikes are most likely. permanently mounted and continuously running cameras on locomotives would provide increased and more informative documentation of ungulate-train interactions; these records should be made available to researchers in order to help mitigate moose-train interactions. standardized data collection is essential to better document encounters, location, timing, train speed, and environmental conditions at the time of the event. the use of data loggers with gps capability would help identify locations where animal encounters are recurrent. reducing speed in identified areas may reduce the risk of strikes because escape is related to slower speed. by undertaking these actions, rail corporations would help mitigate collisions with wildlife, improve operations, and avoid the likelihood of costly derailments as reported in norway (h. korslund, senior information advisor, norwegian national rail administration, personal communication). acknowledgements we thank those who have kindly provided video clips to us and by doing so have provided a rich data source for us to analyze. we thank alces vol. 46, 2010 rea et al. moose-train interactions 187 our research assistant anna dehoop for obtaining and compiling all the video clips for us to view. we thank various railway personnel who answered questions and provided us with valuable insights. we also thank pete pekins and an anonymous reviewer for a thorough review and much work on an earlier draft of the manuscript. references andersen, r., b. wiseth, p. h. pedersen, and v. jaren. 1991. moose-train collisions: effects of environmental conditions. alces 27: 79-84. andreassen, h. p., h. gundersen, and t. storaas. 2005. the effect of scentmarking, forest clearing, and supplemental feeding on moose-train collisions. journal of wildlife management 69: 1125-1132. becker, e. f., and c. a. grouvogel. 1991. relationship of reduced train speed on moose-train collisions in alaska. alces 27: 161-168. bertwistle, j. 2001. description and analysis of vehicle and train collisions with wildlife in jasper national park, alberta, canada, 1951-1999. pages 433-434 in c. l. irwin, p. garrett, and k. p. mcdermott, editors. proceedings of the 2001 international conference on ecology and transportation, center for transportation and the environment, north carolina state university, raleigh, north carolina, usa. bubenik, a. b. 1998. behavior. pages 173-221 in a. w. fransmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. child, k. n. 1987. railways and moose in the central interior of british columbia: a recurrent management problem. alces 19: 118-135. _____, s. p. barry and d. a. aitken. 1991. moose mortality on highways and railways in british columbia. alces 27: 41-49. espmark, y. 1966. railway mortality of reindeer in sweden. särtryck ur zoologisk revy 1: 33. gundersen, h., and h. p. andreassen. 1998. the risk of moose alces alces collision: a predictive logistic model for moose-train accidents. wildlife biology 4: 103-110. geist, v. 1999. moose: behavior, ecology, conservation. voyager press, maine, minnesota, usa. heerschap, a. 1982. big game mortality by trains: cartier to white river, june 1981-june 1982. unpublished report of the chapleau district, ontario conservation officer service, chapleau, ontario, canada. modaferri, r. d. 1991. train-moose kill in alaska: characteristics and relationship with snowpack depth and moose distribution in lower susitna valley. alces 27: 193-207. muzzi, p. d., and a. r. bisset. 1990. the effectiveness of ultrasonic wildlife warning devices to reduce moose fatalities along railway corridors. alces 26: 37-43. rausch, r. a. 1959. the problem of railroadmoose conflicts in the susitna valley, 1955-56. u.s. wildlife service, anchorage, alaska, usa. woods, j. g., and r. h. munro. 1996. roads, rails and the environment: wildlife at the intersection in canada’s western mountains. transportation related wildlife mortality seminar. orlando, florida, usa. wells, p., j. g. woods, g. bridgewater, and h. morrison. 1999. wildlife mortalities on railways: monitoring methods and mitigation strategies. unpublished report. parks canada, p.o. box 350, revelstoke, british columbia, canada. alces34(2)_423.pdf alces 31_27.pdf alces34(2)_311.pdf alces34(1)_83.pdf alces30_25.pdf effects o f black bear predation on caribou--a review warren b. ballard new brunswick cooperative wildlife research unit, faculty of forestry, p. 0. box 44555, university of new brunswick, fredericton e3b 6c2 abstract: i reviewed available literature concerning black bear (ursus americanus) predation on caribou (rangifer tarandus) in an effort to gain insight on the possible impacts of black bear predation on a potential re-introduction of woodland caribou (rangifer tarandus caribou) to north-central minnesota. several case histories were reviewed and inferences were drawn from several black bear moose (alces alces) studies. i concluded that black bear predation on woodland caribou i n the proposed re-introduction area would likely be a secondary source of caribou mortality and that between 6-30% of the calves and 0-5% adults might be killed annually by black bears. alces vol. 30 ( 1 994) pp.25-35 predation has been identified as a major limiting factor in many populations of ungu lates. the majority of these losses generally occur among juvenile age classes. because predation can limit ungulate populations, it can potentially impact re-introductions of ungulate species into areas from which they have been extirpated. the north central caribou corporation is in the process of com pleting plans for a potential re-introduction of woodland caribou (rangifer tarandus cari bou) into north-central minnesota. predation, particularly by wolves (canis lupus) (bergerud and elliot 1986), has been identified as a potential factor which could result in the fail ure of this re-introduction. however, other predator species also occur in the area. black bears (ursus americanus) have re cently been identified as important predators of juvenile ungulates in several areas of north america (schlegel 1976, franzmann et al. 1980, adams et al. 1988). black bears are moderately abundant within the proposed re introduction area (i.e., 159 to 244/1,000 km2 [rogers 1987]), and predation by them could be a significant factor in the success or failure of the re-introduction. because of the above concerns the north central caribou corpora tion desired additional information concern ing the potential of black bears to impact the woodland caribou re-introduction to minne sota. the purpose of this report is to review information concerning black bear predation on caribou in general, and on woodland cari bou in particular. methods i reviewed literature concerning black bear predation on ungulates in north america and relied heavily on the review papers by truett et al. (1989) and ballard (1992), and on the bibliography on rangifer tarandus com piled by kreeger and fleming ( 1 991). i contacted a number of individuals seeking unpublished information on the subject; how ever, this report is not an exhaustive review of black bear predation on caribou. i may have omitted some studies concerning woodland caribou; however, i believe the information presented is representative of the state of knowledge concerning woodland caribou black bear relationships. case history studies most studies on bear-caribou relation ships concern barren ground caribou and griz zly bears (ursus arctos) in tundra ecosystems (page 1976, garner and reynolds 1986, bergerud and page 1987, adams et ul. 1988, whitten et al. 1992). in all cases, predation by wolves and grizzly bears was found to be a significant cause of calf caribou mortality. black bear predation on caribou ballard alces vol. 30 ( 1 994) although these studies did not involve black death were found for 11 cases; coyotes, black bears, some inferences can be drawn concernbears, and golden eagles were thought to be ing the timing of mortality and the potential responsible for 11, 3, and 1 mortalities, re significance of black bear predation on woodspectively. yearling and adult female caribou land caribou. also, there have been several experienced high annual survival between studies which have concerned black bear pre1987 and 1992 (>90%) and no cases of preda dation on woodland caribou. tion were observed during 11,915 caribou days of monitoring. black bear feeding sta newfoundland tions were in operation during 1989 and crcte mahoney et al. (1 990) provides the larg et al. (1 99 1) concluded that the stations could est data set concerning the impacts of black have reduced bear predation rates. bears on woodland caribou. between 1979 and 1984 they determined causes of mortality of 220 radio-collared newborn caribou calves within three woodland caribou populations in newfoundland. twenty-three percent of the calves died during their first year of life with predation accounting for 78% of the deaths. predation by lynx (lynx canadensis) and black bears was each responsible for 35% of the mortalities. of 52 mortalities, 23 (44%) oc curred within 2 weeks and 62% occurred within4 weeks of birth (mahoney etal. 1990). there were no differences in vulnerability of calves by sex or weight. wolf predation was not a factor because they no longer occur in newfoundland (mahoney et al. 1990). no estimates of black bear density were provided and the three caribou populations totaled 17,713 + 15% individuals (mahoney et al. 1990). the observed total mortality rate of 23% for calves aged 1 1 2 months was not sufficient to limit population growth during the study period. grand-jardins park, quebec efforts have been underway to restore caribou to grand-jardins park since 1969 when 41 fawn and yearling caribou were released into the park (h. jolicoeur, unpubl. data). however, only 12 of the animals sur vived or stayed in the area. no caribou were released in 1970 but an additional 25 fawns and yearlings were released in 1971. the latter animals remained in the area apparently because they formed an association with the 12 surviving adult animals from the 1970 release. an additional 12 adults were released in 1972. this herd has increased at an average rate of 5% since 1970 (cantin 1991) and currently numbers 125 animals (h. jolicoeur, unpubl. data). wolf densities within this area were re duced during 1979 and 1980 and are currently 10 wolves per 1000 km2 (h. jolicoeur, unpubl. data). black bear density was estimated at 220 per 1000 km2. the meningeal worm [parelaphostrongylus tenuis) was not present gaspesie park, quebec in this area (h. jolicoeur, pers. commun.). crete and desrosiers reported on crete (pers. commun.) suggested that the ex the status of the gaspesie park caribou herd istence of escape habitats (open landscapes) which was declining due to low recruitment. such as alpine tundra or large bogs could play twelve of 13 calves born to radio-collared a key role in allowing introduced caribou to caribou during 1988 were lost during the increase where bears and wolves exist. summer (majority in july); predation by black ~ ~ bears a n d o r coyotes (canis latrans) was sussoutheastern british columbia pected as the cause of these deaths. newborn seip (1992) reported on the status of two calves were also radio-tagged in 1989 and woodland caribou herds in southeastern brit 1990 with 16 of 25 calves dying during the ish columbia and sought to identify limiting first summer of life. indices on the cause of factors. woodland caribou populations de alces vol. 30 (1994) ballard black bear predation on caribou clined or disappeared from this area during avoid predators by calving or living on islands the 1900's. over-hunting was blamed in (d. r. seip, b. c. ministry of forests, burnaby, many cases, but other populations without pers. commun.). hunting exhibited similar declines. in a non migratory caribou population the distribution of wolves, grizzly and black bears, and cari bou overlapped while in a migratory popula tion wolf and moose (alcesalces) distribution did not overlap that of caribou during sum mer. no mention was made of bear distribu tions but i assume they probably overlapped. within the non-migratory quesnel lake cari bou population wolves and bears (both griz zly and black bears) accounted for 55 and 15%, respectively, of the adult mortality. about half of the calves died of unknown causes during the calving period. seip (1992) suggested that calf survival was related to wolf abundance during summer and that wolf predation was driving the caribou population towards extinction. within the migratory wells gray park caribou population, seip (1992) estimated that the adult caribou mortality rate was 8% with most of the mortality being attributed to bear predation (apparently both species). ap proximately 40% of the adult female caribou had surviving calves and the population was stable or slowly increasing. seip (1992) con cluded that the differences between the two populations were due to differences in wolf predation and overlaps in distribution of alter nate prey (i.e., moose) and wolves. seip (1 99 1) concluded that forest-dwell ing caribou in ontario, saskatchewan, al berta, and the non-mountainous regions of british columbia had declined or were elimi nated due to wolf predation and human har vest. increased wolf densities following the range expansion of ungulates such as moose, elk and deer has resulted in greater predation pressure on resident caribou. caribou persist in large numbers in northern or high elevation areas where predation and hunting pressures are reduced. in southern areas they continue to persist in low numbers where they can baxter state park, maine during 1989, 12 woodland caribou were released in baxter s t a t e park, maine (mccollough and connery 1990). within 6 months of the release 10 of the 12 died; at least five died or "were predisposed" to meningeal worm, one died from abomasal ulcers, one from either a fatal accident or black bear predation, and 3 died apparently from black bear predation. black bear density was estimated at about 232/1,000 krn2, coyote density at 183/1,000 km2, and no wolves were present. the authors pointed out that two major factors (i.e., disease and predation) suggested by bergerud and mercer (1989) as being responsible for both the decline of cari bou in north america and failure of recent re introductions were also responsible for the re introduction failure in maine. effects of bear predation on caribou prior to the use of radio telemetry on neonates, bears were largely thought of as scavengers of ungulates (jonkel 1978). since schlegel's (1976) study of elk (cervus elaphus) calves, black bears have been iden tified as significant predators of neonatal elk (schlegel 1976), moose (franzmann et al. 1980, ballard 1992), deer (odocoileus sp.) (wilton 1983, conger and giusti 1992), and caribou (mahoney et al. 1990). wilton (1983) summarized observations of black bears prey ing on ungulates in north america and con cluded that they should be considered effec tive predators of ungulates throughout their range. black bears are not considered effec tive predators on adult moose (ballard 1992), but mercer (1986, cited in mccollough and connery 1990) and seip (1991) suggest that black bears may account for the majority of adult caribou mortality. also, v. crichton (manitoba dep. of nat. resources, winni black bear predation on caribou ballard alces vol. 30 (1994) peg, pers. commun.) reports an increasing incidence of anecdotal reports of black bears preying on adult moose in manitoba and on tario (austin et al. 1994). in other areas scattered anecdotal accounts of black bears preying on adult caribou exist, but black bears are not considered significant predators of adult caribou (v. crichton, pers. commun; l. g. adams, u . s. national park service, an chorage, alas., pers. commun.). regardless, black bear predation is a significant source of mortality to neonates in many ungulate populations. predation has been reported in many stud ies to be the most significant mortality factor affecting caribou populations (miller and broughton 1974, bergerud 1980). most mor tality of caribou neonates due to predation occurs during the first month of life (bergerud 1971, 1980; miller and broughton 1974; jakimchuck 1979; miller 1987; adams et al. 1988, mahoney et al. 1990, whitten et al. 1992; ). generally the most abundant preda tor species constitutes the largest source of mortality (truett et al. 1989). however, this is not always true, particularly with regard to black bear predation. ballard et al. (1990) examined the causes of mortality to neonate moose in southcentral alaska where black bears outnumbered grizzly bears and wolves by factors of 3.2 and 32, respectively. grizzly bears killed 52% of the calves followed by 9% for black bears and 7% by wolves. thus occurrence of predator species, their relative predation efficiency and density, and prey density probably affect which predator spe cies are the most significant cause of mortal ity. neonatal moose are most vulnerable to bear predation during their first 6 weeks of life (ballard et al. 198 1). subsequently their increased mobility appears to make them in creasingly less vulnerable to bear predation (ballard etal. 1980). newborn caribou calves also suffer their highest mortality during the first 1-2 months of life and become less vul nerable to at least bear predation with increas ing age and mobility (truett et al. 1989, adams et al. 1993). predation by black bears has accounted for the deaths of from 2 to 50% of radio collared moose calves in various areas of north america (table 1). ballard (1992) concluded that black bears were a significant source of mortality to moose calves when they out number grizzly bears and wolves by factors of 10 and 30, respectively, or their densities were >20011,000 km2. although only based on two studies from the kenai peninsula, alaska, ballard (1992) also con cluded that black bear kill rates (functional response) appeared to be dependent upon the densities of moose calves while the percent of the moose population killed (functional and numerical response) was not dependent on moose densities. based upon the aforemen tioned studies the importance of black bear predation as an ungulate limiting factor is dependent upon the density of black bears in relation to the number and density of other predator and prey species. it is not known whether black bear predation is a learned phenomena, or whether it has occurred all along. it may not have been detected earlier because many food habits studies have relied on scat analyses which would underestimate the importance of ungulates (ballard and larsen 1987). potential effects of black bears on woodland caribou re-introductions in the boundary waters canoe area wilderness bergerud and mercer (1989) have sug gested that even in the absence of deer (the source for p. tenuis) when wolf densities exceed 1011,000 km2, caribou re-introduc tions will fail. bergerud and elliot (1986) indicated that in general, caribou populations can not maintain their numbers when wolf densities are26.5/1,000 km2 in the absence of alces vol. 30 (1994) ballard black bear predation on caribou escape habitat. the increasing caribou popu lation at grand-jardins park, qucbec appear to fit their prediction in that the caribou herd has been increasing at 5% annually (cantin 1991), wolf and black bear densities have been estimated at 10 and 22011,000 km2, respectively (h. jolicoeur, unpubl. data), and adequate escape cover exists (m. crete, pers. commun.). wolf densities within the pro posed minnesota caribou reintroduction area have been estimated at 16 to 2011,000 km2 (nelson and mech 1992). nelson and mech (1 992) acknowledged that their reported wolf densities were well above the threshold re ported by bergerud and elliot (1986) and bergerud and mercer (1989), but that caribou coexisted in other areas with higher wolf densities: spatsizi provincial park, british columbia (16 wolves/1,000 km2), lake ni pigon, ontario (1 014 wolves/1,000 km2), and pukaskwa park, ontario (14 wolves/1,000 km2). none of the aforementioned authors mentioned bear predation. bergerud et al. (1983) mentioned that the caribou population at pukaskwa park num bered 20-30 animals and that wolves seldom preyed upon them. black bears also occur in the area, but their densities are unknown (g. eason, ontario ministry of natural resources, wawa, pers. commun.). the caribou in this area have apparently declined, and only a few remained in the area by 1992 (g. eason, pers. commun.). causes for this apparent reintro duction failure are unknown. bergerud etal. (1 990) suggested that cari bou successfully calved on islands to escape predation by wolves. nelson and mech (1992) also suggested that islands at lake nipigon, ontario provided escape habitat for caribou from wolf predation. v. crichton (pers. comm.) indicates that woodland caribou on the east side of lake winnipeg, manitoba traditionally calve on islands in large lakes and on occasion within bogs. these herds are composed of 40-100 individuals. prior to 1978 few black bears were observed on the islands, but since 1978 numerous black bears have been observed, and the caribou popula tion has apparently remained stable. it is unknown whether black bears are preying on caribou calves, but the recent occurrence of bears on the islands suggests that they may be preying on caribou calves. hair from adult caribou has been found in bear feces during july, but it is not known whether these obser vations represent carrion feeding or actual predation. caribou persist at low densities (about 0.02/km2) where wolf density is high (161 1,000 km2) at spatsizi provincial park, british columbia (a. t. bergerud, univ. of victoria, british columbia, pers. comm.). both black and grizzly bears occur in the area, but their importance as predators has not been studied. however, if they prey on caribou the magni tude of such predation would be secondary to that by wolves. pitt and jordan (1992) examined the use of islands by black bears within the boundary waters canoe area by examining usage of bait stations at islands with and without hu man camp sites. they found that black bears appeared to be more abundant on islands with camp sites than those without camp sites. although their study results were limited be cause they could only place bait stations within 1 km of camp sites (this limited the size of the island which could be studied--w. pitt, utah state univ., logan, pers. commun.) their re sults have a direct bearing on bear usage of island sites. islands with camp sites appeared to be occupied less frequently by wolves but more frequently by bears (w. pitt, pers. commun.). apparently bears were attracted to the islands because of the presence of food associated with humans. because woodland caribou use islands as escape habitat during calving, the attraction of bears to island sites occupied by campers will increase the fre quency of bear-caribou interactions. this may result in increased bear predation be cause of the presence of human camp sites. table 1.causes of mortality and survival rates of radio collared moose calves to november in relation to observed kill rates and predator densities in north america (modified from ballard 1992). h south central alaska kenai peninsula, ak + cr areas area 1-3 area 1 area 4 areas 1947 bum 1969 bum southwest eastcentral saskatnew new f2 '3 pooled yukon* alaska chewan bmnswick foundland v, years 1977,1978 1979 1984 1977-84 1977,1978 1981,1982 1983,1985 1984 1982 1983,1985 1983-88 causes of mortality (%) grizzly bear 4 1.9 42.9 52.2 44.0 6.4 2.7 41.9 51.5 black bear w 0 grizzly & black wolf 1.6 6.5 2.5 6.4 1.4 17.9 15.2 unknown predation 2.4 unknown causes 3.2 surviving (%) 46.0 42.9 17.4 39.4 44.6 48.5 18.8 18.2 50.0 8 1.8 70.0 density (no./1000km2) grizzly bear 24 10 28 24-28 12-28 12-28 16 16 0 0 0 black bear 0 0 90 0-90 205 258 1 6w 8-11 200-400 mod.? 570 alces vol. 30 (1994) s i 4 4 3 3 'j? 3 09 2 '-? n 'j? 3 09 3 2 5 ballard black bear predation on caribou black bear predation on caribou ballard alces vol. 30 ( 1 994) however, if wolves are more abundant on islands not occupied by humans, then caribou may be subjected to high levels of wolf preda tion on islands without human camp sites. using the wolf densities reported by nel son and mech (1992), it appears that the area considered for caribou reintroduction exceeds the threshold level of wolf densities that bergerud and mercer (1989) indicated were necessary for caribou to survive. assuming that black bear densities are similar to those reported by rogers (1987) for the superior national forest (1 59 to 244 bears/1,000 km2), bear densities approach the level suggested by ballard (i 992) for black bears to be a signifi cant source of mortality. based upon both the reported bear, wolf, and deer (i.e., <386/1,000 km2, pitt and jordan 1991) densities within the proposed reintroduction area and reviewed literature, black bears might kill between 16 to 30% of the calves produced by adult wood land caribou, and may kill 0-5% of the adults annually. the accuracy of these estimates would depend upon the magnitude of wolf predation. in any case, the evidence suggests that in the presence of wolves, black bear predation will be a secondary source of mor tality which could be additive to other sources of mortality. the success or failure of the caribou re introduction may depend on the timing and numbers of caribou released during the initial reintroduction. if black bear predation is, in part, a learned phenomenon then it would appear that introduction of large numbers of caribou at the initial stages of the project may allow caribou to establish a foot-hold before bears and wolves learn of their presence. it may also be advisable to establish diversion ary feeding stations as was done by crete et al. (1991) and boertje et al. (1992) which may reduce black bear kill rates. at low caribou numbers the increased survival of just a few individuals can make the difference between success and failure of a caribou transplant. although my analysis is based upon reviewed literature, i point out that the interactions between caribou and black bears, particularly woodland caribou, are poorly understood as suggested by the paucity of studies cited in this review. acknowledgements this review was funded by the north central caribou corporation. i thank a. t. bergerud, a. f. cunning, v. crichton, g. j. forbes, s. mahoney, w. c. pitt, d. seip, h. r. timmermann, and the north central caribou corporation for critically reviewing early drafts of this manuscript. h. jolicoeur pro vided unpublished data on the caribou popu lation in grands-jardins park, qucbec. j. l. nelson was most helpful in providing back ground information. l. swanson translated a french article to english. m. cr2te and one anonymous reviewer provided helpful com ments on the final draft. references adams, l. g., b. w. dale, and l. d. mech. 1993. wolf predation on caribou calves in denali national park, alaska. proc. 2nd internat. wolf symp. in press. edmonton, alberta. , and f. j. singer. 1988. neonatal mortality in the denali caribou herd. pages 33-34 in r. d. c a m e r o n and j. l. davis, eds. proc. 3rd n. am. caribou workshop. alaska dep. fish and game wildl. tech. bull. no. 8. juneau. austin, m. a., m. e. obbard, and g. b. kolenosky. 1994. evidence for a black bear, ursus urnericanus, killing an adult moose,alcesalces. can. field-nat. 108:in press. ballard, w. b. 1992. bear predation on moose: a review of recent north ameri can studies and their management impli cations. alces suppl. 1 : 162176. , c. l. gardner, and s . d. miller. 1980. influence of predators on summer movements of moose in alces vol. 30 (1994) ballard black bear predation on caribou southcentral alaska. proc. n. am. moose conf. workshop. 16:339-359. , and d. g. larsen. 1987. impli cations of predator-prey relationships to moose management.'swedish ~ i l d l . res. suppl. 1 :581-602. , and s. d. miller. 1990. effects of reducing brown bear density on moose calf survival in southcentral alaska. al ces 26:9-13. , and j. s. whitman. 1990. brown and black bear predation on moose in southcentral alaska. alces 26: 1 8. , t. h. spraker, and d k. p. taylor. 1981. causes of neonatal moose calf mortality in south-central alaska. j. wildl. manage. 45:335-342. , j. s. whitman, and d. j. reed. 1991. population dynamics of moose in south-central alaska. wildl. monogr. 114. 49pp. beaulieu, r. 1984. moose calf mortality study. saskatchewan parks and renew. resour., wildl. pop. manage informa tion base, 84-wpm-8. saskatoon. 5pp. bergerud, a. t. 197 1. the population dynamics of newfoundland caribou. wildl. monogr. 25. 55pp. . 1980. a review of the population . dynamics of caribou and wild reindeer in north america. pages 556-581 in e. reimers, e. gaare, and s. skjenneberg, eds. proc. 2nd internat. reindeer/cari bou symp., roros, norway, 1979. direktoratet for vilt og ferskvannsfisk, trondheim. , and j. p. elliot. 1986. dynamics of caribou and wolves in northern british columbia. can. j. zool. 64: 15 151529. , r. ferguson, and h. e. but ler. 1990. spring migration and disper sion of woodland caribou at calving. ani mal behaviour. 39:360-368. , and w. e. mercer. 1989. cari bou introductions in eastern north america. wildl. soc. bull. 17: 11 1120. , and r. e. page. 1987. displace ment and dispersion of parturient caribou at calving as anti-predator tactics. can. j. 2001. 65: 1597-1606. , w. wyett, andb. snider. 1983. the role of wolf predation in limiting a moose population. j. wildl. manage. 47:977-988. boer, a. h. 1988. moose, alces alces, calf mortality in new brunswick. can. field nat. 102:74-75. boertje, r. d., w. c. gasaway, s. d. dubois, d. g. kellyhouse, d. v. grangaard, d. j. preston, and r. 0 . stephenson. 1985. factors limit ing moose population growth in game management unit 20e. alaska dep. fish and game. alaska dep. fish and game fed. aid in wildl. restoration prog. rep. w-22-3 and w-22-4. juneau. 5 lpp. , d. v. grangaard, and d. g. kellyhouse. 1988. predation on moose and caribou by radio-collared grizzly bears in eastcentral alaska. can. j. zool. 66:2492-2499. , d . v. g r a n g a a r d , p . valkenburg, and s. d. dubois. 1992. testing socially acceptable meth ods of managing predation: reducing pre dation on caribou and moose neonates by diversionary feeding of predators, macomb plateau, 1990-94. alaska dep. fish and game, fed. aid in wildl. resto ration prog. rep., proj. w-23-5. 28pp. cantin, m. 1991. tendances demo graphiques de la population de caribous, rangifer tarandus, des grands-jardins. qucbec, ministkre du loisir, de la chasse et de la peche, direction rkgionale de qucbec, service de l'amenagement et de l'exploitation de la faune. 26p. conger, s. l., and g. a. giusti. 1992. observations of black bear (ursus americanus) predation on columbian black-tailed deer (odocoileus hemionus black bear predation on caribou ballard alces vol. 30 ( 1 994) columbicanus). calif. fish and game 78:131-132. c r ~ t e , m., and a. desrosiers. 1994. range expansion of coyotes threatens a remnant herd of caribou in southeastern quebec. arctic: submitted. , c. banville, d. le henaff, j. levesque, and h. ross. 1991. high calf mortality endangers the gaspesie park caribou herd. pages 178-179 in c. e. butler and s. p. mahoney, eds. proc. 4th n. am. caribou workshop. st. john's, newfoundland. f r a n z m a n n , a. w . , and c . c. schwartz. 1986. black bear preda tion on moose calves in highly productive versus marginal moose habitats on the kenai peninsula, alaska. alces 22: 139 153. , and r. 0 . peterson. 1980. causes of summer moose calf mortality on the kenai peninsula. j. wildl. manage. 44:764-768. garner, g. w., and p. e. reynolds. 1986. arctic national wildlife refuge coastal plain resource assessment. final report baseline study of the fish, wildlife, and their habitats. vols. i and 11. usdi fish and wildlife service, region 7 . anchorage, alaska. 695pp. jakimchuk, r. d. 1979. an overview of five major herds of barren ground caribou in canada. report to polar gas project by r. d. jakimchuk management associ ates ltd., sidney, british columbia. 84pp. jonkel, c. 1978. black, brown (grizzly), and polar bears. pages 227-248 in j. l. schmidt and d. l. gilbert, eds. big game of north america. stackpole co., hamisburg, pa. kreeger, t. j. and c. fleming. 199 1 . bibliography of rangifer tarandus 1961 present. unpubl. manuscript. t. j. kreeger, cedar creek natural history area, bethel, minnesota. 61pp. larsen, d. g., d. a. gauthier, and r. l. markel. 1989. causes and rate of moose mortality in southwest yukon. j. wildl. manage 53:548-557. mahoney, s. p., h. abbott, l. h. russell, and b. r. porter. 1990. woodland caribou calf mortality in insu lar newfoundland. trans. 19th iugb congress, trondheim, norway. mccollough, m., and b. connery. 1990. an evaluation of the maine cari bou reintroduction project, 1986 to 1989. unpubl. report. maine caribou project, univ. of maine, orono. 53pp. miller, f. l. 1987. management of bar ren-ground caribou (rangifertarandus groenlnndicus) in canada. pages 523 534 in c. m. wemmer, ed. biology and management of the cemidae. smith sonian inst. press, washington, d.c. 577pp. , and e. broughton. 1974. calf mortality on the calving ground of kaminuriak caribou. canadian wildl. serv. rep. ser. no. 26, ottawa. 26pp. nelson, m. e., and l. d. mech. 1992. winter wolf density in the east-central boundary water canoe area in north eastern minnesota, 1989-90 and 1990 9 1. unpubl. manuscript. n. central for. exp. sta., st. paul, minnesota. 13pp. page, r. e. 1976. early calf mortality in northwestern (spatsizi plateau) british columbia. b.sc. thesis, univ. of guelph, ontario. 129pp. pitt, w. c., and p. a. jordan. 1991. a survey for parelaphostrongylus tenuis in the region of a proposed caribou reintro duction site. typewritten report to the north central caribou corporation, du luth, minn. 12pp. , and . 1992. a survey for black bears in the region of a proposed caribou reintroduction. unpublished re port to north central caribou corpora tion, duluth, minn. 17pp. roger, l. l. 1987. effects of food supply alces vol. 30 (1994) ballard black bear predation on caribou and kinship on social behavior, move ments, and population growth of back bears in northeastern minnesota. wildl. monogr. 97. 72pp. schlegel, m. 1976. factors affecting calf elk survival in northcentral idaho--a progress report. proc. 56th ann. conf. w. assoc. state game fish comm. 56:342-355. s c h w a r t z , c . c . , and a . w. franzmann. 199 1. interrelationship of black bears to moose and forest succes sion in the northern coniferous forest. wildl. monogr. 113. 58pp. seip, d. r. 1991. predation and caribou populations. rangifer, spec. issue no. 7:46-52. . 1992. factors limiting woodland caribou populations and their interrela tionships with wolves and moose in south eastern british columbia. can. j . 2001. 70:1494-1503. stewart, r. r., e. h. kowal, r. beaulieu, and t. w. rock. 1985. the impact of black bear removal on moose calf survival in east-central sas katchewan. alces 21 :403-418. truett, j. c., a. t. bergerud, and d. roseneau. 1989. caribou calving ar eas and neonatal mortality: a review. . report to alaska oil and gas associa tion, anchorage, alaska. 148pp. whitten, k. r., g. w. garner, f. j. mauer, and r. b. harris. 1992. pro ductivity and early calf survival in the porcupine caribou herd. j. wildl. man age. 56:20 1-2 12. wilton, m. l. 1983. black bear predation on young cervids--a summary. alces 19:136-147. alces32_15.pdf alces 31_61.pdf alces36_217.pdf alces 31_105.pdf alces vol. 45, 2009 safronov moose in northern yakutia 17 regional populations and migration of moose in northern yakutia, russia valeriy m. safronov institute for biological problems of cryolithozone of siberian division, russian academy of sciences, 41 lenin ave., yakutsk, russia 678891 abstract: following an overall population decline of moose (alces alces) after the 1970s, extensive aerial and ground surveys conducted since 1985 indicated that there were 7 distinct populations in northern yakutia. they are isolated geographically by mountain ridges and major rivers, and are named the leno-olenek, predverkhoyansk, yana, chondon, momo-selenyakh, indigirka, and kolyma populations. although most occupy forest habitat associated with major rivers, some are migratory (40-200 km) moving both n-s and e-w, and certain populations overlap on winter range. population densities generally range from 1-2 moose/10 km2, with higher and lower local densities. the northernmost chondon population is unique by occupying sub-tundra forests and ridges. because protective regulations did not produce measurable population recovery and were abandoned in 2004, management strategies must be adopted to address the ecological differences of these separate populations. effective moose management in yakutia will require further identification of range and habitat use, habitat structure and availability, and population estimates and dynamics of regional populations. alces vol. 45: 17-20 (2009) key words: alces alces, management, migration, moose, population dynamics, yakutia. the moose (alces alces) population in yakutia increased about 15% annually in the 1960s. it was expected to increase fourfold during the subsequent decade, from 60to 240,000 (tavrovsky et al. 1971). however, this projected increase was unrealized as the population was only about 78,000 by the mid-1970s. population growth was curbed by overhunting, predation by wolves (canis lupus), increased disturbance, and extrusion to less favorable winter range. the yakutian moose population did not reach its maximum at the time when the majority of moose populations thrived and reached maximum density in the european part of russia (filonov 1983). in 1997-2000 there was a moratorium on moose hunting in parts of yakutia that was subsequently extended to 5 years along with a geographic extension of protection. due to its futility, the moratorium was canceled in january 2004. management of moose in yakutia should not occur under a single administrative district rather, it needs to recognize and address regional populations characterized by different size, habitat, and reproductive capacity. even initial aerial censuses in the mid1960s showed uneven distribution of moose in yakutia. several areas with relatively high moose density were in northeastern yakutia, including western predverkhoyanie (0.55 animals/10 km2), the yukagirskoe upland (0.5), the kolymskoe and abyiskoe lowlands (0.5), and some areas in the yana basin (0.55). the rest of the northeastern territory and the northwestern area were characterized by low moose density (tavrovsky et al. 1971). however, identifying separate populations was hindered due to short-term aerial censuses and narrow coverage of survey strips. more recent aerial censuses have confirmed that regional populations of variable density occur in yakutia. extensive aerial (250 hr and 37,500 km) and ground surveys undertaken in 1985-2001 moose in northern yakutia safronov alces vol. 45, 2009 18 identified 7 relatively isolated moose populations. the leno-olenek population is the largest in northwestern yakutia at 1.8-2 moose/10 km2 at its core, and inhabits the basins of the siligir, muna, severnaya, motorchuna, molodo, and olenek rivers (fig. 1); the average population density was 0.5/10 km2. the winter range in the north includes the nekekit river, the lower course of the merchimden river, and upstream of the molodo river. in the south, the winter range borders upstream of the tyung and linde rivers, and adjoins the lena river to the east. the leno-olenek population is probably in transient contact with the predverkhoyanskaya population to the south (fig. 1). however, the frequency of contact is probably low due to their different migration routes. the migration route of the leno-olenek population is oriented north-south, whereas the predverkhoyansk population moves east-west. the leno-olenek and leno-viluy populations overlapped in the 1960s at the watershed of the viluy and linde rivers where their density reached 1.2 moose/10 km2. subsequent development of gas fields in this area caused the southern range of the leno-olenek population to shift northward, and the northern range of the leno-viluy population southward; consequently, contact between these populations has become rare. prior to our surveys, it was believed that yakutian moose were not migratory (belyk 1948). however, the leno-olenek population migrates annually through the lower course of the muna river in september-october and returns north in may; the migration is120-200 km long (yazan 1972). the basic winter range includes the basins of the siligir, muna, severnaya, and motorchuna rivers that provide plentiful winter forage. the average snow depth is 40-50 cm which is less than in the north (60-70 cm). the spring migration is probably associated with increasing energy fig. 1. the location of 7 distinct moose populations found in northern yakutia, russia; the 7 populations were the leno-olenek (1), the predverkhoyansk (2), the yana (3), the chondon (4), the momo-selenyakh (5), the indigirka (6), and the kolyma (7). arrows depict the general direction of seasonal migrations. alces vol. 45, 2009 safronov moose in northern yakutia 19 demands and the timing of availability of grass that increases northward in this part of yakutia. the western predverkhoyanie region has a moose population occupying the prilenskoaldanskaya plain and verkhoyanie foothills, separated by the verkhoyanskij ridge to the east (fig. 1). the average population density in the northern prilenskaya area was 1.6 animals/10 km2 ranging locally from 0.53.3, and in floodplains was as high as 5.2-6.8 animals/10 km2. in september-early october moose migrate to mountain taiga forests from the lena river floodplains and its tributaries; in november-december they return to the floodplains. the direct-line migration distance is no less than 40 km. the yana population found in the yana river basin is separated west of the predverkhoyansk population by the cherskij ridge (fig. 1). this population occasionally migrates to the momo-selenyakh depression. single animals occur near the upper timberline of the burkat and the selenyakhskij ridges. in the northern part of the yana-indigirskoe upland adjacent to foothills of the cherskij ridge, moose inhabit watersheds and valleys of the oldgo, abyrabyt, and djanky rivers with an average density of 1.2 animals/10 km2. northward they occur to the kular ridge and are found only in riparian forest along the kazachka and ulakhan-kyuegelir rivers. the chondon population is the northernmost in yakutia. it inhabits the upper chondon river mostly along the valleys of its tributaries the dodoma, ygannya, nagdakha, and nemkuchan rivers, as well as the northern nemkuchanskij, selenyakhskij, and irgichyanskij slopes on the kyun-tas ridge (fig. 1). this region spreads from the sub-tundra plain zone to mountain forests bordering tundra to the north and mountain tundra to the south; these varying habitat features dictate the specific habitat conditions and isolation of the population. the average density of moose in this area was 3.2 animals/10 km2 in october 1986 and 1.3 per 10 km2 in 1988. the momo-selenyakh population occupies the momo-selenyakh depression (fig. 1) with an average density of 0.3 moose/10 km2. density was higher (1.6) in the extended central part of the depression between the indigirka and selenyakh rivers. moose ascended mountains along the forested river valleys but were rarely found in unforested areas; some (0.5 animals/10 km2) grazed near herds of snow sheep (ovis nivicola) at 668 m elevation in the upper pechatnaya and the berezovka rivers in the northeastern part of the momskij ridge. they easily migrated beyond the limits of the momo-selenyakh depression northeast along the selenyakh river valley and lower slopes of the andrey-tas mountain ridge to the abyiskoe lowland and ozhoginskij dale where they come in contact with the indigirka and kolyma populations (fig. 1). the large indigirka population is widespread inhabiting the basin of the middle and lower region of the indigirka river (fig. 1). the population density in autumn-winter was 0.4-0.6 moose/10 km2 in the abyiskaya and yana-indigirskaya lowlands. the uyandinaselenyakh interfluve where moose are rarely found separates the indigirka and momaselenyakh populations. moose occur west of the indidgirka river north to the kondakovskij plateau (0.4 animals/10 km2). contact with the kolyma population to the west is rare except in the ozhogino dale, and the basin of the badyarikha and the ozhogino rivers where density ranges from 1.1-1.3 moose/10 km2. the indigirska population moves north for spring-summer grazing in tundra 50-60 km from the forest line near lake djyukarskoe; in october-november they occupy forested areas only. the most eastern population of moose in yakutia is the kolyma that occupies the basin of the kolyma river and yukagirskoe plateau, stretching to the verkhoyankoe plateau (fig. 1). the average population density varied from 0.8-1.1 moose/10 km2 depending on the moose in northern yakutia safronov alces vol. 45, 2009 20 landscape. here moose regularly migrate during fall-winter from the kolymskaya lowland to the alazeyskoe and yukagirskoe plateaus; this migration was initially documented by egorov (1965). identification of separate, regional populations of moose in yakutia was only possible after conducting extensive aerial and ground surveys at the end of the 1900s in the wake of an overall population decline. despite protective regulations, stability and recovery of the population was not realized; management strategies must be adapted relative to the regional differences in seasonal habitat use, migratory behavior, and relative isolation. effective moose management in yakutia will require further identification of range and habitat use, habitat structure and availability, and population estimates and dynamics of regional populations to develop effective measures for the continued protection and recruitment of moose in yakutia. references belyk, v. i. 1948. fur-bearing animals of yakutia. pages 191-253 in reports of the first scientific session of yakutsk division of as ussr. yakutsk, russia. (in russian). egorov, o. v. 1965. wild ungulates of yakutia. nauka publishing, moscow, russia. (in russian). filonov, k. p. 1983. the moose. lesnaya promyshlennost, moscow, russia. (in russian). tavrovsky, v. a., o. v. egorov, and v. g. krivosheev. 1971. mammals of yakutia. nauka publishing, moscow, russia. (in russian). yazan, y. p. 1972. game animals of pechora taiga. kirov branch of volgo-vyatsk publishing house, kirov, russia. (in russian). alces 31_153.pdf alces vol. 47, 2010 becker et al. moose highway crossings 69 spatial and temporal characteristics of moose highway crossings during winter in the buffalo fork valley, wyoming scott a. becker1,5, ryan m. nielson2, douglas g. brimeyer3, and matthew j. kauffman4 1wyoming cooperative fish and wildlife research unit, university of wyoming, department 3166, 1000 east university avenue, laramie, wy 82071, usa; 2western ecosystems technology, inc., 2003 central avenue, cheyenne, wy 82001, usa; 3wyoming game and fish department, 420 north cache, jackson, wy 83001, usa; 4u.s. geological survey, wyoming cooperative fish and wildlife research unit, university of wyoming, department 3166, 1000 east university avenue, laramie, wy 82071, usa. abstract: to accommodate increases in traffic volume and to address highway safety concerns, transportation managers often need to expand existing travel corridors which may result in increased risk of wildlife-vehicle collisions. by understanding the spatial and temporal characteristics of wildlife crossings, managers can apply appropriate mitigation techniques to reduce collision risk while maintaining habitat linkages. the u.s. highway 287/26 reconstruction project in northwest wyoming provided an opportunity to examine the influence of habitat, landscape, and anthropogenic features that influence highway crossing locations of shiras moose (alces alces shirasi). a model developed to estimate adult (≥ 2 years) female moose winter habitat selection was used at a smaller spatial scale to determine if it could accurately identify moose crossing locations along a 9.7-km section of u.s. highway 287/26 that bisects a high-density moose winter range in the buffalo fork valley. to test our model’s predictive capability, we used 201 moose crossing locations collected previously by independent researchers using snow-track survey techniques. the majority (81%) of moose crossing events occurred in areas classified as high or medium-high relative probability of use. we also examined temporal patterns of moose crossings and the influence of fence types in influencing crossing location. moose crossed the highway more frequently during early to mid-evening and less frequently during mid-day. our findings indicate that preferred habitat and landscape features such as relatively flat, low elevation habitats dominated by deciduous shrubs/trees interspersed with conifers had a stronger influence on crossing location than fences. alces vol. 47: 69-81 (2011) key words: alces alces shirasi, fence, habitat selection, highway crossing, moose, moose-vehicle collisions, spatial, temporal, wyoming. rising human populations create an increasing need to expand transportation corridors to accommodate the concurrent rise in traffic volume. this can lead to sharp increases in the number of wildlife-vehicle collisions (mcdonald 1991, groot bruinderink and hazebroek 1996, farrell and tappe 2007). in the united states, conover et al. (1995) estimated that approximately 726,000 deer (odocoileus spp.)-vehicle collisions occurred in 1991 costing an estimated $1,500 (u.s.) per accident. approximately 4% of these collisions resulted in human injuries with an estimated 211 human fatalities. when collisions occur with larger animals (e.g., moose [alces alces]), the risk of human injury and increased property damage rises significantly (joyce and mahoney 2001). methods aimed at reducing wildlife-vehicle collisions have been marginally successful. mitigation to reduce the number of collisions or prevent animals from entering the roadway (e.g., roadside clearing, 5present address: u.s. fish and wildlife service, p.o. box 2717, cody, wy 82414, usa. moose highway crossings – becker et al. alces vol. 47, 2011 70 fencing, overpasses, and underpasses) appear to be most effective, but maintenance and repair costs often limit their implementation and long-term effectiveness (bashore et al. 1985, feldhammer et al. 1986). wildlife-vehicle collisions can rarely be linked to a single factor, but the spatial and temporal patterns of accidents are not random events and appear to be related to daily and seasonal activity patterns of animals (bashore et al. 1985, gunderson et al. 1998, waller and servheen 2005, dodd et al. 2007). moreover, traffic volume, speed limits, driver awareness, and weather conditions have been implicated as factors influencing the risk of collisions (lavsund and sandegren 1991, joyce and mahoney 2001, seiler 2005). numerous studies have used modeling approaches to identify habitat, landscape, and anthropogenic features related to collision risk (hubbard et al. 2000, nielsen et al. 2003, malo et al. 2004, seiler 2005). oftentimes these models are presented with coefficients in tabular form and a description of the characteristics where wildlife are most likely to cross a road (e.g., waller and servheen 2005, dussault et al. 2007). by expanding on these models and mapping probabilities across a desired study area, transportation and wildlife managers can more easily interpret the likelihood that an animal will cross in a specific location. studies of wildlife-vehicle collisions often examine habitat and landscape characteristics after the frequency of accidents becomes socially unacceptable (finder et al. 1999, seiler 2005, dussault et al. 2006). by examining spatial and temporal patterns of animal movements associated with roadways, proactive engineering can be implemented into roadway design to reduce the likelihood of wildlife-vehicle collisions (groot bruinderink and hazebroek 1996, finder et al. 1999, dodd et al. 2007). the purpose of this study was to evaluate the efficacy of using a winter habitat selection model (becker 2008) for adult (≥2 years) female shiras moose (a.a. shirasi) to predict highway crossing locations in northwest wyoming. the u.s. highway 287/26 reconstruction project in northwest wyoming presented an opportunity to assess this technique because a 9.7-km section of this highway bisects a highdensity moose winter range in the buffalo fork valley (houston 1968). previous research that identified core crossing areas from snow-track surveys (young and sawyer 2006) provided independent data to validate our predictions of moose crossing locations. understanding spatial and temporal characteristics of moose crossings should improve mitigation efforts associated with highway improvement and construction. study area the winter study area was located approximately 50 km north of jackson, wyoming and was defined by the winter distribution of adult (≥2 years) female moose fitted with global positioning system (gps) collars (fig. 1; becker 2008). the area encompassed roughly 1,100 km2 of predominately public land which included portions of grand teton national park (gtnp) and bridger-teton national forest (btnf). all roads within the study area were 2 lane highways with speed limits ranging from 88 km/h in gtnp to 105 km/h outside of gtnp boundaries. from 2005-2007, mean daily traffic volume along u.s. highway 287/26 averaged 509 vehicles/ day during winter (november-april) and 1251 vehicles/day during summer (may-october; wyoming department of transportation [wydot] 2006, 2007, 2008). from 19752004, annual precipitation averaged 56.2 cm (range = 37.9-79.1 cm) with nearly 75% falling as snow from november-may (national oceanic and atmospheric administration 2005). vegetation types occurred along an elevational gradient. riparian areas dominated by willows (salix spp.) intermixed with narrowleaf cottonwood (populus angustifolia) alces vol. 47, 2010 becker et al. moose highway crossings 71 were located in large, relatively flat floodplain environments at lower elevations and along nearly all drainages within the study area. engelmann spruce (picea engalmanni) and subalpine fir (abies lasiocarpa) were found on some mesic sites while big sagebrush (artemisia tridentata) occurred in more xeric locations at lower elevations. higher elevations were dominated by lodgepole pine (pinus contorta), douglas fir (psuedotsugia menziesii), subalpine fir, and engelmann spruce interspersed with aspen (populous tremuloides) (knight 1994). the highway study area was approximately 34 km2 within the moose winter range in buffalo fork valley (fig. 1; see methods for descriptive definition of the study area) where moose density was estimated at 4.0 moose/km2 in 2005 (d. brimeyer, wyoming game and fish department, unpublished data). private land encompassed roughly 11 km2 with the remaining area managed by gtnp and btnf. the majority of private land was maintained for livestock grazing (i.e., herbaceous cover) and tourist accommodations; smaller areas were held in conservation easements to preserve natural vegetative communities. methods moose captures and tracking we darted and immobilized adult female moose from the ground or helicopter on winter range in the buffalo fork valley during february 2005 and 2006. we used 10 mg of thiafenfig. 1. the winter and highway study areas in northwest wyoming, 2005-2007. we used the winter study area to evaluate the frequency and timing of adult female moose crossing events along u.s. highway 287/26 and u.s. highway 26/89/187. the highway study area was located along a 9.7-km section of u.s. highway 287/26 in the buffalo fork valley. we used the highway study area to create and validate a predictive map of moose crossing locations and to evaluate the influence of fence types on moose crossing locations. moose highway crossings – becker et al. alces vol. 47, 2011 72 tanil (a-3080, wildlife pharmaceuticals, fort collins, colorado, usa; kreeger et al. 2005) for capture, and an intramuscular injection of 300 mg naltrexone (trexonil, wildlife pharmaceuticals, fort collins, colorado, usa) as an antagonist after handling. capture and handling procedures were performed in accordance with approved university of wyoming animal care and use committee protocols (approved 2005, 2006). moose were fitted with tgw-3700 global positioning system (gps) collars with store-on-board technology (telonics, mesa, arizona, usa) programmed to release on 1 march 2007. the collars were set to fix hourly locations during winter (15 november-15 june). the high fix-rate success (becker 2008) negated correction for fix-rate bias (nielson et al. 2009). frequency and timing of highway crossings to estimate the number of highway crossing events occurring within the study area, we mapped winter locations of moose from 2005-2007 in arcgis 9.2 (environmental systems research institute, redlands, california, usa), and used the home range tools extension (rodgers et al. 2007) to create movement paths for each individual. we assumed that a crossing occurred when the straight line between 2 consecutive locations crossed either u.s. highway 287/26 or u.s. highway 26/89/187. because locations were collected every hour, a crossing event was assumed to occur within the time period between 2 consecutive locations. the time of crossings was categorized into 4 distinct time periods reflecting daily activity patterns (renecker 1986): 1) 0300-0859 hr (early to mid-morning), 2) 0900-1459 hr (mid-day), 3) 1500-2059 hr (early to mid-evening), and 4) 2100-0259 hr (night). we used the r statistical software package (r core development team 2006) to run 200 bootstrap samples of individual moose (manly 2007) and estimated the mean proportion of moose crossings with 95% confidence intervals within each time period. the bootstrap results were plotted against the expected proportion of crossings to determine if moose crossed in proportion to expected in a given time period. we assumed that a difference existed if the expected proportion of crossings fell outside the range of the 95% confidence intervals. this method treats the marked animal as the experimental unit, thereby eliminating issues related to pooling data across individuals and the potential for spatial or temporal correlation in animal movement (thomas and taylor 2006). predicting crossing location development of the predictive map a resource selection function (rsf) was developed (see becker 2008) to estimate winter habitat selection characteristics of moose following methods outlined by sawyer et al. (2009). this modeling effort identified 7 variables as potentially important predictors of winter habitat selection by adult female moose. these included the proportion of riparian/deciduous shrub, mixed conifer, and aspen habitats, as well as elevation, habitat diversity, slope, and distance to coniferous cover. using population-level coefficients from the rsf, we developed a predictive map of possible moose crossing locations along a 9.7-km section of u.s. highway 287/26 in the buffalo fork valley. we first recorded the gps location of mile markers 3.0-9.0 along u.s. highway 287/26 and plotted these in arcgis. we then digitized the 9.7-km section between these mile markers from a u.s. geological survey (usgs) 1:24,000 scale digital orthophoto quarter quadrangle map, and divided each 1.6-km section into 10 equal segments representing secondary mile markers to the nearest 0.16 km. the highway study area was defined as that area within a 1.5 km buffer around the highway. the buffer distance represented the alces vol. 47, 2010 becker et al. moose highway crossings 73 approximate, average daily distance moved by gps-collared adult female moose during winter (becker 2008). to measure the 7 variables that were potentially important predictors of moose crossing locations, we created circular sample units with 25-m radii that were systematically distributed across the study area. using a 30 x 30-m vegetation layer, we extracted vegetation data from each sample unit using hawths analysis tools (beyer 2004) and calculated the proportion of each vegetation type occurring within each unit. we used spatial analyst to estimate slope from a usgs 26 x 26-m digital elevation model, and the existing vegetation layer to create a distance to cover layer. cover was defined strictly as coniferous habitats that could potentially provide thermal cover during winter. estimates for elevation (m), slope (°), and distance to cover (m) were extracted from the midpoint of each sample unit. we considered a quadratic term to estimate slope in addition to the linear form of the variable. the habitat diversity coefficient in our winter habitat selection model was estimated using a 250-m radii circular sampling unit to capture diversity at a biologically relevant scale (becker 2008). this distance represented the average distance an adult female moose traveled in a 4-hr period during winter. because this was the scale used in the original model, we created 250-m radii circular units centered on the midpoint of each 25-m radii sample unit. we extracted vegetation data from each circular unit and calculated a shannon-weiner diversity index based on the proportion of 6 vegetation classes (i.e., spruce-fir, lodgepole pine, mixed conifer, aspen, riparian/deciduous shrub, and burn/other) within each unit. we used the r statistical software package (r core development team 2006) to estimate the relative probability of use for each sample unit using population-level (adult female) winter habitat selection coefficients (becker 2008). the model predictions were assigned values from 1-4 representing the highest to lowest estimated use probabilities in 25% increments (i.e., highest use probability = 1 [highest 25%], lowest use probability = 4 [lowest 25%]; sawyer et al. 2009). these predictions were mapped across 50 x 50 m pixels in the highway study area. validating the predictive map young and sawyer (2006) recorded 201 moose crossing events from snow-track surveys along the 9.7-km section during winter 2003-2004 and 2004-2005. they recorded the location of each crossing event to the nearest 0.16-km mile marker rather than with a gps location. these crossing data were used to validate our predictive map and assess its accuracy in identifying crossing locations. since we did not know exactly where moose crossed the highway relative to the nearest secondary mile marker, we created 80-m buffers around each 0.16-km marker, extracted relative probability of use class information from each buffer, and estimated a mean relative probability of use class for each secondary mile marker. the 80-m buffer size covered the entire area between each 0.16-km marker without overlap, so each secondary mile marker could be uniquely associated with the number of crossings that were recorded at that location. markers with mean relative probability of use of 1.00-1.50 were assigned as class 1 (high-use area), means 1.51-2.50 were assigned as class 2 (medium-high-use area), means 2.51-3.50 were assigned as class 3 (medium-low-use area), and means 3.51-4.00 were assigned as class 4 (low-use area). we joined the relative probability of use class and the number of crossing events associated with each secondary mile marker from the independent sample. we used the r statistical software package (r core development team 2006) to run 200 bootstrap samples (manly 2007) to estimate the mean proportion of moose crossings with 95% confidence intervals that occurred within each relative probability of use class. we expected that the proportion of moose crossings would moose highway crossings – becker et al. alces vol. 47, 2011 74 equal the proportion of mile markers classified within each relative probability of use class. the bootstrap results were plotted against the expected proportion of crossings within each relative probability of use class. we assumed that a difference existed if the expected proportion of crossings did not fall in the range of the 95% confidence intervals. fence type and moose crossings to evaluate if fence type influenced moose movement, we created a gis layer depicting 3 different fence types found on both sides of the highway: 1) bighorn fence, 2) four-strand, barbed wire fence, and 3) buck-and-rail fence. the bighorn fence was a 2-pole, 2-wire fence that stood approximately 1.1 m high. sections of 4-strand, barbed-wire fence were located mostly along stretches with permanent standing water and were approximately 1.1 m high. a small section of buck-and-rail fencing about 1.5 m high was located west of the gtnp boundary. no fencing occurred within gtnp near the western end of the study area. at the eastern end, no fencing occurred between (approximately) milepost 8.5 and 9.0 on the north side of the highway, and from mileposts 8.0-9.0 on the south side. we assumed that the straight line used to depict moose movement accurately reflected the fence type that was crossed. only those crossing events that occurred within the 9.7km section were used to assess the effect of fence type; 19.4 km of fenced or unfenced area (both sides of highway) could potentially be crossed by moose. we used the r statistical software package (r core development team 2006) to run 200 bootstrap samples of individual moose (manly 2007) and estimated the mean proportion of moose crossings with 95% confidence intervals that occurred at each fence type (including those areas not fenced). we expected that the proportion of moose crossings would equal the proportion of each fence type that occurred within the highway study area. the bootstrap results were plotted against the expected proportion of crossings for each fence type. we assumed that a difference existed if the expected proportion of crossings did not fall in the range of the 95% confidence intervals. results a total of 257 crossing events were recorded; 19 of 22 collared moose crossed the 9.7-km section during the study period. only 8 moose crossed the highway ≥10 times and these moose accounted for 84% of all crossing events (n = 217). because the 4 time periods represented an equal proportion of a 24-h period, we expected an equal proportion of crossings during each time period (0.250). the bootstrapped proportion of crossing events was more than expected during early to midevening (x = 0.351; 95% ci = 0.299-0.401), less than expected during mid-day (x = 0.119; 95% ci = 0.088-0.151), and as expected during night (x = 0.272; 95% ci = 0.234-0.311) and early to mid-morning (x = 0.258; 95% ci = 0.215-0.311; fig. 2). the predictive map indicated that areas classified as high or medium-high relative probability of use occurred on both sides of u.s. highway 287/26 between mileposts 3.2 0.0 0.1 0.2 0.3 0.4 0.5 0300-0900 0900-1500 1500-2100 2100-0300 time of day p ro p o rt io n o f c ro ss in g s fig. 2. bootstrapped mean proportion of crossing events with 95% confidence intervals plotted against the expected proportion of crossings (dashed line) by time of day for adult female moose in northwest wyoming, winter 20052007. time periods represent early to midmorning (0300-0900 hr), mid-day (0900-1500 hr), early to mid-evening (1500-2100 hr), and night (2100-0300 hr). alces vol. 47, 2010 becker et al. moose highway crossings 75 and 4.5, 6.1 and 6.7, and 7.0 and 9.0 (fig. 3). the model predicted that these areas were characterized by a high proportion of riparian/ deciduous shrub and aspen habitat with a lower proportion of coniferous cover. landscape attributes indicated these sections of highway also had greater amounts of habitat diversity, were at lower elevations, had relatively flat slopes, and were moderate distance to cover. mileposts that occurred on either side of the bridge over the buffalo fork river and the bridge over blackrock creek were each classified as high-use areas which indicate a high likelihood that moose utilized these structures to cross u.s. highway 287/26. analysis of the 201 moose crossings documented previously (young and sawyer 2006) indicated that the highest percentage of crossing events occurred in areas classified as high or medium-high relative probability of use (81%, n = 162); fewer crossings occurred in areas classified as medium-low or low relative probability of use (19%, n = 39; fig. 3). the proportion of mile markers classified within the 4 use classes (high, medium-high, medium-low, low) was 0.229, 0.459, 0.164, and 0.148, respectively. moose crossed the highway in proportion to expected for all use classes (high: x = 0.323; 95% ci = 0.199-0.454; medium-high: x = 0.476; 95% ci = 0.339-0.612; medium-low: x = 0.108; 95% ci = 0.045-0.179; low: x = 0.092; 95% ci = 0.040-0.151), although the general trend was that frequency of crossing increased with higher probability of use (fig. 4). there was approximately 6.5 km of fencing fig. 3. relative probabilities and associated classes (low = 0-25%, medium-low = 26-50%, mediumhigh = 51-75%, high = 76-100%) of habitat use for the highway study area developed from a model of winter habitat selection for adult female moose in northwest wyoming, 2005-2007. the circles along the highway represent the number of moose crossings recorded to the nearest 0.16-km mile marker during winter 2003-2004 and 2004-2005 (data provided by young and sawyer 2006). moose highway crossings – becker et al. alces vol. 47, 2011 76 on the north and 6.3 km on the south side of the highway. the primary type was bighorn fence that was along 6.3 km and 4.5 km of the north and south sides, respectively. about 6.5 km (north = 3.2 km, south = 3.3 km) of highway was unfenced, most occurring within gtnp and east of blackrock creek. because buckand-rail fence and barbed wire fence occurred in relatively small proportions, these fence types were grouped into a combined category of “other” for analysis. a total of 311 fence crossings by 19 of 22 moose were recorded during the study period. only 9 moose crossed fences ≥10 times and these accounted for 87% of all crossing events (n = 269). bighorn fence occurred along the greatest proportion of highway (0.558) and unfenced (0.338) and "other" fence less (0.104). crossings occurred at unfenced areas in greater proportion than expected (x = 0.547; 95% ci = 0.341-0.702), in proportion to expected at bighorn fence (x = 0.417; 95% ci = 0.264-0.615), and less than expected in areas with “other” fence (x = 0.035; 95% ci = 0.013-0.068; fig. 5). discussion frequency and timing of highway crossing events approximately 88% of all moose crossing events in the buffalo fork valley occurred from early evening to mid-morning (15000859 hr), coinciding with peaks in daily moose activity patterns (renecker 1986). although traffic volumes generally decrease at night, low light conditions at dawn, dusk, and night increase the risk of collision. in newfoundland, approximately 75% of all moose-vehicle collisions occurred between sunset and sunrise and severe human injury and death were twice as likely to occur after dark (joyce and mahoney 2001). similarly, dussault et al. (2006) noted that moose-vehicle collisions were 2-3 x higher at night. it is unlikely that the relatively low winter traffic volume on u.s. highway 287/26 (wydot 2006, 2007, 2008) impedes highway crossings by moose; however, animals often avoid roadways during periods of high traffic volume. for example, elk (cervus elaphus) shifted use away from an arizona highway when diurnal traffic volume was high, yet returned at night when traffic volume declined (gagnon et al. 2007a). increased traffic volume was implicated in preventing bighorn sheep (ovis canadensis) from reaching important mineral sites in rocky mountain national park, colorado (keller and bender 2007). although many moose crossing events were documented in the buffalo fork valley, only 1 moose-vehicle collision was recorded during the study; it occurred at dusk near 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 bighorn other no fencing fence types p ro p o rt io n o f c ro ss in g s fig. 5. bootstrapped mean proportion of crossing events with 95% confidence intervals plotted against the expected proportion of crossings (dashed lines) by fence type along a 9.7-km section of u.s. highway 287/26 in the buffalo fork valley, wyoming, winter 2005-2007. 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 high medium-high medium-low low mean relative probability of use class p ro p o rt io n o f c ro ss in g s fig. 4. bootstrapped mean proportion of crossing events with 95% confidence intervals plotted against the expected proportion of crossings (dashed lines) by mean relative probability of use class along a 9.7-km section of u.s. highway 287/26 in the buffalo fork valley, wyoming, winter 2005-2007. alces vol. 47, 2010 becker et al. moose highway crossings 77 milepost 7.4 that was classified as a high probability of use area. while some accidents may go unreported, moose-vehicle collisions are relatively rare events in the buffalo fork valley with only 5 documented from 19952004 (young and sawyer 2006). predicting moose crossing locations in the buffalo fork valley moose crossings were not randomly distributed along the 9.7-km section of u.s. highway 287/26 in the buffalo fork valley; rather, aggregations of moose crossings occurred at locations that could be predicted by estimating winter habitat selection characteristics. the spatial aggregation of crossings demonstrated that collision risk was greatest in areas identified as high or medium-high relative probability of use. mitigation could be applied where crossings are most likely to occur in an attempt to reduce moose-vehicle collision risk. moose crossings were aggregated in areas where preferred habitat (deciduous shrubs/ trees) and landscape features occurred on both sides of the highway. adult female moose in northwest wyoming select for low-elevation, riparian habitats that contain abundant deciduous forage in winter (houston 1968, becker 2008). this relationship suggests that location of highway crossings was related to the spatial distribution of available forage; the same relationship was identified for moose in scandinavia (gundersen et al. 1998, seiler 2004, 2005). although moose crossings typically occurred in areas that contained abundant forage, crossing locations also had higher habitat diversity suggesting that the distribution of habitat types across the landscape likely influenced crossing locations. private lands used for livestock grazing adjacent to the highway were composed mostly of herbaceous cover and contained little habitat diversity or preferred forage; few moose crossings occurred in these areas (mile markers 4.5-6.1; fig. 3). in contrast, private lands held in conservation easements were composed of a mix of riparian and coniferous habitats and, not surprisingly, more crossings occurred in these areas (mile markers 6.1-7.0; fig. 3). in other areas of north america where preferred habitat was common and habitat diversity was relatively low, highway crossings and wildlife-vehicle collisions were more randomly distributed (bashore et al. 1985, feldhammer et al. 1986). bridges over the buffalo fork river and blackrock creek were identified as having a high probability of use suggesting that moose may utilize these structures to cross beneath the highway. although location frequency (hourly) was insufficient to confirm whether a moose actually used a bridge to cross underneath the highway, snow-track surveys and remotely triggered cameras documented numerous moose crossing under these bridges (young and sawyer 2006). lengthening existing bridges may facilitate wildlife crossings which should reduce the risk of wildlife-vehicle collisions along short sections of highway (seiler 2004, 2005). however, bridges can act as “edge-creating landscape features” that increase the risk of collisions (e.g., white-tailed deer [odocoileus viginianus]; hubbard et al. 2000). moreover, the low-intermittent traffic volume during winter in the buffalo fork valley (wydot 2006, 2007, 2008) might cause moose to flee from the infrequent, yet sudden auditory and visual stimuli of a vehicle crossing a bridge; this may increase the potential for collision if they cross in a less suitable location (gagnon et al. 2007b). fence type and moose crossings moose tended to cross the 9.7-km section of u.s. highway 287/26 more frequently in unfenced than fenced areas; however, fences within the buffalo fork valley were not designed to prevent moose crossings. seiler (2005) described the highest risk of moosevehicle collisions along sections of road withmoose highway crossings – becker et al. alces vol. 47, 2011 78 out moose-proof fencing. although fencing along an interstate highway in pennsylvania reduced the number of deer observed in the right-of-way, it did little to reduce the number of deer-vehicle collisions (feldhammer et al. 1986). the lack of fencing in areas of preferred moose habitat limits our assessment of the influence of fence type on moose crossings. nonetheless, we believe that preferred habitat and landscape features were most influential in determining where moose crossed the highway because fencing was not high enough to physically deter moose. management implications our study suggests that models of moose habitat selection and associated probability of use maps can be used to identify areas with an increased risk of moose-vehicle collisions. moreover, if time constraints created by highway construction projects prevent data collection and analysis of wildlife crossings, existing habitat-based models can be used to locate areas where mitigation techniques may be most appropriate (clevenger et al. 2002). using a habitat-based model, our results indicate that existing and proposed mitigation in the buffalo fork valley may be adequate unless moose-vehicle collisions increase following highway reconstruction. for example, the expansion of the buffalo fork bridge in 2007 created a wider corridor for moose to travel underneath the highway between high probability use areas, reducing the likelihood that moose would cross the road surface. plans to lengthen existing bridges over rivers and streams that act as natural travel corridors may facilitate additional animal movements under the highway between high use areas (hubbard et al. 2000, ng et al. 2004, sawyer and rudd 2005, seiler 2005); if so, transportation managers might avoid more costly mitigation such as underpasses and overpasses. if further improvements are needed in the buffalo fork valley, vegetation removal along the highway right-of-way to increase motorist visibility may be the most easily applicable and socially-acceptable form of mitigation (gundersen et al. 1998, rea 2003, andreassen et al. 2005). a suite of other ungulates, large and small carnivores, and rodents also cross highways (young and sawyer 2006), hence, prominent wildlife crossings should be identified and mitigation techniques should benefit multiple species (sawyer and rudd 2005). for example, core elk crossing areas (young and sawyer 2006) were similar to those identified for moose in the study area, thus, appropriate mitigation could reduce collision frequency for both species. implementing mitigation efforts that benefit multiple species will likely require detailed scientific data, such as used in our model, but it will ultimately benefit wildlife by maintaining important habitat linkages and reducing highway-related mortality (ng et al. 2004). acknowledgements we extend special acknowledgement to the late dr. s. h. anderson without whom none of this would have been possible. funding was provided by teton county conservation district, wyoming animal damage management board, wyoming department of transportation, wyoming game and fish department, and wyoming governor’s big game license coalition/wildlife heritage foundation of wyoming. we thank bridger-teton national forest, grand teton national park, and yellowstone national park for logistical support and our pilots g. lust (mountain air research, driggs, idaho [retired]), d. savage (savage air services, american falls, idaho), and d. stinson (sky aviation, worland, wyoming) for their expertise in the air. we wish to acknowledge the efforts of numerous personnel who assisted with field, office, and logistical support, especially c. anderson, c. beers, s. dewey, t. fuchs, w. hubert, s. kilpatrick, t. kreeger, f. lindzey, w. long, m. patritch, h. sawyer, and t. thurow. a. barbknecht, alces vol. 47, 2010 becker et al. moose highway crossings 79 t. thomas, e. wald, and one anonymous reviewer improved this manuscript through their constructive reviews. references andreassen, h. p., h. gundersen, and t. storaas. 2005. the effect of scentmarking, forest clearing, and supplemental feeding on moose-train collisions. journal of wildlife management 69: 1125-1132. bashore, t. l., w. m. tzilkowski, and e. d. bellis. 1985. analysis of deer-vehicle collision sites in pennsylvania. journal of wildlife management 49: 769-774. becker, s. a. 2008. habitat selection, condition, and survival of shiras moose in northwest wyoming. m.s. thesis, university of wyoming, laramie, wyoming, usa. beyer, h. l. 2004. hawth’s analysis tools for arcgis version 3.26. (accessed august 2006). clevenger, a. p., j. wierzchowski, b. chruszcz, and k. gunson. 2002. gisgenerated, expert-based models for identifying wildlife habitat linkages and planning mitigation passages. conservation biology 16: 503-514. conover, m. r., w. c. pitt, k. k. kessler, t. j. dubow, and w. a. sanborn. 1995. review of human injuries, illness, and economic losses caused by wildlife in the united states. wildlife society bulletin 23: 407-414. dodd, n. l., j. w. gagnon, s. boe, and r. e. schweinsburg. 2007. assessment of elk highway permeability by using global positioning system technology. journal of wildlife management 71: 1107-1117. dussault, c., j-p. ouellet, c. laurian, r. courtois, m. poulin, and l. breton. 2007. moose movement rates along highways and crossing probability models. journal of wildlife management 71: 2338-2345. _____, m. poulin, r. courtois, and j-p. ouellet. 2006. temporal and spatial distribution of moose-vehicle accidents in the laurentides wildlife reserve, quebec, canada. wildlife biology 12: 415-425. farrell, m. c., and p. a. tappe. 2007. countylevel factors contributing to deer-vehicle collisions in arkansas. journal of wildlife management 71: 2727-2731. feldhammer, g. a., j. e. gates, d. m. harman, a. j. loranger, and k. r. dixon. 1986. effects of interstate highway fencing on white-tailed deer activity. journal of wildlife management 50: 497-503. finder, r. a., j. l. roseberry, and a. woolf. 1999. site and landscape conditions at white-tailed deer/vehicle collisions in illinois. landscape and urban planning 44: 77-85. gagnon, j. w., t. c. theimer, n. l. dodd, s. boe, and r. e. schweinsburg. 2007a. traffic volume alters elk distribution and highway crossings in arizona. journal of wildlife management 71: 2318-2323. _____, _____, _____, a. l. manzo, and r. e. schweinsburg. 2007b. effects of traffic on elk use of wildlife underpasses in arizona. journal of wildlife management 71: 2324-2328. groot bruinderink, g. w. t. a., and e. hazebroek. 1996. ungulate traffic collisions in europe. conservation biology 10: 1059-1067. gundersen, h., h. p. andreassen, and t. storaas. 1998. spatial and temporal correlates to norwegian moose-train collisions. alces 34: 385-394. houston, d. b. 1968. the shiras moose in jackson hole, wyoming. technical bulletin no. 1. grand teton natural history association, moose, wyoming, usa. hubbard, m. w., b. j. danielson, and r. a. schmitz. 2000. factors influencing the location of deer-vehicle accidents in iowa. journal of wildlife management 64: 707-713. moose highway crossings – becker et al. alces vol. 47, 2011 80 joyce, t. l., and s. p. mahoney. 2001. spatial and temporal distributions of moose-vehicle collisions in newfoundland. wildlife society bulletin 29: 281-291. kellar, b. j., and l. c. bender. 2007. bighorn sheep response to road-related disturbance in rocky mountain national park, colorado. journal of wildlife management 71: 2329-2337. knight, d. l. 1994. mountains and plains: the ecology of wyoming landscapes. yale university press, new haven, connecticut, usa. kreeger, t. j., w. h. edwards, e. j. wald, s. a. becker, d. brimeyer, g. fralick, and j. berger. 2005. health assessment of shiras moose immobilized with thiafentanil. alces 41: 121-128. lavsund, s., and f. sandegren. 1991. moosevehicle relations in sweden: a review. alces 27: 118-126. malo, j. e., f. suárez, and a. díez. 2004. can we mitigate animal-vehicle accidents using predictive models? journal of applied ecology 41: 701-710. manly, b. f. j. 2007. randomization, bootstrap, and monte carlo methods in biology. third edition. chapman & hall/ crc, boca raton, florida, usa. mcdonald, m. g. 1991. moose movement and mortality associated with the glenn highway expansion, anchorage, alaska. alces 27: 208-219. ng, s. j., j. w. dole, r. m. sauvajot, s. p. d. riley, and t. j. valone. 2004. use of highway undercrossings by wildlife in southern california. biological conservation 115: 499-507. nielsen, c. k., r. g. anderson, and m. d. grund. 2003. landscape influences on deer-vehicle accident areas in an urban environment. journal of wildlife management 67: 46-51. nielson, r. m., b. f. manly, l. l. mcdonald, h. sawyer, and t. l. mcdonald. 2009. estimating habitat selection when gps fix success is less than 100%. ecology 90: 2956-2962. national oceanic and atmospheric administration. 2005. (accessed october 2005). r core development team. 2006. r 2.4.0: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. rea, r. v. 2003. modifying roadside vegetation management practices to reduce vehicular collisions with moose alces alces. wildlife biology 9: 81-91. renecker, l. a. 1986. bioenergetics and behavior of moose (alces alces) in the aspen boreal forest. ph.d. dissertation, university of alberta, edmonton, alberta, canada. rodgers, a. r., a. p. carr, h. l. beyer, l. smith, and j. g. kie. 2007. hrt: home range tools for arcgis. version 1.1. ontario ministry of natural resources, centre for northern forest ecosystem research, thunder bay, ontario, canada. sawyer, h., and b. rudd. 2005. pronghorn roadway crossings: a review of available information and potential options. western ecosystems technology, inc., cheyenne, wyoming, usa. _____, m. j. kauffman, and r. m. nielson. 2009. influence of well pad activity on winter habitat selection patterns of mule deer. journal of wildlife management 73: 1052-1061. seiler, a. 2004. trends and spatial patterns in ungulate-vehicle collisions in sweden. wildlife biology 10: 301-313. _____. 2005. predicting locations of moosevehicle collisions in sweden. journal of applied ecology 42: 371-382. thomas, d. l., and e. j. taylor. 2006. study designs and tests for comparing resource use and availability ii. journal of wildlife management 70: 324-336. waller, j. s., and c. servheen. 2005. effects alces vol. 47, 2010 becker et al. moose highway crossings 81 of transportation infrastructure on grizzly bears in northwestern montana. journal of wildlife management 69: 985-1000. (wydot) wyoming department of transportation. 2006. automatic traffic recorder report for togwotee pass, 2005. wyoming department of transportation, cheyenne, wyoming, usa. _____. 2007. automatic traffic recorder report for togwotee pass, 2006. wyoming department of transportation, cheyenne, wyoming, usa. _____. 2008. automatic traffic recorder report for togwotee pass, 2007. wyoming department of transportation, cheyenne, wyoming, usa. young, d., and h. sawyer. 2006. wildlife crossing study: u.s. highway 287/26, moran junction – dubois. western ecosystems technology, inc., cheyenne, wyoming, usa. alces vol. 48, 2012 cumberland aerial survey for moose 67 potvin double-count aerial surveys in new brunswick: are results reliable for moose? roderick e. cumberland new brunswick department of natural resources, fish and wildlife branch, p.o. box 6000, fredericton, new brunswick, canada e3b 5h1. abstract: following the rapid decline of deer (odocoileus virginianus) across northern new brunswick in the late 1980s, the new brunswick department of natural resources began to utilize a double-count helicopter survey to estimate deer numbers. although the survey was designed for deer, moose (alces alces) sightings were also recorded; however, no analysis was conducted on the accuracy or usefulness of these data to estimate moose numbers. the survey design was a modification of the potvin double-count survey method for deer which accounts for most caveats to aerial surveys. this double-count (mark-recapture) technique allows calculation of bias for both observers, for single and groups of moose, and individual flights. moose population estimates calculated from 79 flights ranged from 0.17-3.49 moose/km2 and were similar to a variety of estimates throughout north america. population estimates from 2004-2009 correlated well with corresponding 2009 population indices for moose based on number of moose seen by deer hunters (corr. = 0.725, p <0.001). the potvin estimates in wildlife management zone 2 were highly correlated (0.82-0.93, p <0.05) with other indices based on road kill moose, moose sightings, and harvest success rates; moose sightings and hunter success were also correlated in several other zones. this analysis indicates that potvin surveys produce reliable population density estimates of moose in boreal/acadian forests, given that the sighting probability is >0.4 and flights occur before mid-february when moose may occupy denser canopy cover. alces vol. 48: 67-77 (2012) key words: aerial survey, alces alces, deer, density estimate, moose, new brunswick. sustainable and effective management of large ungulates requires a reliable estimate of population size by management unit. gauging the success of annual management actions and harvest prescriptions hinges upon the ability to obtain cost-effective density estimates with precision that allows detection of annual changes. following a rapid decline of deer (odocoileus virginianus) numbers across northern new brunswick, the new brunswick department of natural resources (nbdnr) initiated the use of a double-count aerial survey in 1996 that was developed in quebec to estimate deer densities (rice and harder 1977, potvin et al. 1992, hardy 1994, potvin and breton 1995). although the survey was designed and modified specifically for the observation of deer in boreal forests, observers also recorded sightings of moose (alces alces). however, no further analysis was conducted on the reliability of this survey or how precise or accurate it was to estimate moose numbers. various survey methods have been used to determine ungulate densities including line transects, distance sampling, spotlight surveys, and variable plot surveys (kie 1988, rabe et al. 2002). other methods include track counts, pellet group surveys, and measure of browse abundance that provide indirect counts or indices (kendall et al. 1992, witmer 2005), and although cost-effective, are not without their challenges and inaccuracies (storm et al. 1992, fritzen et al. 1995, focardi et al. 2002, freddy et al. 2004, collier et al. 2007). some of these methods have low power to detect population change of 20-50% (strayer 1999) or produce estimates with large confidence aerial survey for moose cumberland alces vol. 48, 2012 68 limits (jordan et al. 1993). although more expensive, aerial surveys can provide quick, reliable estimates of ungulate density (freddy et. al. 2004); however, aerial surveys have their own set of bias and problems. early in the evolution of aerial surveys, caughley (1974) and caughley et al. (1976) outlined deficiencies that included width of transect surveyed, cruising speed, altitude, height above ground, observer experience, and habitat cover type. while quadrat flights had the advantages of counting a higher percentage of animals and removing the need for strip width, transects were more robust and allowed for ease in calculating observer bias and were more economical (caughley 1977). nearly all of these biases result in fewer sightings and less precise estimates. estimates obtained from aerial surveys improved as these problems were addressed (peterson and page 1993). the use of helicopters addressed aircraft speed (peterson and page 1993) and improved the accuracy of moose estimates by 78%, on average, over fixed winged aircraft (gosse et al. 2002). beasom et al. (1981) tested sightability within the survey strip and found that observers saw 34-73% fewer deer in the outer 50 m of a 100 m survey strip compared to the first 50 m of the strip. deyoung et al. (1989) found that they could improve density estimates with little or no loss in precision by reducing their strip width from 200 to 40 m. in wyoming, 42% of moose missed occurred >50 m from the transect path (anderson and lindzey 1996), evidence that the search area not exceed this width. aerial surveys are typically flown over open habitat types because dense overstories combined with oblique angles of view hide ungulates from observers (anderson and lidnzey 1996). however, floyd et al. (1978) found that modifying search methodology greatly improved observability of deer in coniferous forests, such as altering the viewing angle (potvin et al. 1992, gauthier and cumberland 2000). in addition, background snow greatly improved sightability to as high as 78% in an oak-hickory forest (beringer et al. 1998). because not all animals are observed in aerial surveys, several methods are used to correct for this bias, including calculating correction factors by habitat type (anderson and lindzey 1996) or incorporating some form of mark-recapture using 2 observers, which has proved the simplest and most efficient means (pollock and kendall 1987). a double-count procedure is essentially a mark-recapture technique that improves estimates by accounting for missed animals (magnusson et. al. 1978, eberhardt and simmons 1987, borchers et al. 1998). deyoung et al. (1989) used correction factors to improve their population under-estimate (42.3% less than their bailey’s estimate) to within 6% of the estimated population size. further improvement to precision is possible by calculating visibility bias for both observers, and for both single and groups of deer (samuel and pollock 1981, rivest et al. 1995). with such modifications, doublecount aerial surveys that address these issues arguably hold promise for estimating ungulate density in coniferous-deciduous forests. potvin deer estimates are considered reliable in new brunswick. deer densities estimated in wmz 22 from 2000-2009 were strongly correlated to estimates derived from population reconstruction with the provincial deer model. estimated rates of increase and decrease for potvin estimates, the deer model, and actual harvest changes in wmz 22 were strongly correlated throughout the 2000s (nbdnr unpublished data). the objective of this study was to evaluate the reliability of potvin estimates of moose population density in new brunswick by using >10 years of aerial survey data from multiple wildlife management zones (wmz) of variable moose density across new brunswick. alces vol. 48, 2012 cumberland aerial survey for moose 69 methods study area new brunswick is located on the east coast of canada at 46.00° n and 66.54° w. over half of it borders the atlantic ocean along the chaleur bay, northumberland strait, and the bay of fundy. the majority (>80%) of the landbase is forested, and elevation varies from lowlands at sea level with little relief along the eastern coast, to highlands of the appalachian mountains in the north to elevations of 764 m. woodlands are characterized by acadian forest, with spruce (picea spp.), balsam fir (abies balsamea), hemlock (tsuga canadensis), and white cedar (thuja occidentalis) as climax species on softwood sites, and sugar maple (acer saccharum), beech (fagus grandifolia), and yellow birch (betula alleghaniensis) characterizing mature hardwood sites. early successional forests are predominated by poplars (populus spp.), birches (betula spp.), willows (salix spp.), and cherry (prunus spp.). new brunswick is divided into 27 wmzs (fig. 1) to facilitate localized and sitespecific management through seasonal and zone-specific quotas for antlerless deer and all sexes of moose. field techniques our survey method (hereafter referred to as potvin) was modified slightly (gauthier and cumberland 2000) from the double-count aerial survey used to estimate deer numbers by potvin et al. (1992). we flew a 4-seat, bell 206 jet ranger helicopter with a gps (global positioning system) navigation unit. the aircraft was equipped with side bubble windows on the left (port) side that allowed for a wide field of view to reduce parallax, and greatly reduced the effect of dense conifer cover on sightability of moose by reducing parallax angles <45° in the survey strip. the survey strip itself was limited to a 60 m swath and a flight altimeter was installed to maintain consistent height above undulating topography. flight speeds did not exceed 60 knots and aircraft height above the canopy was kept constant at 60 m. we found that slow speeds and low altitude of the survey aircraft usually initiated some form of movement of deer and moose below the helicopter and greatly assisted with observability. the flight crew consisted of the pilot, a navigator-recorder, and 2 observers seated on the left side of the helicopter. two observers were employed for the double-count estimator; observability bias was calculated for both observers, for single and groups of deer, and for each individual flight. every effort was made to ensure that observers were experienced and remained the same within fig. 1. distribution of wildlife management zones (wmz) in new brunswick; wmzs 1-3, 6, and 10 were used to compare with moose population estimates in adjacent maine. aerial survey for moose cumberland alces vol. 48, 2012 70 each annual survey. to approximate a true mark-recapture study, observers must be isolated from audio to prevent overhearing conversation that might bias sightings. the helicopter was equipped with a modified intercom system designed to allow the navigator-recorder to hear the observers while controlling their access to audio with a toggle switch. although snow is not necessary for the survey, the color contrast created between deer, moose, and snow greatly assists observers and increases sightability. therefore, we only conducted aerial surveys following snowfall that completely covered the ground. because both deer and moose were counted, surveys were flown as long as snow occurred, but this was altered as we experienced behavioral change of both deer (yarding) and moose (restricted movement). survey block selection ballard et al. (1997) conducted flights of the potvin double-count survey and suggested that the survey block size should be >200 km2 with transect lines spaced 1 km apart and of equal length to address spatial distribution of deer, and to meet statistical requirements. therefore, we selected large survey blocks with gross habitat characteristics that matched the wmz as a whole. although the spatial distribution of habitat types could not be evaluated, proportions of habitat types were summarized for each wmz using a gis database. to make selection of survey blocks easier, gis map tiles were used as components to define survey blocks and their boundaries. because gis tiles were 43 km2, combinations of 6 tiles (approximately 258 km2) in a matrix were selected in each zone (n = 20-40 per wmz) as potential survey blocks; each was summarized by the 9 habitat types. survey blocks which most closely matched the total zone habitat characteristics based on simple linear regression analysis were selected (adjusted r² values >85%) and subsequently evaluated for topographic characteristics. the block with topographical characteristics with the fewest obstructions (i.e., large lakes, dense housing, suburban areas) and that was most consistent with the habitat in the wmz as a whole was selected. transect lines were flown systematically beginning at the edge of the block and spaced 1 km apart and oriented either n-s or e-w. transects were approximately 40 km long to encompass as much of the variation in habitat as possible to ensure variation between transects was minimized and precision maximized (caughley et al. 1976). moose in new brunswick are currently managed by tracking various indices of population change through time. indices across and within wmzs include the number of moose seen per hour by moose hunters, the number of moose seen by deer hunters, annual road kill in each wmz, and success rate of moose hunters. density estimates were compared to the relative index of moose seen per deer hunter in each zone to determine if these relative values were comparable. changes in population and indices were determined by the slope of a regression line fitted to each index, and by measuring r by regressing log e of the estimated population size (caughley and birch 1971). these rates were compared between estimated population change by the aerial survey and other indices of change. index values were correlated to potvin estimates over time to determine if changes in moose indices were reflected by related changes in the potvin estimates. our estimated moose densities were also compared to estimates from surrounding jurisdictions and others in the literature. results the nbdnr flew 154 individual flights over various habitat types and in nearly all wmzs since 1996. thirty-five (22.5%) were flown as test flights in february and march 1996 and 1997 to determine appropriate block size, alignment, and location to calibrate the survey and to form sight images for deer and alces vol. 48, 2012 cumberland aerial survey for moose 71 moose; therefore, these test flights were not included in the analysis. since 1997, several flights were flown in february or later to estimate deer numbers, and several others had sighting probabilities <0.40. these additional 14 flights were also excluded from the analysis because moose typically move into more coniferous cover in mid-late winter thereby reducing sightability; simulations indicate a loss of robustness when sighting probability is <0.45 (magnusson et al. 1978, potvin et al. 1992). of the remaining 105 flights, 26 flights were excluded from analysis because they did not produce a population estimate due to too few moose. our 1996-1997 test flights suggested that densities <0.5 per km² resulted in >80% of survey lines without sightings. as expected, central and southern wmzs with lower moose densities and harvests accounted for 17 of these 26 flights. population estimates (unpublished data, nbdnr) from the remaining 79 flights indicated high moose density in northern wmzs where deer populations are substantially lower, and lower moose density in southern wmzs. significant within-zone variability was only evident in 4 (5%) flights. estimated moose densities varied from 3.49 moose/km² in northern zones (table 1) to as low as 0.17 moose/km² in southern zones (table 2). average sightability for front and rear observers was 0.85 and 0.83, respectively. fewer than 8 wmzs were flown yearly due to financial and weather limitations; therefore, few zones have continuous data. because potvin estimates were never obtained in all zones in one year, the most recent potvin estimate in each from 2004-2009 was compared with the 2009 index of moose seen by deer hunters. these two values correlated well (corr 0.725, p <0.001), and better than most other indices with each other except the 2 moose sighting indices that were correlated (corr. 0.80, p <0.001). wmz 2 was flown most consistently (1996-1998, 2000, 2003-2004, 2006, and 2008-2009) and provided the best data set for temporal comparisons of the potvin estimates. the observed rate of increase (r) for potvin estimates and all other moose indices in wmz 2 were positive and of similar magnitude (table 3). potvin estimates were highly correlated (0.82-0.93; all p <0.05) with all other populawmz year # moose density ± sd 1 2008 24 1.39 ± 0.42 2 2009 52 3.49 ± 1.14 3 2006 16 1.18 ± 0.32 4 2008 10 0.62 ± 0.18 5 2008 42 2.66 ± 0.77 6 2009 27 1.70 ± 0.83 7 2007 20 1.28 ± 0.44 8 2008 13 0.82 ± 0.23 10 2009 11 0.69 ± 0.36 12 2009 9 0.54 ± na. table 1. potvin estimates of moose population density (moose/km²) in northern wmzs (higher density wmzs) in new brunswick, 2006-2009; wmzs 1-3, 6, and 10 border maine. a wmz was measured in november or december (see methods). wmz year moose density ± sd 13 2004 18 1.07 ± 0.32 14 2006 17 1.13 ± 0.34 15 2004 29 1.79 ± 0.44 16 2007 7 0.48 ± 0.15 17 2003 9 0.66 ± na 18 2007 21 1.37 ± na 20 2007 11 0.68 ± 0.29 21 2009 43 1.29 ± 0.23 22 2009 3 0.17 ± na 23 1999 4 0.25 ± 0.13 25 2009 10 0.72 ± 0.26 table 2. potvin estimates of moose population density (moose/km²) in southern wmzs (lower moose density wmzs) in new brunswick, 20032009. a wmz was measured in december or january (see methods). aerial survey for moose cumberland alces vol. 48, 2012 72 tion indices in wmz 2 (i.e., road kill, moose sightings, and harvest success rate; fig. 2). potvin estimates were also highly correlated to moose sightings (corr. 0.90, p <0.001) and hunter success (corr. 0.79, p <0.011) in wmz 8. although fewer data points were available for potvin estimates in wmz 4 and 5, they were strongly correlated with moose sightings (corr. 0.84, p <0.001) and hunter success (corr. 0.75, p <0.001) in wmz 4, and to a lesser extent in wmz 5 (corr. 0.61, p <0.082 and corr. 0.46, p <0.21, respectively). discussion thirteen years of deer and moose population estimates have been done with the potvin double-count survey in new brunswick. the survey was designed to address limitations of aerial surveys by flying at slow speed (60 kph) and at low altitude (60 m) to increase the likelihood of deer and moose movement that, in turn, increases the likelihood of sightability. specifically, bubble windows reduced parallax and vastly improved the ability to sight ungulates under closed coniferous canopies. the flight altimeter allowed the pilot to maintain survey height and reduce fluctuations in survey width that can affect population estimates. use of a survey transect width of 60 m also reduced parallax, providing easier observation of a narrower strip, including the portion most likely to have observable animals. using 2 observers in the double-count allowed use of the mark-recapture technique that vastly improves estimates, and calculation of observer bias for each flight as well as singles and groups of animals from which the population estimate is corrected. behavioral and ecological differences of ungulate species may introduce bias and affect the accuracy and use of aerial surveys. for example, mccorquodale (2001) found that the spatial distribution of male and female elk during winter was a potential source of bias in helicopter surveys because females were 9 times more likely to be observed than males. however, anderson and lindzey (1996) successfully observed 59% of moose groups over a wide range of habitat types with strip widths of 150-250 m (3-5 x larger than here). given the size and color of moose, and that potvin estimates were found reliable for deer under similar protocols and conditions, it was presumed that moose had equal or higher sightability than deer, and that bias was negligible. because moose shift from deciduous to conifer habitats as snow depth increases (coady 1974, peek et al. 1976), some caution against flying in late winter (february-april) when moose tend to occupy denser vegetation (lynch 1975, karns 1982, gasaway et al. 1985, crete et al. 1986, anderson and lindzey 1996, ballard et al. 1997). the most optimal time to survey moose is when they occupy more open habitat types (anderson and lindzey 1996); thus, most aerial surveys are flown late fall and into early winter when snow covers the ground and moose occupy relatively open habitat types (coady 1974, lynch 1975, peek et al. 1976, karns 1982, gasaway et al. 1985, crete et al. 1986, anderson and lindzey 1996, quayle et al. 2001). quayle et al. (2001) found that % vegetative cover and snow cover, as well as temperature, affected sightability of moose in british columbia. our estimates of moose density are comparable with those in adjacent maine (l. kantar, maine department of inland fisheries and wildlife, pers. comm.) where the range of density estimates along the new brunswick metric estimated r r² adj. doubling time potvin 0.176 84.9 3.9 years harvest 0.226 70.8 3.1 years road kill 0.276 77.3 2.5 years success 0.358 82.0 1.9 years sightings by mh 0.364 65.9 1.9 years table 3. the estimated rate of increase (r) and the corresponding doubling time for the population (caughley and birch 1971) for moose population indices in wmz 2, new brunswick. alces vol. 48, 2012 cumberland aerial survey for moose 73 border is 1.24-3.05 moose/ km²; density in bordering wmzs 1, 2, and 6 were 1.393.49 (table 1). in nearby new hampshire, the density estimate was 1.19 moose/ km² using the flir method (bontaites et al. 2000). population density of heavily harvested populations or those in poor habitat are typically <1.0 moose/km², but in less-exploited situations, density can approach 3.0/km², and in extreme cases be as high as 5-10/km². collectively, these other estimates indicate that the potvin estimates reported in this study (0.17-3.49 moose/ km²) are reasonable. due to cost and weather conditions, <10 of 25 wmzs were flown in any given year. further, consecutive flights within wmzs were sporadic due to the need for specific data from other zones in successive years. while this limits the continuous nature of the data set, sufficient data existed in several zones to make comparisons to other moose population indices in new brunswick. i compared the 2006-2009 potvin estimates to the number of moose observed by deer hunters because this index was more robust than the observation rate by moose hunters. moose observations by 3,500 moose hunters during the 3-day season was low in some wmzs, versus the number of moose observed 0 0.5 1 1.5 2 2.5 3 3.5 4 0 5 10 15 20 25 30 35 40 45 50 po tv in d en si ty (m oo se /k m 2 ) number of road killed moose y = 0.0679x + 0.1764 r2 adj. = 0.774, p <0.05 0 0.5 1 1.5 2 2.5 3 3.5 4 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 po tv in d en si ty (m oo se /k m 2 ) moose hunter success rate y = 8.1442x 4.7078 r2 adj. = 0.766, p <0.05 0 0.5 1 1.5 2 2.5 3 3.5 4 0.05 0.1 0.15 0.2 0.25 0.3 0.35 po tv in d en si ty (m oo se /k m 2 ) # moose sighted/hour hunted y = 10.358x 0.0384 r2 adj. = 0.859, p <0.05 fig. 2. relationships between potvin estimates (moose/km²) and 3 indices of moose population density in wmz 2 in new brunswick, 1996-2009: a) road kill, b) moose hunter success rate, and c) moose sightings per hour by moose hunters. each index was positively correlated (p <0.05) with the potvin estimates. aerial survey for moose cumberland alces vol. 48, 2012 74 by 50,000 deer hunters over 4 weeks. the potvin estimates correlated well with the relative number of moose observed by deer hunters across wmzs. although the absolute change in the potvin estimates for wmz 2 increased over the course of our study from 1077 to 8731, the estimated rate of increase was less than that calculated for the other indices of population change (i.e., roadkill, hunter sightings, hunter success, and harvest; table 3). the potvin estimates were highly correlated to changes in hunter success and moose sightings in northern zones where a sufficient number of flights allowed for this analysis. although the potvin estimates correlated well with the other indices of moose density, 9 flights did not produce a density estimate in areas where the moose population was presumably high enough to provide an estimate. further, 4 of the 79 flights where estimates were obtained had a fluctuation >50% from the previous or consecutive estimate. preliminary analysis shows little evidence that this was attributed to weather or temperature anomalies. the question remains whether such estimates should be excluded from an annual population analysis, or be treated as a low population and managed accordingly; current analysis has not formed the basis for management decisions. rivest et al. (1995) analyzed visibility bias of the potvin survey and found that doublecount estimates were superior to those derived from independent sightability studies such as gasaway and other single survey types. potvin estimators also have the advantage of reflecting the prevailing conditions during the actual survey. potvin and breton (2005) tested the double-count method simultaneously with infrared surveys and found that the thermal infrared technique provided more variable results; accuracy was 54-89% and was mostly influenced by forest canopy. conversely, 4 of 6 potvin surveys yielded reliable results and the density estimates were within 64-83% of the assumed densities based on population reconstruction. provided sighting probabilities are >0.45, the double count method provides valid estimates of deer density for management purposes; however, if sighting probabilities are <0.40 it might underestimate density (potvin et al. 2004). moose in new brunswick are currently managed by tracking various indices of population change through time, although none have been tested or validated with actual population sizes or rates of change. the potvin method could be a valuable tool in producing more reliable moose population estimates in boreal/ acadian forests provided sighting probability is >0.40 and flights occur prior to february. because unreported harvest of moose by first nation’s people in new brunswick creates a void in age data which limits the ability to implement population reconstruction (e.g., paloheimo and fraser model [1981]), early winter potvin estimates might provide both a population estimate and a basis for understanding any population response associated with unreported harvest. in new brunswick, deer and moose data were collected simultaneously and such an approach could be more cost-effective than traditional, single species surveys in jurisdictions with multiple species of ungulates. however, at northern latitudes where deer tend to occupy closed-canopy mature coniferous forest in winter, the timing of favorable weather, presence of snow, and yarding migration greatly narrows the window when combined surveys could occur. although the potvin estimates were limited to <10 wmzs flown annually and few zones had continuous annual data, the population estimates were reasonable and correlated with other population indices. annual flights may not be necessary to calibrate other modeling and population reconstruction efforts, and this technique may also prove advantageous in high density areas requiring local management strategies. alces vol. 48, 2012 cumberland aerial survey for moose 75 references anderson, c. r. jr., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24: 247-259. ballard, w. b., j. a. kershaw, h. a. whitlaw, d. l. sabine, g. j. forbes, and s. young. 1997. a preliminary evaluation of the aerial double count technique for estimating white-tailed deer densities in new brunswick. unpublished report. university of new brunswick, fredericton, new brunswick, canada. beasom, s. l., j. c. hood, and j. r. cain. 1981. the effect of strip width on helicopter censusing of deer. journal of range management 34:36-37. beringer, j., l. p. hansen, and o. sexton. 1998. detection rates of white-tailed deer with a helicopter over snow. wildlife society bulletin 26: 24-28. bontaites, k. m., k. a. gustafson, and r. makin. 2000. a gasaway-type moose survey in new hampshire using infrared thermal imagery: preliminary results. alces 36: 69-76. borchers, d. l., w. zicchini, and r. m. fewster. 1998. mark-recapture models for line transect surveys. biometrics 54: 1207-1220. caughley, g. 1974. interpretation of age ratios. journal of wildlife management 38: 557-562. _____. 1977. analysis of vertebrate populations. john wiley and sons ltd., new york, new york, usa. _____, and l. birch. 1971. rate of increase. journal of wildlife management 35: 658-663. _____, r. sinclair, and d. scott-kemmis. 1976. experiments in aerial survey. journal of wildlife management 40: 290-300. coady, j. w. 1974. influence of snow on behaviour of moose. naturaliste canadian 101: 417-436. collier, b. a., s. s. ditchkoff, j. b. raglin, and j. m. smith. 2007. detection probability and source of variation in whitetailed deer spotlight surveys. journal of wildlife management 71: 277-281. crete, m., l. rivest, h. jolicoeur, j. brassard, and f. messier. 1986. predicting and correcting helicopter counts of moose with observations made with fixed-wing aircraft in southern quebec. journal of applied ecology 23: 751-761. deyoung, c. a., f. s. guthrey, s. l. beasom, s. p. coughlan, and j. r. heffelfinger. 1989. improving estimates of white-tailed deer from helicopter surveys. wildlife society bulletin 17: 275-279. eberhardt, l. l., and m. a. simmons. 1987. caliberating population indices by double sampling. journal of wildlife management 51: 665-675. floyd, t. j., l. d. mech, and m. e. nelson. 1978. an improved method of censusing deer in deciduous-coniferous forests. journal of wildlife management 43: 258-261. focardi, s., r. isotti, and a. tinelli. 2002. line transect estimates of ungulate populations in a mediterranean forest. journal of wildlife management 66: 48-58. freddy, d. j., g. c. white, m. c. kkeeland, r. h. kahn, j. w. unsworth, w. j. devergie, v. k. graham, j. h. ellenberger, and c. h. wagner. 2004. how many mule deer are there? challenges of credibility in colorado. wildlife society bulletin 32: 916-927. fritzen, d. e., r. f. labiski, d. e. easton, and j. c. kilgo. 1995. nocturnal movements of white-tailed deer: implications for refinement of track count surveys. wildlife society bulletin 23: 187-193. gasaway, w. c., s. d. dubois, and s. j. harbo. 1985. biases in aerial transect surveys for moose during may and june. journal of wildlife management 49: 777-784. gauthier, p., and r. e. cumberland. 2000. a aerial survey for moose cumberland alces vol. 48, 2012 76 new brunswick manual for aerial surveys of white-tailed deer populations utilizing the potvin mark-recapture technique. deer technical report no. 5. new brunswick department of natural resources, fredericton, new brunswick, canada. gosse, j., b. mclaren, and e. eberhardt. 2002. comparison of fixed-wing and helicopter searches for moose in a midwinter habitat-based survey. alces 38: 47-53. hardy, t. 1994. methods of estimating white-tailed deer densities and population trends a review. senior thesis, faculty of forestry. university of new brunswick, fredericton, new brunswick, canada. jordan, p. a., r. o. peterson, p. campbell, and b. mclaren. 1993. comparison of pellet counts and aerial counts for estimating density of moose at isle royale: a progress report. alces 29: 267-278 karns, p. d. 1982. twenty-plus years of aerial moose census in minnesota. alces 18: 186-196. kendall, k. c., l. h. metzgar, d. a. patterson, and b. m. steele. 1992. power of sign surveys to monitor population trends. ecological applications 2: 422-430. kie, j. g. 1988. performance in wild ungulates: measuring population density and condition of individuals. usda forest service general technical report psw106. pacific southwest forest and range experiment station, berkeley, california, usa. lynch, g. m. 1975. best timing of moose surveys in alberta. alces 11: 141-153. magnusson, w. e., g. j. caughley, and g. c. grigg. 1978. a double survey estimate of population size from incomplete counts. journal of wildlife management 42: 174-176. mccorquodale, s. m. 2001. sex-specific bias in helicopter surveys of elk: sightability and dispersion effects. journal of wildlife management 65: 216-225. paloheimo, j. e., and d. fraser. 1981. estimation of harvest rate and vulnerability from age and sex data. journal of wildlife management 45: 948-958. peek, j. m., d. l. uurich, and r. j. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3-65. peterson, r. o., and r. e. page. 1993. detection of moose in midwinter from fixedwing aircraft over dense forest cover. wildlife society bulletin 21: 80-86. pollock, k. f., and w. l. kendall. 1987. visibility bias in aerial surveys: a review of estimation procedures. journal of wildlife management 51: 502-510. potvin, f., and l. breton. 1995. standard for aerial surveys of white-tailed deer populations. quebec department of environment and wildlife, rene-levesque, quebec, canada. _____, and _____. 2005. from the field: testing 2 aerial survey techniques on deer in fenced enclosures – visual double-counts and thermal infrared sensing. wildlife society bulletin 33: 317-325. _____, _____, and l-p. rivest. 2004. aerial surveys for white-tailed deer with the double-count technique in quebec: two 5-year plans completed. wildlife society bulletin 32: 1099-1107. _____, _____, _____, and a. gingras. 1992. application of a double-count aerial survey technique for white-tailed deer, odocoileus virginianus, on anticosti island, quebec. canadian field-naturalist 106: 435-442. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43-54. rabe, m. j., s. s. rosenstock, and j. c. devos, jr. 2002. review of big game survey methods used by wildlife agencies of the western united states. wildlife society alces vol. 48, 2012 cumberland aerial survey for moose 77 bulletin 30: 46-52. rice, w. r., and j. d. harder. 1977. application of multiple aerial sampling to a mark-recapture census of white-tailed deer. journal of wildlife management 41: 197-206. rivest, l-p., f. potvin, h. crepeau, and g. daigle. 1995. statistical methods for aerial surveys using the double-count technique to correct visibility bias. biometrics 51: 461-470. samuel, m. d., and k. h. pollock. 1981. correction of visibility bias in aerial surveys where animals occur in groups. journal of wildlife management 45: 993-997. storm, g. l., d. f. cottam, r. h. yahner, and j. d. nichols. 1992. a comparison of 2 techniques for estimating deer density. wildlife society bulletin 20: 197-203. strayer, d. l. 1999. statistical power of presence-absence data to detect population declines. conservation biology 13: 1034-1038. witmer, g. w. 2005. wildlife population monitoring: some practical considerations. wildlife research 32: 259-263. alces vol. 48, 2012 edlich and stolter – effects of odours on moose 17 effects of essential oils on the feeding choice by moose sabine edlich and caroline stolter department of animal ecology and conservation, biocenter grindel, martin-luther-king platz 3, 20146 hamburg, germany. abstract: moose (alces alces) browse on coniferous tree species to different extents during winter; for example, norway spruce (picea abies) is avoided, scots pine (pinus sylvestris) is preferred, with juniper (juniperus communis) of intermediate use. conifers contain essential oils that may act as feeding deterrents, thereby reducing food intake by herbivores. because essential oils are volatile, our objectives were to determine if 1) odour plays a role in the food choice by moose, 2) whether single monoterpenes act as feeding deterrents, and 3) if this might be a mechanism used to discriminate against unpalatable plants. the essential oils of norway spruce and juniper and 2 monoterpenes (limonene and camphene) predominant in the essential oil of norway spruce were tested for their potential as deterrents in feeding trials. deterrence was assessed in food choice experiments by measuring the time spent feeding on food treated with the different odours associated with these compounds. there was no statistical evidence that food treated with the essential oils of spruce and juniper and single monoterpenes from norway spruce were avoided by moose. however, our data indicate that the essential oil of norway spruce probably has a negative effect on moose foraging because of the large absolute difference in feeding time between treatments and that overall, odour had a significant effect on feeding time. because our experimental design may have influenced the results, we suggest research approaches to better measure deterrence effects. alces vol. 48: 17-25 (2012) key words: alces alces, conifers, essential oil, feeding choice, feeding time, monoterpenes, moose, odour. plants have evolved direct and indirect plant defence mechanisms for protection against pathogens and herbivores (rosenthal and janzen 1979, dicke and vet 1999). defence mechanisms can be mechanical (e.g., burning hair, thorns, spikes, wax coatings) or chemical such as plant secondary metabolites (psm; fox 1981, karban and myers 1989). the largest groups of psms are phenols, terpenoids, and alkaloids; their effectiveness as feeding deterrents is due to their toxicity (postabsorptive effect), inhibition of food digestion (post-ingestive effect), and deterrence through smell or taste (pre-ingestive effect) (bryant et al. 1991, gershenzhorn and dudareva 2007, stolter et al. 2009). however, it is important to recognize that psms are common in the plant kingdom and part of the natural diets of many herbivores. conifers form vast forests distributed widely in the northern hemisphere, with many playing important economic roles in the wood industry, including production of resin and essential oils (kubeczka and schultze 1987). foraging by large herbivores can cause substantial damage to coniferous forests including direct destruction of trees, especially in monocultures and young stands (sjöberg and danell 2001, edenius et al. 2002). conifers have a high diversity of psms that presumably deter feeding by mammalian herbivores (e.g., bryant et al. 1991, epple et al. 1996). effects of odours on moose – edlich and stolter alces vol. 48, 2012 18 specifically, essential oils of conifers have a wide variety of monoterpenes with some acting as deterrents to snowshoe hares (lepus americanus; sinclair et al. 1988), red deer calves (cervus elaphus; elliott and loudon 1987), and moose (alces alces; sunnerheimsjöberg 1992). danell et al. (1990) found a negative correlation between consumption and concentration of the terpenoid pinifolic acid in scots pine (pinus sylvestris); a similar observation was made by sunnerheim-sjöberg (1992) with a different monoterpene [(-) angelicoidenol-2-o-β-d-glucopyranoside] in scots pine. the reluctance of herbivores to feed on monoterpenes might relate to inhibition of microbial activity in the digestive system (postingestive effect). for example, oh et al. (1967) studied the essential oils of douglas-fir needles (pseudotsuga menziesii) and found that oxygenated monoterpenes decreased microbial activity in the rumens of sheep (ovis aries) and deer (odocoileus hemionus columbianus). therefore, avoidance of monoterpenes, hence avoidance of some coniferous tree species, might be learned from negative, post-ingestive effects. furthermore, monoterpenes are characterized by their highly distinctive odour, and because of their volatility, monoterpenes and consequently essential oils might also deter animals prior to ingestion. moose forage on leaves, shoots, and twigs of lignified plants including twigs, needles, and bark of conifers in winter; their food selection can be influenced by psms (danell et al. 1990, stolter et al. 2005, stolter 2008). among conifers, moose prefer scots pine which is known for low concentration, but high diversity in phenolic compounds compared to other coniferous trees (stolter et al. 2010). in contrast, norway spruce (picea abies) that is common throughout europe, is avoided and used only when food resources are scarce. other conifers like common juniper (juniperus communis) vary in utilization among habitats (hörnberg 2001, månsson et al. 2007, pers. observ., c. stolter). because monoterpenes of the essential oils of conifers are volatile, their odour might be one cue in forage selection by moose. however, the role of volatile monoterpenes and essential oils in forage selection is little explored, although smell appears important in food choice (levin 1976, bryant et al. 1991). we investigated whether the essential oils of norway spruce and common juniper, and specific monoterpenes influence forage selection of moose. specifically, we wanted to determine if 1) odour plays a role in forage selection by moose, 2) whether single monoterpenes act as feeding deterrents, and 3) if this might be a mechanism to discriminate against unpalatable plants. based on the assumption that an animal theoretically maximizes its net calorie intake per feeding time (emlen 1966) and that diet optimization is influenced by nutritional value of food (e.g., positive effects of nutrients and negative effects of psms; freeland and janzen 1974, pulliam 1975), we used feeding time to investigate the possible differences in deterrent effects. methods choice of essential oils we assumed our captive moose would have similar forage selection as wild moose that prefer scots pine, typically reject norway spruce, and have intermediate use of common juniper (hörnberg 2001, månsson et al. 2007, pers. observ. by authors). we tested and verified this assumption in a pilot study when we fed captive moose twigs of the 3 species (edlich 2009); our findings were in accordance with previous studies. consequently, we used 6 substances in our experiments: essential oils of norway spruce and common juniper (because we assumed that these odours might be deterrent) and the monoterpenes limonene, camphene, borneol, and eucalyptol. these monoterpenes were chosen because they are predominant in the essential oils of norway spruce, but rare in scots pine and common alces vol. 48, 2012 edlich and stolter – effects of odours on moose 19 juniper (edlich 2009). however, due to restrictions of the zoo, we removed the experiments for borneol and eucalyptol; thus, only 4 experiments are presented here. because it was not possible to extract enough essential oils from norway spruce and common juniper for the experiments, we substituted essential oils purchased commercially. to examine differences in the terpenoid composition of the essential oils of plant samples, and to examine if the commercial oils (spruce oil, juniper oil; shandiin, hamburg, germany) could be used for our experiments, we compared the chemical profiles of the commercial oils with the essential oils extracted by distillation from plant material collected during winter 2008-2009 in lower saxony, germany (53°09’02“n, 9°54’44“e). we sampled up to 5 individual trees per species by clipping the first 4-8 cm of several branches; those samples were combined into a single composite sample per species. plant material was frozen in plastic bags at -20° c until distillation. oils were extracted by steam distillation as described by pfannkuche (2000). about 50 g of frozen needles were distilled for 3 h and the extracted essential oil was analysed by gas chromatography linked with a mass spectrometer (shimadzu gc-ms qp 2010s). the chemical profile of the distillate oils matched that of the commercial oils. experiments feeding trials were used to test the deterrent effect of odour with 3 female and 1 male moose housed together in a 12-ha enclosure in wildpark lüneburger heide, lower saxony, germany; all had previously eaten scots pine and norway spruce as part of their winter diet. they had access to 4 feeding troughs placed adjacently (~1 m apart) of which only 3 were used. typically, moose used the feeding troughs in the same arrangement; often 2 moose (usually a cow and her yearling) fed together at 1 trough. because they fed voluntarily, not all animals participated in each experiment (1-h feeding trial); therefore, experiments were repeated 10 times for each odour. toward the typical feeding times, we placed 2 plastic boxes (40 x 33.5 x 8.5 cm) with a known weight of food pellets in each of the 3 troughs (wildkraftfutter sommer für wiederkäuer, nösenberger pferdefutter; brackel, germany). one of the boxes was perforated and underneath had an unreachable pad of cotton wool soaked with 5 µl of an essential oil or monoterpene; this low concentration was used to mimic the odour of a non-damaged tree. the plastic boxes were cleaned with 2-propanol and equipped with an unused pad after each trial. to prevent preference for a specific box, the positions of the boxes were changed randomly. moose were allowed to acclimate to the feeding protocol and boxes for 2 days, after which the daily feeding routine consisted of 2 daily feeding times: 1000-1100 and 1500-1600 hr during which we carried out the 1-h experiments. the experiments consisted of 4, 10-day periods partitioned into 2, 5-day periods separated by a 2 day break. in each 10-day period there were 10, 1-h feeding experiments of the 4 treatments (spruce, juniper, limonene, or camphene). treatments differed between the morning and afternoon feedings. because consumption of the food pellets was nearly complete in each trial, we used feeding time to assess deterrence. further, because 2 moose often fed together at 1 trough, we were unable to measure the amount of food consumed by an individual moose. therefore, a video camera (sony full hd camcorder) was installed at each of the 3 used troughs to record consumption time and identify moose. water and branches of deciduous trees were available ad libitum throughout the experiments. statistical analysis statistics were performed with pasw 18 (spss 2010, ibm cooperation). the wilcoxon-test was used to test for differences effects of odours on moose – edlich and stolter alces vol. 48, 2012 20 in feeding time between the boxes with and without odour. we calculated a general linear model (glm) for repeated measurements to test for differences among the different odour experiments. before using this statistical approach, we tested against violation of sphericity using the maucleystest. we included the differences in feeding time between the boxes with odour and without odour as a dependent variable. to gain a balanced design, we included only the first 4, 1-h measurements of each animal; because not every animal participated in every experiment, 4 days was the maximum participation of 1 moose per treatment. further, by using only the first 4 measurements, we presumably minimized the possibility of habituation influencing the results. we used each treatment and individual moose as within-subject effects. we tested for differences between the treatment with essential oil of spruce and the other treatments using within-subject contrasts; this was also done to test for differences between animals (p-values are bonferroni-corrected). results comparison between feeding boxes with and without odour we tested for differences between the boxes with and without odour for each treatment separately. the absolute mean feeding time was higher for boxes with odour of the essential oil of norway spruce (80%) and common juniper (19%); the opposite occurred for limonene (13% lower) and camphene (4% lower) (table 1, fig. 1). although no treatment was statistically different (all p >0.05; table 1), there was a strong tendency with norway spruce. comparison of different odours because of our repeated measurement design, we conducted further analyses with a glm for repeated measurements using the difference between boxes with and without odour (time feeding on the box with odour – time feeding on the box without odour). again, we restricted our data to the first 4, 1-h experiments of each moose, and calculated the mean differences within a treatment (fig. 2). odour was a significant (p = 0.003) inner-subject factor (table 2). the contrasts (within subject-contrasts) between spruce vs. limonene (p = 0.010) and spruce vs. camphene (p = 0.017) were also significant (table 2); odour box without odour box with odour mean se mean se z p spruce 212.94 22.28 385.50 28.61 -1.82 0.068 juniper 298.00 71.41 355.25 79.29 -0.73 0.465 limonene 269.88 44.15 236.13 48.97 -1.10 0.273 camphene 292.63 79.26 282.06 78.65 -0.37 0.715 table 1. mean feeding time [s] of moose (n = 4) at the boxes with and without odour. repeated measurements (4, 1-h experiments) were pooled for each animal before statistical analyses. we used wilcoxon-test to test for significant differences. fig.1. mean feeding time (± se) of moose (n = 4) on boxes without odour (black circles) compared to boxes with odour (open circles). repeated measurements (4, 1-h experiments) were pooled. alces vol. 48, 2012 edlich and stolter – effects of odours on moose 21 though not significant, the spruce vs. juniper contrast showed a strong tendency (p = 0.086; table 2). moose were not significant in withinsubject contrasts (table 2). discussion most psms are not acutely toxic (e.g., phenols and terpenes) but have negative effects at certain concentrations (bryant et al. 1983, 1991, mcarthur et al. 1991, mcintosh et al. 2003, stolter et al. 2005). therefore, many animals have evolved mechanisms to detect these compounds and regulate their intake (dearing et al. 2005), which presumably reflects the variable use of coniferous species by herbivores (hansson et al. 1986, roy and bergeron 1990, eppele et al. 1996). given their volatility, monoterpenes are detected through chemical sensory perception like smell (chapmann and blaney 1979), and because the essential oil of each coniferous species has characteristic composition and concentrations of volatile monoterpenes, species-specific odours result (e.g., norway spruce vs. scots pine). because norway spruce is not a preferred browse of moose and used only when forage is limited, our aim was to determine whether its essential oil or one of its monoterpenes has a deterrent effect on moose. we found no statistical difference between the feeding time spent on treated and untreated samples. further, none of our treatments with single components acted as an absolute deterrent indicating that there was no strong individual effect of the psms on food consumption. however, because all food was consumed in each trial, it is possible that our ability to measure deterrence was masked by the experimental protocol. in contrast to a similar study with red deer (elliott and loudon 1987), we found no difference in the feeding time between boxes with or without odour for all treatments. however, we found a strong tendency (80% difference, table 1) that moose fed longer on boxes treated with glm (repeated measurements) within-subjecteffects df f p odour (3/9) 10.02 0.003 moose (1.07/3.20) 01.60 0.295 interaction (2.48/7.44) 01.12 0.390 within-subject contrasts f p spruce/juniper 06.39 0.086 spruce/limonene 34.51 0.010 spruce/camphene 23.07 0.017 moose 1 / 2 00.02 0.906 moose 2 / 3 05.52 0.100 moose 3 / 4 02.32 0.225 table 2. results of a general linear model (glm) for repeated measurements (n = 4, 1-h experiments) using the difference in feeding time between the boxes with and without odour as a dependent variable. odour (essential oil of norway spruce and common juniper, limonene, and camphene) and individual moose (n = 4 animals) were used as within-subject effects. we tested for differences between the spruce treatment and the other treatments and for differences between the animals using within-subject contrasts (p-values are bonferroni corrected). 300 200 100 0 -100d iff er en ce s be tw ee n bo x w ith /w ith ou t o do ur [s ] spruce juniper limonene camphene fig.2. mean differences in feeding time between the boxes with and without odour (± se) for 4 moose. values above zero indicate longer feeding times on the boxes with odour compared to the boxes without odour. repeated measurements (4, 1-h experiments) were pooled. effects of odours on moose – edlich and stolter alces vol. 48, 2012 22 the essential oil of norway spruce indicating its possible negative effect on consumption. we suggest that moose fed longer on the box with odour due to deterrence because animals should maximize their food intake (emlen 1966); slower feeding might indicate that animals are more cautious or avoid a particular food (e.g., post-ingestive effect). the glm supported this idea because odour had a significant effect on feeding time between the boxes, and comparative testing showed that these differences were pronounced between the essential oil of norway spruce and both monoterpenes, but not between the essential oils of both conifers (table 2). interestingly, an opposite result was found for limonene; moose fed longer on the box without odour resulting in a pronounced negative difference (fig.2), suggesting that limonene had no or positive influence on foraging. elliott and loudon (1987) tested the odour of essential oils of sitka spruce (picea sitchensis) and lodgepole pine (pinus contorta) and selected monoterpenes on red deer by measuring mass of food consumed rather than time spent feeding. they found a deterrent effect for the essential oil of lodgepole pine on male deer, but not for sitka spruce or with female deer (but see duncan et al. 1994, 2001); in contrast to our results, the monoterpene limonene acted as a feeding deterrent for female red deer. these differences might relate to experimental design, specifically, by measuring food intake and using higher concentrations of monoterpenes given that the effects of psms are dose-dependent (harborne 1991). our objective was to determine if consumption of the plant species was deterred by odour, not to measure the concentration limit where monoterpenes might act as a deterrent. furthermore, the use of plant material instead of essential oils (our experiment) might be more “natural” experimentation, although cutting plant material causes an increase in odour due to the damage of resin ducts. assuming that animal experience might also influence such studies, we note that our moose had previously fed on norway spruce and may have been “olfactory-adapted.” because all animals were housed together during the experiment, individual feeding might have been affected by dominance, although we found no significant effect with individual moose (table 2). the odours of essential oils function to signal chemical composition of a plant to an animal, which in the case of norway spruce is determined by its variety of terpenoids and high concentration of specific phenolics (stolter et al. 2010). these compounds might (in addition to terpenoids) have negative influences on digestion (stolter et al. 2009). however, our results did not indicate that a strong deterrent effect exists because no treatment acted as an absolute deterrent. interestingly, we found differences between the influence of monoterpenes and essential oils. chemical defence mechanisms are complex, and essential oils are a combination of individual components (i.e., specific monoterpenes) that can act synergistically to provide greater toxicity or deterrence than the equivalent amount of a single substance (gershenzon and dudareva 2007). in his assessment of the environmental effectiveness of terpenoids, harborne (1991) showed that the concentration and universally dependent dose were important. herbivores could not “smell out” certain monoterpenes from an essential oil, but if a monoterpene was predominant or overbalanced the concentration, the essential oil might provide an avoidance or deterrence effect. because the study moose were not deterred by the treatments, we assume that smell alone probably plays a minor role in the relative use and avoidance of norway spruce; taste and texture (chapmann and blaney 1979) and/or the concentrations of other compounds (e.g., phenols) are likely important. conducting experiments with captive moose is logistically difficult and has degrees of compromise because of limited animal number and size of treatment groups. acalces vol. 48, 2012 edlich and stolter – effects of odours on moose 23 knowledging such limitations, future research concerning the effects of psms on food use and choice by moose might test treatments on individual moose, examine effects at different concentrations of psms, use mass of food consumed in combination with feeding time as indicators of preference, control for common experience with psms of experimental moose, and consider the influences of sight and taste. acknowledgements we would like to thank wildpark lüneburger heide and wildpark schwarze berge in lower saxony, germany for their cooperation, and irene tomaschewski for technical support. references bryant, j. p., f. d. provenza, j. pastor, p. b. reichardt, t. p. clausen, and j. t. du toit. 1991. interactions between woody plants and browsing mammals mediated by secondary metabolites. annual review of ecology and systematics 22: 431-446. , g. d. wieland, p. b. reichardt, v. e. lewis, and m. c. mccarthy. 1983. pinosylvin methyl ether deters snowshoe hare feeding on green alder. science 222: 1023-1025. chapmann, r. f., and w. m. blaney. 1979. how animals perceive secondary compounds. pages 161-198 in g. a. rosenthal, and d. h. janzen, editors. herbivores their interaction with secondary metabolites. academic press, new york, new york, usa. danell, k., r. gref, and r. yazdani. 1990. effects of monoand diterpenes in scots pine needles on moose browsing. scandinavian journal of forest research 5: 535-539. dearing, m. d., w. j. foley, and s. mclean. 2005. the influence of plant secondary metabolites on the nutritional ecology of herbivorous terrestrial vertebrates. annual review of ecology and systematics 36: 169-189. dicke, m., and l. vet. 1999. plant-carnivore interactions: evolutionary and ecological consequences for plant, herbivore and carnivore. pages 483-520 in h. olff. v. k. brown, and r. h. drent, editors. herbivores: between plants and predators. blackwell science, oxford, uk. duncan, a. j., s. e. hartely, and g. r. iason. 1994. the effect of monoterpene concentration in sitka spruce (picea sitchensis) on browsing behaviour of red deer (cervus elaphus). canadian journal of zoology 72: 1715-1720. _____, _____, m. thurlow, s.young, and b. w. staines. 2001. clonal variation in monoterpene concentrations in sitka spruce (picea sitchensis) saplings and its effect on their susceptibility browsing damage by red deer (cervus elaphus). forest ecology and management 148: 259-269. edenius, l., m. bergman, g. ericsson, and k. danell. 2002. the role of moose as a disturbance factor in managed boreal forests. silva fennica 36: 57-67. edlich, s. 2009. einfluss des geruches auf das frassverhalten von elchen (alces alces), untersucht am beispiel dreier koniferenarten. diploma thesis, university hamburg, hamburg, germany. elliot, s., and a. loudon. 1987. effects of monoterpene odors on food selection by red deer calves (cervus elaphus). journal of chemical ecology 13: 1343-1349. emlen, j. m. 1966. the role of time and energy in food preference. american naturalist 100: 611-617. epple, g., h. niblick, s. lewis, l. d. nolte, d. l. campell, and j. r. mason. 1996. pine needle oil causes avoidance behaviors in pocket gopher geomys bursarius. journal of chemical ecology 22: 1013-1025. fox, l. r. 1981. defense and dynamics in effects of odours on moose – edlich and stolter alces vol. 48, 2012 24 plant-herbivore systems. american zoologist 21: 853-864. freeland, w. j., and d. h. janzen. 1974. strategies in herbivory by mammals: the role of plant secondary compounds. american naturalist 108: 269-289. gershenzon, j., and n. dudareva. 2007. the function of terpene natural products in the natural world. nature chemical biology 3: 408-414. hansson, l., r. gref, l. lundren, and o. theander. 1986. susceptibility to vole attacks due to bark phenols and terpenes in pinus contorta provenances introduced into sweden. journal of chemical ecology 12: 1569-1578. harborne, j. b. 1991. ecological chemistry and biochemistry of plant terpenoids. pages 399-426 in j. b. harborne, f. a. tomás-barberán, and a. francisco, editors. proceedings of the phytochemical society of europe, clarendon press. oxford, uk. hörnberg, s. 2001. the relationship between moose (alces alces) browsing utilization and the occurrence of different forage species in sweden. forest ecology and management 149: 91-102. karban, r., and j. h. myers. 1989. i nduced plant responses to herbivory. annual review of ecology and systematics 20: 331-348. kubeczka, k.-h., and w. schultze. 1987. biology and chemistry of conifer oils. flavour and fragrance journal 2: 137148. levin, d. a. 1976. the chemical defenses of plants to pathogens and herbivores. annual review of ecology and systematics 7: 121-159. månsson, j., c. kalén, p. kjellander, h. andrén, and h. smith. 2007. quantita-2007. quantitative estimates of tree species selectivity by moose (alces alces) in a forest landscape. scandinavian journal of forest research 22: 407-414. mcarthur, c., a. e. hagerman, and c. t. robbins. 1991. physiological strategies of mammalian herbivores against plant defenses. pages 103-114 in r. t. palo, and c. t. robbins, editors. plant defenses against mammalian herbivory. crc press, boca raton, florida, usa. mcintosh, f. m., p. williams, r. losa, r. j. wallace, d. a. beever, and c. j. newbold. 2003. effects of essential oils on ruminal microorganisms and their protein metabolism. applied and environmental microbiology 69: 5011-5014. oh, h. k., t. sakai, m. b. jones, and w. m. longhurst. 1967. effect of various essential oils isolated from douglas fir needles upon sheep and deer rumen microbial activity. applied and environmental microbiology 15: 777-784. pfannkuche, a. 2000. einsatzmöglichkeiten der mikrodestillation zur gewinnung und fraktionierung kleiner mengen ätherischer öle. ph.d. thesis, university of hamburg, hamburg, germany. pulliam, h. r. 1975. diet optimization with nutrient constraints. american naturalist 109: 765-768. rosenthal, g. a., and d. h. janzen. 1979. herbivores their interaction with secondary plant metabolites. academic press, new york, new york, usa. roy, j., and j.-m. bergeron. 1990. role of phenolics of coniferous trees as deterrents against debarking behavior of meadow voles (microtus pennsylvanicus). journal of chemical ecology 16: 801-808. sinclair, a. r. e., m. k. jogia, and r. j. andersen. 1988. camphor from juvenile white spruce as an antifeedant for snowshoe hares. journal of chemical ecology 14: 1505-1514. sjöberg, k., and k. danell. 2001. introduction of lodgepole pine in sweden -ecological relevance for vertebrates. forest ecology and management 141: 143-153. alces vol. 48, 2012 edlich and stolter – effects of odours on moose 25 stolter, c. 2008. intra-individual plant response to moose browsing: feedback loops and impacts on multiple consumers. ecological monographs 78: 167-183. _____, j. p. ball, r. julkunen-tiitto, r. lieberei, and j. u. ganzhorn. 2005. winter browsing of moose on two different willow species: food selection in relation to plant chemistry and plant response. canadian journal of zoology 83: 807-819. _____, _____, p. niemelä, and r. julkunentiitto. 2010. herbivores and variation in the composition of specific phenolics of boreal coniferous trees: a search for patterns. chemoecology 20: 229-242. _____, p. niemelä, j. p. ball, r. julkunentiitto, a. vanhatalo, k. danell, t. varvikko, and j. u. ganzhorn. 2009. comparison of plant secondary metabolites and digestibility of three different boreal coniferous trees. basic and applied ecology 10: 19-26. sunnerheim-sjöberg, k. 1992. (1s, 2r, 4s, 5s)-angelicoidenol-2-o-β--dglucopyranoside--a moose deterrent compound in scots pine (pinus sylvestris l.). journal of chemical ecology 18: 2025-2039. tributebubenik.pdf alces34(1)_139.pdf alces30_81.pdf alces35_11.pdf alces30_57.pdf alces30_127.pdf alces vol. 48, 2012 duetsch and peterson sexing moose by pelvis 1 using pelvis morphology to identify sex in moose skeletal remains jason a. duetsch1, 3 and rolf o. peterson2 1department of fish, wildlife, and conservation biology, colorado state university, fort collins, colorado 80523-1474, usa; 2school of forest resources and environmental science, michigan technological university, houghton, michigan 49931, usa. abstract: the only published method for sex determination in even-toed ungulates (i.e., cervidae) through the use of skeletal remains (excluding the skull) is pelvic suspensory tuberosity presence/absence in white-tailed (odocoileus virginianus) and black-tailed deer (odocoileus hemionus columbianus). tuberosities are not easily distinguishable on moose (alces alces) pelvises, even when a large number are available for comparison. unlike in horses (equus caballus) with similar skeletal structure as moose, pelvic inlets of moose show no distinctive sex characteristics on an individual level. several linear angular (n = 5) and linear (n = 3) measurements were made on isle royale moose pelvises (n = 35). results showed statistically significant differences between male and female pelvises for all angles, with unambiguous data collected from the angle created by the ischiatic arch (ventral brim of the ischium). as a rule of thumb, males and females exhibit an ischiatic arch angle of <90° and >90°, respectively. two of the length measurements were also statistically different; however, overlap of these measurements would prevent their practical use. learning more about sexing techniques will increase our forensic, archeological, and anatomical knowledge of moose anatomy and benefit sex determination in the field when only headless, scavenged, or partial carcasses remain. alces vol. 48: 1-6 (2012) key words: alces alces, anatomy, moose, pelvis, isle royale, sex, tuberosity, ungulate, ischiatic arch. taber (1956) provides a method to determine sex of cervidae from skeletal remains excluding the skull for white-tailed (odocoileus virginianus) and black-tailed (o. hemionus columbianus) deer. remains of males (>2 years old) are distinguished from those of females by the presence of tuberosities where ligaments attach that support the penis. these tuberosities are small bony projections found on the caudal border of the ishium bone of the pelvis (fig. 1). tuberosities are not easily distinguishable on moose (alces alces) pelvises, even when a large number of pelvises are available for comparison. we found that college undergraduate students (n = 9) correctly identified the sex of moose from its pelvis in only 61% of trials (n = 15), despite training that included diagrams (taber 1956) and examining tuberosities on moose pelvises. this problem is frequently compounded when pelvises originate from wolf-killed or scavenged carcasses that may have significant chewing and/or other deterioration on the posterior end of the pelvis where tuberosities are located. other techniques attempted with white and black-tailed deer include an assessment of the general shape of the pubic symphysis. todd and todd (1938) described the sexual distinction as the male showing a symphyseal face bi-convex in outline and the female having an outline concave ventrally and convex dorsally. edelmann (1943) stated similarly: “on the pubic symphysis the pelvis of the 3colorado division of wildlife, 6060 broadway, denver, co 80216, email: jason.duetsch@state. co.us sexing moose by pelvis duetsch and peterson alces vol. 48, 2012 2 buck is thick and like a protuberance; that of the doe is thin, flat in front and slightly hollowed.” however, this technique requires that the pelvis in question be sawed in half along the pubic symphysis. in species with a somewhat similar skeletal structure to moose (e.g., horses [equus caballus]), detailed anatomical information exists which suggests that there are significant observable differences between male and female pelvises. dyce et al. (2002) describe how the horse pelvis inlet of females, when viewed from the front of the pelvis, is wide and rounded while that of the male is more angular and cramped, particularly ventrally. in both sexes, the outlet from the pelvic cavity is much smaller than the inlet. the difference is so noticeable in cattle and sheep that one can determine the sex of carcasses from which all other identifying organs or structures have been removed (bone 1988). however, a view of the pelvic inlet of moose shows no distinctive sex characteristics on an individual level and no published data exists for moose which states otherwise. for example, a sample may show adult male pelvises with larger pelvic inlets than females or vice versa. the same is true when observing the overall space (a characteristic not easily quantified in moose) of the canal formed by the 2 halves of the pelvis. the purpose of this study was to assess and develop a useful technique to identify sex of moose from pelvis characteristics. such a technique would have forensic, archeological, and anatomical applications. methods pelvises from moose of known sex were collected in 2003 from isle royale national park (48ºn, 89ºw) under permit from the united states national park service as part of a study of wolf-moose ecology (peterson 1977). isle royale includes 544 km² of natural habitat that supports moose at a density of 1-2/km². several linear angular (n = 5) and linear (n = 3) measurements were made (fig. 2). all angles had their vertex along the pubic symphysis. angle 1 is the ischiatic arch (ventral brim of the ischium), angle 2 is the inside of the lesser sciatic notch, and angle 3 is the fig. 1. pelvic girdle of the white-tailed deer, viewed from the rear, showing the suspensory tuberosities for the attachment of the penis ligaments in the male and their absence in the female (after taber 1956). alces vol. 48, 2012 duetsch and peterson sexing moose by pelvis 3 pelvic brim. angles 2 and 3 were measured twice, first with vector lengths of 53.98 mm (2.125 in) from the vertex to the first point of contact, then with vectors of 76.2 mm (3.0 in) from the vertex to the first point of contact. the length measurements evaluated were total length, length of pubic symphysis, and outside width at the lesser sciatic notch. our initial hypothesis was that significant differences between sexes would exist for all 8 measurements. pelvises from 17 male and 18 female moose >2 years old (confirmed by counting tooth cementum annuli) were used in this study. a classification system of 1-4 was used to record the condition of each of the 3 angles studied. class 1 status had no physical flaws which could impair the measurement of the given angle. class 2 status was not flawless but capable of a non-impaired measurement. class 3 status was very chewed and/or deteriorated to the point where angle measurement may be impaired. class 4 status was when measurement was not attainable for the given angle due to extreme chewing and/ or deterioration. angle 1 was measured with a viewing device we constructed that would standardize the perspective and distance of the observer. the device was clamped level onto the ventral side of the pubic symphysis. we then sighted perpendicular to the plane created by the pubic symphysis with the vertex at the intersection of the ischiatic arch and pubic symphysis (fig. 3). a 152.4 mm (6.0 in) diameter cutout center 180º protractor was used on this vertex, and the angle measured to the point of first contact with the ischiatic arch on either side of the vertex. angle 2 was measured with a 406.4 mm fig. 2. moose pelvis anatomy with lengths and angles 1, 2, and 3 from (a) ventral view and (b) dorsal view. sexing moose by pelvis duetsch and peterson alces vol. 48, 2012 4 (16.0 in) graduated shape-retaining flexcurve ruler. the ruler was pressed into the inside of the lesser sciatic notch of the pelvis, then removed for measurement. the shape obtained was traced onto paper and the vertex of the created hyperbola found. using the protractor, vectors of 53.98 mm (2.125 in) and 76.2 mm (3.0 in) were drawn from the vertex out to the point of first contact. the different side lengths for each angle were measured to help ensure angle differences between male and female pelvises were not missed or misrepresented due to selection of an arbitrary vector length (fig. 4). angle 3 was measured similarly as angle 2, the only difference being that the graduated flexcurve ruler was pressed into the pelvic brim located on the anterior portion of the pelvis. the 3 length measurements were 1) total length, 2) length of the pubic symphysis, and 3) outside width at the lesser sciatic notch. total length measurements were unattainable in most pelvises as a result of wolf chewing to the anterior and/or posterior portions of the pelvis. when possible, total length was measured from the furthest anterior edge of the iliac crest to the furthest posterior edge of the ischium. on all pelvises the length of the pubic symphysis and outside width of the lesser sciatic notch were measured with calipers (no chewing or deterioration was usually present in these areas). after measuring angles 1, 2, 3, and the 3 lengths, the sample was mixed with others and the process repeated again to evaluate measurement error. afterward, a random sub-sample was taken and measured by persons other than the primary researcher. these persons had no access to the sex or previous measurements of the moose pelvises. this was done to evaluate inter-observer measurement error. results the angle measuring procedure (including inter-observer measurements) proved to be repeatable to within ±5° on all angles with the majority being ±2°. all 5 normally distributed angular measurements were higher in females than males based on two-tailed ttests (p <0.05, table 1). however, only angle 1 allowed unambiguous identification of sex (fig. 5). two of the 3 linear measurements (total length and length of the pubic symphysis) were longer in males than females based on two-tailed t-tests (p <0.05, table 1). the width at the lesser sciatic notch was not different between sexes (male mean = 141.78 mm, female mean = 142.17 mm, p = 0.8065, table 1). no differences (p >0.05) were found between the 5 angular measurements by the primary researcher (observer 1) and all other observers (observer 2) on male and fig. 3. proper use of the angle 1 device on moose pelvis (lateral view). fig. 4. (a) measuring angle 2 on an isle royale moose pelvis using the graduated flexcure ruler formed to the inside of the lesser sciatic notch of the pelvis. (b) outlined sciatic notch with 53.98 mm (2.125 in) and 76.2 mm (3 in) sides drawn to form measurable angles. alces vol. 48, 2012 duetsch and peterson sexing moose by pelvis 5 female pelvises (n = 16, table 2). only the sub-sample of pelvises measured by persons (collectively termed observer 2) other than the primary researcher was compared against those measurements collected on the same pelvises by observer 1. discussion the angular data collected on pelvis morphology indicated that differences exist between the sexes (table 1). angle 1 was the best measurement because no overlap existed between the 15 male and 14 female pelvises sampled (fig. 5); there was 11° of separation between the largest male and smallest female measurement. although angles 2 and 3 were significantly different between sexes (table 1), both had overlapping data ranges. although total length and length of pubic symphysis were significantly different between males and females, there was enough overlap to create uncertainty (one-third overlapped in total length and two-thirds overlapped in length of pubic symphysis). total length was not measureable in most pelvises due to deterioration from feeding/ scavenging animals. overall, linear pelvic data were ineffective in distinguishing sex of a moose. measurement error was not problematic despite the complexity of the angular measurements. the angle of the ischiatic arch (angle 1) was the only measurement that definitively identified the sex of an unknown moose. as a rule of thumb, males and females exhibit an ischiatic arch angle of <90° and >90°, respectively. angle 1, the ischiatic arch (the most robust sexing technique of the study) was measured with a device that appears cumbersome and fig. 5. box plot data depicting the minimum, lower quartile, median, upper quartile, and maximum numerical values for angles 1, 2 (53.98 mm/76.2 mm) and 3 (53.98 mm/76.2 mm) in isle royale moose. the different vector lengths for angles 2 and 3 created 4 measurements each, and were measured to help ensure angle differences between male and female pelvises were not missed or misrepresented due to selection of an arbitrary length. male ± se female ± se p-value angle 1 67.63 ± 1.51 107.50 ± 1.49 <0.0001 angle 2 (53.98 mm) 114.62 ± 0.75 118.83 ± 0.78 0.0004 angle 2 (76.20 mm) 100.21 ± 0.65 106.19 ± 0.67 <0.0001 angle 3 (53.98 mm) 124.83 ± 1.51 111.92 ± 1.98 0.0026 angle 3 (76.20 mm) 97.60 ± 0.99 90.28 ± 1.11 0.0015 total length (mm) 473.88 ± 3.79 450.83 ± 5.62 0.0041 width at lesser sciatic notch (mm) 141.78 ± 1.10 142.17 ± 1.12 0.8065 length of pubic symphysis (mm) 159.38 ± 0.93 153.00 ± 1.01 0.0000 table 1. summary of means for 5 angular and 3 linear measurements of pelvis morphology for isle royale moose (17 m:18 f). p-value was determined from a t-test of null hypothesis that the sexes had the same pelvis morphology. sexing moose by pelvis duetsch and peterson alces vol. 48, 2012 6 difficult to use without proper instruction. we believe that easier and more accurate methods to measure the ischiatic arch angle are possible, and increased use of this measurement will yield such improvements. that said, with training the current measurement is applicable for field research because it is practical to employ in the field. identifying new (non-skull) sexing techniques of skeletal remains, specifically headless, scavenged, or partial carcasses, will benefit wildlife, forensic, archeological, and anatomical studies. we expect that these angular measurements are applicable to other moose populations, the length measurements less so. regardless, we have identified a simple pelvic measurement to distinguish between male and female moose that has a variety of applications in research, management, and forensics. acknowledgments the authors would like to thank isle royale national park (co-op agreement no. ca-6310-9-8001), the national science foundation (deb-9903671), and noe marymor (natural resources conservation service private lands biologist) for their help and support. references bone, j. f. 1988. animal anatomy and physiology. 3rd edition. prentice-hall, inc. englewood cliffs, new jesey, usa. dyce, k. m., w. o. sack, and c. j. g. wensing. 2002. textbook of veterinary anatomy. 3rd edition. saunders, inc. philadelphia, pennsylvania, usa. edelmann, r. h. 1943. textbook of meat hygiene. 8th revised edition. lea and febiger. philadelphia, pennsylvania, usa. peterson, r. o. 1977. wolf ecology and prey relationships on isle royale. scientific monographs series no. 11, u.s. national park service, washington, d. c., usa. taber, r. d. 1956. characteristics of the pelvic girdle in relation to sex in blacktailed and white-tailed deer. california fish and game 42: 15-21. todd, t. w., and a. w. todd. 1938. the epiphysial union pattern of the ungulates with a note on sirenia. american journal of anatomy. 63: 1-36. male means female means observer 1 observer 2 p-value observer 1 observer 2 p-value angle 1 72.25 70.83 0.7968 106.42 107.75 0.7943 angle 2 (53.98 mm) 115.08 116.17 0.4945 118.57 119.00 0.7882 angle 2 (76.20 mm) 100.70 101.80 0.5552 106.96 107.07 0.9734 angle 3 (53.98 mm) 126.33 130.00 0.4518 117.56 120.50 0.6667 angle 3 (76.20 mm) 98.65 100.20 0.6391 92.19 94.00 0.6553 table 2. summary mean data and their comparison by observer for 5 angular measurements of pelvis morphology in isle royale moose performed to test for inter-observer measurement error (see text). the p-value was determined from a t-test of the null hypothesis that there was no inter-observer measurement error or difference. alces34(1)_181.pdf alces34(1)_239.pdf alces29_55.pdf alces vol. 48, 2012 samuel et al. – deer keds on moose 27 review of deer ked (lipoptena cervi) on moose in scandinavia with implications for north america william m. samuel1, knut madslien2, and jessica gonynor-mcguire3 11226 kane wynd nw, edmonton, alberta, canada t6l 6x7, 2norwegian veterinary institute, pb 750 sentrum, n-0106 oslo, norway, 3warnell school of forestry and natural resources, and southeastern cooperative wildlife disease study, university of georgia, athens, georgia 30602, usa. abstract: the deer ked (lipoptena cervi) is an old world dipteran ectoparasite of moose (alces alces) and other cervidae. it has undergone significant expansion in distribution on moose of scandinavia in recent decades. this has been accompanied by much published research dealing with the range expansion and possible factors involved, problems for moose, exposure of northern populations of reindeer (rangifer rangifer tarandus), and public health issues. apparently, lipoptena cervi was introduced into northeastern united states in the late 1800s, presumably on an unknown species of european deer, and it soon spread to white-tailed deer (odocoileus virginianus). we review the current situation in scandinavia and north america and document the first record of l. cervi on moose in northeastern united states. alces vol. 48: 27-33 (2012) key words: alces alces, distribution, deer ked, effect on host, lipoptena cervi, moose, north america, odocoileus virginianus, scandinavia, white-tailed deer. the deer ked, lipoptena cervi (insecta, diptera, hippoboscidae) is a widely distributed, blood-sucking, reddish-brown, dorsoventrally flattened ectoparasite that occurs on old and new world members of the cervidae. notable hosts include red deer (cervus elaphus), roe deer (capreolus capreolus), fallow deer (dama dama), and especially moose (alces alces) in finland, sweden, and norway. lipoptena cervi was described by linnaeus (1758) as pediculus cervi, probably basing his description on examination of published figures, not specimens (see bequaert 1957, pp. 488-489). it is an ancient fly found associated with the remains of a late neolithic human mummy in an italian glacier (gothe and schöl 1994). lipoptena cervi has undergone rapid and ongoing west and northward expansion of its distribution in scandinavia in recent decades (välimäki et al. 2010). this has been accompanied by emerging public health and conservation issues, mostly in finland, including dermatitis on increasing numbers of rural people bitten by deer keds (härkönen et al. 2009), exposure of northern reindeer (rangifer rangifer tarandus) herds (kaitala et al. 2009), as well as an epizootic of hair-loss and deaths of moose in southeastern norway and mid-western sweden in 2006 and 2007 (madslien et al. 2011). the objectives of this paper are to briefly review current issues involving deer keds on moose in scandinavia and to provide information about its occurrence in northeastern united states. life cycle all species of deer keds, including l. cervi, are viviparous and produce one larva at a time. winged young adults of l. cervi emerge from pupae on the ground from late summer through autumn. they do not fly far in seeking hosts, which in scandinavia is usually moose. wings of males and females are lost once on moose. adults feed on blood and interstitial fluid, mating and overwintering on the host. deer keds on moose – samuel et al. alces vol. 48, 2012 28 the developing larva is retained in the uterus of the adult female. females extrude white mature larva, the prepupa, one at a time, from autumn until the next summer (härkönen et al. 2010). the skin of prepupa fattens and hardens into a seed-like dark case as the prepupa transforms to the pupa and drops to the ground on vegetation or snow. pupae remain on the ground until autumn (for details see bequaert 1953, haarløv 1964, and a life cycle diagram in samuel and madslien 2010). status in scandinavia expanding distribution and consequences for hosts lipoptena cervi has expanded its distribution westand northward in scandinavia in recent decades (see distribution maps in kaunisto et al. 2011 and välimäki et al. 2010). summarizing this expansion, välimäki et al. (2010) state that l. cervi “has been resident in sweden for more than two centuries, whereas in finland (~50 years) and norway (~30 years) it has established itself relatively recently.” there are two fronts of ked expansion: one that began in 1960 in southernmost finland (59-60o n) with source keds from russia, and one that began on or perhaps slightly before 1983 in southeastern norway (~59o n) with source keds from sweden. by the end of 2008, deer keds in norway had spread from the first known source south and east of oslo and the oslo fiord (~59 o n), west and north at ~7 km/yr. northward expansion along the border of sweden and norway reached ~62o n by end of 2008. northern expansion in sweden is less well documented, but has been relatively slow. in the early 1960s keds were as far north as 59o, ~420 km from the southern tip of sweden, and by 2008 keds were found as far north as ~62.5o. northern expansion in finland has been more rapid, spreading at a rate of 11 km/yr, and now is close to the arctic circle (~66o n) where it overlaps with the southern edge of reindeer herding areas (kaitala et al. 2009). thus, deer keds currently occupy a small part of norway near and north of oslo, the southern half of sweden, and approximately the southern 75% of finland. there has been progress in identifying factors involved in the expansion of l. cervi distribution. changes in moose numbers and climate are likely key factors. numbers of moose began to increase rapidly in norway, sweden, and finland in the 1960s, owing to increased clear-cutting practices by the forest industries and changes in management strategies to sex and age-specific harvests (lavsund et al. 2003). lavsund et al. (2003) documented that harvest of moose, which likely reflects population density, was highest in the early 1980s (sweden and finland), again in finland in the late 1990s and early 2000s (also see selby et al. 2005), and in the 1990s in norway (also see lykke 2005). in 2003, moose densities in all 3 countries “were lower in the north than in the south and higher in norway and sweden than in finland” (lavsund et al. 2003). the 2008 moose population in finland stabilized at 90,000 (pusenius et al. 2008, in kaitala et al. 2009). kaitala et al. (2009) suggested that increase in moose densities in finland has been the main reason why deer keds have been able to expand their range and increase in numbers. välimäki et al. (2010) point out that there was little or no range expansion of l. cervi in fennoscandia when densities of moose were relatively low. northern range expansion of l. cervi should be limited by colder temperatures and shorter growth season that would most affect off-host life stages, because it is the pupae on the ground and recently emerged young winged adults, not adult keds on moose, that are exposed to changing and potentially adverse northern environmental conditions (härkönen et al. 2010, välimäki et al. 2010). härkönen n et al. (2010) studied development of pupae and timing of adult emergence along a latitudinal gradient from boreal taiga in central finland to arctic tundra in northern alces vol. 48, 2012 samuel et al. – deer keds on moose 29 finland, and found success of pupae emerging to young adults was higher (19% emerged at the southernmost site at 62o n) and earlier in the south and much lower (<2% emerged at the high arctic site at 70o n) and later in the far north. however, a few pupae emerged in the high arctic, nearly 500 km north of its current northern range. thus, at 70o n, fewer young winged adults had less time before winter to find a host, but the results from 5 sites at different latitudes suggest that “spread of deer keds to cervids in the southern parts of the reindeer herding area seems inevitable.” this indicates that l. cervi has broad ecological tolerances (kaunisto et al. 2011) and suggests that the colder and shorter growing season in the north may slow, but not stop the northern spread of deer keds. moose are numerous in southern parts of finnish lapland (pusenius et al. 2008, in kaitala et al. 2009) and are considered the main host for deer keds (kaitala et al. 2009). thus, it is likely that reindeer will be attacked by adult flies; in addition, they are a suitable host. kynkäänniemi et al. (2010) experimentally infested reindeer with l. cervi, and a few keds survived and reproduced. kaunisto et al. (2009) found that deer keds occasionally infest semi-domesticated reindeer and wild finnish forest reindeer, r. t. fennicus. bequaert (1957) reported that reindeer introduced to scotland from lapland became “heavily infested with l. cervi” shortly after their arrival. keds can be numerous on moose and they might be associated with mortality of moose. madslien et al. (2011) found up to 16,500 keds on moose during an outbreak of ked-caused hairloss in moose in eastern norway and midwestern sweden, 2006-2007. these numbers were considered conservative because keds prefer hair-covered skin and many moose had severe alopecia. during the outbreak over 100 moose were observed; most with severe loss of hair and varying degrees of emaciation, and some with atypical behaviour (e.g., in farm buildings, unresponsive to humans). some were found dead and others in obvious severe distress were humanely killed. madslien et al. (2011) felt the main cause of the outbreak, high numbers of deer keds, was in response to extremely high summer and autumn temperatures in 2006, providing good development and survival of pupae. in finland, numbers of keds averaged 10,616, 3,549, and 1,730 on bulls, cows and calves, respectively, with full coats of hair (n = 23) (paakkonen et al. 2010). public health issues newly emerging young adult keds attack humans as well as ruminants, especially in finland. keds do not reproduce on humans, but infestation is a nuisance for rural people such as hunters, berry pickers, and forestry workers (härkönen et al. 2009, kortet et al. 2009). apparently it is not the ked bite that is the problem, but rather the inconvenience of removing “dozens of keds from hair and clothes.” more serious health issues have emerged and increasing numbers of people in finland suffer from chronic dermatitis and occupational allergic rhinoconjunctivitis following bites by deer keds. it is estimated that several thousand people have ked-caused dermatitis and growing numbers of forest workers have become sensitized to ked bite. a recent twist has been added to this subject. several of the many newly-described species in the bacterial genus bartonella cause diseases in humans (jacomo et al. 2002, chomel et al. 2009). bartonella henselae, for example, was identified in 1990 and is now known to cause several clinical diseases including cat scratch disease. it is transmitted from cats to humans by the bite of an infected cat flea or by a cat bite or scratch. dehio et al. (2004) found that l. cervi collected from roe and red deer in germany were infected with bartonella schoenbuchensi. the twist is that dehio et al. (2004) noticed that clinical signs and other aspects of cat scratch disease were similar to that of dermatitis in humans deer keds on moose – samuel et al. alces vol. 48, 2012 30 caused by the bite of l. cervi and suggested that bartonellae are “pathogens that should be considered possible etiological agents of deer ked dermatitis.” status in northeastern united states apparently in the late 1800s l. cervi was introduced to northeastern united states on an unknown deer transported from europe, and later established on white-tailed deer (bequaert 19531). summarizing records from bequaert (1937, 1942, and 1957), l. cervi was collected from white-tailed deer in new hampshire (grafton county in 1907, sullivan county in 1950) and a captive wapiti (cervus canadensis) in sullivan county in 1942. it was collected from white-tailed deer in new york (albany, cattaraugus, and hamilton counties in 1938, 1949 and 1954, respectively), massachusetts (dukes county in 1924), and pennsylvania (pike and clinton counties, no date; mckean and cameron counties in 1953). keds also were reported as being common on wapiti in cameron county, pennsylvania. other reports of l. cervi in the united states are from white-tailed deer in new york (bump 1941), west virginia (kellogg et al. 1971), and worcester county, massachusetts (matsumoto et al. 2008). given the above, l. cervi is probably more prevalent on white-tailed deer and moose of the northeastern united states than currently known or assumed. if l. cervi survived on wapiti sympatric with white-tailed deer in cen1in what we think is the only discussion of this introduction, bequaert (1953) states “this parasite was, in my opinion, brought from the old world by man with european deer. from this host it strayed to the native virginia deer, odocoileus virginianus, on which it now breeds at several localities in the northeastern united states......the exact date of the introduction is unknown; it was first recognized in the united states in 1907, when it was described as a new species by coquillet (l. subulata) [later changed to l. cervi]. if it were truly native, or autochthonous, it would be found over a much wider territory and be more abundant; while its presence on native deer could scarcely have escaped the early american entomologists.” tral pennsylvania (bequaert 1957), it should presumably occur on moose sympatric with white-tailed deer. exact numbers of keds on these hosts are unknown. anecdotally, bequaert (1957) mentioned a heavily infested white-tailed deer from sullivan county, new hampshire in 1950, and one of us (samuel)2 saw many keds on an old male white-tailed deer killed in mckean county, pennsylvania by his father around 1950 (samuel and madslien 2010). matsumoto et al. (2008) found only 6 keds on 4 of 27 white-tailed deer in worcester county, massachusetts. more recently, keds were collected opportunistically from several moose and whitetailed deer at hunter check stations by one of us (gonynor-mcguire) in 2007 and 2008. in october 2007, gonynor-mcguire thoroughly examined harvested moose for ticks and keds in new hampshire; examinations were 5-10 minutes in length. several dozen keds were collected and tentatively identified as l. cervi, but were subsequently lost. in october 2008 in cooperation with new hampshire fish and game department personnel (k. rines and k. gustafson), gonynor-mcguire supplied kits for collecting keds at harvest check stations. fourteen keds were collected from 11 white-tailed deer at stations in cheshire, coos, grafton, and sullivan counties, and 1 ked from each of 2 moose from stations in coos and grafton counties. keds from both species of host were identified by samuel using keys and descriptions of bequaert (1937, 1957), and were subsequently compared with l. cervi from moose in norway. the keds from the 2 moose from new hampshire were deposited in the university of alberta parasite collection; 2hunters in northeastern united states must occasionally see keds on their killed deer, maybe their moose. obviously my father did, but he thought they were ticks. hunters mistaking keds for ticks is mentioned on fact sheets for this parasite on the internet (e.g., http://ento.psu.edu/extension/factsheets/deer-keds). in 2007, many hunters at a new hampshire deer check station near berlin, nh said they thought keds were ticks (gonynor-mcguire, personal observation). alces vol. 48, 2012 samuel et al. – deer keds on moose 31 accession numbers uapc #11566 and #11572. several keds from each of 2 moose from the madslien et al. (2011) study in norway were also deposited: uapc #11563 and #11564. keds from 6 white-tailed deer were deposited in the uapc (#11565, #11567-11571), and from 4 white-tailed deer in the university of alberta e. h. strickland entomological museum (uasm213577-213580). this is the first report of this parasite from moose in north america. keds are observed annually on harvested moose and white-tailed deer in new hampshire, with keds more common in recent years, but prevalence and numbers have not been monitored (k. rines, pers. comm.). while studying winter ticks (dermacentor albipictus) at the university of new hampshire, d. bergeron (pers. comm.) observed small numbers of keds on a few of the many moose he surveyed at 3 regional harvest check stations in northern new hampshire, 2008-2010. surveys of moose and white-tailed deer for deer keds in various jurisdictions in the northeastern united states would be worthwhile, particularly if potential public health problems associated with bartonella species and deer keds become a reality. matsumoto et al. (2008) found dna of b. schoenbuchensis in 5 of 6 keds from 4 white-tailed deer at 3 different locations in worcester county, massachusetts3. if check stations for deer and moose are part of surveys, the following information might be relevant. madslien et al. (2011) examined fresh carcasses of moose in sweden and norway and found keds aggregated in neck, axillae, groin, and perineal regions. keds were located on the skin surface, their head oriented towards the skin. as the carcass cooled, they crawled away from the skin and onto protective hairs. samuel and trainer (1972) examined a standardized area (medial 3these authors raise the possibility that b. schoenbuchensis was possibly introduced to the northeast with the original importation of ked-infested cervid(s). surface of the hind leg and inguinal region) of white-tailed deer in southern texas for lipoptena mazamae. they found no significant migration of keds to or from the area during the 2 hours following death of deer. dead deer were hung by the back legs, which were spread using a gambrel. after 2 hours, keds began moving to extremities such as the nose, ears, and lower legs. given time constraints at hunter check stations, the technique used by sine et al. (2009) for detecting winter ticks on moose might be useful for detecting deer keds. references bequaert, j. 1937. notes on hippoboscidae. 5. the american species of lipoptena. bulletin of the brooklyn entomological society 32 (new series): 91-101. _____. 1942. a monograph of the melophaginae, or ked-flies, of sheep, goats, deer and antelopes (diptera, hippoboscidae). entomologica americana 22 (new series): 1-220. _____. 1953. the hippoboscidae or louse-flies (diptera) of mammals and birds. part i. structure, physiology and natural history. entomologica americana 36 (new series): 211-442. _____. 1957. the hippoboscidae or louseflies (diptera) of mammals and birds. part ii. taxonomy, evolution and revision of american genera and species. entomological americana 36 (new series): 417-611. bump, g. 1941. bureau of game. 30. annual report new york state conservation department (1940): 213-260. chomel, b. b., h-j. boulouis, e. b. breitschwerdt, r. w. kasten, m. vayssiertaussat, r. j. birtles, j. e. koehler, and c. dehio. 2009. veterinary research 40: 29. dehio, c., u. sauder, and r. hiestand. 2004. isolation of bartonella schoenbuchensis from lipoptena cervi, a blood-sucking deer keds on moose – samuel et al. alces vol. 48, 2012 32 arthropod causing deer ked dermatitis. journal of clinical microbiology 42: 5320-5323. gothe, r., and h. schöl. 1994. deer keds (lipoptena cervi) in the accompanying equipment of the late neolithic human mummy from the similaun, south tyrol. parasitology research 80: 81-83. haarløv, n. 1964. life cycle and distribution pattern of lipoptena cervi (l.) (dipt., hipposc.) on danish deer. oikos 15: 93-129. härkönen, l., s. härkönen, a. kaitala, s. kaunisto, r. kortet, s. laaksonen, and h. ylönen. 2010. predicting range expansion of an ectoparasite – the effect of spring and summer temperatures on deer ked lipoptena cervi (diptera: hippoboscidae) performance along a latitudinal gradient. ecography 33: 906-912. härkönen, s., m. laine, m. vornanen, and t. reunala. 2009. deer ked (lipoptena cervi) dermatitis in humans – an increasing nuisance in finland. alces 45: 73-79. jacomo, v., p. j. kelly, and d. raoult. 2002. natural history of bartonella infections (an exception to koch’s postulate). clinical and diagnostic laboratory immunology 9: 8-18. kaitala, a., r. kortet, s. härkönen, s. laaksonen, l. härkönen, s. kaunisto, and h. ylönen. 2009. deer ked, an ectoparasite of moose in finland: a brief review of its biology and invasion. alces 45: 85-88. kaunisto, s., l. härkönen, p. niemela, h. roininen, and h. ylönen. 2011. northward invasion of the parasitic deer ked (lipoptena cervi), is there geographical variation in pupal size and development duration. parasitology 138: 354-363. _____, r. kortet, l. härkönen, s. härkönen, h. ylönen, and s. laaksonen. 2009. new bedding site examination-based method to analyse deer ked (lipoptena cervi) infection in cervids. parasitology research 104: 919-925. kellogg, f., t. p. kistner, r. k. strickland, and r. g. gerrish. 1971. arthropod parasites collected from white-tailed deer. journal of medical entomology 8: 495-498. kortet, r., l. härkönen, p. hokkanen, s. härkönen, a. kaitala, s. kaunisto, s. laaksonen, j. kekäläinen, and h. ylönen. 2009. experiments on the ectoparasitic deer ked that often attacks humans; preferences for body parts, colour and temperature. bulletin of entomological research 100: 279-285. kynkäänniemi, s., r. kortet, l. härkönen, a. kaitala, t. paakkonen, a. mustonen, p. niemen, s. härkönen, h. ylönen, and s. laaksonen. 2010. threat of an invasive parasitic fly, the deer ked (lipoptena cervi), to the reindeer (rangifer tarandus tarandus): experimental infection and treatment. annales zoologici fennici 47: 28-36. lavsund, s., t. nygrén, and e. j. solberg. 2003. status of moose populations and challenges to moose management in fennoscandia. alces 39: 109-130. linnaeus, c. v. 1758. systema naturae, 10th edition. laurentius salvius stockholm, sweden. lykke, j. 2005. selective harvest management of a norwegian moose population. alces 41: 9-24. madslien, k., b. ytrehus, t. vikören, j. malmsten, k. isaksen, h. o. hygen, and e. j. solberg. 2011. hair-loss epizootic in moose (alces alces) associated with massive deer ked (lipoptena cervi) infestation. journal of wildlife diseases 47: 893-906. matsumoto, k., z. l. berrada, e. klinger, h. k. goethert, and s. r. telford, iii. 2008. molecular detection of bartonella schoenbuchensis from ectoparasites of deer in massachusetts. vector borne zoonotic diseases 8: 549-554. alces vol. 48, 2012 samuel et al. – deer keds on moose 33 paakkonen, t., a. m. mustonen, h. roininen, p. niemela, v. ruusila, and p. nieminen. 2010. parasitism of the deer ked, lipoptena cervi, on the moose, alces alces, in eastern finland. medical and veterinary entomology 24: 411-417. pusenius, j., m. pesonen, r. tykkyläinen, m. wallén, and a. huittinen. 2008. hirvikannan koko ja vasatuotto 2006. (moose population size and calf production in 2006). pages 7-14 in m. wikman, editor. riistakannat 2007: riistaseurantojen tulokset (game populations 2007): results of game monitoring). riista-ja kalatalouden tutkimuslaitoksen selvityksi 5/2008 (in finnish). samuel, b., and k. madslien. 2010. are louse flies potential problems for moose and humans of maritimes canada and northeastern united states? the moose call 25 (june): 1, 3-5. samuel, w. m., and d. o. trainer. 1972. lipoptena mazamae rondani, 1878 (diptera: hippoboscidae) on white-tailed deer in southern texas. journal of medical entomology 90: 104-106. selby, a., l. petäjistö, and t. koskela. 2005. threats to the sustainability of moose management in finland. alces 41: 63-74. sine, m., k. morris, and d. knupp. 2009. assessment of a line transect field method to determine winter tick abundance on moose. alces 45: 143-146. välimäki, p., k. madslien, j. malmasten, l. härkönen, s. härkönen, a. kaitala, r. kortet, s. laaksonen, r. mehl, l. redford, h. ylönen, and b. ytrehus. 2010. fennoscandian distribution of an important parasite of cervids, the deer ked (lipoptena cervi), revisited. parasitology research 107: 117-125. alces35_73.pdf alces32_51.pdf alces29_27.pdf alces34(1)_41.pdf gasaway.pdf alces vol. 45, 2009 minoranskiy et al. – moose of the rostov region 21 history and status of moose in the rostov region, russia viktor a. minoranskiy1, viktor v. sidelnikov2, and elena i. simonovich3 1department of zoology, southern federal university; 2center of hygiene and epidemiology, federal state of health care; 3scientific research institute of biology, southern federal university. rostovon-don, taganrog, russia. abstract: moose (alces alces) disappeared from the rostov region in the 19th century due to agricultural development, hunting, and deforestation. they reappeared in the second half of the 20th century due to broad conservation measures including intensive forest management, and by the 1970s numbered >1000 and were found throughout the region. although hunting was regulated, the population became stagnant in the 1980s presumably from trophy hunting that skewed the sex and age structure, as well as measurable wolf (canis lupus) predation. political reform in the 1990s further caused population decline due to increased and less regulated hunting, increased poaching without punishment, reduced predator control, decline in forest management, and large forest fires. currently the population is at a 50-year low and occupies 1/3 of its range in the 1980s. moose are no longer considered a commercial species, rather a species of concern. alces vol. 45: 21-24 (2009) key words: alces alces, management, moose, population dynamics, population recovery, predation, social impacts. historically, moose (alces alces) inhabited the flood plain forests and small forested steppe ravines in the territory of the modern rostov region. moose occupied this region in the 17th and 18th centuries into the early 19th century; in the 1660s moose were so common in the don river region that hides were a major export product to the muscovy (i.e., moscow state; kirikov 1959). moose gradually disappeared from the lower don area in the 19th century due to intensive agricultural development, hunting, deforestation, and increasing populations of boar (sus scrofa l.), red deer (cervus elaphus l.), and roe deer (capreolus capreolus l.). moose reappeared in the rostov region in the second half of the 20th century, in large part, to broad conservation measures. since the middle of the 20th century the steppes experienced extensive reforestation, largely due to establishment of pine (pinus spp.) plantations. for example, the area of state forest nearly doubled to >1.8 million ha from 1947 to 1975, with the peak activity in the rostov region during the 1970s. pine plantations provide basic moose forage for about 10 years. further, in 1967-2005, >2.2 million ha of forest were protected including >1.2 million ha of forest shelter belts. this region currently consists of a dense, connected network of forests and forest belts that provide moose optimal shelter, forage, mobility, and population distribution. the moose population also responded positively to concurrent, intensive agricultural activity in the region during the 1970s. the primary horticultural emphasis was on apple (malus spp.) production that also provided moose an additional forage resource. moose gradually expanded southward inhabiting riparian habitat of the don river and steppes, and utilizing new forages including sunflower (helianthus annuus) tops and heads of reed (phragmites australis). during and after their spring, summer, and autumn dispersal, moose occupy a wide range of forest habitat and agricultural areas, particularly farms with tall-stalked crops (i.e., sunflower and corn). moose of the rostov region – minoranskiy et al. alces vol. 45, 2009 22 another effective form of conservation was the maintenance of strictly protected natural areas, including state game parks that preclude hunting of rare and endangered species. since the 1960s the number of state game parks has increased annually with the main objective of protecting wild ungulates. by 1982 there were 21 regional and 1 national wildlife reserves in the don river area; the rostov state forest and hunting facility is being established currently. additionally, in the 1950-80s many measures were implemented to improve the distribution and size of game animal populations including harvest strategies and regulations, establishment of various hunting facilities, and improved protection, forage/forest management, and biotechnical methods and strategies. moose populations were also enhanced through predation control. wolves (canis lupus) have been killed regularly throughout the region for a number of reasons including compensation to the gosstrakh, the dynasties of wolf-hunters living in the region (fig. 1). wolf reduction via helicopter gunning was arranged with aerotaxation on helicopters for the direct purpose of exterminating wolves. the wolf population is minimal because about 70-80% of the professional wolf kill occurs at dens. an effective regulatory system for controlling moose hunting is the glavokhota – rsfsr, legislation that provides criminal liability for shooting of moose. further, personal access to rifles/weapons was minimal as only certain people were allowed such weapons in their possession. widespread educational efforts in schools and universities in the 1970-80s focused on conservation and ecological concepts to promote such appreciation in society. thus, a multitude of factors had positive influence on the appearance, distribution, and increase of moose in the don river region. natural recruitment of moose in the rostov region began in the late 1960s although the first report of moose was recorded in 1950 in the veshenskiy area of the voronezh region (fertikov 1975). afterward moose regularly entered the don river area from ukraine through forests alongside the seversky donets. in 1966, 423 moose were counted in 15 of the 37 regional districts; by 1969 a separate population had inhabited the flood-plain forests of the don and seversky donets. by 1970, >1000 moose inhabited 22 districts, this population expanded to about 1300 in 1972 (table 1, fig. 1), and buy the end of the decade moose were throughout the region. moose were seen in the adjacent forest at the outskirts of rostov, and at the popular city beach. some crossed the banks of manych-gudilo lake, walking for hundreds of kilometers through the open steppe, rarely entering forest belts. the dramatic expansion in both number and range of the moose population was evidence that favorable environmental conditions existed in the region. reproductive analyses in the 1970s confirmed such as 80% of adult figure 1. the change and relationship between the moose population and wolf harvest in the rostov region, 1964-2007; data from the department of okhotnadzor of the federal service of rosselkhoznadzor, rostov region. moose population wolf harvest alces vol. 45, 2009 minoranskiy et al. – moose of the rostov region 23 cows had 2 embryos, and occasionally 3. however, despite intensive population control of wolves, predation loss was estimated as about 100 moose. and, by the 1980s, the population was obviously stagnant and a myriad of influencing factors became starkly evident. the incentive system to harvest and deliver meat to the state encouraged shooting of the largest, most productive individuals. trophy hunting was in favor, and adult bulls with well-developed antlers were desired and shot preferentially. the population changed radically from a well-balanced sex and age structure to one predominated by young males and cows, and productivity declined. the regional population declined rapidly and migrations ended from the volgograd and voronezh regions (table 1). the country-wide crisis in the 1990s had much negative influence on the regional moose population and its management. it caused impoverishment of the human population, reform in the state management system of hunting, cancellation of public hunting inspections, and removal of legal, economical, and other regulatory mechanisms to maintain and manage state-associated hunting facilities. the personnel situation at regional hunting facilities was greatly affected because many experts left and replacements were often without adequate education and experience. hunters appeared to have easy access to automatic rifles of the scs type (self-loading carbine of simonov) with shell types typically used in the armed forces. these weapons decade population (min-max) average population/yr year of maximum 1964-1969 300-979 597 1969 1970-1979 1075-1540 1301 1977 1980-1989 523-907 692 1980 1990-1999 209-925 439 1990 2000-2008 166-234 232 2005 table 1. chronological change of the moose population in the rostov region, russia, 1964-2008. data are from the department of okhotnadzor of federal service rosselkhoznadzor in the rostov region. fig. 2. change in the distribution of moose in the rostov region, 1980-1990s to 2007. moose of the rostov region – minoranskiy et al. alces vol. 45, 2009 24 presumably caused excessive wounding and low recovery rates due to rapid-fire shooting across long distances with low-power bullets. further, the cancellation of criminal liability for shooting moose that probably intensified poaching, and the termination of subsidies for shooting wolves that increased the wolf population and predation of moose (fig. 1), also combined to reduce the moose population. subsequent reformation of the forestry system that terminated reforestation activity caused sharp decline in the amount of young pine suitable as moose forage as older plantations matured and no longer provided adequate winter forage. a large number of forest fires occurred, and the loss of pines was up to 1000 ha annually in the north. all of the above had great influence on the moose population, and in combination resulted in rapid and tremendous decline of the population in the 1990s. in the past 10-12 years the population has stabilized to a level about 25% of that in 1990 (fig. 1), and its geographical range (8 districts) is about half that in the 1980-90s (20 districts) (fig. 2). as a result, moose in the rostov region no longer have commercial value and are considered a species of concern and are subject to strict, protective management. references fertikov v. i. 1975. restoration of the geographical ranges, acclimatization of wild hoofed animals and pheasants in the rostov region. thesis abstract. rostovon-don university, rostov-on-don, russia. (in russian). kirikov s. v. 1959. changes in the animal world in the native zones of the ussr: steppe zone and forest-steppe. academy of sciences of ussr, moscow, russia. (in russian). alces35_143.pdf alces34(2)_435.pdf alces34(2)_319.pdf alces vol. 47, 2010 moen et al. moose home range 101 using cover type composition of home ranges and vhf telemetry locations of moose to interpret aerial survey results in minnesota ron moen1, michael e. nelson2, and andy edwards3 1natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811-1442; 2u.s. geological survey, northern prairie wildlife research center, 8711– 37th street se, jamestown, north dakota 58401 (retired); 31854 treaty authority, 4428 haines road, duluth, minnesota 55811 abstract: although home ranges of radio-collared moose are typically used to establish habitat requirements and range size of moose, they can be useful in the implementation of aerial surveys. a survey area is usually stratified into low, medium, and high moose density blocks, and radio-collared moose can provide data to improve the stratification procedure because cover type composition in home ranges could help stratify survey blocks. vhf telemetry locations and home range data can also be used to evaluate survey results. in minnesota high moose density survey blocks contained more of the conifer forest cover type and less of the wet bog cover type than was present in moose home ranges or vhf telemetry locations. proportionately more moose were observed in the mixed forest and regenerating forest cover types during the aerial survey, even though vhf telemetry locations indicated moose were using the wet bog cover type. the survey will be biased and underestimate the moose population if undetected moose are not corrected for by a sightability correction factor. further evaluation of survey data and increased resolution of moose locations is required to resolve this issue. alces vol. 47: 101-112 (2011) key words: alces, home range, cover type, survey, stratification. home range size of moose reported in the literature from vhf radio-collars varies from 4-> 250 km2 (hundertmark 1997); seasonal home ranges are usually 10-20 km2. the cover type composition of home ranges can be calculated based on a gis layer derived from classified satellite imagery or aerial photograph interpretation. home range size and cover type composition could be used when designing or evaluating aerial surveys. for example, stratification could be based in part on how much of a preferred cover type is in a survey block. annual and seasonal home ranges have historically been calculated using the minimum convex polygon (mcp) method (mohr 1947). more recently, kernel estimators (worton 1989, laver and kelly 2008) are the preferred method to calculate home range, especially when gps locations are available. most radio-telemetry research with moose was done before the advent of gps collars, and telemetry studies focusing on survival still commonly use vhf collars. therefore, the mcp remains a common denominator for comparison of home ranges across space and time because of its long history of use. most moose aerial surveys are done using a stratified random block (srb) technique (gasaway et al. 1986). pre-survey stratification flights are part of the protocol, but not always done because of the cost and timing of such flights. most agencies have background knowledge of moose distribution from prior aerial surveys that can be used to stratify blocks into low, medium, or high density. the initial stratification is typically based upon knowledge of moose distribution and previous surveys. in minnesota, survey blocks are restratified annually if an unexpected number of moose home range moen et al. alces vol. 47, 2011 102 moose are observed in specific blocks during the previous year’s survey (lenarz 1998). variance in population estimates of moose generated from aerial surveys is usually so high that a >20% change in the population estimate is required to produce a statistically significant change between years (gasaway et al. 1986). it follows that attempts to reduce sources of variance would be beneficial for moose management. low precision in survey results is often caused by incorrect stratification (lenarz 1998); for example, variance increases if many moose are observed in a low density survey block during the survey. sightability is an additional factor affecting moose population estimates. moose are less visible in thicker cover types such as conifers. logistic regression sightability models originally developed in western states correct for the reduction in sightability (e.g., anderson and lindzey 1996, unsworth et al. 1998, quayle et al. 2001). if the difference in cover type composition among strata is assessed, it could theoretically correct for sightability. one a priori approach when habitat use by moose is known is to use cover type in the stratification procedure. for example, in british columbia existing satellite imagery and past research on moose behavior was used to stratify blocks as high or low density (heard et al. 2008). block shape and size was variable and blocks could be stored in gps units in the helicopter. this was possible because the survey area included cover types identifiable by satellite imagery that were rarely used by moose in winter. an alternative a posteriori approach to evaluate stratification is to use vhf telemetry locations and home range to interpret survey results. logical predictions are that home ranges would contain proportionately more preferred cover types, moose would be more frequently located in preferred cover types, and survey blocks stratified as high density would contain more area of preferred cover types. this should be especially true in low density moose populations because moose should not have to use marginal habitats. we used home range and satellite imagery to interpret aerial survey results in minnesota. we first calculated home range size and cover type composition for moose wearing vhf radio-collars in minnesota. next, we compared habitat use by males and females, and contrasted cover type for the different home range calculation methods. finally, we tested whether cover types of home ranges corresponded to the cover type composition in low, medium, and high-density survey blocks, and whether cover types of moose seen during the survey were consistent with vhf telemetry locations. study area our study area was in northeastern minnesota where moose are currently surveyed (fig. 1). forests in northern minnesota are transitional between canadian boreal forests and northern hardwood forests to the south (pastor and mladenoff 1992). historic and n 40 0 40 80 120 kilometers fig. 1. study area in northeastern minnesota in which moose are currently surveyed (lenarz 2010). each block is ~35 km2 and about 10% of blocks are flown in each annual survey. shading represents low (white), medium (gray), and high (green) stratification levels. the outline of the composite 95% kernel home ranges of radio-collared moose is shown with a heavy black line. alces vol. 47, 2010 moen et al. moose home range 103 recent land use has reduced the proportion of upland conifers (white spruce [picea glauca] and white pine [pinus strobus]) in northern minnesota forests (frelich 2002, wolter and white 2002). northern minnesota has a continental climate with moderate precipitation, short warm summers, and severe winters. snow cover is usually present from december-march. land ownership in the study area is mostly public within the superior national forest and the boundary waters canoe area and wilderness; state, county, and tribal lands are also part of the landscape, and blocks of industrial forest land exist outside of the superior national forest. methods aerial survey moose in northeastern minnesota are surveyed annually by the minnesota department of natural resources (mdnr) and tribal biologists (edwards et al. 2004); here we evaluate survey results from 2004-2011. an aerial survey is used following the gasaway technique with survey blocks stratified as low, medium, or high density (lenarz 2011); presurvey flights to stratify survey blocks were not done. stratification into low, medium, and high density blocks is based on knowledge of past moose density in the area and collective knowledge of local managers about current moose numbers and recent habitat changes; classification is based on the expected number of moose to be observed in each block if surveyed within a 5 year period. search effort in low, medium, and high density blocks is 1.4 ± 0.3, 1.6 ± 0.3, and 1.7 ± 0.3 min/km2 (mean ± sd), respectively (m. lenarz, mdnr, pers. comm.). the helicopter survey is flown annually in early january and includes 40 plots with at least 6 high density; the remaining plots are allocated based on variances from the previous year’s survey (m. lenarz, pers. comm.). moose observed within a survey block are identified to sex and age class, and a gps waypoint is taken directly over each observation point. in addition, a visual obstruction covariate (voc) estimate is made relative to how much vegetation would prevent observation of an adjacent moose (lenarz 2006, 2011). the voc is similar to the vegetation cover class described in a moose sightability model developed in wyoming (anderson and lindzey 1996). one exception was that rather than use the criteria of a 3-m perimeter, observers were instructed to estimate percent visual obstruction within 2 moose body lengths of where a moose is first observed; both approaches evaluate the relatively same size area. animal capture and radio-collars in 2002-2005 moose were captured and fitted with vhf radio-collars (advanced telemetry systems, inc., isanti, minnesota) in february or march. moose were netgunned or darted from a helicopter (lenarz et al. 2009, 2010); animal capture and handling procedures met guidelines recommended by the american society of mammalogists (gannon and sikes 2009). we divided the vhf telemetry locations into periods relevant to the survey: all year, winter (1 december-30 april), and near the aerial survey dates (1 december-31 january). radio-collared moose were monitored weekly from a cessna 185 aircraft by a single observer with the same pilot on each flight. location was recorded on a gps unit carried by the observer or by the pilot in the aircraft. mean location error was 300 m in a blind test (m. lenarz, unpubl. data). we tested the effect of location error with simulated data sets. we assumed no directional bias in location error and estimated parameters of the error distance distribution using maximum likelihood with built in non-linear optimizers in program r (r development core team 2006). we used 30 replicates of location error for each vhf location (j. fieberg, mdnr, unpubl. data) to moose home range moen et al. alces vol. 47, 2011 104 test for the effect of location error on cover type composition. home range was calculated for all moose with >15 locations per year with at least 285 days between the first and last locations. when locations were available for >1 year from the same animal, we calculated a second or third home range for the same individual. we did not use 14 home ranges that had disjunct areas because we wanted to maintain a 1:1 relationship between minimum convex polygon (mcp) home ranges and fixed kernel-based home ranges. we calculated home range in arcview 3.3 with the animal movement analyst extension (hooge and eichenlaub 2000). we compared the size of male and female mcp, 50% kernel, and 95% kernel home ranges with an unpaired t-test. we first tested for homogeneity of variance using a folded f-test; we used a t-test if variances were not different between groups and a satterthwaite t-test if variances were different between groups. cover type analysis we used the land use land cover (lulc) raster data set that was based on source imagery from june 1994 with an overall classification accuracy of >95% (mdnr 2007). only 6 terrestrial cover types comprise >90% of the area where moose occur in northeastern minnesota. the mixed forest type (~50% of the area) has a mature canopy that includes aspen (populus tremuloides), paper birch (betula papyrifera), white spruce, and balsam fir (abies balsamea). the conifer forest type has at least 67% conifer species in the canopy and is primarily upland conifers. the deciduous forest type has at least 67% deciduous species including red maple (acer rubrum) and oaks (quercus spp.) in the canopy and is primarily in upland areas above the lake superior shoreline. the wet bog type has black spruce (p. mariana) or tamarack (larix laracina), although trees may be at low density in this cover type. the regenerating forest type identifies disturbances occurring in 1973-1994. managed forests often have regenerating aspen and red (p. resinosa) or jack pine (p. banksiana) plantations. the marsh and fen type has small marshes and fens and comprises a relatively small portion of the area. water covers about 10% of the area. other cover types included shrubby grassland, human developments, and gravel pits that cumulatively represented <5% of land area and were not analyzed. we used arcview to determine the lulc type for vhf telemetry locations, simulated location error of vhf telemetry locations, home range estimators (mcp, 50% kernel and 95% kernel), and the 35 km2 survey blocks used in the aerial survey. cover types of actual locations and simulated error locations were compared using each moose as the experimental unit. we calculated cover type composition of the 50% kernel, the 95% kernel, and the mcp home ranges for each moose. we also calculated cover type composition of the composite of the 95% kernel home ranges buffered by 10 km, and the cover type composition of the area surveyed annually for moose (lenarz 2011). we compared cover type use by cow and bull moose in winter using home ranges, vhf telemetry locations, and moose locations from the aerial survey. for the home ranges (mcp, 95% kernel, and 50% kernel) we used a t-test for cover type composition differences between cow and bull home ranges. to test for differential habitat use in winter, we used a proportion test with h0 being that cow and bull moose did not use habitat differently for the vhf telemetry locations. we also used anova to test for a difference in cover type composition between home range estimators and point locations. we tested for the effect of error in position locations on cover type composition by comparing the mean percent composition of all simulated error locations to the cover type compositions of the vhf telemetry location. we also compared the cover type composialces vol. 47, 2010 moen et al. moose home range 105 tion of the vhf telemetry locations against the cover type composition of 5 replicates of the position error data set, using the percent of area in each cover type as the test variable, and locations from a single moose as the experimental unit. we used several comparisons to determine the relationship between moose seen during the aerial survey and moose located via vhf telemetry. first, we compared cover type composition of the low, medium, and high density survey blocks to the cover type composition of vhf telemetry locations from december and january. next, we compared the cover type composition of locations of moose taken during the survey to the cover type composition of the vhf telemetry locations from december and january using a χ2 test. finally, we tested whether the voc (lenarz 2011) and group size of moose were affected by cover type in the survey using anova. we used statistix (v. 4.1, boca raton, florida) for all statistical tests. significance level for all tests was set at p = 0.05. gis analysis was done with either arcview 3.3 or arcmap 9.1 (esri, redlands, california). results we calculated annual home ranges of 84 cows and 47 bulls (table 1). bull home ranges were larger than cow home ranges (satterthwaite t-test, mcp: t54.8 = 2.22, p = 0.03; 95% kernel: t64.4 = 2.87, p = 0.01; 50% kernel: t67.0 = 2.50, p = 0.01); an unpaired ttest was also significant for each home range type (t115 > 2.57, p <0.001). the satterthwaite t-test was preferred because the folded f test for homogeneity of variance indicated there were differences in variance between bulls and cows (f39,76 = 2.64, 1.50, and 1.48 for mcp, 95% kernel, and 50% kernel, with p-values of 0.001, 0.07, and 0.07, respectively). because bull home ranges were larger than cow home ranges, we compared the percent of each cover type within their home ranges. cover type composition was not different between bulls and cows in 23 of 24 possible comparisons of mcp, 50% kernel, 95% kernel, and point locations (t-test (|t115| < 1.03, p >0.31 for equal variance or |t>52| <1.14, p >0.36 for unequal variance). the exception was that cows had more marsh and fen type in their home range than bulls (t-test, |t113.4| = 2.50, p = 0.01 for unequal variance); this cover type comprised only 3% of the landscape and contained ≤3% of moose locations. we grouped bull and cow data together since cover type composition of locations and home range estimates were not different between sexes. the percent area in each cover type was not different whether annual mcp, annual 95% kernel home range, annual 50% kernel home range, or all vhf telemetry locations were used to estimate cover type use (anova, f3,463 <1.86, p >0.14) (table 2). we tested sensitivity of cover type classification of each moose location to error in estimating true position. the mean percent of each cover type from simulated location errors was within a 95% confidence interval for vhf locations of both cows and bulls, except for the marsh and fen type (table 3). when we compared cover type composition using individual animal as the experimental unit, there were no differences in cover type mean ± sem median min max mcp females 28 ± 3 19 2 141 males 45 ± 6 30 10 243 95% females 40 ± 3 31 4 163 kernel males 61 ± 5 46 19 158 50% females 7 ± 1 4 1 38 kernel males 10 ± 1 7 2 35 table 1. annual home range size (km2, mean ± se) of adult male and female moose in northeastern minnesota. kernel home ranges and mcp calculated with animal movement analyst (hooge & eichenlaub 2000). number of locations was 25 ± 1 for females and 24 ± 1 for males. duration of time period was 342 ± 2 days for females and 350 ± 3 days for males. moose home range moen et al. alces vol. 47, 2011 106 composition in 5 different replications of the error data set compared to the actual locations in any comparison (t-test (|t162| <1.52, p >0.17 for equal variance or |t>141| <1.52, p >0.17 for unequal variance). survey stratification and home range moose did not use cover types in proportion to availability whether the comparison was based on mcp, 95% kernel home range, 50% kernel home range, or vhf telemetry locations. moose used the wet bog, conifer forest, and regenerating forest types proportionally more than their respective availability. in wet bog, conifer forest, and regenerating forest types, the percent of vhf locations and all home range estimators was more than twice the sem calculated for the respective aerial survey (fig. 2). for the survey blocks, stratification led to clear trends in some cover types. proportions of conifer forest and regenerating forest increased as stratification increased from low to high moose density survey blocks (fig. 3). in contrast, the deciduous forest and mixed forest types declined as stratification increased from low to high moose density. the wet bog cover type increased from low to medium density survey blocks, and then decreased. cover type composition in high moose density survey blocks was somewhat similar to the cover types of moose vhf locations in december and january, but there were differences (fig. 3). of most biological importance were the differences between the wet bog and conifer forest types. moose had more vhf telemetry locations in the wet bog type, while the high density survey blocks had more area in the conifer forest type (fig. 3). the regenerating forest and the mixed forest types were also used by moose relatively more in the high density survey blocks compared to availability. cover types that moose were observed in during the survey were different from cover types of vhf telemetry locations in december and january (fig. 4). moose were less likely to be seen during the aerial survey in conifer forest and wet bogs than they should have been according to cover type use from vhf telemetry locations across the survey area (χ5 2 = 34, p <0.0001) or the cover types of vhf telemetry locations within the 95% kernel home range area (χ5 2 = 40, p <0.0001). instead, both bulls (χ5 2 = 45, p <0.0001) and cows (χ5 2 = 88, p <0.0001) were observed more often in mixed forest during the aerial survey than indicated from vhf telemetry locations. cover type composition of locations of cows without calves and cows with calves was not different (χ5 2 = 3.2, p = 0.66). the differences in cover types between survey locations and vhf telemetry locations were also present within the medium and high density survey blocks separately (table 4). in the medium and high density survey blocks (χ5 2 = 68, p < 0.001, and χ5 2 = 40, p < 0.001, respectively) there were more moose observed in the mixed forest and regenerating forest types during the aerial survey. fewer moose were seen in the wet bog and conifer forest table 2. results of binomial probability tests comparing cover type composition of vhf telemetry locations of cows and bulls in winter; cows n = 1,457 for cows and n = 606 for bulls. cover type cow (%) bull (%) cow:bull ratio z p regeneration/young forests 15.9 16.7 0.95 -0.39 0.70 mixedwood forest 42.8 43.6 0.98 -0.29 0.77 deciduous forest 1.9 1.3 1.46 0.77 0.44 coniferous forest 20.3 17.2 1.18 1.59 0.11 wet bog 15.7 19.5 0.81 -2.01 0.04 marsh and fens 2.1 1.3 1.61 1.05 0.29 alces vol. 47, 2010 moen et al. moose home range 107 cover types, yet more moose were present in those cover types based on vhf telemetry locations. using data from 2004-2011, the voc varied among cover types (anova, f5,1343 = 2.87, p = 0.01) and among years (anova, f6,1343 = 6.54, p <0.001). however, the only difference in cover types was that voc in regenerating forest was lower than voc in mixed forest; all other pairwise comparisons were not different. similarly, there was only one difference among years; 2011 had a lower voc than all years except 2007. the overall mean voc was 38%, with a range of 0-90% for most cover types. group size is another possible confounding factor in analysis of cover type use; however, group size did not vary among cover types (anova, f5,1594 = 1.27, p = 0.27), although it varied among years (anova, f7,1594 = 4.22, p <0.001). no increasing or decreasing trend was evident in group size across years. overall mean group size during the aerial survey was 1.97 ± 1.09 (sd) moose. discussion cover type composition of seasonal home ranges or vhf telemetry locations did not correspond with survey block stratification. one reason for this may be the different decision rules under which moose behave and manag 0 5 10 15 20 25 30 35 40 45 50 deciduous forest mixed forest marsh and fen wet bog conifer forest regenerating aerial survey area mcp home range 95% kernel 50% kernel vhf locations cover type % o f a re a or % o f lo ca tio ns fig. 2. area in terrestrial cover types on northeast moose range in minnesota that is covered in the aerial survey compared to area of minimum convex polygon, 95% kernel, and 50% kernel home ranges of vhf radio-collared moose. the cover type of vhf telemetry locations from which home range was calculated is shown. the se is for percent of area in each cover type or for the number of vhf locations by moose. the se on the aerial survey area is across the 453 possible survey blocks. vhf locations error locations cover type cows bulls mean ± 95%ci regeneration / young forests 16.0 ± 1.4 15.4 ± 1.8 15.0 ± 0.4 mixedwood forest 41.1 ± 1.9 41.1 ± 2.4 41.8 ± 0.5 deciduous forest 2.3 ± 0.5 2.8 ± 1.1 1.7 ± 0.1 conifer forest 17.7 ± 1.5 19.5 ± 1.8 19.6 ± 0.3 wet bog 17.7 ± 1.7 19.6 ± 2.2 17.5 ± 0.2 marsh and fens 3.2 ± 0.5 1.5 ± 0.4 2.4 ± 0.1 table 3. cover type of simulated locations with position error (n = 30) compared to mean percent of vhf telemetry locations in each cover type for cows and bulls. the se for vhf locations is based on individual animals (84 cows and 57 bulls); for the simulated error data set se is across the 30 replicates. moose home range moen et al. alces vol. 47, 2011 108 ers operate. home ranges were smaller than survey blocks, and moose are not restricted to rectangular home ranges like the survey blocks. at the spatial scale of home range, a different mix of cover types can be used by moose than is possible within a survey block that could contain areas (habitats) with very low or high moose density. more precision and reduced costs were achieved in british columbia when fixed rectangle survey blocks were modified to match habitat features (heard et al. 2008). the most striking difference between the aerial survey and the vhf telemetry locations was the distribution of moose among cover types in january when the aerial survey occurs. high density survey blocks contained 4x more area in the conifer forest type than in the wet bog type, yet both cover types had the same number of vhf telemetry locations in december and january. furthermore, relatively few moose were observed in the wet bog and conifer forest types during the aerial survey, with disproportionately more observations in the mixed forest type. the contrast between the relative use of 0 10 20 30 40 50 60 deciduous forest mixed forest marsh and fen wet bog conifer forest regenerating low density block medium density block high density block vhf locations in december/january cover type % o f a re a or % o f lo ca tio ns fig. 3. cover types in low, medium, and high density survey blocks compared to the cover types of moose locations obtained in december and january. standard errors on survey blocks are based on cover types across the aerial survey, while standard errors on vhf locations are based on individual moose. stratification: low density medium density high density cover type survey vhf available survey vhf available survey vhf available regeneration / young forest 12 17 7 13 18 10 24 15 10 mixedwood forest 58 43 45 64 47 41 55 42 39 deciduous forest 3 6 9 3 1 4 5 0 4 conifer forest 10 17 15 10 19 20 11 23 29 wet bog 15 18 12 6 14 15 3 17 7 marsh and fen 0 0 3 1 1 3 2 3 4 table 4. comparison of cover type percentage where moose were observed during the aerial survey, moose located by vhf telemetry in december and january, and the area available in each cover type across the survey area for low, medium, and high density survey blocks. survey locations are from 2004-2011 and vhf telemetry data is from 2002-2005. alces vol. 47, 2010 moen et al. moose home range 109 cover types identified in the aerial survey and vhf telemetry has important implications for survey design. stratification of survey blocks is intended to generally reflect moose density (lenarz 1998) and account for sightability; however, moose are difficult to observe in conifer cover during aerial surveys (karns 1982, peterson and page 1993). the lower sightability of moose in conifer cover would be compensated for by a sightability correction factor (scf) if the voc varied among cover types; however, voc did not vary among cover types. rather, voc reflects the conditions at a location where moose are observed not visibility differences among cover types (lenarz 2006). if unobserved moose in heavy conifer cover are not accounted for, the population size will be underestimated. this was recognized in the development of the voc method in minnesota (lenarz 2006), and is relevant in any region where moose use conifer cover in winter unless the scf accounts for sightability differences among cover types. the regenerating forest, conifer forest, and wet bog types were important components of home range estimates, vhf locations, and locations during the survey corroborating current understanding of moose habitat needs (peek 1997, thompson and stewart 1997). one surprising outcome of the home range and vhf telemetry calculations was the consistency of cover type composition at different scales. cover type composition was not different in the mcp, the 95% kernel, the 50% kernel, and the actual vhf locations. patches of habitat in northeastern minnesota are small enough that areas as big as a moose home range would include both used and unused cover types. the cumulative movements around a home range result in inclusion of each cover type, but not in proportion to availability on the landscape. home range sizes of moose in minnesota were consistent with published literature (hundertmark 1997) even though the number of locations for many moose was relatively small for calculating an annual home range. sensitivity analysis using gps data in quebec indicated that an mcp home range for moose was usually underestimated with <100 annual or <30 seasonal locations (girard et al. 2002). however, the kernel home range size would 0 10 20 30 40 50 60 70 deciduous forest mixed forest marsh and fen wet bog conifer forest regenerating entire moose survey survey within 10 km of vhf moose home ranges vhf locations in december/january cover type % o f lo ca tio ns o r % o f a re a fig. 4. cover types of vhf telemetry locations compared to the cover types in which moose were located in during the 2004-2011 aerial surveys in minnesota. the cover types of vhf locations are from december and january. moose locations during the survey are either from the entire survey or from locations during the survey that were within 10 km of the composite 95% kernel home range boundary. moose home range moen et al. alces vol. 47, 2011 110 not change greatly if more locations were available each year in our data (r. moen, unpubl. data). fine scale evaluations possible with gps locations (leblond et al. 2010) cannot be accomplished with vhf telemetry, but our estimated mean location error of 300 m would have little effect on home range calculations >10 km2. although such error could affect cover type composition, the cover type composition of simulated locations with position error was very close to the cover type composition of actual locations. differences were unlikely to be biologically significant given the range of cover type composition among individual animals. even rarely used cover types that were also rare on the landscape remained in the location data set when location error was simulated. we evaluated many possible confounding factors that would have affected our conclusions. the years were different within the data set as the vhf telemetry data were collected in 2002-2005, and the survey data in 2004-2011. the same patterns remained when we subset the data creating a much smaller sample size for overlapping years. group size of moose did not vary among cover types or over time; if it had this might have changed moose distribution estimates. patterns remained consistent whether we used the entire survey area or we restricted the analysis to the part of the survey area with radio-collared moose. the analysis could be expanded by using an independently derived cover type classification, and reaching the same conclusions from independent satellite classifications increases confidence in results (moen et al. 2008). management implications if a pre-survey flight is not possible to stratify survey blocks, home range size and cover type composition data can be used to improve survey stratification. even if a presurvey flight is possible, any procedure that will increase the ability to detect change in a moose population would help agencies better manage moose. finer scale analysis of cover types and cover type composition within a survey block could be one factor affecting the decision of stratification level for each survey block, in addition to using historical understanding of moose presence. the ideas presented in this paper can be further developed and tested with ongoing and future radio-telemetry research in minnesota. a correction factor for cover type could be developed from observation rates of radio-collared moose in different cover types. we believe that accounting for moose that are probably underestimated in wet bog and conifer forest during aerial surveys would produce a higher population estimate of moose in northeastern minnesota. acknowledgements we would like to thank various individuals for their contributions to this research. al buchert, mdnr conservation officer pilot, flew the cessna 185 for locating moose. john fieberg, mdnr biometrician, created the location error data set. mike schrage, fond du lac resource management division, was instrumental in obtaining funding from the tribal wildlife grants program. mark lenarz, leader, mdnr forest wildlife populations and research group, provided gps locations for the aerial survey and discussed the aerial survey methodology. amanda mcgraw provided comments on an early version of the manuscript. this is contribution number 529 of the center for water and the environment, natural resources research institute, university of minnesota. references anderson, c. r., jr., and f. g. lindzey. 1996. moose sightability model developed from helicopter surveys. wildlife society bulletin 24: 247-259. edwards, a. j., m. schrage, and m. lenarz. 2004. northeastern minnesota moose alces vol. 47, 2010 moen et al. moose home range 111 management: a case study in cooperation. alces 40: 23-31. frelich, l. e. 2002. forest dynamics and disturbance regimes. cambridge university press, cambridge, united kingdom. gannon, w. l., and r. s. sikes. 2009. guidelines of the american society of mammalogists for the use of wild mammals in research. journal of mammalogy 88: 809-823. gasaway, w. c., s. d. dubois, d. j. reed, and s. j. harbo. 1986. estimating moose population parameters from aerial surveys. biological papers of the university of alaska, fairbanks, alaska, united states. no. 22. girard, i., j. p. ouellet, r. courtois, c. dussault, and l. breton. 2002. effects of sampling effort based on gps telemetry on home range size estimations. journal of wildlife management 66: 1290-1300. heard, d. c., a. b. d. walker, j. b. ayotte, and g. s. watts. 2008. using gis to modify a stratified random block survey design for moose. alces 44: 111-116. hooge, p. n., and b. eichenlaub. 2000. animal movement analyst extension to arcview. version 2.0. alaska science center biological science office. u.s. geological survey, anchorage, alaska, united states. hundertmark, k. j. 1997. home range, dispersal, and migration. pages 303-335 in a. w. franzman and. c. c schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., united states. karns, p. d. 1982. twenty plus years of aerial moose census in minnesota. alces 18: 186-196. laver, p. n., and m. j. kelly. 2008. a critical review of home range studies. journal of wildlife management 72: 290-298. leblond, m., c. dussault, and j. p. ouellet. 2010. what drives fine scale movements of large herbivores? a case study using moose. ecography 33: 1102-1112. lenarz, m. s. 1998. precision and bias of aerial moose surveys in northeastern minnesota. alces 34: 117-124. _____. 2006. 2006 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, united states. http:// files.dnr.state.mn.us/outdoor activities/ hunting/ moose/moose_survey_2006. pdf. _____. 2010. 2010 minnesota moose harvest. minnesota department of natural resources, st. paul, minnesota, united states. http://files.dnr.state.mn.us/ outdoor_activities/hunting/moose/moose_ harvest_2010.pdf. _____. 2011. 2011 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, united states. http:// files.dnr.state.mn.us/outdoor_activities/ hunting/moose/moose_survey_2011. pdf. _____, j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013-1023. _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated survival in northeastern minnesota moose. journal of wildlife management 73: 503-510. minnesota department of natural resources (mdnr). 2007. landsat based land useland cover. (accessed january 2008). moen, r., c. burdett, and g. j. niemi. 2008. movement and habitat use of canada lynx during denning in minnesota. journal of wildlife management 72: 1507-1513. mohr, c. o. 1947. table of equivalent populations of north american mammals. american midland naturalist 37: 223-249. moose home range moen et al. alces vol. 47, 2011 112 pastor, j., and d. j. mladenoff. 1992. the southern boreal-northern hardwood forest border. pages 216-240 in h. h. shugart, r. leemans, and g. b. bonan, editors. a systems analysis of the global boreal forest. cambridge university press, cambridge, united kingdom. peek, j. m. 1997. habitat relationships. pages 351-375 in a. w. franzman and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., united states. peterson, r. o., and r. e. page. 1993. detection of moose in midwinter from fixed wing aircraft over dense forest cover. wildlife society bulletin 21: 80-86. quayle, j. f., a. g. machutchon, and d. n. jury. 2001. modeling moose sightability in south-central british columbia. alces 37: 43-54. r development core team. 2006. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. isbn 3-900051-00-3, (accessed december 2006). thompson, i. d., and r. w. stewart. 1997. management of moose habitat. pages 377401 in a. w. franzman and c. c. schwartz, editors. ecology and management of the north american moose. wildlife management institute, washington, d. c., united states. unsworth, j. w., f. a. leban, e. o. garton, d. j. leptich, and p. zager. 1998. aerial survey: user’s manual. electronic edition. idaho department fish and game, boise, idaho, usa. wolter, p. t., and m. a. white. 2002. recent forest cover type transitions and landscape structural changes in northeast minnesota, usa. landscape ecology 17: 133-155. worton, b. j. 1989. kernel methods for estimating the utilization distribution in home-range studies. ecology 70: 164168. alces30_37.pdf alces32_25.pdf alces manuscript guidelines � ���������� � �� ��� ��� ��������������������������� �������������������������������� �������� ���� �������������� ���� ���� ������� ����������� ��� ������������������ �������� �������������������� ������������ ��� ����������� ���������������� ���������������������� ���������������������������������� � ���������������������������� ��� !"��#�$����������������%���&�'� ����(��������������������������� � ���������������������������� ��� !"���������������������� �������������������)�)�*!��+�� �,��������� � ����������������������������-.�!/��������� ��� �� ���������� �� ��� � ��� ����� �� ������ ���� ��������� �� ����� ����� ������ ��� � ��� ���� �� ���� ������ ��������� ���� ����� !" �� �#$$% ���� ���� � ������ �� � �������� ���������� �� ��� � ��� ����� �� ������ ������ ����% �� �&������ ��� ����������� � ������ ���� ����� � ������ ������ ������ ���� �� ��� �� ��� ��� ��'�� � ������� �� ����� ����������% ��������� � ��� ��� � ��� ���� ��������� ������ ����� �� ���� ����� �� �� ���� ����������� ���� �������� ������ ���� � ��� ����� ��������������������� !"������ ""�# ����!��$ ��" ��� ������������%��&������ #� ����$' �����"� ��"������(��������������%�# "�� ��$$�&� "��� )*������ !"� +� #�$ ���� ��� ���� �� ��� �� ��������� �� ����� �, �-�"" � ��#� -�"" � ���� � � ������ %� #�$ ��� ���� ����� �!��"��"������������������"� ��� "���� "�� ����� � � .�&�(��/� "����� ���� �� � ����������# ���������� ���"'$����#������" ��"&���� ������ ��#����� �� ��� ������� ������� ����� � � � ����� ��� "����� �0��!� " ���� ���� ��"�#� �� #�� "��� ����"� ��(��� �� ����� �������������/���% �� �%�& "����$��� 1�� �� ��� ��2�"�"�����������������/�$��� ��" ����$��# ������������"&����"���"&��3���� ��$�/� �$$� ��� &� ��� ��4� ��� �# " �%� ��# �������"" �%����������� !"��"����� �(������ � �"���'�����"'$����#������"�����%�!�!��� �!!��� �%� ��"���3�����$ ���� ��&��5� ���� �0!�������#��$�&��"���!��$ ��" ���!������ ���������/� &�� ��(�� #�(�$�!�#� ����� #�� "� $�#�%� #�$ ��������"���!��!���" ����������� ������� !"� �����"� ��"��������$#�����5�"�� ���"������"� ����������������������%���"��" ��'���!����#��"�����%� #�$ ���/����&�$$��� "����������#��##��������"���������"��# "�� ���� ( �%� ������� !"� � � 6!#�"��� "�� "���� %� #�$ ���� ���� �$��� !��"�#� ��� "��� ����� 7��$#� 7 #�� 7��� � "� � � 8�% �� �%� & "� ��$���� 1� �� ��9/�"���������&��� "���#� #����� ��% (������"��� �� #�����5���(����� ����� �������#�����$$���!� �"����"� ��#�# � ���"$'���������"� ��" �%���"���� ���������� ���"� ������ &��� "�� �##����� �� �""!�:: & & & $ � 5 � � � � # � � � : ; � $ � � � & & & : �$��� �"�$ ��������� ����� ������ !��$ ����� �� % ��$� ������� !"� #���� � �%��"�# ������"���� �$�%'���#����� �%����"� ��� ������ �������������� "����%�� ��"� "�� �� � ����!�$��� # �"� ��" �� � � ���� ������� !"���� % ��"�����!�!����!�����"�# �"�"��������$�<��"������ ����*��������� ����������#�7��5���!����"���=�"����" ���$ *������'�!�� ��/���"�&��5����'�������� � ""�#�# ���"$'�"��"����# "�����"���'�" �� ����� �� �$$���"�#� "�� ��� ����� �"�� �# "�� &������ %���"���������� !"�"���"�$���"�"&� ��( �&���� 5��&$�#%���$�� ����"� "��� ���� 3��" � � -�( �&���� 3�#%�� ���� ""�#� ����� ��� !"�����#�"���� % ��$ "'/� #���/����$'���/ �"��!��"�" ��/��������'/����� ������/��$��� "'/��!!��!� �"�����3��"���""��/���#����"�� � ���"� ��" ���"���0 �" �%�5��&$�#%� �>�����"��##�������?1/�<�$�$�/���"�� �/�����#��>@��1a@ ��������������������������������������������� ������������������� � �� ��� ��� ! ������� � �� ������� =��#�(�$�! �%�"�����%� #�$ ����&����(� ������#�"��"�!���!��" (�����"� ��"������(� ������� "�� �� ���!�"��� &��#� !��������� ��� ������� !"�!��!���" �� ��=�������� �� ��� � !��#���#������$$'/���"���������$#��" $$��#� �����"��"�����%� #�$ ��������$���$'����!��� � �$� ����"�����( � ��/�����!"�#�������� !"� ���"�������� ""�#� ��# % "�$������� � /��� # �5�""��/���$$�& �%�"�����%� #�$ ��� ����� "���������$#�������"���������"� ������������� ���� #�"� $�� ��� $�'��"/� ��!�� �$$'� ���� "��$�� ��#�����������$ �"� ��*������ !"��"��"�#����" �������� "�� "��� %� #�$ ���� ��"$ ��#� ��$�& ��'������"����#�"��"�����"����������# � ��� " �� ������� ����������� ������� ��b��������!�!�����! ������"�� ������� !"� & "�� �$$� $$��"��" ���� ���"� �� !��( #�#� ���� "��� ��( �&� !������ � � c�� ��" ���� "���! ��������!����4��$ "'�#�"���"� 0 !� �"��� ���� ����� � >� �"�"���������� !"��� ����� #�����%��#�4��$ "'�&� "��!�!��/�1 d�0 1� @������ d�0� � ������������"� ��� e���� ��� ������� � *� �"� ��1 d����� � ��������% �������$$�!�%��/� ��$�# �%�"��$�� ��#� $$��"��" ��� ���� ���!���� � �0��!"� ���� "��� #�"� ��#�������!��# �%���"���� ������" ������� ��$�&�/��$$�!��"�����"���������� !"����"��� "'!�#�#���$���!���# "#���$�!������ � �$$� "�0"� ����$#� �� $��"�3��" � �#/��0��!"�����!�%������������# "���������(�$���� ������" ����������$�&� &� �������� %�"�3��" � �# %&�'�������� � c�� ��"� ����5� ��# �'!����"�� &��#�� �"� "��� � %�"� ���% � � � �'!������'��!!�����"�"���� %�"����% ����$' �� "� ��!��"�������'!����"�#�&��#����!����� �� % /�1�'�����$#���$$�� (��� ��) $��� ����� � 6����������� ���"/� ����� ��� � ���� -����� ��� .�$(�" ��/ & "�������"�� e������"�$���"� @�!"��� �!"�� � ������"$'� !�������#� � � ���� ����� ���"� ��# ���"�� e������$#�������#�"����%���"������$$ "�0"/� ��$�# �%�"��$�� �&�� ��&��� � c����"���#��$ ���&��#� "�������"� �� "�$ �� ��6���"���"�0"������"" �% ���"�����(� $��$�� ��&��#�!������ �%����"� &���� "�� "'!�� "��� "�0"� �� "�$ ��� &����� ��� 4� ��# ��="�$ �������$#�������#������� ��" � � �����/���" ��&��#����#������( �" ������ % / ������/���"���"�� % ���� � �/��"�" �" ��$��'���$� �� % /� /� �/� x /� �/� �/� �"� �/� ��#� ������ �� !��$ ��" ����% (��� ��"���"�0"��� % /������� c����"����� "�$ ����������������!��$ ��" ��� % (��� �� "��� -�b�-�<���� ���" ��� ���� ��$�&� ��6!!��������$�""������#�*��)�����# "'!�� ����$#� ��$'� ��� ���#� ��� �# ��"�#� � "�����%� #�$ ��� ��)����� � ����� ��"�$ �����������!���� %��!�����"���� �#��"�#�d��!���� �����" ��' ���# �%�������$��� �#��"�#�d��!��������!��" ��� �� !���%��!�� ����� ��$�&� � � ���� �����#/ ��#� �����4���"� $ ���� ��� "��$�� " "$��/� ���� ���# �%��& "� ��"��$��/����"��"��/���#������� ������ "�" ��������$#���"���� �#��"�#����� ��$�&�� ������ ""�#�������� !"� ��=�#��"� ����"�����"'!������"�0"�& $$�����##�#�#�� �% � ��$�!��#��" �������!� �" �% ��� �# *���� � �0��!"�����"���� ��" !�%�����"���������� !"/��$$�!�%������"��� �������#� �������" (�$'/� ��$�# �%� "��$�� ��#�� %������!" ��� ��8�% �� �%����"������� ��#� !�%�/� "��� ���� �%� ���#� ����� ��$�&� ����$#����"'!�#� ��"����!!���$��"����������# "��� !�%�� ������� ��"��" �%� �"� )1,�� ����$# �!!���� ��"����!!���� %�"������� (��������� � ��0"� ���"��"��� ����$# ��$'�������#��"�"�����""������"���� ��"�!�%� "��!��( #��"���������"��##�������������"��� &���� "�# �����������"����##������"�"���" �� ���"����"�#' ���������"��"���!!����� ���� # �"�$'���$�&���$��"�3��" � �#���$ #�$ ������ @ ������"���/���#���% ���& "��"�������� ��$ ��!����� !"� ������!��# �%� "�� "��� ��"���f� �������$$�&�#��'�)>�����"��##�����, ����� ���"��"�� "���� ���" ����� & "�� "��� �##���� ������" �� ��=������##����� ��$��%���"����� � �%$��$ ��/�#����"� �#��"�"��������#���#���' ( ���������� � �� ��� ��� ��������������������������� �������������������������������� �����4���"�$ ��� ��=��"�����������1���"���� & "�� �##����� ����%��/� ����� ��&� �##���� ����$#���% ���������&�$ ��/���#� ��!����#�# �'�"�����!����� !"�������!��# �%�"��"����!� !��!� �"����"������$$�&�#��'�)>�����"��#� #�����,���#�"����##����� ������" �� �>��"�$ ��� e !� ��#��� ����$#� ��� ��$�#�# � � ���$� ���"��"�������# ������#���$�& ������������ ���� ��) !��������)��� �#�'��� � ���� ������� !"� ����$#� ��% �� ��� "��� � ��" !�%�� & "�� "��� #�"�� �����%�#� & "�� ���� ��( � ���/� ������!��# �%� ��"���f�� ����/ �##����/� ��#� "�$�!����� ������/� � �%$�� �!���#� ��"����!!���$��"������� ��=���(� $��$�/ "�����"���f����0����������#��$��"��� ���� $ �##���������$#��$������!��( #�# ���������� "��/��$$�"�0"� ��#���$���!���# �#����� '��)� � ���� ���� �%� ���# �-.�� ��$��"�3��" � �#���#��!!����������� �%$� $ �����$$�& �%�"���������!��# �%���"���� �� �����" �� ��8�% ��"���$ ���& "��-.���$$�&�# �'�����$�����#���� �%$���!��� ���������� �� #��� ��� "��� $ ��� �� $ � "�#� "�� �d� ������"��� � ��$�# �%��!�����/�"'!�#� ���!!��������$�"� "��� ������-.� ��$�#������� ���#���� !" ����� "���!�!�����$$�&�#��'����'!������#�"���$��" �������� ��� � ��� 1� ��"���� � � =�� "����� ���� 1 ��"����/� ��!���"�� "�� �� $��"� ������ & "� )�� +6=c��=<�����-�c+�-��������� ������ � ����" "$����% ���$��"�3��" � �# ���"�����0"�$ �����$$�& �%�"���-. ������" "$� ����$#����"'!�#� ���!!��������*��)�$�""���/ ��#�����$#���"� ��$�#�������( �" ��� ����� " "$�����"��������"���� @�&��#�����#���!��� ���"�" (�����"�����" �$�f�����"��" �����%�� " "$��� ��'� ��� ����!"��$�� �� ����"��� " "$�� �������&5&��#�����"���" �������� $�"������ ��� ��"�����"��" �#�'�� �� �+�,� � ������0"�$ ���!��� ( #��� "��� �������� ��� "��� ��"������/� $��"� 3��" � �#/� �� �!!���� ��#� $�&�������� *��) $�""��� ��=��"���������1���"����/���!���"��"�� � $��"�������& "��)��#,�� ��*��)�$�""���� ��=� "����� ���� �� � ��"����/� !����#�� "��� $��" ��"���f�������& "��)/���#,�� ��*��)�$�""���� =��"�����������1���"�����& "��# ������"��#� #������/�"��������� �# ��"�#��'���!����� !"� �"� "��� ��#� ��� ����� ��"���f�� $��"� ���� � � � � �%$����!����� !"/�����������#���!����� !" ��!���"�#������"���� ��"���$'��'�������� �����!����/���$$�& �%������"���f��$��"����� ��'� �$��� ��� ���#� "�� �# ��"�� ��� �##���� ����%��% (��� �������"��"����������(�� �#�'�� �))����+��,� � 8�% �� �%��� "��� ��0"� $ ��/� "'!�� "��� �##��������� ��� "�� ��"������� �"� "��� " ��� ��� "��� �"�#'� $��"� 3��" � �#� ���!!������#�$�&����������%�$�� $�""����� � /�#����"�������$#�"'!�� ��=��"���� ���� �� 1� ��"����� & "�� # ������"� �##������/ "�� ���##�����������$#����% (��� ��"������� ��#������ ��"�����"������������$ �������(�� ����� �##����� ����$#� ��� !����#�#� �'� "�� ��!����� !"�������!��# �%�"��"����!!��!� � �"�� ��"������/� ��#� ����� �##����� ����$#� �� ��!���"�#� & "�� �� ��� ���$�� � � >�����" �##�������������$#���� �# ��"�#� �������"� ��"����������(��/� ��# ������"������"���" �� ���"����"�#' ��=��$�#��!��"�$����e !���#�� �*����!�� � ��"��� $��( �%� �� � �%$� �$��5�$ ��/�"'!���8��-���/�$��"�3��" � �# ���!!����������%�$���$�""���/���$$�&�#��'�� ��$�����#���� �%$���!��� ��8�% ��"'! �%�"�� ���"���"���"���"���� �%$���!�������"������� $ �� ���������"���"�����$#����������� ����� !��� �$� ��>�����"� "� ������!���%��!����#�#� ��"����������( �" �������$ "���"����� "�" ��� ���� ���"���"� ����$#� �# ��"�� "��� !���$�� �"�# �#����"����'!�"��� ��"��"�#/�"������" �!��"��"� ����$"�/� ��#� ��'� ��3��� ����$�� � ������� �"��!��"�" ����#��&�������"���&��5 c���� !" ��������"��#������$#������ ������ $������&��������� �!��(�#�"���� 4������� ��!��"�# -��# � �)����$�!������ � ��������� (�$���� #��" � ��" ��� ��� %�"�3��" � �#����"�� $ �����$$�& �%�"������"���" ���'!������� ��� �@@��@@@@��!! �@@@���@@@����"� ��$ ��/ ��������������������������������������������� ������������������� � �� ��� ��� " � %�"�3��" � �# .�& /��)�� � ��"���$��( �%���� �%$� �$��5�$ ��/�"'!��a�'�7��#����$$�&�#��'�� ��$��� ��#� �� � �%$�� �!���/� $��"�3��" � �#� � �!!���� ��#� $�&��������*��)� $�""��� � � b�$� $�& �%�"���� �%$���!���/�������%�$���"'!��"� !��( #�� 9� 1� &��#�� �� �$!����" ��$� ��#�� "��"� ���"� #���� ��� "��� ��3��� "�! ��� !��� ���"�#� ��"���!�!�� ��8�$�&�"���5�'�&��#�/ #��&�����$ #��$��5�$ ���� �!"� ��& #"��������� "��� !�%�� ��"&���� "��� ���% �� � � ���(�� � �$��5�$ �����$�&�"� ����$ #�$ �����#���% � "'! �%�"���� ��"�!���%��!�����"�0"�/� �#��"�# d��!����������"���$��"����% ��� � /�#����" ����"� �� ���" ��� ���# �%� ��� !�%�� ����5� "� �"��"� �� ��&� !�%�� %��)���� ��) �0�� ��!������ � � � $���"��"� ��!�!��/� �$�(�$��������# �%� ����������$'����#� ��������������� !"�� � ��!� ���'����" ������# �%���������"��#/� � �!!��������*��)�"'!�g��1�������#��'����#� �%������$��"�3��" � �#/� ��*��)�"'!�/�& "����$' "���� ��"�$�""�����������&��#� ���!!�������g ��#� � �� "��" ��'� ���# �%�� ���� �#��"�#� d �!��������!��"������!���%��!�/� ��*��)�"'!�/ & "����$'�"���� ��"�$�""������"���� ��"�&��#� � �!!�������� ��0��!"� &����� !��!��� ����� �������#�/���#�������$$�&�#��'���!�� �#���# 1��'!���� ���0��!"�����"��� �"��#��" ��/��$$ ��3������" ��������$#���� #��" � �#��'�!� � ���'����# �%� ������!��$ �����!�!���������& #�����%� ������3��"�� ��$�# �%���"���$�� �"��'��� % / ���#���� "�/���� "�"�����/����!��$�%'/�"�0� ����'/�!�'� �$�%'/�!���� "�$�%'/�!�!�$�" �� #'��� ��/���#�$$ �%/��(�$��" ��������"�" �" � ��$�"���� 4���/������������"��#�/�����'�� "���#'��� ��/�����%����"/�$�&��������� ���"/��#���" ��/������� ��/��#� � �"��" ��/ !� $���!�'���#��"����� � $���"�! �� �������� 4���"$'/�"����� ������ #����$���$�0 � $ "'� � "�����%�� ��" �����#�$���$$ �%����������� !" ���" ��� � � =�� ���"� �����/� ��&�(��/� ����� ��� !"������$#������%�� ��#� ����"��# " ���$ �����"�"��"� ��$�#������ �"��#��" ��/���6c2 �-��� � �� !��" ���"�/� *��.�c�/� -�� �6���/� c=��6��=��!��� ������� �%��!�� � ��� "�������$#� ��$�#��� ��!�$���" �%�"����"�#'������& "� ������% ��/ ����"�'/�������" ���" ��6���"���!��"�"�����"� !�����"��"�#'������#���� !" ��� ����*��.�c�����" �������$#��� ��$' #���� ��� ��$�(��"� !����#����/� �4� !���"/ ��#�"���� 4��� ��c�"��/����!$ �%�!�� �#�/ �������������0!�� ���"�$�#�� %�/���#���"�� �#�� ��� #�"�� ���$'� �� ����$#� ��� ��$�#�# 7�����!��� �$�/���"���������$#�������"��"�� $ "���"���� ���� ��"��#�� �$���#'� !��$ ���#/ "���� �# ��"����'�#�( �" ����"��'���(����#� �����"�������"��#� ��<�&�"���� 4��������$# ��� #��" � �#� ��#� �0!$� ��#� �� ���� � ��" #�"� $� "�� ��5�� "���� ��!��"��$� � � 8���# ���������������� �$$'��(� $��$���4� !���" ��� ���� ��$� !��#��"�� �& "�� "��� ���!��' �������#�$���" ��/���!���"�#��'��������/ ��!����"�����������$#����!��( #�# ��*�"�� ) ���������� � �� ��� ��� ��������������������������� �������������������������������� �#������$#����!�����"�#� ��"���!��"�"���� 6���"���-��6�������" ���"��� %�$ %�" � �# �%��!�����"�#� ��� %��������"��$��/��(� #� �%� ��!�" " ��� ��� ������" ��� "��"� ����$# �$���#'� ��� �$��� � � ��$'� "����� 4���" ��� �� ��#� ��"���!��!�������"���&��5�����$#��� �##�����# ������� �# �%������$#������%��� ��#� ��"��������$�% ��$���4��������� ��"�� �"��#��" �����#�*��.�c�����" ��� ��=� ���"������/�����$"������$#����!�����"�#� � "��� !��"� "���� ����c=��6��=�<�����$#� �# ��"��"�� �� �����"� ��" �������"����"�#'/� �"��!��"�� " ��� ��� "��� � �# �%�/� ��#� ���!�� ����� "� �"����!��$ ���#�&��5 ��-���$"������$#���"��� ��!��"�#���#���$'�"������"� �!��"��"�� �#� �%������$#�����##�����# ���'�"���" ��# �� ���� �������(��'���!��"����"������������ � ����������' ��-�������$���!���$�" �����# # ���" �����������"���������������'���� �� �$�#�# ���������!�/�� %� � �����/���#�%��� ���$�����$�� �������"����"�#'�����$#���#�"�� # ����� �� � � "� "��� ��a<�7��c+�*�<�� "��"�����&�����(�����"� ��"�#�����"��" �$$' "��"����� ��" � ����#�"���� ��$���!��"�����"�� ��������/� %���"�#� � ���� �$� ��!!��"/� �� ��$!�#� �!��(�� "��� 4��$ "'� ��� "��� ����� ��� !" ��"����� ���� ���!��� �$�� ���� "��� ����� ���'�����$$� ������" ���% (��� ��"���-�b�-� �<���� ���" �� � � �$$� ����������� ���"� �� ����5�#��%� ��"�"����� % ��$���" �$����#����" ����������#�"�� ��"���"�0"��'�"����������#� '�����'�"����������$�&� ��-����������"'$�� ���� $ "���"���� � "�" ���� �� ������ ���� % (�� ��$�& ��*���� � ���$��� ����$#� ��"� ��!��" #�"��!�����"�#� ��� %���� ��-�����"��"��$��� � "���"�0"��� �%�)���$�,���$$�&�#��'�"���"��$� ������ �����$�������$#���"����%�����"�#���� ���$$�#�"����"�/�"��������"� � �%����'��$��5 ��"� ��/�e����/���!�" " �������"������������ ����/� ��� "����� & "�� ��&� ��� ��� � %� � ���" #�"� � � ����� #�"�/� ��� �� ������'� ��� "���/ ����$#����!$���#� ��"���"�0" ����"���������$# �����$"� �� �����"� ����� ��� ������ ��� ��� �� ��� ������������� ����� ������%� #� ����������"" �%��!�"��$�� >�����"������"��$���������!���"��!�%�/ ��$$�& �%�"���-�b�-�<�������" �� ������ " ����"��!��( #��"������� �%����#���#����� ����" (�� !�%�� �������� �"� "��� "�!� ��� ���� !�%� ���$$�"��$�������$#�����������#���# !�����"�#� �� "��� ��#��� �� &� ��� "��'� ��� � "�#� ��"���"�0" ��>��!����"��$��� ��"������� ���"���#����"�� e��������#� ��"���"�0" ��� "$�� ��#��$$�!��"�����"��$������"����"'!�#�#���$�� �!���# �����$�������$#��������"���"�#�"��� " "���& #"�����"���!�%���1 d����/�$��( �%�1 d� ������% �������$$�� #���� � /� 9 d����& #�� 7 #��� "��$��� ��'� ��� �������#�"�#� �' "��� �%�"���!�%��� #�&�'��� � /�)$��#���!�, !�%���� ��"�" ������#�� "" �%�"���"��$���"��"�� $��%"�����"���!�%���1� @����/��%� ��$��( �% 1 d�������% �������$$�� #���� � /�1 @��� & #�� ����"��������"��$���$'� �# ��"��&��� ��"��$�������"�������"���� �!�%���� % /�"'!� )���$�� ����" ���#h ,��"�"�����""�����# "�!���������!�%�� ���$��" "$�����% ��& "��"���&��#�)���$�, $��"�3��" � �#� ����%�$���$�""���/���$$�&�#��' "��� "��$�� ������� ��#� �� !�� �# � � ���(�� � � �%$�� �!���� ��� "��� ����� $ ��� ��#� ��% � "'! �%�"���" "$� ����#�"���" "$��& "����!�� �# =��"���" "$�� ��$��%���"������� �%$��$ ��/�#����" �#��"�"��������#���#���'������4���"�$ ��� ���� " "$��� ���"� ��� ���� ��� ��#� �$���� ��� � ���#��� ���� ��#���"��#� "��� "��$�� & "���" ������ �%� "�� "��� "�0"� � � /� ����$#� )�"��# �$���,� ���'! ��$$'/�"���" "$��& $$� ��$�#��"�� ������ ��� (�� ��$��� ��#� ��%�� ���� ����� ���#/�"�������������"��� "����� ��!����"��� ����/� ��#� !$����� ��#� #�"��� ��� ���!$ �% �����( �" ��������$#���"�������#� ��"��$� " "$�� ��b��"��"�������$#�������#�"����#��� "������!$�0 "'����"��$��" "$�����#�!��( #� ���"����#�"� $���������$�&� ���"���$ �����$�&�"���"��$��" "$��#��&�� � �%$�� ��� e��"�$� $ ��� ������� "��� !�%�� ��� "&����"������% �� ���## " ���$���� e��"�$ $ ���� ���� "���� ���#� "�� ��!���"�� ��$��� ��������������������������������������������� ������������������� � �� ��� ��� * ���# �%�� ����� "��� ��#'� ��� �� "��$�/� ��" ����$#� ��"� �!!���� �� "��� ��#'� "��$� � � � � �%$����� e��"�$�$ ��� ���$���#��&�������� "��� !�%�� ��"&���� "��� ���% ��� ��$�&� "�� $��"���&������"��$� ��<��(��" ��$�$ ��������$# ���!�����"� ����"��$� ���$����$��������"����%�����"�#�& "� "��� ��"" �%�� ��� �� "��$�� �# "�� � � c�� ��"� ��� �!����� � � /� "��� �!���� ����� "�� ��$����"� ��"� �� ����$������#���&����# �%������$# ��"����� �"��#�"��� �$#� ����! "�$ ���"���� ��" &��#� �������# �%���#�#����"���#����# �%� & "�� �� !�� �# � � -�&� ���# �%�� ����$#� �� "'!�#��$����& "��"���$��"����% ����#����� $�(�$� ��� ������# �%� ����$#� ��� �#��"�#� 1 �!�������$�" (��"��"�����&����(� ����$��� ���# �%������$#�������"��#� ��"�� �����!��� " (����$��� ��7��# �%� ����$���������$# �!!���� �$���� & "�� "��� $��"� ����#��'� ��� � ��$��� ��=�"�%��������$#����"'!�#��$����& "� "��� � %�"� ����#��'� ��� �� ��$��� � � �"��� ����� ��$���"� �������$#����(��" ��$$'��$ %��# �� �%�"���#�� ��$�!� �"�������"�����$!����� ��� �� ������"��� �� % /� i/� j/� �/� �/� �/� �"� � ��"��������"���������$$��������� ������$� ����������!��"�#�& "��"����!!��!� �"��$�(�$ ���!��� � ���� � /�� %� � ���"�# % "�� ��c����� ��'�������#�"�� �# ��"��� �� �%�(�$��� ��c� ��"� ���� e����� ��$���� ��� ��"��$� (�$��� ��� @ &����������#/���#�"���� �# ��"��"���$�(�$ ��� !��� � ��� �'� ��!��" �%� "��� �!!��!� �"� ������� ��� � %� � ���"� # % "�� �� % /� @/� @ @/ @ @@/��"� � b��"��"��� �� "��$��� ����$#� ��� #�� %� ��"�#�& "����"�� �5������!����� $ "'�$�(�$� � � /�k�l@ @d/�kk�l@ @ /�kkk�l@ @@ ���� ����� ��$� ��!����� !"� � � �"��" �%� �"� ) ,/ ����� ��$���!����� !"������$#���$$�&����� ����" (�$'� "����%�� "��� " "$�/� "���� $��"�"�� � %�"/���#�"����#�&� ����"��������"������� "��"� �$$� ��!����� !"�� �� "��� " "$�� ��#� "��$� ��"��� ��� �!!��!� �"�� ���"��"�� ��$�&� "�� "��$� ������� ��"����"��"���!!�����$��"�3��" � � �#� ���# �"�$'���$�&�"������ e��"�$�$ ����" "�����""������"���"��$�/���#���% ���& "���� ��"�� �5��������� ��$���!����� !" �����$ �� ��"�������" ���#�& "��"������"��"�� ������� " �� ��=�������"��"�� ��$��%���"������� �%$��$ ��/ #�� ��"� �#��"� "��� �����#� ��#� ��'� ������ 4���"�$ ��� ���������&����"��"����% ������� ��&�$ ��/���#� ��!����#�#��'� "��������!��#� �%� ��!����� !" � � c���� !" (�� ��"�� �$� ��" ��4� � �%����!�� � �����"��"����'����!$���# ��#�����"��$�������%�����$���"�/�"����#��� ���!$�0 "'����"���" "$����#�"��$� ��=��!$������ ��� ��"�� �5� ��� ����� ��$� ��!����� !"/� "'!� )<�"�,� ��$$�&�#� �'� �� ��$��� ��#� �� � �%$� �!���/� $��"�3��" � �#� �� �!!���� ��#� $�&��� �����*��)�$�""���/�"����"'!��"��� ������" �� (��#�� !������� ��) ���#���������� � =$$��"��" ���� ����$#� ��"� ��!��"� #�"�� !��� ���"�#� ��"��$�� ��-�����"��� %����� ��"���"�0" �� �%�)b % ,���$$�&�#��'�"���� %���������� 8�% ��"'! �%�"���� %������!" �����������& !�%�� ��$$�& �%� "��� $��"� "��$�� ���� -�b�-� �<���� ���" ��/� �� "����� ���� ��� "��$��� ���" ����"��!��( #��"������� �%����#���# �������" (�� !�%�� �������� �"� "��� "�!� �� ����� !�%� � � �$$� � %���� ��!" ���� ����$#� �� �������#� ��#� !�����"�#� �� "��� ��#��� � &� ��� "��'� ���� � "�#� �� "��� "�0" � � b %��� ��!" �������"����"'!�#�#���$���!���#���# ����"�����������"���#����"�� e��������#� � "����� ��"�0" b %������!" ������% ��& "��)b % ,�$��"� 3��" � �#� ����%�$���$�""���/���$$�&�#��'�"�� � %�������������#���!�� �# �����(����� �%$� �!�������"��������$ �����#���% ��"'! �%�"�� � %������!" �� ��=��"�����!" ��� ��$��%���"��� ��� �%$��$ ��/�#����"� �#��"�"��������#���# ��'������4���"�$ ��� �������� %������!" �� ����$#���% ���������&�$ �� ����!" �������" ������� �����#��$������������#���������#��� �"��#�"��� $$��"��" ���& "���"������� �%�"��"�� "�0"�� � /�����$#�)�"��#��$���,� ��b��"��"�� ��#� �����( �" ���� ����$#� ��"� ��� ���#� � � %������!" ��� �����!$�0��'���$�����5�'� ����$#���� ����!���"�#� �������� ���$�%��# ���"��� $$��"��" ��� "��$�/���"����"���� ��"�� � %������!" �� =$$��"��" ��������� "����!��"�%��!����� $ ���#��&��� %���� ��b���"�����( �&�!������/ + ���������� � �� ��� ��� ��������������������������� �������������������������������� %��#�4��$ "'�!��"���! ������$�����!� �"���� $ ���#��&��� %���������������� ""�#���#�"�� �� % ��$����"� ��# ��b %�����!� �"�#����#�" ��"� 0�!� �"����������"�����!"��$� �����!�� "���%�����"�#�%��!���(��'� ��4��$ "' ������ ��'����3�#%�#����� "��$�/���#��� % ��$����' ��(��"�����!��!���#��� �%�"��# " ���$�%��!� � ��"�"���� 4��� ��m���%��!� ����! ������!��� "�%��!��� ���� ����$$'� ��"� ����!"��$�/� �(�� ����"�����( �&�!������ ��=��� "��������/���� � ��$�!� �" �%����"���3�����$��������$�&�/��$$ $$��"��" �������"�������"���� %���"�!������ � ���$�4��$ "'�"���������!��!�����!��#��" �� ��"����$��"��� ������� �% >��!�������� $$��"��" ���!���!�%� ��=#��� " �'������ $$��"��" ����'�!� �" �%�"�����"���f� �������#�"���� %��������������"������5� � ���"�!��� $ ��=�� "� ����"���( ���/��$��� �# ��"� "��� �� ��"�" ��� ��� "��� $$��"��" ��� ��� "�� ���5 ������� $$��"��" ����� "������!��"�%��!� ���$ ���#��&��� %����/����%���!���� $$��"��� " ���/�����$#����!$����#�"��� "� �"��"������� ���� "���� ��9n��������1�� ��������$���� ���"�0" ��=$$��"��" ����"��"����������#���#�"� � �%$����$����& #"������!�������# ����$#�� ����$#�"���������������"������������&��$� !�%������ � $$��"��" ������0 ����$��%"�/� �@ ��� ��$'������" �$�$���$$ �%�����$#�������# ���$ ���#��&��� %����/�& "��#�"� $�#� ����� ��" ��� % (��� �� "��� ��!" �� � � ���# �%� � � %���������$#����# �" ��" ���$$�$ �������"��� ���� � ��"$'�"� �5�����%��#���!��#��" ��/���# �$$��'���$�/���!����� !"�/������� !"�/�#�� � ��$�!� �"�/���#�!�� �#�����"����&�$$�!��� !��" ���#�"��"������"����"���� %������#�$��%� ����%��"���$$�&�������#��" �� ����""������# �������� ��� ��#���#� � %����� ���"� ���� � $�% �$����#�������$����"���� d����� %����"�� ��#��" �� � � 7����(��� !��� �$�� ���� �!!��� ����� $�""���� ���� $���$$ �%� � ���� "��'� ��� ����� $�% �$�� &���� ��#���# � � 6��� �� �$��� ����� ��� �� ���"� �� % /� .�$(�" ��� ��� �� �$� ��#��& "����!� �" �%�#�( ���������������"� ���!� �"�#�������"��� ��c����"�������"'!�� &� "�� ������������ e����#����"����$�""�� �% ����$#�������#������$$�� %����� ��"�������� ��� !" >��"�%��!��� ���"� ��� ��� � %�� ���"���" ��#�!� �"�#�& "������""��� � �� ����""�� �%��� ������ �%� ���"� ��� ��#�� ����� ����"�� �� !� �"�#�������"������#����"����"���"�� %�$' &���� ��!�� �!���#� ��� !��"�%��!�� � � c� ��"��������"'!�&� "�� �������$����������$#��� ���#� "�� �# ��"�� ��%� � ��" ��� �� � e�� � �!��"��" � � � � $��� "�� $ ���#��&�� � %����/ � e����#�!��!��" �����������"�" �������"��� ������$$'����� #���#�� ����!��"�%��!������" �$��������#���#�"��� "�� "���� ��9n��������1 � ��������$��������"�0" ������� ��&�� ��) ����� �# *��� ��) #����� � 6���# % "����� ���������� % /� /�1n����$����"���������� � "���� ��"�&��#���������"����/� ��&� ������� "� ���!�$$�#���" ���!�$$���"���# ��$�������� �� % /�� ��"/�"� �#����#������������#����!��� ������ �� % /� ����� ��� �#(���� � � .'!����"� ��������� "�!����������#�����#3��" (����� % / 1�'�����$#� ��$$��/� ��"� ��"� "����� ���#� �� !��# ��"���#3��" (����� % /�!$�"��&����d���� =����"�������� ������������ /@@@/��0��!" ���� !�%��� �� ���5�/� �$��5� " ��/� ��� '��� #�"�� � � c�� ��"� ����"� �� ������ ��� �'!��� ��"&�����������" (�/���!���"���������� � �� !������ �� % /� 1d� ���� !$�"�� � � <�(��� ��� ��5�#� #�� ��$�� � � /� ���� @ @ /� ��"� @ � 6����'���$����������( �" ������ % /�o/���� ��������������"��� "��"��"���$$�&���������/ ��$����"���������� �� �#�� � "���� % /�"���� ���#��������"�����/� ����)@,��e������"��# �% �$���/���� ��"���� ��"�&��#� �������"���� ��=� ����������/��!�$$���"�"������������#��� " ���� � � 6��� ����" ���� ��$'� &����� ���(��� � ���"��#�� ��$�!��!��" ����� ���!�����"� !��� � �� �� �� ��) )����� � 6���"���1������ �'�"����@@@ �"����%��1�@@��������� #� %�"� c�"����4������ ��#�'����"��'���/�& "���" !���"��" ��� �� % /� �� �!� $� ��n� � � 6��� �� �!��"��!�������!$���$�#�"����� % /� ��@f�� �!�$$� ��"� ���"��/� �0��!"� �� !����"�����/ ��������������������������������������������� ������������������� � �� ��� ��� $ "��$��/���#�� %����/�&����� �$�""��������( �� " �����������#�& "�����!�� �#��� % /����!� ��n� ��������!�� � -����� $�""���� ���#� �� �'���$�� ���� �"�" �" ��/� "��"�/� ��� (�� ��$�� �� % /� /� x /�����/��/� /����!/��"� ������$#��� "'!�#� �� "�$ ��/���"���#��$ ��# ��<������/ ���������"� %�����"� ����#�"�������#��"�$ ����" ���/� ��� �����( �"�#� �"�" �" ��$� "���� �� % /�$�/��/��0!/�$ �/�� � /���0 /��c/���/���/ #�/� �"� �� ����$#� ��"� ��� "�$ � ��# � � +���5 $�""������ % /��/��/��/��/��/��"� ������$#��� "'!�#� ��� ����� �� �%� "��� �!!��!� �"�� ���" �(� $��$�� �� ���"� ���!�"��� &��#� !������ ���� =����"����!���������"��� #�������'���$� ���#� ��� )���3���" ���,� �� % /� � j� 1d/��� j @ @@ �/� ��"� �$���� "��� �!���� &���� ���#� �� )�#3��" (��,� �� % /� p @� ������ � � 7���� !��� �$�/���!��"��0��"�!����� $ " ����� % /���j @ @d1/� ��"��� p� @ @d� � � ������ !"�� !����#� ��!����� !"���� % /�! � �����$����"��������� !" ��$�#��� p1� ������"���� �� % /�!� ��� � -�!��"� ����$"�� ��� �"�" �" ��$� "��"�� �' % ( �%�"�����$��$�"�#�(�$������"���"��"��"�" �� " �/�"�������� �"�#�#�%������������#����#��/ ��#�"���!����� $ "'������"� � �%�"�����$��� $�"�#�(�$������"���"��"��"�" �" ���� % /���j�1 �n/ �#�/���j�@ @ /���j� @g� /� 1�#�g���j�@ @ / �"� � � � ��$��$�"�#� (�$���� ��� "��"� �"�" �" �� ��#������ �"�#�!����� $ " �������$#������� !��"�#�& "������0 ������������"�����$$'�1� � %� � ���"�� %���� ���� ��" � ����"�" �������$# ������#�"����!��"�(��'�$��%��������$$�(�$��� �� % /���j�@ @@@@1�����$#������!��"�#������j 1 @�0� @��� ���������������# �!��� �����#�"�� ���!$��� e������$#���� �# ��"�#�& "������� �����������"��$��"��#���'��� % /� x �j�d /��� j�1 /� �j� d� �1#������� � -�%���� ��� �4��" ��� ���� ��� ����!���"�#� �� "��� "�0" � � ��� ��$� ������ �� ��%���� ��� �4��" ���� ����$#� �� �!�$$�#���"� ��"��'���(����"������!��( ���$' #�� ��#� ��"���!�!�� ���"�����4��" ��������$# ��� ���"��#� ��� "��� !�%�� ��#� "'!�#� "� !$�� �!���# �����'����"���� #��" � �#��'��������� ��!����"������!$���#��$����& "��"���� %�" ���% � ��=��"�������������!��" ��$��$'�$��% �4��" ���"��"���'���(��"���������p �$ ��/ ��"���������$#�����5�"����4��" ���������$� ����& #"��!� �" �%��9n���� ���#�� ��� #����� � �$&�'�� ��� �'�"q��� =�"����" ���$� #f6� "r�� ��=�� �� "� ��#��'���$� ��8� " ����� "������$#����% (�� ��!����"��������$$�& �%������(��"�#/���"� � ���� "�4���" "'�"��"���'�� ���!�����"�"�� !��� � ��� ��� �� ��� ��$/� "��#�� # ���� ��g � % /�) &����� $"�& "��d @���0� @ 9�����1� 0��� �����$����� ,��������$$�& �%������= �� "�������$���!��� ""�#� �����s����"��������� ��"��#���� @����g ����%'�s���$�� �����$�� ��"��#����t��$� �t�g "��!���"���� s� ��$� ��� ���� ��"��#� �� a�$( ���a�g " ��� s� � ��"�� �� ��/� ����� ����/� #�'/ �"� � ��"��#���������#����������$'g (�$����s�$ "������� ��"��#����#�� �**��2�������3 �& *���3 ��) �!��4 �& �� � �����( �" ������#��'���$������" ��������"�� �"����" ���$���������#�" ��� ��"�������'������$"��������"� ������������� ��� ���� �� ��� �� ��������� �� ����� � ����%� #�������������( �" ������#��'���$� -�"" ���#�-�"" �� ������$���!��( #���!!��# � ���� ��� �"��#��#� �����( �" ��� � � �0"��� (� $ �"���������%� ��#������( �" ������#��'�� ��$�� ���� ��� ����#� �� ���� ��������"���� � � ����#� ���� �$�� �� ���� � �� ���� ��� ���� ����� ��%�&����������8���"'$��*����$ ���� ""��� ���� <����"��#��#� �����( �" ���/� �'���$�/ ��#������'������"����#�� ��#������$$'� � !����"�������&����"��'������ ��"����#� ��"�� "�0" �������( �" ���/��'���$�/���������'�� & "��p ����� �%�����$#�����(� #�# ��c����" �"��"� ���"������ & "�� �����( �" ���/� �'�� ��$�/���������'�� #�!�#������ � ���"�����!�� �#����� ��$$�&�#� �'� 1� �!���� � � =����"� �� ����� ������� "��� ���3���" ��� �� �� ��� ��� ��� �� # ���������� � �� ��� ��� ��������������������������� �������������������������������� "������ % /���$(��/���&�/���#���$$�� ��c����" �'!����"��!��� 0��/����� 0��/�������� � �% ��������$������4� ��#�"���(� #�� ����# �% �$�� �%�4��"�" ������5������$#����!$���# ��"��� !�� �#�� ��� ������/� ��"� ��'� �!!��� � "����������������"����"����!���"��" �� ��# ������� ������ �$ ��� �� � =� � �!$����� ���"'!������ ���������� ��!����� "����� � � �� ��$��� ���"� !����#�� "��� ���� ����#� "���� ��$���� !����#�#� �'� �� (���� �� !��!�� " �� ��<������#� "����������!���"�# & "����� ���$��� ������$��"� "��� �������� ����#� ��� ��� �� !����#�#� �'� )��#, � � �� �0��!$��������������#���� ��� ��% (������ #���).��# �%����#���3������" ���,�����(�� 7�����������" �%�!���%��!���������!$�0$' !���"��"�#���� ��/�!$����"������������"�"�� $��"����% �/�& "��!�� �#����"���"�!����"��� ��� �� �� ��) �!�����$�! �� ��� � �� ��" � �������/� "�$ � ��#���#� ��!����"��� ���/���$$�&�"���� ��"����" �������������� ����/� �0��!"� �� "��� " "$� � � =�� �� ��" � � �����/�"���� ��"�$�""������"���%����� ���!!��� ����/�"������� �#��/���#�"����!�� ������� ����$�&������� ��+��������������������( � �"�#� & "�� "�� �� � ��"� $�""��� &���� ��!��"�# & "� ����!���%��!�/�!��( #�#�"������� �% �����"�����������#�& "�����"����%�����& "� "��������� ��"�$�""��g�� % /�) #�� ���"�"��� �!�� ��� ��$�#�#���#�! ������ ������� ��� ��#� 3��5� ! ��� ���� &� '��� �� ,� � c�� ��" !��( #������!�� �����������$����"��'���� �!��"��"�"��"����"�#' ��c����"� ��$�#��"�� "�0���� �� ��"���f�� ���� � � 6��� )�! ,� "� �# ��"����� �%$����5��&���!�� ��/����)�!! , ����!$���$ ��c����"�% (���� ��" � ���������� #����" ��"�#� �� ��$�� ��� ��$" (�"�#� !$��"� ���$����"���!$��"��!�� ��� ����#�� �/�& #�$' ����!�#��������$" (�" ��/���� ����(�� �"'���" �#�4��"�$'�#���� ��#��'� "�������������� c�� ��"� ��! "�$ ��� ������� �����/� �0��!" &��#��"��"�����!��!��������g�� % /�)�$��5�� ���&��������(��������� ��,����)�$��5����� �(����� ������� ��� , ��������� � ��"����� ���� ���!��� �$� ���� ���� �"���'� ��#� �������'� �� �!�$$ �% ��" �� &��#�/� ��!�� �$$'� �� ��" � �� �����/ ��#� �!�� �$ ��#� "���� � � � "���� 8� " ��� �� ���� �����!�$$ �%� ������!"��$�/���"����" ������#����� �"��"$' ����"���������$#���� "���)�!�$$�����5,����"�����(� $��$�� �����" ���!�"���&��#�!����������"���0�� ���"�� � ������� !"������������'���#����� �"���'��� �!�$$ �%�������� "� ��!� �"�#���#����� ""�# ������( �& ������ �������#�� �� ��5�� � ��"���� ���"� ������� "��"� �$$� !��$ ���#� $ "���"��� � "�#� ��"���"�0"�������������!��# �%�� "�" �� ��"���-�b�-�<�������" �� ����"�������� �$��� ���!��� �$�� ���� "��� �������'� ��� ��� "���f����������#�#�"������!��$ ��" ����� "�# ��"���"�0" ����"�������'�������"���������" ������������������#�"� $������ " �%�$ "���"��� ��"���"�0" 6���"����������#�'�����'�"���"��� "� !��$ ���#� $ "���"���g� � % /� 8�$$��#� � ��d�/ *���������#�>�"������� ��d� �������( �� " ������'�������#� ��"�����"���� �������%�� � ��" ��g�� % /��*<-�� ���� ��6���"���� ��" ��"���f��$��"��������$$�&�#��'�)������,/� � "���������� ���"����g�� % /��"�!������������� � ��d� ��b���� "�" ��������#��'�!����"�����/ #�� ��"� ��!���"�� "��� ��"���� ��#� #�"�� �'� � �����g�� % /��8�$$��#� ��d� ��6��������� "�� ��!���"�� �� ��� ��� ��� � "�" ���� % (��� � !����"��������#�!�"�"����� ��������$�% ��$ ��#��g�� % /���#�� ��� ���/�8�$$��#� ��d� =��� "�" ���� ������� �����(��"��������'���/ �����$!����" ��$���#���& "� ��������$�% ��$ ��#��g�� % /���#�� ��� ���/�. �#�$��%���# >�"������ ���/�7 $"���������� ���/�8�$$��# ��d/�7 $"��� ��d� ��b���� "�" ���� ������� �� & "��p �����������"��"����������"������� � ��1�'����/�% (��"���������������/���!���"� "���'�����& "���������/���#���!���"��"�� � "�" ���� & "�� ��� ��$���g� � % /� �8�$$��# ���/� ��dg� �#�� ��� ���g� 8�$��"� ��dg *���������#�>�"������ ��d� ��6���$�""��� ��/��/��/��"� ��"��# �" �%� �����$" !$��� "�" ��� ���"����������"������� ��"��������'���g�� % / �8�$$��#� ���/� ��dg�����c'5�� ��dg���� ��������������������������������������������� ������������������� � �� ��� ��� �, c'5��������� ��d�/��� ��=��"�����������# ���" 4��"�" ���/� �$���� !���!������/� ��� $��%"�' !��$ ��" ���/�!��( #��"�����"���������#�'���/ "���� ����"� �� ��$��� ��#� % (�� "��� !�%� ���������g� � % /� � ��������� ��#� 8��� � ��d�������������c'5��������� ��d���d���� u<�"�� "��"� "��� � "�" ���� �� "� �� !���%��!� &����#��&�������!��( ���� �������������� ����"���!��!�������!��( # �%��0��!$�����# ������"�$ �"�#� ��"���-�b�-�<�������" �� ���"� ��!�!�� v c������"����"�$�%��#� ����3���$ ����� ��� ��'� ��� � "�#� ��� !��$ ���#� $ "���"���/ ��$�# �%�"��������#�# ����"�" ���/��'�!�� � �� !�����# �%�/� ��#� %�(������"� ��!��"� 6�!��$ ���#� ������" ������!������$����� ��� ��" ��������$#�����(� #�#/���"� ��"��' �������#/�!��( #��"����������#���� $ �" ����� "���������/�"���� �# ��"��"�����"�������"�� � "�" ��g� � % /� �. � � � ��������/� ��" * � � <�" � -�� /� %����� � ����� ��� �* � 7 ���5��"��/���5����#�6� ( /�� %�&�������� � � $��$'/� ������" ��� ��"� ��#� ( �� "�� =�"����"� ����$#� !��( #�� "��� &��� "�� ���� ��#��##����g�� % /��* �7 ����5��"��/������ .���� >�%�/� ���%#))***���'���������) +�����***)����������� � � c�� ��"� !��( #� � "�" ��������"������������� ��"���-�b�-� �<���� ���" �� � � ��" �$��� )���� ""�#,� �� ) ��!��!���" ��,�����$#����� "�#� ��"���"�0" ��$'���#��������#�"�������!��$ ���#� ����� ��" �� ��*������ !"������!"�#�����!��$ ��� " ���� � /�) ��!����,������$#���$'�����������# "��&��������$�"�$'������" �$/���#������ "�# ��� ��"��'�&�����$���#'�!��$ ���#/��� �%�"�� ��" � !�"�#�!��$ ��" ���'��� ��=��"���#�"�� � ��"�5��&�/���!$����"���'����& "��)� �%����,g � % /� �-�#%���� ������/� � �%����� � � ��"���� ���"����!��!���#�"��!��( #���( #�����"��" !�!���� � "�#� ��� ) �� !����,� ��(�� ����� ��� ��!"�#�����!��$ ��" �� ��$����!� ��&��� ��"����� ���� ���!��� �$�� ���� "��� ����� ���'�����$$� ������" ���% (��� ��"���-�b�-� �<���� ���" �� � � -���������� ���"� �� ����5�#�& "���� % ��$���" �$�����#��������� ���"�����������#�"�� ��"���"�0" �����'�����$# ���$ �"�#� ���$!����" ��$���#��������# �%�"� "�����"���f������������ ��b�����$" !$��� "�� " ����& "��"����������"������/�"�����4����� �� ������$�% ��$ � � -���������� �� �� % (�� '����& "��"����������"������/����"����# �� " �%� ���#� �'� $�&�������� $�""���� ��/� �/� �/ �"� � ����"���f����������#� � " �$������"'!�# ���!!��������$�""��� ��=� " �$��������!���"�# �'� �� � �%$�� �!��� � � =�� "��� ��"���� �� �� �%���'/� "��� ��$$� ����� �� !����#�#� �'� "�� �����( �" ������#� ��"���"�0"�� "�" ���& "� � !����"����� ��b���1���"����/���!���"��"�� � ������ & "�� )��#,� � �� $�&�������� $�""���� =�� "����� ���� �� � ��"����/� !����#�� "��� $��" ��"���f�������& "��)/���#,�� ��$�&������� $�""���� � � c�� ��"� �#��"� "��� �����#� ��# �����4���"� $ ���� ��� ���������� � "�" ��� 7�������$" !$��������������& "��"������� ��"������� ���� ���#/� ��!$���� "��� ��"���f� �������� & "�� �� ��$ #� $ ��� ��� d� ������"���/ ��"���"���� ��"�� "�" �� �0��!"��������5����"��� ��" "$�����#� � &��#�3�����$��������� % /��� � �"�/������� ( �" ��������$#�������#�&����(���!��� �$� ��$ "���"����� "�" ��� ����"���������$#����� ��$"��������"� ��������������������� �� �� ������������� ����� �� ���� %� #����� �� �����( �" ���� ���#� �� $ "���"���� � "�" ��� �����( �" ���� ���� ������ ��� !��$ ��" ��� ������$'�� "�#� ������������% (��� �����$� ��-�"" ���#�-�"" �� �����!��( #������!!��� # 0���������( �" �������#�����" "$������!��$ � ��" ��� ��=�� ��#���"/���"���������$#�&� "��"�� ��������"���!��$ ��" ��� ����$$ ��� "$�����# ��$�� (��!�%����������������4� ��#� ������ ��������"��!�!���� ��!�� �# ��$����#����5� � "�" �������������� !"������!"�#�����!��� $ ��" ���� � /�) ��!����,�����������""�#���� � "��'�&�����$���#'�!��$ ���#/���"�), ������, ����$#����"'!�#���"���"���(�$������������� �"�"�����#����"���� "�" �� ��=��"�����" � !�"�# #�"�����!��$ ��" ��� ����"�5��&�/���!$����"�� '���� & "�� ), � �����, � � =�� �� �����������# ��!��"�����"����#������"�& "��$ � "�#�� ���� �� ���������� � �� ��� ��� ��������������������������� �������������������������������� $�" ��� �� � "�#� �� "��� "�0"/� "��� ��������� ����$#� ��$�#������##�����&����� "���'��� ��"� ��# ���0��!$�����������������"'$������ "������"��������$ "���"����� "�" ���� ������� ����% (�����$�& "�#���� ����!���� � ������� $�� ��6 -=��<.����-/�a �� �� ��9 ��7 �"�� ��" ( "'�!�""�������������� �� �"�� ����$��5� t �7 $#$ �*���%� �d@�n1n�n � <�"��������( �" �����������������3���� ��$��������$'�� "�#� ������������% (��� � ���$�� "�#���� ����!���� � �� �����3 &��� ��) 2��# � 7��/�6 ��<+�������/� =�� /� ��� c �%�/��� �� � �!! 8��7� � ��� �'�� 9 �)�����6 w�-/�t �. �� ��� ��8 ��"�" �" ��$����$'� � �����#� �# � � >���" ���.�$$/� =�� / ��%$�&��#��$ ���/�� !�� " �� �% � � ��� � +>�� ����� / b��#�� �"��/�<8 ��d9n�!! �'����� �� *��7� � ������� $�� ��6 b � < w * � < < / � � � 7 / � � � # � � � � ��.7�-�w �� �� ��*���%����"��� <��"�� ���� ���� ������ !�!�$�" ��� >�%���d n�d11�� �� �* �7��������# � 8 �$�%'� ��#� ����%����"� ��� "�� ���( #�� ���� "���� ���=��" "�" ���>����/ 7��� �%"��/�c� <�"���"�"�$�!�%����������������"�% (�� �& ����� ��) !��$����!� ���!��)����� � �� ����� 2��# �6 �-=�����=/� � /� � � t � �*��<�-/� t/ ��#� * � � <�6*���� � �� �"���"� =�" ��'�! �8 �"�$���"�'/�7 $$ ������%/ ��/�*�����19� /� ��d ��� 9�!! <�"���"���!$������#�#�"������"������" �% ���"� ��� !��( #�# � � �����( �"�� &��#� $ 5�� >�����# �%�� �>��� �/� �'�!�� �� ��'�! �/� ��#� �������" ���� ������ � �����( �" �����������" �%�!��$ ��" ��� ������$'� � "�#� �������� ���� % (��� � ���$�� �& ����� ��) !��$����!� ���!��)����� � ��)�2�)#�� ����!��6 -�c+�-�/�� �/��� �-�*>��/���#�a b ��8-�.�* �� ��d ��b �$#�"� �$������ ��&� +>������#� "�$���"�'� �'�"�� >�%��� n � n�� � � � � �� �"�$$ /� � � t ��$����/�t� /���#�* ��<��������#� � 8 �"�$���"�'� m=== � � >��� � �� �"���"� =�" ��'�! �8 �"�$���"�'/�7 $$ ������%/ ��/�*�����19� /� ��d <�"���"���!$������#�#�"������"������" �% ���"����!��( #�#/���"�"�"�$�!�%������ �����������"�% (�� �& ����� ��) !��$����!� ���!��)����� � ��� �$ � �# *���) ������6 ����� � < � �� � *���� ���� �7��5���! �� 1� n � �@ �'����� � �����:� ��) '���6 8����-c/�7 �8 �� �� ��c���%��!� ��/ ��(����"�/� ��#� !��#�" ��� ��"��� �� &�$(��� �� ���"�&��"� �$��5� � � >� c ���� �/�6� ( ��� e���/������� �� n��!! <�"��� ��$�#��"���!��( ��������"�"������ �� "� ����"�!��"����"��� ��" "�" ���" "$� ��2��� ��� �#*��!������ � ������� $�� ��6 8=����/�� ��� �� ���"��#��#����#�%� #�� $ ��������������!�!�$�" ��� �(��"��'� � ��"�� � ����" �* � �<�" �-����� /�7 $#$ 8�����/������"� �� n�!! ��2��� ��� �#*��!������ � ��� �$ � �# *���) ������6 .�2��/��c /�� �* �8�2�-/���#�c �+ ��-��< �� �� ��>�!�$�" ���#'��� �� ��#�!��'���$�" ���� !���������0!$� "�# ��#�����(�� �%�&�$��!�!�$�" ��� ��"�� ���"�����2�5�� ��2�5���b �����#�7 $#$ 8�����/�b ��$�-�! ��-�� � ��9n�!! ��2��� ��� �#*��!������ � ����!& �� �#�'��6 ��*<-�� �<��-=�� *=<=��-2� �b <��6-���-���6-��� �� ��� ��� �� ��������%����"�%� #�$ ��������"���!��� ( � ��������������� "�" ����" �* � �<�" -����� /������"�/��< �� �!! <�"���"��������( �"�#��%���'������ � !����"������!����#���"�����"�������� ��#� ������!��#�� "�� "��� � "�" ��� �� "�� "�0"g� � % /� � "�#� �� "��� "�0"� ��� �*<�( ���������� � �� ��� ��� ��������������������������� �������������������������������� � ����������*<-� ���� #������ !�������� $�� �'� �� � �#�'��+�,� � ������� $�� ��6 8����-c/�7 �8 �� ��1 ��8����!��#�" �� ��� ������� �� ��( �&� ��� �����"� <��"� ���� ���� �"�# ��� ��#� "�� �� ����%�� ���"� �!$ ��" ��� ���$������!!$ � � 91� n9 /� � � � � +�-c<�-/� ��#� � � c *=���� � ��@ � � =��$������ ��� !��#�� "����������������(����"����������� � ���"����"��$� �$��5� � � >��� � < � �� *��������� �7��5���! �� 9� �� d� /���#�c �+ ���-��< �� ��n ��=�!$ � ��" ���� ��� !��#�"���!��'� ��$�" ���� !� "������������%����" ���&�# ���7 $#$ -�� ���!!$ � �d� �9@1 /���#�� �c �*=����� ��@ �������"� ��� ��#�� �%� ���&�� ����� #��� "'� �� ������ ��$�� ���( (�$� �� ���"����"��$ �$��5� ���$����19��� /� /� ��#� t � � � 7.=�*�< ��@ ��8��&����#��$��5������!��#�" �� ��������� �����"����"��$��$��5� ���$��� 19� �� /� � � . � �>-�a�-/� ��#� c � a � > ��2���� �� �����������������"�$ ������ ��$�� ���"�$ "'� �� ���"�����"��$ �$��5� ��t �7 $#$ �*���%� ��d� d� �1 /�t �� �7.=�*��!�$�" ���#'��� ������������ � ���"�����"��$��$��5� ��7 $#$ �*���%� � �����!! (���� ��8 ������ ��"��� ��( � ��/� ����!"�#� ������� !"� ���"� ��� ���� ""�#� �� ���#��!'� �����"� �1 ��! ���� ��#� �� # % "�$� ����� ��� ����� � d� ����� # �5�""� � � c �5�""��� ���"� ��� �$���$' $���$$�#� & "�� "��� ��"����f� ����� � � ��0" ����$#����!��( #�#���"�� �����������&��#� !������ �%������"��=8*����!�" �$�g�!���� ����$'� 7��#>�����"� ��� 7��#�� ��#� ��� �� ���==� � $� � � =#��" �'� "��� &��#�!������ �% ���"&���� ��#� (��� ��� ������� ��� "��� # �� 5�""� ����"���������$#�����5�"������"���� ���"� �����������������"���&��� "��"��#�"��� � ���&� ���&��#�!������ �%����"&�������� ��"������������"$'�����!"�# ���������#��! �� ��#� # �5� � $�� ��� "��� ������� !"� ���"� �� #��" ��$ ������!���"���& $$������"�"�������� ��� !"� �"��"���#���$����$����$�'��"������ � �����"� ������ �������� b���� ��$�!� �" �%����"���3�����$/��$$� $$��� "��" �������"����!��!���#������"$ ��#����(� �����)b %������!" ������#� $$��"��" ���,�/��� !������ ���$�%��!� ���4��$ "'/���#���#���# "��"�����"��$�� e���!!��� �%� ��"���3�����$ � � /� 9n� ��� �� ��� & #�� � � �� �� ���"� �� #�����'�"�����"������ ��=���!��"�������� ��$ "���������>*�����#��" ��������#�����'�"�� 3�����$/�"�����"�������& $$���� �(� ��#�����"�� &��5 ��� � $��$'/���"�������'��������%�# ���� ��$��"���� !��#��" ��� �� ��(���$� !��"�� %��!����������# ��=��� "��������/���'��� % � ��$������ #���#� $$�% �$����"�����#��" ���& $$ ��� ��"����#� ��#� !��$ ��" ��� ��� "��� ����� ��� !"���'����#�$�'�# ������� �� ���� �����%%��"�#��'�-�"" ���#�-�"" �� ����/ ���'�!�!����& "�� �!��"��"�#�"����#� #��� ������"�!��$ ���#������������!����&� " �% �"'$� ���$"���%���# "������'����!�" ��"���# !��( #�� ��$!��$� ��%%��" ���/� ��������� ��� $����"�$����"����!����&� " �% ���� ������$��# "������������ $'���%�" (����( �&� ��=� " �$ ����" ���"��"�������( �&��"��#��"�� ��$�#� "���%�"��$ 5���)�"�! #���( �&���,/�)"��'�# # ��"���#���"��#,/����)"��'�# #���"����#�"�� !�!���������$$', ��8�"�"���/�&��f�����$"� �� " �� "��� ��( �&���� # #� ��"� ��#���"��#� "�� &��5/���#�&��f�����!��� � $ "'� �� "�"����$! "������#���"��#y��=��"����������� ��� �$�#/ "��������"�"���(��'�$���"�"��"�"���!����� � ��"����� � ��"$'��$��� ����!�!��f�������%� ����$#� ��� �$���/� ��� $'� �!!��� �"�#/� ��# ���( �� �% ��*������ !"������$#����# ���" ��#����� �� ����"���������$#������$"������� "��� ���'� ����������� �(� $��$�� ���� �#( ��g � % � ���� ��������"���� ��� ����#������$� �� ���� � �� ���� ���� ���� ���� � �� %�&� ��������������������������������������������� ������������������� � �� ��� ��� �" �������� ��8�� �"'$�� *����$� ���� ""�� ���� ��-�"" ���#�-�"" �� �����!��( #��"��$�� ��� ������� �0!���� ���� & "�� ��!���$���� &��#����#�������$'�� ����#�&��#�/�"��" ���������$!��$� ��&� " �%���!�!�� ������%� #�$ ����!$������%���"�#��$��� ���!��� � $ "'������"���� ��.�&�(��/��#���� ����� "�� "����� %� #�$ ���� �� ��������'� "� ����������� �"���'���#�"���0!�# "��"���!��� $ ��" ���!��������������� ����������!���" �� ������"� ��"���� �������" �$ ��.��;����� ���� 7��& ���"��"���5��#��## ���/�7����� 8�$$��#/�-r����������"� �/�� ������ ��"��/ �$�b���e����/�>�"�a����/�����5����&��"e/ � �#��� �e5��/���#��#���$��������"�� ������ ���"����#���%%��" ������%��# �%�"����"'$� ��#� �����"� ��� ������ ��" �$��/� ��#� ���$ �� #���"�����"�����%� #�$ ��� ��(������� �8�� ��2��� *�<6��� ��**=���� ��� ���� ��" � ���"'$����#������"��"�� �8�� �����$� ���� ��"����/� �# "���/� ��# !��$ ����� ��� 0"���# ������� $����8 �$� �%'��# "���/�=�� /��� ��%�/�=� ���1d�!! -���=/� t � � /� ��#� � � 7 � -���= � � ��� *������ !"�%� #�$ ������������ �� �� �� ��������� �� ����� � � � t � 7 $#$ *���%� �d1�� /���!!$ � ��11�!! alces 31_77.pdf alces34(1)_91.pdf alces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 113 characteristics of post-parturition areas of moose in northeast minnesota amanda m. mcgraw1, ron moen1, and mike schrage2 1natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811-1442; 2fond du lac resource management division, 1720 big lake rd, cloquet, mn 55720 abstract: habitat used 3-4 weeks post-parturition is important to survival of moose (alces alces) calves because neonates are vulnerable to predation, and cows require adequate forage when calf mobility is limited. radio-collared cows were located and visually observed from helicopters from 21 may-5 june, 2004-2007 to identify post-parturition areas in northeastern minnesota that were defined as 100 ha surrounding the cow-calf location. we determined cover type composition in post-parturition areas compared to the 95% kernel home range of moose. buffers of 5, 10, 25, and 50 ha were created around post-parturition areas to determine cover type composition at smaller spatial scales. post-parturition areas were also compared to equivalent areas surrounding cows without calves. post-parturition sites had more lowland conifer and shrubland or regenerating/young forest cover types than random locations within the home range. cows with calves selected areas with larger proportions of lowland conifer, shrublands, and regenerating forests than did cows without calves. these cover types could have been used for cover and for foraging, respectively. there was no difference in the amount of water available in post-parturition areas (3.5% ± 0.8) when compared to home ranges (3.5% ± 0.8). distances between consecutive post-parturition locations (1.7 ± 0.4 km) were less than expected when compared to distances to random points within the home ranges (3.3 ± 0.4 km). alces vol. 47: 113-124 (2011) key words: alces alces, calves, habitat, moose, minnesota, post-parturition. moose (alces alces) calving site studies have typically been conducted by searching for parturition sites 3-4 days after birth (leptich and gilbert 1986, langley and pletscher 1994, bowyer et al. 1999). identifying parturition sites is important because calf mobility is limited for the first 3-4 weeks after birth (altmann 1958, 1963). further, cows occupy a post-parturition area within their home range where the cow-calf pair lives for 3-4 weeks. this area is used during the period when calves are most vulnerable and presumably it facilitates calf recruitment into the adult population. identifying habitat characteristics of post-parturition areas of moose in northeast minnesota is important given that recruitment rates are currently declining (lenarz et al. 2011). cow moose give birth at sites that provide some hiding cover but do not necessarily have the highest quality or quantity of forage available (leptich and gilbert 1986, langley and pletscher 1994, bowyer et al. 1999). this is often interpreted as a trade-off between avoiding predators and meeting nutritional requirements (bowyer et al. 1999), and may be important to consider as influencing choice of calving habitat in minnesota where black bear (ursus americanus) and wolves (canis lupus) occur. the variability in vegetative cover and density, visibility, and proximity to water has made describing calving sites difficult (addison et al. 1990) because births may occur from hilltops to islands (wilton and garner 1991, addison et al. 1993, chekchak et al. 1998). undisturbed lowland areas dominated by conifers and near water were associated with calving in maine (leptich and gilbert 1986), as were areas with mature, mixed, and moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 114 coniferous forests when water and islands were not available in new hampshire (scarpitti et al. 2007). a boreal forest mix is the matrix from which moose in northeast minnesota choose a parturition site. important habitat types in the home ranges of moose are young mixed conifer and deciduous forests, including aspen (populus tremuloides), paper birch (betula papyrifera), and balsam fir (abies balsamea). early successional forests (11-30 years post disturbance) are used because forage is within reach of moose (kelsall et al. 1977). summer ranges consist largely of black spruce (picea mariana) lowlands as well as uplands and cut over areas dominated by paper birch, aspen, and balsam fir (peek et al. 1976). in early summer, moose generally use upland, lowland, and plantation areas in proportion to their occurrence (peek et al. 1976). if parturition sites are chosen from within a cow’s home range, limits exist as to where a calf can be born, and the availability of suitable habitat for parturition could be important. further, characteristics of the larger post-parturition area used by the cow-calf pair during the following 3-4 weeks have not been studied in detail. the objective of this study was to identify and describe post-parturition habitat of cow moose in northeastern minnesota. study area lake and cook counties (47°30’n, 91°21’w) in the arrowhead region of northeastern minnesota are part of the northern superior uplands (fig. 1) (minnesota dnr [mndnr] 2010). the southern boundary of the northern superior uplands coincides with the boundaries of the canadian shield as it extends into minnesota. upland vegetation consists of fire-dependent forests dominated by a mix of white (pinus strobus) and red pine (p. resinosa), aspen, paper birch, white spruce (picea glauca), balsam fir and white cedar (thuja occidentalis). jack pine (p. banksiana) stands and conifer swamps of tamarack (larix laricina) and black spruce are also present. northeast minnesota has a humid continental climate with cold winters and warm summers. precipitation occurs as snow (180 cm annually) with snow cover typically in december-april, and rain (70 cm annually) of which 40% occurs during the growing season. methods adult female moose (n = 36) were captured beginning in february 2002 and fitted with radio-collars (lenarz et al. 2009). these moose were monitored weekly for mortality from february 2002-march 2008. annual 95% kernel home ranges were based on locations from radio-telemetry flights. the average home range size of cows used in this study was 40 ± 5 km2 (x ± se), ranging from 8-312 km2 (moen et al. 2011). calving begins in northeast minnesota and ontario around 10 may, with peak calving during the following 3 weeks (addison et al. 1993, bowyer et al. 1998, lenarz et al. 2005). the parturition site was unknown because it was not known when or where calves were born prior to the time they were sighted from the helicopter during the survey period from 21 may-5 june. we assumed that if calves were present, they were observed. if no calves were observed, it was unknown whether the cow was barren, or had given birth and already lost the calf. we assumed that cows without calves (whether barren, still pregnant, or lost calf) would not operate under the same impulses as cows with calves, and therefore would exhibit different habitat preferences. because cows restrict movement after birth (poole et al. 2007), we assumed the parturition site was near the post-parturition location of the cow-calf pair. we defined the post-parturition area as 100 ha (565 m radius) surrounding the post-parturition location (poole et al. 2007). we used 100 ha as the post-parturition area because we did not know exactly when or where calves were born, but alces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 115 a 100 ha area around the point location likely included much of the post-parturition area used by the cow-calf pair (poole et al. 2007). helicopter flights from 21 may-5 june, 2004-2008 were used to locate radio-collared cows to identify presence of calves. if a cowcalf pair was observed, a waypoint was taken to establish the center of the post-parturition area; those cows were not relocated during subsequent search flights. if no calf was observed in the subsequent flight, it was assumed that the cow was barren or the calf was lost (lenarz et al. 2010). for these cows, the location from the first flight was used as the point location for a cow without calf. cows without calves were analyzed separately and compared to cows with calves. the position error of cow-calf locations was estimated as <100 m based on past experience (m. schrage, unpublished data). post-parturition habitat composition cover types of post-parturition areas were identified within the home ranges of cows using 2 independent satellite imagery classification systems; the gap analysis program (gap), level ii and the land use land cover (lulc) classification system. both are raster datasets derived from landsat thematic mapper (tm) images with 30 m resolution (mndnr 2007). gap level ii classifies 10 different cover types and lulc defines 16. we used the 2 coverage datasets available in minnesota that had the highest accuracy and similar land cover classifications. the gap coverage data was collected in 1991-1993 while lulc was collected in 1995-1996. because of the elapsed time between the 2 coverage datasets, and because both gap and lulc have similar cover type classifications, it was important to check for consistency between them. some forest harvest occurred within the study area since the gap and lulc coverage data were collected, but other major disturbance was limited. the study moose were south of a large blowdown in the boundary waters canoe area wilderness in 1999, and south of 2 large fires in northeast minnesota in the past 20 years. we first identified habitat characteristics of post-parturition areas and then tested with anova for variability of cover type composition near the post-parturition area at the 100 ha scale by generating 16 additional points ## ## # ## # # # ### # # ## ## # ### ## # # # # # # ## # # # # # ### # # # # # # ### # ### # # ### # # # ## # # # # # # êú êú êú êú êú finland isabella duluth silver bay two harbors st. louis county lake county # # cook county 0 30 60 90 kilometers ## ## # ## # # # ## # # # # # # # # ## # # # # # # # # # # # # # # # # ### # # # # # # ## # # # ## # # # # # # # # ## # # # # # # êú êú êú finland isabella silver bay st. louis county lake county # # counties home ranges # cow/calf locations êú cities 0 30 60 kilometers n ew s fig. 1. home ranges of radio-collared cows are outlined in black within the arrowhead region of northeast minnesota (moen et al. 2011). dots indicate cow-calf locations obtained during post-parturition helicopter flights, 2004-2008. moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 116 within the defined area. the 16 points radiated from the center of the known cow-calf location at 90º angles and 100 m intervals within the post-parturition area. these 16 points were also buffered to 100 ha. to compare cover type composition of post-parturition areas to the home range, 25 random points were generated within each home range using arcview 3.3. buffers of 100 ha were applied to each random point and represented potential post-parturition areas. to test whether cows with calves selected for post-parturition cover type characteristics at finer spatial scales, buffers of 5, 10, 25, and 50 ha were created around all points and locations, and each set of random points was compared to the cover type composition within post-parturition areas with anovas. the estimated position error was used to set the smallest buffer size for characterizing cover type composition to 5 ha (126 m radius). we also estimated cover type composition when cows were not observed with calves during flights. location of cows without calves and random locations within their respective 95% kernel home ranges were examined at the same spatial scales as those of cows with calves; differences in cover type composition were determined with anova and χ2. water we measured water bodies classified as lakes, rivers and streams, beaver ponds, and other available water within the 100 ha area surrounding each location on screen using farm service administration (fsa) color orthophotos from 2003-2004. the fsa photos were used because the satellite imagery data did not include fine scale water features that may be important to moose. we randomly distributed 200 points across the spatial extent of the composite home ranges of all cows. water bodies within 100 ha surrounding each random point were used to estimate overall water body availability within the study area. water body type and availability within the post-parturition areas were also compared (t-tests) to water body type and availability within 100 ha around locations of cows without calves. we also compared straight-line distance to water between cows with calves and cows without calves. distance was measured from the known location of the animals to the nearest water body using fsa color orthophotos from 2003-2004. finally, we compared distance between consecutive post-parturition locations to the distance from post-parturition location to 30 random points within the home range to test whether consecutive post-parturition locations were closer together than would be expected of a random distribution. statistics cover types within the post-parturition areas and within home ranges at different buffer sizes were compared with anova and linear regression using statistix (version 9.0; analytical software, boca raton, florida). post-parturition areas of cows with calves were also compared to areas surrounding the location of cows without calves using a χ2 test for independence. the amount of different water types were compared using anovas. differences in percent water type available between post-parturition areas and random areas were compared with t-tests. means are presented throughout as x ± se.we used excel 2007 (microsoft corp., redmond, wa) for t-tests and χ2 tests. results habitat composition the 100 ha post-parturition area covered 4 ± 6% of the average home range. in both the gap and lulc only 4 cover types comprised >90 ± 7% of the area within home ranges (table 1). gap cover type categories covering >90% of the area included lowland conifer, shrublands, upland conifer, and upland deciduous forests. in lulc the 4 cover types alces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 117 covering >90% of the area were conifer and mixed forests, regenerating/young forests, and bog (fig. 2). cover type classifications for gap and lulc describe similar habitats with different names. we identified 70 post-parturition locations, and all were in the home range of the maternal cow; certain cows had multiple births during the study. cows with calves selected areas with more lowland conifer and shrublands (gap), or conifer and bog (lulc) than cows without calves at all spatial scales we examined (χ2 2 >5.9, p <0.004). the selection of lowland conifer and shrublands (gap) and conifer and bog (lulc) was stronger as the buffer surrounding the post-parturition location was reduced from 100 to 5 ha. cows without calves had less lowland conifer and shrublands and more upland conifer and deciduous in the gap coverage (fig. 3), and had more mixed forest and regenerating/young forests and less conifer and bog in the lulc coverage. cover type composition did not change as areas surrounding post-parturition and random locations were reduced incrementally from 100 to 5 ha (f2, 70< 0.7, p ≥0.69). lowland conifer was about 22% of the area in the random locations compared to 24-30% of the postparturition areas in the gap coverage (fig. 4a). similarly, in the lulc conifer was about 18% of the area in random locations compared to 17-21% of the post-parturition areas (fig. 4b). mean area of lowland conifer increased as the buffer around the post-parturition location was reduced from 100 ha to 5 ha, but differences were not significant because of high variability among cows. the contrast between random locations and post-parturition locations suggests that some non-random actions were occurring associated with the lowland conifer cover type. upland deciduous forest may be negatively correlated with gap percent home range cover lulc percent home range cover lowland conifer 22 ± 0.6 conifer 18 ± 0.5 upland conifer 13 ± 0.4 bog 17 ± 0.7 shrubland 19 ± 0.5 regenerating/ young forest 14 ± 0.6 upland deciduous 42 ± 0.8 mixed forest 42 ± 0.8 sum 96 ± 2.6 91 ± 7.0 table 1. cover type composition of the home ranges of 36 radio collared moose in northeast minnesota using the gap and lulc classified cover types from landsat satellite imagery. fig. 2. cover type composition in 100 ha areas around random areas in the home range and in 100 ha areas around the post-parturition locations using gap and lulc datasets, northeastern minnesota. moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 118 lowland conifer (p = 0.051, n = 70), however the increase in lowland conifer (r2 = 0.006, p = 0.12) and subsequent decrease in upland deciduous (r2 = 0.0014, p = 0.48) were not correlated as buffer size declined. we divided cows with calves into 3 groups based on lowland conifer within postparturition areas compared to lowland conifer within home ranges (fig. 5). the 3 groups were defined as to whether lowland conifer in the post-parturition area was above, within, or below the 95% confidence interval of the random potential post-parturition areas within the home range of each cow. the variability among groups indicates why we found no significant differences in mean value of lowland conifer between random and post-parturition areas. within the group with higher than expected lowland conifer in the post-parturition area (41% of cows), there was a subgroup (5 of 29) with >80% lowland conifer in the postparturition area while home ranges had about 20% lowland conifer available. water cows with calves were closer to water than cows without calves (p = 0.011) (fig. 6). the amount of water bodies in post-parturition areas was not different from the amount in random areas distributed across the study area (fig. 7). lakes, rivers and streams, beaver ponds, or other water were present in 80% of post-parturition areas, 60% of random areas, and 70% of areas surrounding cows without calves. water covered about 3.5 ± 0.8% (x ± se) of the area within 100 ha areas randomly distributed throughout the spatial extent of all home ranges. this was the same as the 3.5 ± 0.8% water measured within 100 ha post-parturition areas. cows without calves had half as much water within 100 ha (1.6 ± 0.5%); the largest difference being the absence of lakes. post-parturition areas and random areas consisted of approximately 2.5 ± 0.8% lakes, while areas around cows without calves were about 1.5 ± 0.3% lakes. beaver ponds were present in 27% of 100 ha post-parturition fig. 3. change in cover type composition as the area surrounding known cow-calf locations and locations of cows without calves was incrementally reduced from 100 ha to 5 ha, northeast minnesota. vertical error bars represent 95% confidence intervals. alces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 119 areas, 18% of 100 ha random areas, and 37% of 100 ha areas surrounding cows without calves. beaver ponds were more common in areas surrounding cows without calves than in random areas (χ2 2 = 8.48, p = 0.0036). distance between post-parturition sites of 36 cows observed with calves, 23 gave birth in ≥2 years, resulting in 34 paired parturition events. post-parturition locations in consecutive years were closer to each other than distances between post-parturition locations and random locations in the home range (observed: 1.7 ± 0.4 km, n = 34; random: 3.3 ± 0.4 km, n = 34) (fig. 8). the minimum distance between consecutive post-parturition locations was 39 m and the maximum was 4,333 m. many (34%) post-parturition locations were within 1 km, and 60% were within 2 km of the previous year’s location. b) a) fig. 4. change in cover type composition as the area surrounding the known cow-calf locations (ppa) and random locations were incrementally reduced from 100 ha to 5 ha in the gap cover type (a) and the lulc cover type (b), northeast minnesota. moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 120 fig. 5. lowland conifer in the 100 ha post-parturition area was compared to lowland conifer in the home range of moose cows in northeast minnesota. the 1:1 line would indicate similar proportions of lowland conifer in the post-parturition area as in the home range. cows either selected for lowland conifer (triangles), used lowland conifer in accordance to its availability (squares), or avoided lowland conifer (diamonds). a subset of cows had post-parturition areas with a very high proportion of lowland conifer relative to the home range (circled triangles). fig. 6. frequency distribution of distance of cows with calves and cows without calves to water bodies; distance was determined using aerial photograph interpretation. water bodies included lakes, rivers, beaver ponds, and other water in northeast minnesota. alces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 121 discussion post-parturition areas had more lowland conifer and shrubby areas with a water component than areas used by cows without calves in gap coverage at all spatial scales. lowland conifer would likely provide hiding cover that is important in other regions with predators (stringham 1974, leptich and gilbert 1986, langley and pletscher 1994, bowyer et al. 1999). it is suggested that cover types providing better foraging and less hiding cover are used more by cows without calves. because the cows were visually located only 1-2 times each survey period, we cannot be certain that cows without calf had not previously lost a calf, were still pregnant, or barren. we found high variability among cows in use of cover types after calving, which is consistent with previous studies. despite this high variability, we identified trends in cover type use by cows fig. 7. amount of each water body found within post-parturition areas, random areas, and areas surrounding cows without calves in northeast minnesota. fig. 8. distance between consecutive post-parturition locations of moose cows (circles) compared to average distance to 30 random points within their home ranges (squares), northeast minnesota. vertical error bars represent 95% confidence intervals of distance to random locations within the home range. moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 122 with calves and cows without calves. defining and identifying specific characteristics associated with the parturition location and post-parturition area is important because this area must support the cow-calf pair for at least 3 weeks. based on home range measured for these study animals, cows with calves use 1-4% of the home range for 5-10% of the year when calf mobility limits movement. we used a post-parturition area of 100 ha around the calf-cow location (poole et al. 2007) to compensate for not knowing the actual parturition location. the 100 ha area used by poole et al. (2007) was the minimum convex polygon encompassing locations of cows with calves for approximately 9 days post-parturition. using this definition of post-parturition area, cows were variable in cover type use. yet lowland conifer tended to increase at smaller spatial scales compared to random locations. possible factors contributing to this variability could be the age of calves (unknown) or how cow-calf movements change during the first 3 weeks post-parturition. the cow-calf pairs found in areas with >75% lowland conifer (gap coverage) during the first week of search flights may have represented neonates and preferred birth sites. when we identified areas with high levels of lowland conifer across the home ranges, most post-parturition areas contained this habitat. if lowland conifer is used for post-parturition habitat, then post-parturition habitat is probably not limiting within home ranges of cows in northeast minnesota. however, because >90% of the 70 cow-calf pairs were found during the first week of flights and parturition dates are unknown, this was a weak test for cover type selection. if cow-calf pairs had been relocated and seen during the second week of flights, a stronger test of cover type selection against cow-calf use would have been possible. reports from observers in the helicopter indicated association of cow-calf pairs with beaver ponds, wetlands, small lakes, and rivers during post-parturition flights. the amount of water in defined post-parturition areas, in simulated post-parturition areas, and in 100 ha buffers around cows without calves was relatively low (generally <3.5%); however, cows with calves were generally observed nearer these small water features than cows without calves. while moose have used islands as calving sites in other areas (peterson 1955, bailey and bangs 1980, stephens and peterson 1984, addison et al. 1990, wilton and garner 1991), there were few islands available in the study area. each post-parturition location was observed on an aerial photograph, and these locations were not identified as islands in any habitat type, including islands in bogs. different degrees of calving site fidelity have been defined in multiple calving studies. in algonquin provincial park in ontario fidelity was observed in the repeated annual use of calving areas on islands (addison et al. 1990). cows in managed forests following moose habitat guidelines in ontario had higher degrees of parturition site fidelity (<3 km between consecutive parturition sites) relative to progressively clear-cut forests (4.87 km). parturition sites within 1 km of the site used the previous year were used by 25% of cows (n = 35) (welch et al. 2000). using similar criteria to compare distance between consecutive post-parturition locations, we found 60% of cows were within 2 km of the previous year’s location and 34% were within 1 km. even with potential movements of cows away from parturition sites in the weeks following parturition, and without knowing the exact parturition location or age of calves, distances between post-parturition locations of individual cows in consecutive years were closer than expected from a random distribution. despite these sources of variation, an element of fidelity to a post-parturition area was still observed and may have a basis in cover type composition. the study was designed originally to acquire a sample of newborn calves to folalces vol. 47, 2011 mcgraw et al. – moose post-parturition areas 123 low to adulthood to obtain an estimate of calf survival; therefore, helicopter flights occurred after peak calving to maximize the number of observed calves. however, this resulted in a range of age among calves, and likely introduced some of the variability in cover type composition and presence of water. a study employing gps technology and vaginal implants or more frequent visual observations of cows during the calving period would provide more precise measurements of calving sites, timing of birth, post-parturition movements, and habitat use. management implications at this point the variability observed among cows indicates that the presence of multiple cover types in close proximity to one another may be an important characteristic of post-parturition areas in northeast minnesota. the most common cover types within the home ranges were also important post-parturition habitat. this indicates that moose may adapt to local conditions when selecting post-parturition areas. cows with calves tended to use more areas of lowland conifer, shrubland, and bogs and were nearer water than those cows without calves. lowland conifer, shrubland, and bog cover types are among the most common within the home ranges of radio-collared cows, after mixed forests. availability of lowland conifer, shrubland, and bogs is likely not a limiting factor for moose in northeast minnesota; however, managers should consider these cover types and the presence of water when planning timber harvests or road construction. acknowledgements we thank mndnr pilot john heineman and observers lance overland, andrew edwards, charlie nahgahnub, john erb, john mcmillan, and angela aarhus. funding for this work was provided by the tribal wildlife grants program, the fond du lac band of lake superior chippewa, the mndnr, the university of minnesota duluth and the natural resources research institute. summer support for the senior author was provided by the integrated biosciences graduate program, university of minnesota duluth. this is contribution number 531 from the center for water and the environment at the natural resources research institute, university of minnesota duluth. references addison, e. m., r. f. mclaughlin, d. j. h. fraser, and m. e. buss. 1993. observations of preand post-partum behavior of moose in central ontario. alces 29:2733. _____, j. d. smith, r. f. mclaughlin, j. d. h. fraser, and d. g. joachim. 1990. calving sites of moose in central ontario. alces 26:142-153. altmann, m. 1958. social integrations of the moose calf. animal behavior 6: 155-159. _____. 1963. naturalistic studies of maternal care in moose and elk. pages 233-253 in h. l. rheingold, editor. maternal behavior in mammals. wiley & sons, new york, new york, usa. bailey, t. n., and e. e. bangs. 1980. moose calving areas and use on the kenai national wildlife range, alaska. proceedings of the north american moose conference and workshop 16: 289-313. bowyer, t. r., v. van ballenberghe, and j. g. kie. 1998. timing and synchrony of parturition in alaskan moose: long term versus proximal effects of climate. journal of mammology 79: 1332-1344. _____, _____, _____, and j. a. k. maier. 1999. birth-site selection by alaskan moose: maternal strategies for coping with a risky environment. journal of mammology 80: 1070-1083. chekchak, t, r. courtois, j. p. ouellet, l. breton, and s. st-onge. 1998. charac-1998. charac-characteristics of moose (alces alces) calving moose post-parturition areas – mcgraw et al. alces vol. 47, 2011 124 sites. canadian journal of zoology 76: 1663-1670. kelsall, j. p., e. s. telfer, and t. d. wright. 1977. the effects of fire on the ecology of the boreal forest, with particular reference to the canadian north: a review and selected bibliography. occasional paper 323. canadian wildlife service, ottawa, ontario, canada. langley, m. a., and d. h. pletscher. 1994. calving areas of moose in northwestern montana and southeastern british columbia. alces 30: 127-135. lenarz, m. s. 2011. 2011 aerial moose survey. minnesota department of natural resources, st. paul, minnesota, usa. (accessed march 2011). _____, j. fieberg, m. w. schrage, a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013-1023. _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2005. moose population dynamics in northeastern minnesota. minnesota department of natural resources summaries of wildlife and research findings 34-38. _____, _____, _____, and_____. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. leptich, d. j., and j. r. gilbert. 1986. characteristics of moose calving sites in northern maine as determined by multivariate analysis: a preliminary investigation. alces 22: 69-81. minnesota department of natural resources (mndnr). 2007. ecological classification system. minnesota department of natural resources, st. paul, minnesota, usa. (accessed february 2011). _____. 2010. (accessed february 2011). moen, r., m. e. nelson, and a. edwards. 2011. using cover type composition of home ranges and vhf telemetry locations of moose to interpret aerial survey results in minnesota. alces 47: 101-112. peek, j. m., d. l. urich, and r. t. mackie. 1976. moose habitat selection and relationships to forest management in northeastern minnesota. wildlife monographs 48: 3-65. peterson, r. l. 1955. north american moose. university of toronto press, toronto, ontario, canada. poole, k. g., r. serrouya, and k. stuartsmith. 2007. moose calving strategies in interior montane ecosystems. journal of mammology 88: 139-150. scarpitti, d. l., p. j. pekins, and a. r. musante. 2007. characteristics of neonatal moose habitat in northern new hampshire. alces 43: 29-38. stephens, p. w., and r. o. peterson. 1984. wolf-avoidance strategies of moose. holarctic ecology 7: 239-244. stringham, s. f. 1974. mother-infant relations in moose. naturaliste canadien 101: 325-369. welch, i. d., a. r. rodgers, and r. s. mckinley. 2000. timber harvest and calving site fidelity of moose in northwestern ontario. alces 36: 93-103. wilton, m. l., and d. l. garner. 1991. preliminary findings regarding elevation as a major factor in moose calving site selection in south central ontario, canada. alces 27: 111-117. alces vol. 46, 2010 mcgraw et al. – moose advisory committee 189 an advisory committee process to plan moose management in minnesota amanda m. mcgraw1, ron moen1, grant wilson2, andrew edwards3, rolf peterson4,louis cornicelli2, mike schrage5, lee frelich6, mark lenarz7, and dennis becker6 1natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, mn 55811, usa; 2minnesota department of natural resources, division of fish and wildlife, 500 lafayette road, st. paul, mn 55155, usa; 31854 treaty authority, 4428 haines road, duluth, mn 55811, usa; 4school of forest resources and environmental science, michigan technological university, houghton, mi 39931, usa; 5fond du lac resource management division, 1720 big lake road, cloquet, mn 55720, usa; 6department of forest resources, university of minnesota, st. paul, mn 55108, usa; 7minnesota department of natural resources, 1201 east highway 2, grand rapids, mn 55744, usa. abstract: c������ ���� ��� ������� �� ����� �� m�������� ��� �� � l���������� s������ l�� ����c������ ���� ��� ������� �� ����� �� m�������� ��� �� � l���������� s������ l�� ���� ������ ���� ��� d�p������� �� n��u��� r���u���� (dnr) ������p � m���� m��������� ��� r������� p��� (mmrp). p���� �� ������p��� ��� mmrp, ��� dnr ��� ��qu���� �� ���� � m���� a������y c�������� (mac). t�� mac ��� 8 ����� ���� au�u�� 2008�ju�y 2009 ��� p������� ���������� ��� �������� ��������������� �� ��� dnr �� � 45�p��� ��p��� ������b�� �� ��� ��������. t��� p�p�� details the mac process and serves as a reference for agencies that find themselves in a similar man� agement circumstance. procedural decisions, planning needs, and development of the final report are ����u���� ������. alces vol. 46: 189�200 (2010) key words: alces alces, m��������, �����, m���� a������y c��������, m���� m��������� p���. t�� 2008 m�������� l�������u�� ����u�� s������ l�� (c��p��� 368, s������ 76) �������� ���� ��� m�������� d�p������� �� n��u��� r���u���� (dnr) ������� � m���� a������y c�������� (mac) �� ��k� ����� (alces alces) ���������� ��� �������� ������ ���������� �� ��� �����y. t���� ��������� ������� ���� b� u��� by ��� dnr �� ������p � ������������y �������� m���� m��������� ��� r������� p��� (mmrp) ������� by ��� dnr p��������. t�� pu�p��� �� ���� p�p�� �� �� ������b� ��� mac p������, p������ ������� �� �� ��� ���u�� ���� ���u��u���, ��� �������y p��b���� ��� p�������� ��p��������� ���� ��� p������. r�������������� ���� by ��� mac ��� ����u��� ���� ��������y �� p������ ��� ������ ���� ���qu��� b��k���u�� �� �����p��� ���������� ��� ������������� �� ��� p������. t�� b��������� ��� ���������� b���� ��� ��������������� �� ����u���� �� ��� mac ��p��� (p������� �� ��. 2009) ��� ���� b� �u����� ��������� �� ��� mmrp. a�� �u����� �� ���� p�p�� ���� ���b��� �� ��� mac, ���� ��� �x��p���� �� ��� ���� �u���� ��� ������ �� ��������y �� ��� mac. h��������� ������� �������� ���� ����� ���� b��� p������ �� �pp�����b�� �u�b��� �� �������� m�������� ����� b����� 1885 (p��k �� ��. 1976). a� ��� ������ ��� �������, u���� �������� �u����� ��� ��b���� ������� ���������� ���� ������� ���u��� ����� �u�b��� �� ���y ��� ������, ���u����� �� ����u�� �� ��� ����� �u����� ������ �� 1922. m���� p�pu������ �u���y� ���� b��� ����u���� ����� ��� 1920�, ���� ������ �u���y� ��� p�����y ���������� ������ ����� 1959 (k���� 1982). m���� ��������� ���� ��� 1930��1970� ��� ������ 2 ���ju������ p�pu������� ������� �� ��������� ��� ��������� m�������� (f��. 1). a b������� ����� �u����� ������ ��� ����b������ �� 1971, ��� ������� ���� ������� ��p������y ��� ��� ��������� ��� ��������� p�pu������� moose advisory committee – mcgraw et al. alces vol. 46, 2010 190 (mndnr 1990); ��� �u����� ������ ��� ������� ���� b������� �� ���u�� �� 1993. t�� �u����� ������ �� ��������� m�������� ��� ������ �� 1997 b���u�� �� � �������� ������� �� ��� p�pu������. m���� �u����� ��� ������u�� �� ��������� m�������� bu� ������� by �������������� �u����� ��� ���������� �� bu������y �� 2007. t�� p�����u� ����� ���������� p��� ��� ������� by ��� dnr d������� �� f��� ��� w������� �� 1990. t�� ��������� ����� p�pu������ ��� ��������� �� ���� ��������� ���� 2,631±989 �� 6,558 ±3160 (���� ± 90% ci) ���� 1971�1986; ��� ��������� p�pu������ ��� ��������� �� �b�u� 4,000 ������� �� 1990 (m�������� m���� m��������� p��� 1990). i� ��� 1990 p���, ��� dnr �������� ������ ��z�� ���� ����� p�pu������� ��� ��������� ����u���u� ��� ����1900� ��� ���� p�����u� ������ ����� ��u�� ��k��y ��� ������u� �� ���� u�� ��� ��b����� �������. t���� �bj������� ���� �u������ �� ��� 1990 p��� b���� �� p�pu������ ��������� ���� 1986: 1) ���������� ����� p�pu������� 15�20% by 1992, 2) ������u��� � ����� ������� �� � b�� ������ b����, ��� 3) ���������� �pp���u������ ��� ��������u�p���� u�� �� �����. s��������� �� ������� ����� p�pu������ ����� ����u��� habitat management efforts specific to each ������, pub��� ��u������, ��� b�������� ����� ������� ���� ���������� �u���� ������. a���� ������p���� �� ��� 1990 p���, ��� ��������� ����� p�pu������ (~ 4000) �������� �� �b�u� 100 ������� �� 2007 (l����z 2007); �u����� �� ��� ��������� ���� ��� ���������� u�� ����� ��� 1996 ������. m�������y �� ��� ��������� p�pu������ ��� ���������� p�������y ���� p�������� ��� ��������u� �������� (mu���y �� ��. 2006). d��p��� ��� ��������� �� �u����� ��� ��������� ��b���� ���������� �������� to benefit moose, the population in northwest m�������� ��� ������ ����pp�����. t�� ������� �� ��� p��� 20 y���� ��� �������� ���������� ���� � ������� ���p����u�� ����� (mu���y �� ��. 2006). t�� ��������� ����� p�pu������ ��� ��������� �� 7,593±1761 (���� ± 90% ci) �� 2009, ��� 5,528±1318 �� 2010 (���� ± 90% ci). o���� ���� ���� �u����� ���� ���� p�pu������ �� ���������; ��� �x��p��, ��� p��� p������ �� ������ ��� ����:��� ������ �b������ �u���� �u���y� ���� �������� �������y ����� 2001 (l����z 2009�). a vhf ��������������y p��j��� �� ��������� ��� ���� �� �����u����� ��������y ��� ��������� �� 2002 b���u�� �u������ ����� ��� ��u��� �� ��������y ���� u�k����. by 2008 ���� �������� ��������� ���� ��� �����u����� ��������y ���� �� ��u�� bu��� ��� ���� ��� �ub���������y ������ ���� ��p����� ��������� �� n���� a������, ��� ��� ������� �� ��� ���� ��������y ���� �� ��u�� ���� �u���� ��� ������� �� ��������� m��������. m������� �� �������y ��� ��������y ��������� ���� ��� p�pu������ ��� ��������� �� ������� �� 15% ���u���y (l����z �� ��. 2010), y�� u���� 2010, ��� p���� �������� �� ��� ���u�� �u���y ��� ��� �������� �������. a� �� ��� ���������, �u�� ��������y �� ��u�� ����� �� ��� ��������� �pp����� �� b� �������u����� ��� p��b�b�y ������ �������. i������ �� j��u��y ��� ���� f��. 1. m���� ����� �� m��������, 1965�2009. d��k ���y ��� ����� ���y ����� �������� ���� ��� ��� ����� p�pu������ ������y, ���p�������y (l����z, mndnr, u�pub������ ����). alces vol. 46, 2010 mcgraw et al. – moose advisory committee 191 �p���� ���p����u��� ���� ���������� ���� ��� �����b����y �� �������� ��� ���u�� �u������ �� ��u�� ����� (l����z �� ��. 2009). hu����� �� ��� ��������� ���� ������u�� by b��� �������������� ��� ���b�� �u�����. s������������� �u����� ���� ��������� �� ������� �� 109 bu��� ���u���y ����� ��� dnr ���p��� bu������y �u����� �� 2007 (l����z 2009b). t��b�� ������� ��� �������� �b�u� 45 �����, �����y bu���, �� ��� ���� p����� (m. s������, f��� �u l�� r���u��� m��������� d�������, p���. ����.). hu����� �u����� ��� �������������� �u����� ��� �������y �������� ���� >80% �� 2001 �� 48% �� 2008 (l����z 2009�, b); �u����� ���� �� ���b�� �u����� ��� ���� �������� (m. s������, f��� �u l�� r�� ��u��� m��������� d�������, p���. ����.). h������ qu���� ��� �������������� �u����� ��� �u������y ���y ������������, �������� �� bu���, ��� ������ �b�u� 2% �� ��� ��������� ���� ��u�� p�pu������ (p������� �� ��. 2009). t�� ������� �� ��� ��������� p�pu������ ��� ������� ��� ���� �� ��� ��������� ��� �� ��������� pub��� ��������� ��� ������ ��� ����������. t�� dnr ��� ����� p������� ��� ������y ��������� ������������� ��� ���u�����y ������ �� ������� ����� ��������. h������, the intensified concern for the future of moose �� m�������� ��� �� ����������� ��������� ��� �u����� �����y ������. du���� ��� ������ �� 2007�2008 ��� m�������� d��� hu����� a�� ��������� �p��������� ����������� ����������� ���� ���y �xp������ ������� ���� ��� ������� �� ����� �� ��� ��������� ��� p����b�� ������� �� ��� ���������. i� ���p����, ��� l�������u�� �������� ��� dnr �� ������p � p��� ������� �y��� ���� p������y ���������� ��� �������� ����� ��� �����, ��� �� p������ � p������� ��p��� �� ��� l�������u�� by 15 j��u��y 2009 (2008 mn l�������u�� c�.368, sf 2651, a�� ����� 2 s������ 76). t�� ��� ���� �������� ��� dnr c����������� �� ���� � ��������� �� ����� �xp���� �� p������ ���������� ��� �������� ��������������� �� ��� dnr. t�� ���u�� �� ����� ������� ��� ��� mac, ������ �� au�u�� 2008. moose advisory committee a dnr �� ��� �������� ��������� ���������� �� dnr �����, n����� a������� ���u��� ����u��� ��������, ��� u����� s����� f����� s������, ��� ����� ���k�������� �������� mac ���b��� �� ju�� 2008. m��b��� ���� ������ �� ����u�� �xp������ �� �������, �����, ��� ���b�� �������� ��������, ��u��y ���� �������� �� ����� �����, ������ ����������, ��� ���k������� ���up� ���� ������ �������� m��������. t�� ��������� ��� ��k�� �� p������ ����������� ��� ��������������� ��� ��� mmrp �� b� ������p�� by ��� dnr. t�� mac p������� ��� dnr ���� ������� ���� ��� ����� p��b��� �� b��� ��������� ��� ��������� m��������, ��� p��p���� p��� ��b�� ���������� ��� �������� ���u����� �� its final report (peterson et al. 2009). there ���� � ����� �� 18 mac ���b��� ����u���� 5 dnr ��p��y���, 2 ���b�� b���������, ��� 4 �������� ����������, 3 �� ���� ���� ������� ���� ���� ��� u��������y �� m�������� ��� 1 ���� m������� t������������ u��������y. t�� ��������� 7 ��������� ���b��� ���� ���� ���k������� ���up� ����u���� t�� n��u�� c���������y, ��� sup����� n������� f�����, ��� m�������� f����� r���u���� c�u����, ��� l�k� c�u��y l��� d�p�������, ��� m�������� d��� hu����� a����������, ��� m�������� c��p��� �� ��� w������� s�����y, ��� � ������ �����. t�� dnr �������� ��������� ��k�� d�. r��� p������� (m������� t������������ u��������y) �� ����� ��� mac b���u�� �� �� � ���p����� b�������� ���� � ���� ������y �� ����� �������� ��� ��� ��� ���������� ���� ��� dnr. t�� dnr �������� ��������� b�� ������ ���� � ����m��������� ���b�� �� ��� mac ��u�� ���� ��� p����������� ������� �b�u� ��� p��b���� ���������� ���� m�������� ����� ���/�� p�������� ���u�����. t�� c���� ��� �b���� ���� ��� �������� �� ����� d�. r�� m��� (u��������y �� m��������) ��� ��� p����������. b���u�� �� ���� ��� ��� ����� ���k ��� ��� mac, d�. p������� ����������� dr. moen as co-chair and the mac affirmed ���� p�������. moose advisory committee – mcgraw et al. alces vol. 46, 2010 192 t�� ��������u�� �������� ���� ��� dnr ����u�� ���� k�y ���k�������� �� ������p � ���������� p��� ���� ����u���, bu� ��� ��� ������� �� ��� ��������� p�pu������. a���� �����u�� ����u����� by ��� mac ���������� ��� ���b����y �� ���� p�pu������, �� ��� ������ ����� ���� ��� ��� �u�b��� ������� ������� ���� ��� �������� �p�����, ����� ���������� ��� �������� �� ����� �� ��������� m�������� became the main focus of the mac. the first mac ������� ��� ���� �� s�p���b�� 2008 ��� ������y �������� ���u���� �� j��u��y� ju�y 2009 �� ����u�� ��������������� ��� the format of the final report for the dnr. t�� ��p��� ��� p�������� �� au�u�� 2009 ��� ��������� b�����y, �u�����, ��u����, �u��u��� ���u�, ��� ����� ��p���� �� ��� ����� p�pu��� ����� �� ��������� ��� ��������� m��������. t�� mac ������ � su���� �� 8 d����� b�� 2008 �� ������ ����������� �b�u� ����� ���� b��������� ��k�� �� p������p��� ���� ����� ju����������� ����u���� n���� d�k���, m����� ���, n�� h��p�����, m�����b�, ��� o������. p���� �� ��� su���� ���y ���� p������� � ���� �� qu������� ���� ��������� ������� p�������� ����� �� ���u� (t�b�� 1); ����� u��b�� �� ������ �ub������ ������� �������. t���� b��������� p������� ����������� �b�u� p�pu������ ����u�, ���������� p��������, ��� ������� ������ ��� ��������� ����� �� ����� ju�����������, ��� ��� b������ ���u�� ��k� ������� ������ ���� �����p������ ���� ����� ���������� p����. t��y ���� �u�����z�� ���� �b�u� p�pu��� ���� ������, ������� �� ������bu����, �������, ����:��� ������, �u������ �����, ������ ���u��, ��� ���� (odocoileus virginianus) �u�b��� �� ����� ���p������ ���������� ������� ��� ��� p��� 10 y���� �� ������. qu������ ��� ������ �������� �p�� �� ��� ������� �u����, dnr �����, mac ���b���, ��� ���b��� �� ��� p���� ������� �������� �������������, �����������, b���������, ��� ����� ���k�������� �� ����� ���� ������ p���p������� ��� ������� ���� ��� ����� ���u����� �� m��������. p���� �� ��� ����� �� ��� su����, ���b��� �� ��� p���� ���� ������� �� � ����� ������� �� �� ������ �� �p�� ��� ����� �� ����u�������� ���� ��� pub��� �b�u� ��� ����� ���u����� �� m�������� ��� ��� ���������� �� ��� dnr ��� ��� mac. t�� ����� ������� ��� �������� by 4 p���� �����, 1 �����, ��� 1 ���������� �������. c�����u�� p���� �������� ��� � ������ ����� session at the presentation of the final report to ��� dnr ���� p������; �������, ����� ���� �� �u����� p���� �������� ��� ���y ��� ������ ����� ������� ���u����. o� 9 d����b�� 2008 ��� mac ��� ���� ��� ������� b��������� �� ����u�� ���u�� b��u��� ����� �� ��� p�����u� ��y’� �������. qu������ ��� ������ �������� ��������� p������������ �� ��� su���� ���� ������b�� �� ��������� �������� �� �������y p�������� ����� �� ���u�. o�� �u����� �� ��� su���� ��� ��� ���� ���� �������u� (����� b��� mac ���b��� ��� ������� b���������) ���� m�������� ��� �ub���������y ���� ���� ������b�� �� ����� population fluctuations and survival rates ���� ���� �u���u����� ������� b���u�� �� ������� �� ������ �u���y�, ���������� �� ���� ���� ����, ��� ��������� (mu���y �� ��. 2006) ��� ��������� �������� p��j���� (l����z �� ��. 2009). m��b��� �� ��� mac ���� ���� ���� ��� p��b���� �u������y ����� by ����� ������� ��� �������� ��� ���� ������u�� ���������� �� ��� p�pu������ ��� ���������. d���u����� �� ��� ������� ������ ��y ���������� ��� ��������� ��� �������� ��� p����� ����� �� ������� �� ��� �������p� ��� ���� � ���u� �� ��� mac ��� �������� �������������. both the physiological influences on moose ��� ��� �������� �� ��� ������� ������� �� ����������� ���� ���������� ���p����u��� ��� ����� �������� ������� ���� ���������; ���� ��� ���������� �������� �� �������� �������� ��� �����. b�������� (parelaphostrongylus tenuis), � p�������� �������� �������y ��u�� �� ������������ ���� (a������� 1964), ��� ����u���� ���� b���u�� b�������� ��������� �� �������y ��p����� �� ��u�� ��������y �� �����. a��������� ����u����� p����� ����u��� ������p���� �� ��b���� ���������� p���� ���� ����� �� ��� ����� �� ����� ����� ��� p����b�� alces vol. 46, 2010 mcgraw et al. – moose advisory committee 193 ��p���� �� ������� ������, ��� ��� �����b����y �� ������u�� ������� ����������� b��������� ���b����y ��� pub��� �p�����. c������ ������ �������� � ����� ���� �� ��������� �u���� ��� �u���� ��� ����u���u� ������p���� �� ���� ������������ b���u�� ����� �� m�������� ��� �� ��� ��u����� ���� �� ����� ������p��� �����. t�� ��p������� �� ���������� �������� �� ���� �u������ ��� ��� p����b�� ��������� �� �����������y ���� �������� �� �������� �� ������� ������ ���� ������ �u���� ����u�����. f�������� ��� �u����, 6 �ub������� ���� ���� 4�6 ���b��� ���� ���� ������ �� focus on areas of importance identified at the su����: h������, d��� m���������, s����� d���������, h�b���� m���������, r�������, ��� c���u��������. mac ���b��� ���� u������� �� ����� �� �ub���������� b���� �� ����� �xp������ ��� p������� ��������, ��� ������ �� ���� ���� ��� �� ���y ������. t�� mac co-chairs identified members whom ���y ���u��� ���u�� ���� ��� �ub���������� ���� ��pu� ���� �ub��������� ���b���; ����� ���b��� ���� ���� ��k�� by ��� c��c����� �� ���y ��u�� ����p� ��� ���� �� �ub��������� �����. t�� �xp�������� ��� ���� ���� �ub���� ������ ��u�� ����u�� k�y ���u�� ��k��y �� ����� �� ��� �u�u��; ���� �ub��������� ������p�� ��������������� ��� p������� b��k���u�� ����������� ������p������ �� ����� ��p��. sub� ���������� p����� ������ �� ����� �������� �� � ������� b���� �� �������� ���u����� ����� mac ���b��� p���� �� mac ��������. sub� ���������� ��� ����u������ �xp������� ����� �������� �u���� mac �������� �� ����� ��� ������ mac �� ������� ��� ��k� �u��������� �b�u� ��� ��������. subcommittee recommendation development n� ������ �u�������� ���� u��� by �ub���������� ��� ��� ����� �������� �� ��� 1. sp��k��� �� ��� �����x� �� ����� �� y�u� �����y’� ���������� ���� (�����, p�������, �� ���b�� ����� ����) �� ��� p��� (p�����u����y ��� p��� 10 y����, bu� ������ �� �����) ��� p������, p����� ������b� ��� ���������, ����, �� ��p�, �� �pp��p�����: �) p�pu������ ������ b) ������bu���� ������� �) ������� �) ����:��� ������ �) �u������ ����� �) ���� �u�b��� ��� ������� 2. d�����p���� �� ��� ��� �b��� ����������� �� ���������� �� y�u� ����� �� p�������. 3. briefly describe moose management, monitoring, and research strategy in your state or province 4. w��� ������ ���u�� ��� �� ������� ��� ����� �� y�u� ����� �� p�������? 5. w��� ������ �� ��b���� ��� ��p������ ��� ����� �� y�u� ����� �� p�������? 6. w��� ��� ��� �u����� ����� ��b���� ��� ���������� ���������� �� y�u� ����� �� p�������? 7. p����� ����, �� ����� �� p������y, ��� ��p 3 ���u�� ��� ����� ���������� t�b�� 1. qu������� p������� �� p��������� p���� �� ��� m���� su���� �� du�u��, m��������, 8 d����� b�� 2008. moose advisory committee – mcgraw et al. alces vol. 46, 2010 194 ��p��� ���u�� b� �������. m��� ����u�������� ����� �ub��������� ���b��� ��� ��� ������, ���� ���������� ����� �� p������� �������� �� ������. t�� �ub��������� ������ ����� ���� �� ��� ������� ����� �� ��� ��p��� ��� ��������� ���� �������� �� ��������� �ub��������� ���� b���. i������ ������ �� ��������������� ���� ����u����� ����� �ub��������� ���b��� ��� ��������� ��� ��������� b����� p����������� �� � mac �������. r�������������� ���� b���� �� ���u�� ������ �u���� ��� su���� ��� ���� ��� ����������� ��� �xp������ p������� by �ub��������� ���b���. a���� ��k��� ����� ����� b���� �� ����b��k ���� mac ���b���, subcommittees submitted final drafts to the mac ��� ����u����� ��� �pp�����. r����� ���������� �� ��� ��p��� ���� ������z�� by �ub��������� �������. harvest �� t��� �ub��������� ������ ������ ���� ����� �u����� �� ��� ��������� p�pu������ ������u� b���u�� �� �������� pointed toward a significant impact on the ��������� p�pu������ �� 7600 �����. t��� ��� ��� ��p�y ���� ������� ������ ���� �u�� �����b�� b���u�� �� �u����� p�pu������ ������ ������u�, ������� ���� �������. u�����y��� ��� ��������������� ��� ��� ��p����� ������ �� ������u� ��� ����� �u�� �� ����� ������ ����� ����������� �u���� ������������. t�� �ub���� ������ ����������� b��������� ��� ������ ���������� ����� ������� ���u�� b� ���pp�� ������ ���������� u���� ��� ���������. f�� �x��p��, �� ��� ����������� ���� ��� ����� ���u�� ���p �u����� �� ��� bu��:��� ����� ��� <67 bu���:100 ���� ��� 3 ������u���� y����. n����� ���� ����� �u����� �� � ��������������� ���� �pp���u���y, ���y ����������� ������� ��� ������ �� ������� �u���� �u����� ���pp�� b���� ���������� �� �������u�� ���������� u���� �� ������ ��� ������ �����. hu���� �u����� �� �yp�����y ����� �� z���� ������ ��� b�u��� ��y w����� c���� a��� (bwca); ���������, ��� mac ����������� ������� z���� ������ ��� bwca �� ������� �u���� �u����� ���� 3 ������u���� y���� ��� <10%. i� ����� z���� ����� �u���� �u����� �� �yp�����y ������, ������ ����u�� ��� ����������� �� �u����� ��� <20% ��� 3 ������u���� y����. c���u�� �� ��� �u����� ������ ������ ��� �� ������������ m�������� ��� ����������� �� �u����� ��� <30% ��� 3 ������u���� y���� (p������� �� ��. 2009). i� ��� ���� �u������� ���� �u����� b� p����� �b�u� ����� �xp�������� �� �u�����. r�������������� ���� ��� ��� ���� ������� ��� ��� ���� �u����� ��u�� ���u�� �� ������� ���������� ���� ���. deer management �� r�������������� by ���� �ub��������� ���� ������� ���� ���� ������� ���������� ��� �������� �� ����� ����� ������������. r������z��� ���� u����� �����y ������� �b�u� ��� �����������p b������ p. tenuis ��� ���p���� �� ����� p�pu�������, ��� �ub��������� ���k � p����u������y �p� p�����. h������ ���������� ����������� ����� ����u��� �������� ���� p�pu������� �� ���� p����u�� ��������� �� <10 ����/��2 ����� ���� ��� ����� ������ ������p, ��� �� b�� ���� ������� ������ ����� �����. t�� ���� �ub� ��������� ���� ������ ��� ���u� �� ���������� �������� ���������� ���������� ������������; ��������������� ���� �����p������ by ��� �������� �ub���������. t�� ���� p�pu������ �� m�������� ��� ������� ��� ������ b�������� �� ��� ���� 1970� ���� � �y���� ���� ������� ��� ���u�� �u����� �������. t��� �y���� p������� ��� ������� �������� ���� �������� ���� ��� ��u���� and a finite number of antlerless permits avail� �b�� ����u�� � ������y. s������� �� 2005, ���� �� ��������� m�������� ���� b��� ������� �� ���u�� ����� p�pu������ b���� �� ���k������� ��� pub��� ��pu� ���� ���� p�pu������� ���� ��� ����. t�� ������ �� ��� ���� p�pu������ ������ ��� ����� �����, ��� ��� u���������y �u���u����� ��� �x���� �� ����� ����� �u�� �u�b �� �������� ��� p�������� ���������� ���� ���� �� ��� �������p�, ��� ��� mac �� ��������� �u����� �������� ���������� moose-deer interactions. specific research ����� ����u��� �������y��� ��u��� �� ����� ��������y, ��� ��� ������� �� ���� ������y ��� ������p�� p�pu������� ������ ��� p��������� alces vol. 46, 2010 mcgraw et al. – moose advisory committee 195 �� b��������. a ������ ��� �� ����������� ����� �� ���������� ������������ �� ��������� m�������� ��� ����� �� �� �pp����x �� ��� mac ��p���. social dimensions �� t��� �ub��������� ���� ��������� �� ��� ��������, ������, ��� �u��u��� ��p���� �� ��� mmrp. m���� ���� an iconic status in minnesota and benefit local ��������� ���� ��p��� p����y �� �u����� ��� ��u����� �������� ����� �����, ��� ���y ��� p��� �� ��� �u��u��� �������y �� m�������� ��� �������� n����� a������� ���b��. m���� ��� �������y ��� �u��u����y ��p������ �� ��������� ��� �������� ���� ���� ���� �� �xp��� ����� �� ��� ����� �����; ������ ��� �pp���u���y �� ���� ����� ��u�� ���u�� �� � �ub�������� ���� �� ��u���� ������ ��� �u��u��� ���u�. t�� economic effects are difficult to quantify but ����u�� ��� ���� �� �����u� ��������� by ������� �����, ��� ���� �� �u���� �� ��u�����, ��� ����� ���. i�p����� �� ��� ��������� �� ��� mac ��� ��� ����������� �� ��� ��p������� �� ����� �� m��������; �������, �� �ub��������� ��� specifically devoted to economic issues asso� ������ ���� �����. t��� �pp����� ��������� ���u���� �u���� �ub��������� ������p����, �����u�� ��� ������ ���������� �ub��������� ��� ��������� �������� �������� ����u���� � �u���y �� ��������� ��� �������� ��� �u��u��� ���u� �� ����� �� m��������. habitat management �� t��� �ub������� ��� ����������� ���� ��b���� ���������� ��� ����� ���u� �� ����� �� ������� ����� ������y. du� �� ����������� �� �������p� ��� ���� u�� b������ ��������� ��� ��������� m��������, specific habitat recommendations were made ��� ���� ������. a����u�� ����� ��b���� �� m�������� �� ��������y ��� �������� �� ������ ���, ��� ��������� ����������� ���������� ���������� ���� ���u� �� ������� �p����� �����. specific habitat recommendations were rather �������, �� ���� ��� � k�y �������� ����. research �� t��� �ub��������� ������p�� �������� ��������������� �b�u� p�pu������ �y������, ��������, ���������� ������������, ��� ��b���� ��qu��������. t���� ����������� ����� ���� ��� p�������z�� b�y��� ��� ���� ��� ������u�� �u���y� �� ��� ��������� p�pu������ �� � �����u�. i� ��� �������z�� ���� �������� ��� ��� ��� � ���������� ���u����, ��� ���� ��p�������� �� ���������� ������qu�� ��� ��qu���� ��� ���qu���, ��������� p�pu������ ����u�����. t�� mac ����������� ���� ����� �������� ������� ��p p������y �� ��� dnr, ��� �u�� �� ��� p��p���� �������� ��� ���u��� �� qu������� ������ by �������� ���p����b�� ��� ����� ��� ����� ��b����. t�� mac �������z�� ��� ��p������� �� ������� ���� ��������� �� ��� b���� ��� ��u�� ��������; �������, �������� ��� ���� �u��� ���������� ���� p�pu������� �������. f�� ���� ������, ��� mac ��������� ���� ��p����� ���u�� b� p����� �� �������� � ��u�� b��������� b���� ��� ���������� ���������. communication �� t��� �ub��������� ������p�� ��������������� �� ��� �������� ���� ���u�� b��� b� p�������� ��� ������������ �����y �� ��� pub���. a ������� ��� ���� ��� pub��� ��u�� ���� ������� mac ����������� ����� ��� ������� by ��� dnr �� ������������y. f�� �x��p��, ��� �������������� �� ������u� �u����� �� ��� ��������� p�pu������ ��u�� b� ������ �� � ������������y ���p���� �� ���� ����� � ��������� p�pu������. i����������y �p������z�� ��b���� ���������� ��� ����� ��� ��� p��p���� ���u����� �� ��� ���� ���� ���� �pp����� �� ���� ���� ��� �������������� ���� ����� �u����� �� ����� b����������y �upp����b�� �� ��������� m�������� (p������� �� ��. 2009). t�� ���� �� ��� �ub��������� ��� �� �u����� ��� �� ���u����, ju����y, ��� ����u������ ����� ���������� ����������� �� ��� pub���, ��� ���y b���������, ������, ��� �������� ���� tors influenced their recommendations. the �ub��������� ���� ����� �� ��������� ����� �� p����b�� �u�u�� ����������y, ����u���� p������ ��� ��������� �� � ���u�� �� u���������� pub��� expectations, conflict between consumptive ��� ��������u�p���� u����, ��� ��� �������� ���� �� ��� pub��� �� p�y �� �������� ����� �� m��������. moose advisory committee – mcgraw et al. alces vol. 46, 2010 196 additional recommendations sub���������� ��� ��� ������� ��� ���u�� ���������� ����� ����������, ��� ������� ���u�� ������� �ub��������� b�u�������. fu����� ��� ����� ���������� ��� ������� �� ��������� �������y��� ����� �� � sp����� �� sp����� c������, �� � ����� ������ t��������� �� e��������� �p����� ���� ��������� by ��� ������ mac. funding -funding was not identified ��������y �� � ������� �� ��� ��p��� ��� �� ������� �ub��������� ��� ������; �������, ��� mac �������z�� ���� ����� ����������� ����� ��� ���������� ���������� ��� �������� ������� ��u�� ��qu��� �u�����. c�������� ���b��� ����u���� ������� ������� �u���� ������y �������� ��� ��� ������� ��� ����� ����� b���� �� ���� ��pu�. t���� �� �u������y �� ��������� �u����� ��u��� ��� ����� �� m��������; ����� ���������� �u��� ��� ��� ������� ����u�� ��� dnr g��� ��� f��� fu�� ���� �� �u���� ����u�� ��� ���� �� �u����� ��� fishing licenses. t�� mac ����������� ���� ���������� ������ ��u���� �� �u����� b� ��u��� �� ��������� ����� ���������� ���� ��������� ���������� ��� ��������. mu�� �� ��� ��������� ��� ���� �������������� ����� ���� ��� ���� ���� ����� �u����� �� m�������� ���� ��� �������� �ub�������� �����u�, ������� ��������u�p���� u��� ���up� �������u�� � �u�� ������ p��������� �� ��� p�pu������ ��� ������y, ��� ���u�b�y ���u�� ������ �� �u����� ����� ���������� ��� ��������. t�� mac ����������� ���� ��� dnr ���k ������y ���� ��� ��������u�� �� ���u�� ���� ����� �x��� �� m�������� ��� ��� ��������b�� �u�u��; ������� �u�����, ���������� �� ����� ��� ����u����, ��� ����� ���������� �� �������� ��u�� b� ���������� �� ��� ���u� �� ����� �� m��������. legal status of moose �� m�������� ��� 3 ���������� �� ����� ������ �p�����: e���������, t���������, ��� sp����� �� sp����� c������ (ssc). s���u� �� ssc �� u������� ��� ����� cates a species with unique or highly specific ��b���� ��qu�������� ���� �������� �����u� ����������. sp����� �� ��� p���p���y �� ����� ����� ��� �p����� ���� ���� ���� t��������� �� e���������, bu� ��� ����������, ��y ���� b� ������ �� ssc. t�� dnr up����� ��� ����� ������ �p����� ���� �b�u� ���� � ������. t�� ������ p������ ��� u������y �u���� ��� mac p������, ��� ����� ���� b���� ���������� ��� �������. a� � ���u��, ��� mac ��� �� �p� p���u���y �� ��k� � �������������� �� ��� ����u� �� ����� �� m�������� ��� ������� �� ��� ����� ���u�� b� ������ �� ssc. d���u����� ��������y �������� ���u�� ������� �� ��� ��� mac ���u�� �������y ��� ����� ��� ���u� ��� ����� ��������������. t�� dnr e��������� sp����� c���������� p��� ����� ����������� �b�u� ��� ������� p������ ��� ���� ��� ����������� �� ssc, t���������, �� e��������� ��u�� ���� ��� �����. b���u�� ��� ����u����� ���u���� �� ������ p�������, �� ��� ������� ���� 2 ��������� ���b��� ��u�� p��p��� ���������� ���u��� ��� �� ������� ��� �������������� �� ���� ����� �� ssc. t�� ���������� ���� ����u���� ��� ����� �� �� ��� ��x� mac ������� ����� �� ��� ����������� �� ���� ����� �� ssc by � ���� ��j����y (9 ���, 8 �������, ��� 1 �b�����); ��� final mac report contains a section expressing ��� ��j����y �p����� ���� ��p������ �������. t���� ��� � �������u� ����� ��������� ���b��� ���� ����� ���u�� ��� b� ������ �� t��������� �� e��������� b���� �� ��� definitions in minnesota statutes (2007: sec� ���� 84.0895). t�� ������ p������ ��� ������� �p����� �� �� �����qu��� ���� mac ���b��� ������� � ����������� ��u�� ��� b� �������b�� �� ��� ����� �� �������� ����u�������� �� u��������� �����qu�����. s��� ��������� ���b��� ���� ���� ������� ����� �� ssc ��u�� lead to management decisions influenced by p�������� �������, ��� ���� ���u��� �� ��u�� ���p p������ ����u���� ��� �������� ��� ����������. completing the mac report �� p���� �� ��� m�y 2009 �������, �ub��������� ���� ����� ���� ���p���� ���� � ������ ���u���� ��� ������ by ��� mac c��c����� �� ������ alces vol. 46, 2010 mcgraw et al. – moose advisory committee 197 ����������y ����� ��������. a� �x��u���� �u����y ������� by ��� mac c��c����� �u�� ����� � b���� ������y �� ����� �� m��������, ��� ����� �� ��� mac, ��� ��� ������ ������ ����. s���� �u������� �� ���� �ub��������� ������� ���� ������� ��� � ��������� �������� ��� ��p����������� �� ��������������� ��� p������� �� ��� ���u����. t�� ��������� ������ ���� p�������� ����� ��� ������� �u�� ������ �� ���� ������� ��u�� ��� u������������ by � ������� �u������. i� ��� ������� ���� p���������y ������������� ��������������� should be discussed briefly at the beginning �� ��� ��p���. t�� ������ ���u���� ��� p����� ��� ��� ��������� ���b��� �� ���� ��� ���� through a cycle of review; final editing was by ��� c�������� c��c�����. conclusion of mac a� ��� ��qu��� �� ��� dnr, ��� mac scheduled a meeting where the final document ��� p�������� �� 18 au�u�� 2009. sub���� ������ ������ ���� ��k�� �� p������ b���� p������������ �u������� ����� ��������, ��� mac ��� dnr ��p������������ ����u���� ��� ������� �� ��� ��p��� ��� ����� ���������� �� m��������. m��y ���b��� b������� ��� ���� ������ ���u�� ������ ������ ��� b� ������b�� ��� qu������� ���������� ����� ����������� ����� ��� ��� ��������k �� ��� p��� u���� ��� mmrp ��� ���p�����. mac ���b��� ���� believed it would have been beneficial to have b������ p������p����� by ��� dnr ���������p �� ���� �������. u�����u�����y, ��� dnr ������������ p������ �� ������ bu� ��� u�� �b��; �� � ���u��, ����� ��� ���y � ������ dnr ��p����������� ��� ��� ��� ���u����y �������� mac ��������. t�� ���� �� ��� mac �u���� ���� p������ ��� �� p������ ��� dnr ���� � ��������� ���� �� ��� �u�u�� ��� ����� �� m�������� ����� ��� ���������� ����� ���� �u������ �� ��� ��u����� ���� �� ����� ����� �� ��� ����� �� ������� ������. o�� ����������� ����u���� bu� ��� ����u��� �� ��� ��p��� ��� ��� ����������� ���� �� ��y ��� b� �����������y p����b�� �� �������� � ����� p�pu������ �� m�������� ���������. h������, ��� mac b������� ���� ����� ��u�� b� p������ �� ��� ��������b�� �u�u��, ��� ���u�� b� ������� ���u���� �u��, ��� ���� ����� ��� ��������������� ���u�� b� ���u��� ����������y. t�� ������� �����u�� of the mac at the final meeting was best ������b�� �� ��p��u�. m��� ���b��� ���� �� ��� �p����� ���� ��� ��������������� ���� ��������� ��� ��u�� p���� ���p�u� �� p�������� �u������ ��� �u�u�� ���������� ��� �������� �� ����� �� m��������. a� �� j��u��y 2010, ��� dnr ��� ������� ��� mmrp b���� �� ��������������� �� ��� mac ��p���; ��� mmrp ���� b� ������b�� ��� pub��� ������� ���� ���p�����. discussion t�� mac ��������� �� b��� dnr ��� ���� dnr ���b��� ��� p������� p���p������� ���� � ����� �� �����p����� ����u���� �������� ���������� ��� ��������, ������ �������, ���� ����y, ����� ����������, ���b�� ���u��� ����u��� ��������, ��� ��� ��u���� ���u���y. n���dnr ���b��� p������� k�������� ��� p���p��� ����� ���� ����u������ ��� ���k�������� ��� ������ by dnr ����� ����������. mac ���b��� ���� ��qu����� �� ��p������ ����� ��� �p������ ��� ��� ����� �� ����� ���p������ �������� ��� ������z������, bu� ���� dnr ��� ����dnr ��������� ���b��� ��u�� ���� difficult. specifically, the dnr representatives provided input for the final product, but were ��� ����y� ��������b�� �� ���� ����. i����p������� mac ��������������� ���� ��� mmrp �� �� ��� ���������� �� ��� dnr. m��y mac ���b��� �xp��� ���� �u�� �� ��� ��p��� ���� b� ����u��� �� ��� mmrp, �� p��� b���u�� �ub��������� ��������������� ���� �����u�� ��� ��� b���� �� ��� �������u�� �������� p����b��, ��� ��� dnr ��p�������� ����� p������� ��pu� �� �����y p�������� ��� ����������� ���� ��������������� ���� ����u����. a����y p�������� ��� ����������� ���� ��� ������ �� b������ �� �������� �� ��� p������ by mac ���b���. p����b�� ������� moose advisory committee – mcgraw et al. alces vol. 46, 2010 198 ����� �� mac ��������������� �� ��� mmrp may be due to fiscal limitations, legal require� �����, �� ����� ������� ���� ��p��� ��� dnr. p���u��b�y, �ub�������� ���������� b������ mac ��������������� ��� ��� mmrp ��u�� require strong justification given that the mac ��� ������ dnr ��p�����������. a�� ���������� p���� ��� ����� �������� ��� �ubj��� �� ���������, ��� ��� mac p������ is likewise open to specific criticisms. media ���� ������� �� ��� su���� ��� ��� p�������� tion of the final report to the dnr. monthly ��������, �������, ���� ���y �������� by mac ���b��� ��� ��� dnr p������� p��� ������ ��� p������� p������ ������������. t�� �������� ���� ����������y �p�� �� ��� pub��� bu� ���� ��� ����������, ����������y ��k��� ���� ������. t��� �pp����� ���� �� possible to work efficiently, and also allowed mac members to focus on specific discussion ��p���. t��� �������� ��� ���� by ��� mac ���� �u������ by ��� dnr p������� p��������; ���� mac ���b��� ���� ���� ���� �pp����� ���u��� �pp���u������ ��� ����� ���������. t�� s����� d��������� �ub��������� ���� ������ ���� �� ��� ���� ��� ���������� ��������� �� b� �����p�����, ����� ������ ������������y �� ��� mac p������ ���� ��� ��� ���������� �p���y. t�� mac p������ ��� �������� �� p������ dnr ��pu� ��� ��������� ��� ��� mmrp ���� a specialized group with specific knowledge �b�u� �����. t�� dnr ������ ��� mac �� the first step in a larger process with regard to ������p��� ��� mmrp; ����� ��� ���p������ pub��� ������� ���� ���u� ��� ��k��y ������ p������� k��������, ��� mmrp, ��� mac ��p���, ��� ��������� ����� ����. r�������������� �� ��� mac ���� �������� �� ��� dnr, ����� �� �b������� �� ���p��� �� ��� ������� pub��� �b�u� ��� ����� p�pu������ ��� ��� ����������. i� ���� b� ����������� �� ����u������ ���� ���k�������� not trained in wildlife or fisheries manage� ���� �� �� ��y ������u�� ����� �u����� by �������������� ��� ���b�� �u����� ���� ���� ������ ������ �� ��� ����� p�pu������ �� ��� �������. u���k� ���y ���� �p����� �� m��������, �u�� ������ �������� �� ����� �� ������� �� �u��u�� ��� ��u����. i�p�������y, ��� mac ����������� ��u�y �� ��� �������� ��� �u��u��� ���u� �� ����� �� m�������� ��� �� �����p����� ����� ���u�� ���� ��� mmrp. fu����� ���� b� � k�y ���u� ��� ��p������ ������ �� mac ��������������� �� ���y ��� ���p��� �� ��� mmrp; u�����u�����y, ���k�� ������ ���up� ������ �� �������z� � ��������� �u����� ��u��� ��� ����� ��������. h������, 2 u��qu� ��u���� �� ����y ��� ������b�� �� m�������� ����u���� ��� l�����������c���z��� c��������� �� m�������� r���u���� (lcc� mr) ���� ��� ��� �pp��x������y $25 ������� ���� m��������’� e���������� ��� n��u��� r���u���� t�u�� fu�� �� �p��� ���u���y �� � ������y �� ��b���� ��� �������� p��j����. t�� ������ ��u���, ��� l�������s��� ou����� h������� c��������� (l�sohc), �� �������� 33% �� � ��x �������� ���� � �������u������ ��������� p����� �u���� ��� 2008 �������� (~$90 ������� ���u���y) ���� ��y b� �p��� ���y �� �������, p������, ��� ������� ��������, prairies, forests, and habitat for fish, game, and ��������. t�u�, p��j���� �������� �� ��������/ ���������� ��u�� b� �u���� by lccmr ��� ��b���� �����������/����������� ��u�� b� �u���� by b��� lccmr ��� l�sohc. t��b�� ���u��� ����u��� �������� ���� ���� ��� n����� a������� ������� �� ����� u���� �x������ �����y ������ �� m�������� ���� ��� �b������� �� �������� ��������������� �� ��� mac ��p��� b���u�� �����y �u����� ������ ��� ��� ���u����� by ��� �����, ��� ���b�� ���b��� ��� ��� ��qu���� �� ������ �� ����� �u����� ���u�������. d���u����� �� mac �������� ���� ������� ���b�� �������, ������� �u�� ����u����� ��� ������y �������� ��� ��� mac ���������� �� ��� ��� �b������� �� ������� ���b�� ������� �� ��� ���������������. t��b�� ����������� ��� ��� dnr �����������y ���� � ���� ���k��� �����������p ��� ���� ���p������ �� ����� ���������� ��� ������� p������� (e������ �� ��. 2004). t�� ���b�� b��������� ���� mac ���b��� b���u�� �� ���� �����������p ��� ����� alces vol. 46, 2010 mcgraw et al. – moose advisory committee 199 �xp������ �� ����� b�����y. c��p������� b������ ���� ���������� �������� ������ ����� ����� �� m�������� ���� b� �������� �� ��� �u�u��. mu�� �� ��� �u����� ����� �� p��� �� ��� sup����� n������� f�����, ���� ��j�� �������� by �����, ��u��y, ��� p������ ����������. f����� ���������� practices by the usfs and the dnr influence ����� ����� ��� ��b����. t�� mac ����� ������� ��� ��p������� �� ���p������� b������ ����� ��������, ��� ��� ���� ��� ��� f��� ��� w������� d������� ��� ��� f������y d������� ������ ��� dnr �� ���p�����, ��p������y �� ����� ���������� b��� ����� ��b����. t�� mac p������ ��� �� � �������� ��� �� ��������������� ��� ����� ������� ���� �� ������������ m�������� by p������ k���������b�� ����� b��������� ��� �������� p������������ �� � ������ �������. t��� ���up ��� �����u������ �� ������p��� ��� p��p��� ��� ��������������� ��������y ��� ��� dnr �� ������p ��� mmrp (p������� �� ��. 2009). mac ���b��� ���� ���� �u����� ���������� ������� ���� �� ������u� �� m��������, ��� ���� ���������� �������� ��� ���������� ������� ��� ���������y ������ b����� ������� ���u�� �� ��� ������������ p�pu������ �u�� �� ���u���� �� ��� ��������� (mu���y �� ��. 2006). t�� mac p������ ������b�� ���� p������� � �������b�� ��������k ��� pub��� �������� �� ������� ������������� �������� ���������� ���u�� �� ���� pub��� �������, ��� ��p�������y, p���u��� ��� mac ��p��� ���� p������� ��� ��u������� �� ������� ��� ������� �� ����� �� m��������. acknowledgments w� ��u�� ��k� �� ����k ��� c����� ��� u�b�� ��� r������� a������ �� ��� u��������y �� m�������� ��� �������� � ����� �� r�� m��� ��� g���� w����� ��� ����u��� ��u���� a����� m�g��� �� ��� �� �� ��������� �� ��� mac. au���� ����� ��� ���� p�p�� ��� ������� �������y ����� ��� ����� �u����. references anderson, r. c. 1964. n�u������� ������� �� ����� �������� �xp����������y ���� pneumostongylus tenuis ���� ������������ ����. p���������� v���������� 1: 289�322. edwards, a. j., m. w. schrage, ��� m. s. lenarz. 2004. n����������� m�������� ����� ���������� – � ���� ��u�y �� ���p�������. a���� 40: 23�31. karns, p. d. 1982. t����y�p�u� y���� �� ������ ����� ����u� �� m��������. a���� 18: 186�207. lenarz, m. s. 2007. 2007 a����� ����� �u�� ��y. m�������� d�p������� �� n��u��� r�� ��u����, s�. p�u�, m��������, usa. <���p:// �����.���/�����.��.u�/�u�����_����������/ �u�����/�����/�����_�u���y_2007. p��.> (�������� o���b�� 2009). _____. 2009a. 2009 a����� ����� �u���y. m�������� d�p������� �� n��u��� r�� ��u����, s�. p�u�, usa. <���p://�����.���. �����.��.u�/�u�����_����������/�u�����/ �����/�����_�u���y_2009.p��> (��� ������ o���b�� 2009). _____. 2009b. 2009 m�������� m���� h��� ����. m�������� d�p������� �� n��u��� r���u����, s�. p�u�, m��������, usa. <���p://�����.���.�����.��.u�/�u�����_��� ��������/�u�����/�����/2009�������.p��.> (�������� f�b�u��y 2010). _____, j. fieberg, m. w. schrage, ��� a. j. edwards. 2010. l����� �� ��� ����: ���b����y �� ����� �� ������������ m��� ������. j�u���� �� w������� m��������� 74(5): 1013�1023. _____, m. e. nelson, m. w. shrage, ��� a. j. edwards. 2009. t��p����u�� �������� ����� �u������ �� ������������ m����� ����. j�u���� �� w������� m��������� 73: 503�510. m�������� d�p������� �� n��u��� r���u���� d������� �� f��� ��� w������� (mndnr) 1990. m�������� m���� m��������� p���. pub��� d���� r�����. s�. p�u�, m��� ������, usa. m�������� l�������u�� ����u�� s������ moose advisory committee – mcgraw et al. alces vol. 46, 2010 200 l��. 2008. c��p��� 368, s������ 7 6. < � �� p � :// � � �. ��� �� � �. �� . �� � / ����/?y���=2008&�yp�=0&k�y����_�y p�=���&k�y����=�����&����yp�=���p ���&��=368> (�������� o���b�� 2009). murray, d. l., w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, ��� t. k. fuller. 2006. p����� ����, �u��������� ���������y, ��� ������� ����u����� �� � ��������� ����� p�pu��� ����. w������� m������p�� 166: 1�30. peek, j. m., d. l. urich, ��� r. j. mackie. 1976. m���� ��b���� ��������� ��� ����� �������p� �� ������ ���������� �� ������ ������� m��������. w������� m������p�� 48: 1�65. peterson, r., r. moen, r. baker, d. becker, l. cornicelli, a. j. edwards, l. frelich, g. huschle, m. johnson, a. jones, m. s. lenarz, j. lightfoot, t. martinson, g. mehmel, s. perich, d. ryan, m. w. schrage, ��� d. thompson. 2009. r�p��� �� ��� m�������� d�p������� �� n��u��� r���u���� (dnr) by ��� m���� a������y c��������. <���p://�����.���.�����.��.u�/ ����_��������/��������/�����/���/������ p���.p��.> (�������� j��u��y 2010). alces 31_111.pdf alces 31_181.pdf alces 31_167.pdf alces34(2)_375.pdf alces29_115.pdf alces34(1)_149.pdf alces32_109.pdf alces30_91.pdf alces30_137.pdf alces35_31.pdf alces 31_233.pdf alces34(2)_261.pdf alces29_63.pdf alces30_9.pdf alces34(1)_189.pdf alces32_61.pdf alces29_35.pdf alces35_91.pdf alces32_173.pdf alces34(2)_445.pdf alces vol. 47, 2011 maskey liver fluke in north dakota moose 1 giant liver fluke in north dakota moose james j. maskey, jr.1 department of biology, 10 cornell street, stop 9019, university of north dakota, grand forks, nd 58202, usa abstract: the giant liver fluke (fascioloides magna) is a parasite of white-tailed deer (odocoileus virginianus) and wapiti (cervus elaphus) that can cause extensive and conspicuous liver damage in moose (alces alces), a dead-end host. the implication of f. magna as a factor in the long-term decline of moose in northwestern minnesota has raised concern that a concurrent decline in moose in northeastern north dakota may also be linked to this parasite. i reviewed data collected from moose hunter check stations in1977-1984 and necropsy reports of non-harvested animals examined in 1983-1992 to estimate past prevalence of f. magna in moose in north dakota. i also collected livers from harvested moose in 2002 and 2003 to investigate the current occurrence of this parasite. i also surveyed 78 wetlands at 12 sites in 2003-2006 to examine the potential for f. magna transmission based on the occurrence of aquatic snail intermediate hosts. flukes or signs consistent with fluke infection were observed in 19.6% of harvested moose (n = 158) in 1977-1984, and in 18.8% of moose necropsied (n = 32) in 1983-1992. fascioloides magna was not recovered from any of the 78 moose livers collected in 2002 and 2003. however, lymnaeid snails were found at 10 of 12 sites in the aquatic gastropod surveys indicating that the intermediate hosts for this parasite occur widely throughout the range of moose in north dakota. while this represents the first known report of f. magna in north dakota, this parasite occurs at relatively low prevalence, and there is no evidence that it has been an important factor in recent moose declines, nor that it noticeably impairs the health of moose in north dakota. transmission may be limited by the transient availability of wetlands capable of supporting the life cycle of f. magna. alces vol. 47: 1-7 (2011) key words: alces alces, fascioloides magna, intermediate hosts, lymnaeid snails, moose, parasite, population decline, prevalence. white-tailed deer (odocoileus virginianus) are the normal host for 2 parasites that may cause fatal disease in moose (alces alces). the best-known of these is the meningeal worm (parelaphostrongylus tenuis), a nematode long implicated as a limiting factor of moose populations (lankester 2001, 2010). the other is the giant liver fluke (fascioloides magna), a large trematode that occurs in pairs or groups within fibrous capsules in the liver parenchyma of its normal hosts, white-tailed deer and wapiti (cervus elaphus) (pybus 2001). fascioloides magna has an indirect life cycle, requiring aquatic snails in the family lymnaeidae (hereafter lymnaeid snails) to serve as intermediate hosts (pybus 2001). in dead-end hosts such as moose, juvenile flukes migrate much more extensively than in normal hosts before becoming encapsulated, and as a result, cause considerable destruction of liver tissue. extensive fibrosis of the migratory tracts and capsules containing adult flukes can damage 50-90% of the liver, and sometimes be suspected of causing death of the host (pybus 2001). recently, f. magna was implicated in the long-term decline of moose in northwestern minnesota (fig. 1) where 89% of moose examined in 1995-2000 were infected with f. magna (murray et al. 2006). the north dakota game and fish department (ndgf) conducts annual winter aerial surveys of moose populations in 3 survey areas (turtle mountains, drift prairie, and pembina hills; fig. 1). survey data collected 1present address: department of biology, university of mary, 7500 university drive, bismarck, nd 58504 liver fluke in north dakota moose maskey alces vol. 47, 2011 2 over the past decade indicate that while moose populations appear to be stable to increasing in the turtle mountains and drift prairie areas, moose have declined considerably in the pembina hills (johnson 2002, 2007; fig. 2). during this same period, white-tailed deer in the state have increased considerably, suggesting increased transmission potential of deer parasites (namely f. magna and p. tenuis) to moose (jensen 2007, smith et al. 2007). additionally, because the pembina hills area is adjacent to the declining moose population in northwestern minnesota, concern existed that the north dakota decline also may be related to f. magna infection. this study addressed this concern by 1) examining historical data to estimate past prevalence of this parasite in the moose in north dakota, 2) investigating the current occurrence of f. magna infection in moose, and 3) determining whether suitable intermediate hosts for this parasite occur in the state. methods to estimate the historical prevalence of f. magna in north dakota moose, i reviewed 2 data sets collected previously by ndgf. the first data set consisted of hunter check-station records for 158 moose harvested in 1977-1984. during these first 8 years of the moose season, hunters were encouraged to bring entire carcasses to check stations where the animals were weighed and the viscera examined to assess reproductive status and parasitic infection. the second data set was historical necropsy reports of non-hunting related deaths. these included full necropsies on 32 such moose conducted by the ndgf wildlife veterinarian in 1983-1992 as part of targeted surveillance for wildlife diseases. i reviewed check station data sheets and necropsy reports for evidence of liver fluke infection based on the recovery of flukes from liver tissue or comments in reports that suggested fluke infection including unspecified cysts or capsules in the liver, fibrous areas, migratory tracts, detritus, necrosis, congestion or “bad” or “questionable” livers. additionally, examination of the necropsy reports from targeted surveillance allowed me to compare the relative frequency of f. magna infection with that of other pathogens. a clopper-pearson binomial confidence interval was calculated for the historical estimate of f. magna prevalence obtained from the check station and necropsy data (rosza et al. 2000). in addition, i estimated the current occurrence of f. magna infection in moose by examining 78 moose livers collected from hunters during the 2002 and 2003 moose seasons. livers were sectioned into approximately 2-cm wide slices and examined for the presence of adult or juvenile f. magna and signs associated fig. 1. moose aerial survey units in north dakota (turtle mountains, drift prairie, and pembina hills) and study areas of murray et al. (2006) in adjacent northeastern minnesota (agassiz national wildlife refuge, red lake wildlife management area, thief lake wildlife management area, beltrami island state forest). alces vol. 47, 2011 maskey liver fluke in north dakota moose 3 with f. magna infection such as fibrous tissue, migratory tracts, detritus, liver necrosis, and congestion (lankester 1974). i also investigated the occurrence of intermediate hosts for f. magna by sampling permanent and semi-permanent wetlands, small lakes, and streams for lymnaeid snails during 4 summer periods (2003-2006). i sampled for the presence of lymnaeid gastropods in 78 wetlands that included small lakes and streams at 12 sites (11 in northeastern north dakota, 1 in northwestern minnesota; fig. 3). each site was sampled by a series of 10 1-m sweeps with a dip net approximately every 10 m within 1-2 m of shore. after each sweep the contents of the net were examined for aquatic gastropods; lymnaeid snails observed floating on the surface were collected opportunistically. snails were placed in 70% ethanol or frozen, and were subsequently identified to species using the criteria of clarke (1973) and cvancara (1983). results based on the review of check station records and necropsy reports, the past prevalence of f. magna infection in north dakota moose during the period 1977-1992 was 19.5% (95% c.i., 14.1-25.8%, n = 190). there was evidence of f. magna infection in 31 of 158 (19.6%) harvested moose (table 1). liver flukes were recovered from 18 (11.4%) of these moose, while signs suggesting possible f. magna infection were observed in the remainder (8.2%; 6 with unspecified cysts, 5 with bad livers, 2 with fibrous tracts). only the northeastern area of north dakota (unit m1c, fig. 3) was open to moose hunting from 1977-1982, thus 138 of 158 harvest samples originated from this area. in m1c, liver flukes were recovered from 16 (11.6%) moose, and signs consistent with f. magna infection were observed in the remainder (9.4%). hunting for moose was initiated in units m4-m10 in 1983; liver flukes were recovered from 2 of 20 moose harvested in units m4-m10 in 1983 and 1984 (table 1). six of the 32 (18.8%) non-hunting related fatalities (1983-1992) exhibited pathology 0 50 100 150 200 250 300 drift prairie turtle mountains pembina hills fig. 2. number of moose observed within 3 survey units in north dakota by the north dakota game and fish department during winter 1980-2006. winter surveys were not completed in the drift prairie area prior to 1987. data are from johnson (2002, 2007). fig. 3. north dakota moose hunting units (m1-10) and sites sampled for lymnaeid snails (● = lymnaeids present; o = absent). numerals represent number of wetlands from which snails were recovered/ number of wetlands sampled at each site. liver fluke in north dakota moose maskey alces vol. 47, 2011 4 suggesting f. magna infection (2 with liver congestion, 1 with liver unspecified infection, 1 with fibrosis of the liver, and 1 with fibrous capsules; table 1). only a single moose was believed to have died as a result of f. magna based on the amount of liver damage caused by the fluke infection. flukes or signs of infection were not seen in any of 78 moose livers collected in 2002 and 2003. the moose hunting unit of origin was known for 56 of the samples; however, unit of origin was not available for 22 samples (table 1). a total of 418 lymnaeid snails representing 3 species (lymnaea caperata, l. palustris, and l. stagnalis) were recovered from 10 of the 12 sites, and 55 of 78 wetlands sampled (table 2, fig. 3). the 2 sites where snails were not found were represented by only a single wetland sampling area. all 3 species collected are known hosts for f. magna (swales 1935, foreyt and todd 1978, laursen and stromberg 1993); lymnaea palustris was the most common occurring at 9 sites, and l. stagnalis and l. caperata were found at 5 and 4 sites, respectively (table 2). discussion to my knowledge, this study represents the first report of f. magna in moose in north dakota. however, because the historical data were collected by previous investigators, they were subject to a degree of interpretation. first, i assumed that flukes collected by past investigators were actually f. magna, as this fluke has been recovered from moose in adjacent northeastern minnesota (karns 1972, murray et al. 2006), and the only other large liver fluke in north america, fasciola hepatica, has not been reported in north dakota (pybus 2001). second, i interpreted all signs suggestive of f. magna infection as actually being caused by this parasite; however, certain described signs may have been due to injury (lankester and samuel 1998), bacterial infection (leighton 2001), echinococcus granulosis, or taenia hydatigena cysts (jones and pybus 2001). as a result, the true prevalence of f. magna in moose in eastern north dakota may have been lower than the 19.5% estimated from check station records and necropsy reports. unfortunately, data from white-tailed deer were not available to corroborate the presence harvested 1977-1984 unharvested 1983-1992 harvested 2002-2003 unit examined infected examined infected examined infected m1c 138 29 (21.0%) 6 2 (33.3%) 4 0 m4 4 20 1 0 22 0 m5 6 20 2 1 (50%) 1 0 m6 5 20 21 3 (14.3%) 4 0 m8 4 22 (50.0%) 0 0 16 0 m9 1 20 0 0 7 0 m10 0 20 0 0 2 0 unknown 0 20 2 0 22 0 total 158 31 (19.6%) 32 6 (18.8%) 78 0 table 1. incidence of f. magna infection in harvested moose and unharvested moose in 7 management units in north dakota. species no. collected sites (present/sampled)* lymnaea caperata 48 4/12 lymnaea palustris 271 10/12 lymnaea stagnalis 99 5/12 combined total 418 10/12 table 2. survey data for lymnaeid snails in north dakota and northwestern minnesota, 2003-2006. *includes 78 wetlands at 11 sites in north dakota and 1 wetland in northwestern minnesota (see fig. 3). alces vol. 47, 2011 maskey liver fluke in north dakota moose 5 of f. magna in north dakota. nonetheless, f. magna appears to be enzootic in moose in eastern north dakota, although at a much lower prevalence than in nearby northwestern minnesota. for example, murray et al. (2006) reported 89% prevalence of f. magna in moose in northwestern minnesota in the late 1990s, and karns (1972) reported 87% prevalence in the same region in the 1970s. the failure to detect f. magna in 2002 and 2003 may have been due to the geographic distribution of my sampling. while the majority of historical reports of f. magna infection originated from unit m1c (fig. 3), my ability to sample this area was limited. only 10 moose tag were issued annually in this area in 2002 and 2003, compared to 150 tags issued in 1977-1984; i obtained only 4 samples from unit m1c (table 1). nonetheless, my recent data confirm that f. magna is not highly prevalent in north dakota moose suggesting that the parasite has not experienced a marked increase in prevalence since prior surveys. for example, based on binomial probability, i had a 95% chance of detecting f. magna in unit m1c with only 4 samples if the current prevalence was at least 53%, and a 90% chance if the current prevalence was at least 44%. also, given my sample size of 25 moose in the 4 units where historical data suggests f. magna occurs (m1c, m5, m6, m8), i had a 95% probability of detecting this parasite even if it occurred at a moderate prevalence (11.5%). in addition, all of the 22 samples from unknown locations were also negative for signs of f. magna infection. because several of these unknown samples were received in november when only units m5 and m6 (fig. 2) remained open for hunting, it is probable that a substantial proportion of these livers originated from the eastern part of the state, and f. magna likely infects a relatively small proportion of moose in eastern north dakota. the 2002-2003 hunter returns provided a more complete sampling of the western part of moose range in the state (m4, m9, m10; n = 31), and these results suggest that prevalence of f. magna is low in these areas as well. while my surveys for lymnaeid snails were by no means exhaustive, results indicate that at least 3 species of suitable intermediate hosts for f. magna are widespread within the primary range of moose in the state. natural or experimental infections with f. magna have been reported in l. caperata, l. stagnalis, and l. palustris (foreyt and todd 1978, lausen and stromberg 1993, pybus 2001). although white-tailed deer, the normal host for f. magna, are abundant in north dakota and at least 3 species of intermediate hosts for f. magna appear to be widely distributed in the state, my results suggest that actual transmission of f. magna to moose is currently limited. this may be due to lack of available wetlands. in central and north central north dakota (hunting units m4, m8, m9, and m10) the range of moose lies within the larger “prairie pothole” region of the great plains (usfws 1955) where wetlands are abundant but subject to seasonal dry down and long-term drought cycles (todhunter and rundquist 2004). thus on a seasonal or annual basis, environmental conditions may limit the availability of intermediate hosts or aquatic vegetation, prevent embryonation and hatching of eggs, and reduce survival of metacercariae (swales 1935, pybus 2001). and based on my data, a large part of the primary range of f. magna in moose in north dakota (units m1c, m5, and m6) is within the northern red river valley that is part of the lake agassiz plain ecoregion (u.s. environmental protection agency 1996; fig. 1). this area includes a number of permanent rivers and streams associated with the red river that are known to support lymnaeid snails (clarke 1973, cvancara 1983). however, compared to the prairie pothole region of central and north central north dakota, the red river valley has relatively few permanent or semi-permanent liver fluke in north dakota moose maskey alces vol. 47, 2011 6 wetlands. because riparian habitats and suitable wetlands make up a relatively small proportion of the overall landscape, they may not be capable of sustaining high levels of f. magna infection in cervids. since recent conditions in eastern north dakota apparently support only a moderate level of f. magna transmission, the parasite is unlikely to represent a major source of mortality in moose. the historical data reviewed in this study were collected during a period of moose population growth, and the prevalence of f. magna certainly has not increased since that time. since the completion of the current study, there has been only a single report of an f. magna infected moose in north dakota, a sick adult cow moose collected in unit m6 (ndgf 2004, unpublished); this moose was also infected with p. tenuis. additionally, while only 18.8% of moose necropsied as part of targeted surveillance showed signs of f. magna infection, 75.0% were infected with p. tenuis (manuscript in preparation), suggesting that other mortality factors may be more important than f. magna. although murray et al. (2006) concluded that f. magna was the major source of mortality and morbidity in the declining moose population in northwestern minnesota, it should be noted that the prevalence of f. magna had not increased since the pre-decline period (karns 1972). further, because the nearly simultaneous decline of 2 moose populations in close proximity to each other in northwestern minnesota and in northeastern north dakota cannot both be attributed to f. magna, other factors common to both areas presumably play a larger role in influencing these populations. as a result, future investigations in north dakota and minnesota should consider how other stressors or pathogens such as p. tenuis affect moose population dynamics in the region. acknowledgements i would like to thank north dakota epscor, the ndgf, the u.s.d.i. bureau of reclamation, the u.s. fish and wildlife service, the north dakota chapter of the wildlife society, the und biology department, and the wheeler scholarship for funding this project. i would also like to thank bill jensen and roger johnson for allowing me to collect samples from moose hunters and for providing access to historical data. thanks also to scott peterson for facilitating field work for this project. additionally, i am grateful to rick sweitzer, and to brett goodwin, bob newman, brad rundquist, and jeff vaughan for their assistance throughout this research and for their helpful input on this manuscript. finally, gratitude is expressed to jason smith and eric pulis for their assistance in the field and lab. references clarke, a. h. 1973. the freshwater mollusks of the canadian interior basin. malacologia 13: 1-509. cvancara, a. m. 1983. aquatic mollusks of north dakota. report of investigation number 78, north dakota geological survey. kayes inc., fargo, north dakota, usa. foreyt, w. j,. and a. c. todd. 1978. experimental infection of lymnaeid snails in wisconsin with fascioloides magna and fascioloa hepatica. journal of parasitology 64: 1132-1134. jensen, w. f. r. e. johnson, and b. stillings. 2007. study number c-1: deer population studies. report number a-172. north dakota game and fish department, bismarck, north dakota, usa. johnson, r. e. 2002. moose and elk population study. north dakota game and fish department report number a-155. north dakota game and fish department, bismarck, north dakota, usa. _____. 2007. moose and elk population study. north dakota game and fish department report number a-155. north dakota game and fish department, bismarck, alces vol. 47, 2011 maskey liver fluke in north dakota moose 7 north dakota, usa. jones, a., and m. j. pybus. 2001. taeniasis and echinococcosis. pages 150-192 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. second edition. iowa state university press, ames, iowa, usa. karns, p. d. 1972. minnesota’s 1971 moose hunt: a preliminary report on the biological collections. north american moose conference and workshop 8: 115-123. lankester, m. w. 1974. parelaphostrongylus tenuis (nematoda) and fascioloides magna (trematoda) in moose of southeastern manitoba. canadian journal of zoology 53: 235-239. _____. 2001. extrapulmonary lungworms of cervids. pages 228-278 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. second edition. iowa state university press, ames, iowa, usa. _____. 2010. understanding the impact of meningeal worm, parelaphostrongylus tenuis, on moose populations. alces 46: 53-70. _____, and w. m. samuel. 1998. pests, parasites, and diseases. pages 479-518 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington d. c., usa. laursen, j. r., and b. e. stromberg. 1993. fascioloides magna intermediate snail hosts: habitat preferences and infection parameters. journal of parasitology 79 (6; supplement): 302. leighton, f. a. 2001. fusobacterium necrophorum infection. pages 493-496 in e. s. williams and i. k. barker, editors. infectious diseases of wild mammals. third edition. iowa state university press, ames, iowa, usa. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1-30. pybus, m. j. 2001. liver flukes. pages 121-149 in w. m. samuel, m. j. pybus, and a. a. kocan, editors. parasitic diseases of wild mammals. second edition. iowa state university press, ames, iowa, usa. rosza, l., j. reiczigel, and g. majoras. 2000. quantifying parasites in samples of hosts. journal of parasitology 86: 228-232. smith, j. r., r. a. sweitzer, and w. f. jensen. 2007. diets, movements, and consequences of providing food plots for white-tailed deer in central north dakota. journal of wildlife management 71: 2719-2726. swales, w. e. 1935. the life cycle of fascioloides magna (bassi 1875), the large liver fluke of ruminants of canada. canadian journal of research, section d, zoological sciences 12: 177-215. todhuter, p. e., and b. c. rundquist. 2004. terminal lake flooding and wetland expansion in nelson county, north dakota. physical geography 25: 68-85. united states environmental protection agency. 1996. level iii ecoregions of the continental united states, map m-1, various scales. national health and environmental effects research laboratory, united states environmental protection agency, corvallis oregon, usa. united states fish and wildlife service (usfws). 1955. wetlands inventory of north dakota. united states department of the interior, fish and wildlife service, office of river basin studies, billings, montana, usa. alces32_31.pdf alces34(1)_15.pdf alces34(2)_329.pdf alces 31_35.pdf alces vol. 48, 2012 mcgraw et al. cover type temperature 45 effective temperature differences among cover types in northeast minnesota amanda m. mcgraw1, ron moen1, lance overland2 1natural resources research institute, university of minnesota, 5013 miller trunk highway, duluth, minnesota 55811-1442; 2fond du lac resource management division, 1720 big lake rd, cloquet, minnesota 55720, usa. abstract: climate is probably one of the ultimate influences on the southern boundary of moose (alces alces) distribution because moose are sensitive to warm temperatures in both summer and winter. in 4 different cover types in northeastern minnesota we compared ambient temperatures to black globe temperatures that measures mean radiant temperature of the environment. the 4 cover types were mixed forest, treed bogs, coniferous forest, and deciduous forest that comprised ~85% of home ranges of radio-collared moose in northeastern minnesota. ambient temperature measurements taken from a weather station within the study area exceeded assumed physiological thresholds of 14 and 20º c for 50 and 33% of the study period, respectively. black globe temperatures varied among cover types and temperature differences increased within cover types as ambient temperature increased. the greatest difference between deciduous and conifer cover was 2º c in black globe temperature and occurred during warm periods when skies were clear. the biological significance of these temperature differences is not clear and suggests the presence of alternative cooling mechanisms of cover types, such as water and possibly soil and duff layers acting as heat sinks. use of these potential alternative cooling mechanisms should be considered in future research. alces vol. 48: 45-52 (2012) key words: alces alces, cover type, home range, minnesota, moose, temperature. food supply, habitat composition, and climate all influence the distribution of moose. climatic influences on moose survival and distribution may be most pronounced and play a larger role in limiting populations near the southern edge of the range (kelsall and telfer 1974). regions where temperatures frequently exceed 27° c during summer do not support moose populations unless shaded areas, rivers, or lakes are present (demarchi 1991). as a circumboreal species, moose are well adapted to cold but are intolerant of warm temperatures and show both physiological and behavioral responses in warm weather (renecker and hudson 1986, 1990). temperature thresholds of 14 and 20 c that induce physiological responses by moose are from data collected on 2 captive moose in alberta that increased respiration rate when exposed to ambient temperature >14º c and began open-mouthed panting when temperature exceeded 20º c (renecker and hudson 1986). these thresholds were accepted and applied to free-ranging animals uncritically for more than 20 years; for example, they were used as thresholds in measuring moose response to heat in the boreal forest of quebec (dussault et al. 2004). recently, these thresholds were tested with thermometer data loggers deployed in cover types used by moose in southeastern ontario. because ambient temperature frequently exceeded these thresholds, lowe et al. (2010) concluded that they were too low or not applicable for southeastern ontario. however, the literature indicates that moose do respond to high ambient temperatures. in warmer periods during winter and summer, forest cover types within home ranges have lower temperatures and reduced solar radiation compared to open habitats (black cover type temperature mcgraw et al. alces vol. 48, 2012 46 et al. 1976, schwab and pitt 1991, demarchi and bunnell 1993). denser canopies of mature forests were used by moose during hot temperatures in british columbia (demarchi and bunnell 1995), and behavioral responses in summer include shifting activity to night and early morning hours and using cooler forest cover types (dussault et al. 2004, broders et al. 2011). unexpectedly, use of cover type during both summer and winter by moose wearing gps radio-collars in southeastern ontario did not change with increasing temperatures (lowe et al. 2010). however, temperature differences among forest cover types were <2º c, implying that different forest cover types may not have varied in quality as thermal refuge. further work is required to resolve these conflicting results. if moose are sensitive to high ambient temperatures, then it follows that moose should exhibit thermoregulatory behavior, especially at the southern edge of their range such as in northeastern minnesota and southeastern ontario. in this study we compared the length of time that temperature was above the 14 and 20˚ c thresholds during summer in northeast minnesota, and used black globe thermometers to compare the thermal environment among forest cover types during variable weather conditions. our objective was also to establish baseline information about how these cover types reflect different thermal conditions during summer. study area this study was conducted in lake county in the arrowhead region of northeast minnesota that has a humid continental climate with severe winters and short, warm summers (frelich 2002). precipitation is moderate with an average of 65-75 cm during spring, summer, and fall. average snowfall along the north shore of lake superior in northeast minnesota is about 180 cm annually with snow cover usually present from december-april. the average july temperature is ~17.5° c and the average january temperature is -17° c (noaa 2009). topography of the region is relatively flat with elevation 460-610 m above sea level. northeast minnesota has near-boreal forests, which are the southern extensions of boreal forest from canada that also contain stands of more southern species not typical of true boreal forests (frelich 2002). these forests are classified into 5 main stand types: 1) fir-birch (abies balsamea, betula papyrifera) forests found on good soils, 2) jack pine (pinus banksiana) and black spruce (picea mariana) on coarse, shallow soils, 3) red maple (acer rubrum), aspen (populus tremuloides), birch, and fir in moist areas, 4) red pine (pinus resinosa) on shallow rocky soils, and 5) birch and white pine (pinus strobus) along lakes and streams. conifer swamps dominated by tamarack (larix laricina) and black spruce are also present. methods we measured temperature in 4 different forest cover types throughout the spatial extent of 95% kernel home ranges of vhf-collared moose in northeast minnesota (moen et al. 2011). capture protocols and survival data for the vhf project are presented elsewhere (lenarz et al. 2009, 2010). random points along roads and trails within the home ranges were generated using arcview 3.3 with a 200 m buffer. black globe temperature data loggers were used to determine differences in radiant heat load of local environments. black globe thermometers measure the thermal environment by incorporating ambient temperature, wind velocity, and radiant energy (bond and kelly 1955). black globe temperature data loggers were placed >25 m and <200 m from road edges within the study area. black globes were constructed from 15 cm diameter copper bulbs painted matte black. hobo® pendant or pro v2 temperature-relative humidity data loggers were attached so sensors hung alces vol. 48, 2012 mcgraw et al. cover type temperature 47 in the center of each globe. hobo pendant temperature data loggers measure air temperature with an accuracy of ± 0.47° c from -20 to 70°c, and hobo pro v2 temperature data loggers measure air temperature with an accuracy of ±0.02° c from 0 to 50° c. temperature data loggers were synchronized to begin taking temperature samples at 6 minute intervals for 120 days during summer. stored temperature readings were downloaded every 30 days to ensure data loggers were working properly and that data were not lost. arcview 3.3 and land use land cover landsat thematic mapper (tm) images were used to identify and locate forest cover types that potentially offered thermal relief within the home ranges. land use land cover (lulc) satellite imagery was used to estimate cover type frequency and distribution in the study area. it is a raster dataset with 30 m resolution, >95% classification accuracy, and 16 defined cover types (minnesota department of natural resources [mndnr] 2007). the 4 cover types were mixed forest consisting of approximately 50% mature deciduous and 50% mature coniferous canopy (41 ± 1%), wetlands consisting of treed bogs (18 ± 2%), coniferous forest (19 ± 1%), and deciduous forest (3 ± 1%); together they composed 80 ± 1% ( x ± se) of the 95% kernel home ranges. forty temperature data loggers were placed within the 4 cover types (10 per cover type); one at each of the random points generated using arcview 3.3 (fig. 1). black globes were secured to the trunk of a tree by a steel eye-bolt attached 75 cm above the ground, the approximate shoulder height of a moose while lying down, and extended 15 cm from the tree. black globes were attached by the eye-bolt extending from the top of the black globe to the steel bolt extending from the tree. all globes were attached to the northeast side of the trees for standardization and to reduce the chance that direct sunlight would be shining on the black globes during the warmest parts of the day. topographic and aspect variation were controlled for by placing black globes at flat locations within the defined cover types. we measured the amount of time when ambient temperature exceeded 14 and 20° c. we also defined a hot day when maximum ambient temperature reached or exceeded 24.4°c, the average maximum daily temperature during july (noaa 2009), to determine if there were greater differences between cover types when ambient air temperature exceeds normal conditions experienced by moose in northeast minnesota. cloudy days were defined as days when the cloud cover index was ≥7, and clear days were defined as days when the cloud cover index was ≤3 (minnesota climatology working group 2010); the cloud cover index is a scale ranging from 1-10. ambient air temperature and cloud cover data for the region were retrieved from noaa and the minnesota state climatology archived data for the isabella weather station located within the study area (fig. 1). temperature data were analyzed with repeated measures anova using statistix (version 9, analytical software, boca raton, florida). bonferroni comparisons were used to test for differences among cover types; significance level was set at p = 0.05 for all tests. sub-samples of the data were analyzed to determine the degree to which cover type temperatures differed due to time of day, season, and climatic events such as hot, cloudy, clear, and hot/clear days. results ambient temperature exceeded 14º c for 50% of the time at the weather station in isabella from 15 june-15 october, 2009 (fig. 1). on days when ambient temperature exceeded 14°c, it typically remained there for about 15.3 h; the longest period was 120 consecutive hours (5 days) during august. during those 5 days, ambient temperature was >20º c between 08:00-20:00 hr, with a mean temperature of 24.5 ± 0.5º c ( x ± se). cover type temperature mcgraw et al. alces vol. 48, 2012 48 ambient temperature exceeded 20º c for 33% of the study period when the mean temperature was 23.0 ± 0.09º c. on average, temperatures remained above 20º c for 6 ± 0.49 h, with the longest continuous period being 18 h. highest black globe temperatures were recorded in the afternoons with the greatest difference in temperature between coniferous and deciduous cover types (17.9 ± 0.4°c vs. 19.6 ± 0.4°c, x ± se; f3,39 = 4.47, p <0.001). black globe temperatures in mixed and bog cover types were intermediate between deciduous and conifer cover types, with the treed bog cover type having slightly cooler temperatures at night (fig. 2). as ambient temperature increased, differences in temperature among cover types were greater, with temperature differences from 1.1º c (above 14º c) to 2.1º c (above 24.4º c). the temperature difference between cover types was greatest between deciduous and coniferous cover types when ambient temperature was >14º c (table 1). during 3 separate 11 hour periods (2 in august, 1 in june) when ambient temperature was >24.4ºc, temperature differences between the deciduous and conifer cover types ranged from 1.5-2.4º c, with warmer periods resulting in larger differences in temperature. temperature differences between cover types were greatest during the afternoons (f3,39 = 4.59, p <0.009) (fig. 2). differences in temperature between cover types were smaller on cloudy days compared to clear days. on clear days (cloud cover index <3) deciduous and bog cover types had the largest difference in temperature (table 1). when we restricted the sampling to afternoon (12:00-16:00 hr), the deciduous cover type again had the highest temperature while the conifer cover type had the lowest temperature (fig. 3, table 1). the difference was greatest during the afternoons of days when temperature was >24.4 º c without cloud cover (fig. 3, table 1). on cloudy days (cloud cover index >7) there were smaller differences in temperature fig. 1. locations of temperature data logger in northeast minnesota. darker gray areas seen in the small insert of northeast minnesota indicate 95% kernel home ranges of vhf-collared moose. forest cover types are from land use land cover coverage classifications. the ambient temperatures used to compare with black globe temperatures were collected at a weather station in isabella, minnesota. alces vol. 48, 2012 mcgraw et al. cover type temperature 49 between cover types, although the deciduous cover type still had the highest temperatures and the coniferous cover type had the lowest temperatures. when the sampling period was restricted to the afternoon, differences became greater once again (fig. 3, table 1). days when ambient temperature exceeded 24.4º c were not cloudy. discussion the uncritical acceptance of 14 and 20°c thresholds for moose thermoregulatory response was recently raised because of temperatures within moose range in southeast ontario (lowe et al. 2010). summer ambient temperatures in cover types used by moose in that study were often above the 14ºc and 20°c thresholds in both day and night. similarly, we found that ambient temperatures were above the thresholds in northeastern minnesota. in both ontario and minnesota there were periods of 3-5 days in the summer when ambient temperatures remained higher than the thresholds. in addition to using ambient temperature as a reference, we measured the operative temperature in different cover types. operative temperature is determined by a black body with the same convection conditions as its environment and produces a net heat flow similar to the heat flow on the surface of an animal (bakken 1980). black globe thermometers measure the thermal environment by incorporating ambient temperature, wind velocity, and radiant energy (bond and kelly 1955). forest canopies filter solar radiation, which causes the greatest difference in equivalent black body temperatures within different cover types (schwab and pitt 1991). increasing crown closure decreases operative temperature as summer thermal cover shelters animals from both heat and radiation. conifer forests often have high levels of crown closure and have the highest degrees of thermal shelter (demarchi and bunnell 1993). differences in operative temperature among cover types used by moose were <2º c during summer, similar to the small differences in temperature among forest cover types measured in southeastern ontario (lowe et fig. 2. average temperatures in deciduous, bog, conifer, and mixed forest cover types over a 24 h period (a). temperatures among cover types showed the greatest divergence during the afternoon (b), typically the warmest part of the day. vertical error bars represent standard errors. cover type temperature mcgraw et al. alces vol. 48, 2012 50 al. 2010). it seems likely that a 2° c difference among cover types is not biologically significant to moose with regard to selecting cover type. rather, selective use of cool areas within cover types may be a thermoregulatory strategy of moose. for example, moose bed in water to cool down (ackerman 1987, renecker and hudson 1990). further, selection and use within cover types for bed sites/microhabitats might relate to the effectiveness of duff layers and soils to act as heat sinks. it is possible that there are differences in cooling capability in different cover types that could lead to cover type selection. moose used conifer stands >30 years old and in which conifer represented ≥75% of the basal area, and also shifted active behavior away from high temperatures of late afternoons (dussault et al. 2004). denser canopies of mature forests were used by moose during hot temperatures in british columbia (demarchi and bunnell 1995, ackerman 1987), and moose in nova scotia reduced movement and occupied cooler habitats during hot weather (broders et al. 2012). also, we think it is worth considering the implementation of critical temperature thresholds. recent papers citing the renecker and hudson papers used 14 and 20° c as thresholds, in part because of the explicit identification of increased breathing rates and panting at those temperatures (renecker and hudson 1986). however, in a later paper, an upper critical temperature range of 14-20°c is specified (renecker and hudson 1990). in addition, figure 3 of renecker and hudson (1986) shows variability in the increase in respiration rate. these differences suggest that renecker and hudson did not intend for the 14 and 20°c to be used as predictors of fig. 3. average temperatures in cover types in northeast minnesota during summer afternoons when ambient temperatures exceeded 24.4° c without cloud cover, during hot afternoons, and during afternoons with cloud cover. vertical error bars represent standard errors. cover types significance weather conditions deciduous bog conifer mixed p-value f3,39 >14°c 19.2 ± 0.04 a 18.4 ± 0.04 b 18.1 ± 0.03 b 18.3 ± 0.03 ab 0.011 4.33 >20°c 23.8 ± 0.06 a 22.7 ± 0.05 ab 22.0 ± 0.04 b 22.3 ± 0.04 ab 0.019 3.81 >24.4°c 26.6 ± 0.11 a 25.4 ± 0.08 ab 24.6 ± 0.08 b 24.8 ± 0.07 b 0.012 4.30 clear days 15.8 ± 0.19 a 14.8 ± 0.18 ab 15.0 ± 0.18 b 15.2 ± 0.18 b <0.003 5.87 cloudy days 12.3 ± 0.17 a 12.0 ± 0.16 a 11.8 ± 0.16 a 12.0 ± 0.16 a 0.054 2.83 cloudy afternoons 15.1 ± 0.16 a 14.5 ± 0.16 ab 14.0 ± 0.15 b 14.3 ± 0.15 ab 0.035 3.23 clear afternoons 23.6 ± 0.44 a 22.4 ± 0.44 ab 21.6 ± 0.44 b 21.7 ± 0.44 b 0.012 4.29 hot/clear afternoons 26.6 ± 0.13 a 25.4 ± 0.10 ab 24.6 ± 0.10 b 24.6 ± 0.09 b 0.012 4.27 table 1. temperature differences measured within cover types in northeast minnesota during summer afternoons under different weather conditions. letters indicate homogenous groups. alces vol. 48, 2012 mcgraw et al. cover type temperature 51 change in habitat use, but rather temperatures at which initial responses might become evident. when we consider the equivocal results of cover type selection, the difference between initiation of a physiological response and implementation of a behavioral response like cover type selection needs to be considered. high temperatures do result in a response by moose, and the ultimate limiting factor for expansion to the south probably remains high temperature (kelsall and telfer 1974). recent declines in moose populations in minnesota have been correlated with increasing summer and winter temperatures. the decline in northwest minnesota was correlated with increases in both summer and winter temperatures (murray et al. 2006). in contrast, high average january and late spring temperatures explained more of the variability in moose survival in northeast minnesota (lenarz et al. 2009, 2010). summer temperatures are predicted to increase 1.5-2° c by 2025 (union of concerned scientists 2003), and by 3-4° c by 2100 (ipcc 2007). if average summer temperatures continue to increase, then the likely result will be more and longer periods during which temperatures are continuously higher than the upper critical limit for moose. this study provides information on the potential for available vegetation types in northeast minnesota to serve as thermal cover. if moose at the southern edge of their range are encountering high temperatures at an increasing and prolonged rate, then their behavior should change. also, there is a need to measure alternative cooling mechanisms at very fine spatial scale, such as the potential heat sink that cool soils may provide while moose are bedded. future research should concentrate on how cover types are used when temperatures are high during both summer and winter and how they function to provide thermal relief. acknowledgements we thank mike schrage for his support of this project. funding for this work was provided by the tribal wildlife grants program, the fond du lac band of lake superior chippewa, the minnesota department of natural resources, the university of minnesota duluth, and the natural resources research institute. summer support was provided by the integrated biosciences graduate program, university of minnesota duluth. this is contribution number 536 from the center for water and the environment at the natural resources research institute, university of minnesota duluth. references ackerman, t. n. 1987. moose response to summer heat on isle royale. m.s. thesis. michigan technical university, houghton, michigan, usa. bakken, g. s. 1980. the use of standard operative temperature in the study of the thermal energetics of birds. physiological zoology 53: 108-119. black, t. a., j. m. chen, x. lee, and r. m. sagar. 1976. relationships of rocky mountain elk and rocky mountain mule deer habitat to timber management in the blue mountains of oregon and washington. pages 11-31 in s. r. heib, editor. proceedings of the elk-logging-roads symposium, moscow, idaho, usa. bond, t. e., and c. f. kelly. 1955. the globe thermometer in agricultural research. agricultural engineering 36: 251. broders, h. g., a. b. coombs, and j. r. mccarron. 2012. ecothermic responses of moose (alces alces) to thermoregulatory stress on mainland nova scotia. alces 48: 0-0. demarchi, m. w. 1991. influence of the thermal environment on forest cover selection and activity of moose in summer. m. s. thesis. university of british columbia, vancouver, british columbia, canada. cover type temperature mcgraw et al. alces vol. 48, 2012 52 _____, and f. l. bunnell. 1993. estimating forest canopy effects on summer thermal cover for cervidae (deer family). canadian journal of forest research 23: 2419-2426. _____, and _____. 1995. forest cover selec-1995. forest cover selection and activity of cow moose in summer. acta theriologica 40: 23-36. dussault, c., j. p. ouellet, r. courtois, j. huot, l. breton, and j. larochelle. 2004. behavioral responses of moose to thermal conditions in the boreal forest. ecoscience 11: 321-328. frelich, l. e. 2002. forest dynamics and disturbance regimes: studies from temperate evergreen-deciduous forests. cambridge university press, cambridge, united kingdom. ipcc (intergovernmental panel on climate change). 2007. fourth assessment report of the intergovernmental panel on climate change, 2007: synthesis report. cambridge university press, cambridge, united kingdom. kelsall j. p., and e. s. telfer. 1974. biogeography of moose with particular reference to western north america. naturaliste canadien 101: 117-130. lenarz, m. s., j. fieberg, m. w. schrage, and a. j. edwards. 2010. living on the edge: viability of moose in northeastern minnesota. journal of wildlife management 74: 1013-1023. _____, m. e. nelson, m. w. schrage, and a. j. edwards. 2009. temperature mediated moose survival in northeastern minnesota. journal of wildlife management 73: 503-510. lowe, s. j., b. r. patterson, and j. a. schaefer. 2010. lack of behavioral responses of moose (alces alces) to high ambient temperatures near the southern periphery of their range. canadian journal of zoology 88: 1032-1041. minnesota department of natural resources (mndnr). 2007. landsat based land use-land cover. http://deli.dnr.state.mn.us/ metadata.html?id=l250000120604 . minnesota climatology working group. 2010. historical climate data retrieval. state climatology office. dnr division of ecological and water resources. http:// climate.umn.edu/doc/historical.htm . moen, r. a., m. e. nelson, and a. edwards. 2011. using cover type composition of home range and vhf telemetry locations of moose to interpret aerial survey results in minnesota. alces 47: 101-112. murray, d. l., e. w. cox, w. b. ballard, h. a. whitlaw, m. s. lenarz, t. w. custer, t. barnett, and t. k. fuller. 2006. pathogens, nutritional deficiency, and climate influences on a declining moose population. wildlife monographs 166: 1-30. noaa (national oceanic and atmospheric administration). 2009. climatological data for ely, minnesota. national climatic data center, ashville, north carolina, usa. renecker, l. a., and r. j. hudson. 1986. seasonal energy expenditures and thermoregulatory responses of moose. canadian journal of zoology 64: 322-327. _____, and _____. 1990. behavioral and thermoregulatory responses of moose to high ambient temperatures and insect harassment in aspen-dominated forests. alces 26: 66-72. schwab, f. e., and m.d.pitt. 1991. moose selection of canopy cover types related to operative temperature, forage, and snow depth. canadian journal of zoology 69: 3071-1077. union of concerned scientists. 2003. confronting climate change in the great lakes regions: impacts on our communities and ecosystems. http://www.ucsusa.org/greatlakes/glchallenge report.html . alces 31_173.pdf alces 31_125.pdf alces 31_185.pdf 1 forage and habitat limitations for moose in the adirondack park, new york samuel peterson1, david kramer2, jeremy hurst2, donald spalinger3, and jacqueline frair1 1state university of new york college of environmental science and forestry, 1 forestry drive, syracuse, new york, usa 13210;2new york department of environmental conservation, division of fish and wildlife, 625 broadway, albany, new york, usa 12233; 3university of alaska, 3211 providence drive, anchorage, alaska 99508, usa abstract: we used browse availability models to estimate the number of reproductive female moose (alces alces) that could be supported during summer and winter in the predominantly forested 23,000 km2 adirondack park and forest preserve (park) in northern new york state, usa. we developed allometric equations to predict available browse biomass for individual plants and subsequent biomass estimates in 6 major cover types to estimate the moose carrying capacity within the park. our model incorporated the differential availability and nutritional quality of woody browse species within each cover type and changes in local browsing intensity due to competing vegetation under two different foraging constraints – protein and digestible energy. we estimated the carrying capacity as 8 (protein constraint) and 135 × (energy constraint) greater in winter than summer. spatially-explicit estimates of summer range capacity (animal use days, aud) based on the protein constraint correlated best with variation in local moose density derived from winter aerial surveys (r2 = 0.75, p < 0.01, n = 18). protein availability was limiting in summer (aud = 457 moose) with sparse patches of regenerating forest (< 20 years old) on privately-managed lands estimated to support 86% more moose than the dominant matrix of wetlands and mature mixed deciduous forest. the small and patchy moose population in the park reflects the relative scarcity of regenerating forest and optimal foraging habitat. given statutory constraints of timber harvest in the majority of the park, active forest management on private inholdings will play an outsized role in managing the moose population. alces vol. 58: 1–30 (2022) key words: adirondacks; alces alces; aud; crude protein; energy; forage; habitat; moose; range capacity. the northeastern united states has one of the largest regional populations of moose (alces alces americana) in north america (jensen et al. 2018). however, within that region are pockets of low density moose as in the adirondack park and forest preserve in new york state that essentially represent the entire state population with density estimated as 0.03 moose/km2 (j. hinton, suny college of environmental science and forestry, unpublished data). in contrast, density in adjacent maine, new hampshire, and vermont has been ≥ 0.3 moose/km2 in recent decades (wattles and destefano 2011). unlike in nearby states, moose in the reserve have limited access to large-scale anthropogenic forest disturbance associated with timber harvesting (hicks 1986). moose require an abundance and dense concentration of quality woody browse to meet their nutritional requirements (illius and gorden 1987, shipley et al. 1994), adirondack moose biomass – peterson et al. alces vol. 58, 2022 2 and large-scale disturbances (fire historically) that set back forest succession creates optimal foraging habitat for about 20 years (peek 2007). in the northeastern united states, the natural, large-scale return interval of inland forests is 1000–7500 years, and more often involves localized winter damage (e.g., ice storms) in small patches rather than large swaths of blown-down trees (lorimer and white 2003, millward and kraft 2004). small scale canopy disturbances from pathogens (e.g., hemlock woolly adelgid [adelges tsugae]) and storms (i.e., winter blowdown) drive localized gap dynamics in the region (runkle 1982, seymour et al. 2002) producing diffuse patches of early seral vegetation. the local scale of these natural disturbances contrast with the more temporally and spatially predictable disturbances arising from timber harvest operations in the contiguous commercial forests of maine, new hampshire, and vermont. the geographic and population expansion of new england moose in the 1970– 1990s was associated with unprecedented clear-cutting in response to a regional spruce budworm (choristoneura fumiferana) infestation that created a contiguous swath of regenerating forest/optimal foraging habitat (bontaites and gustafson 1993, wattles and destefano 2011); moderate high populations have been maintained through continual timber harvest (dunfeyball 2019). research points to moose use and preference of regenerating forest habitat and early successional browse year-round throughout this area (thompson et al. 1995, scarpitti et al. 2005, bergeron et al. 2011, millette et al. 2014). understanding the relationship between availability and nutritional quality of forage resources is imperative to accurately assess the nutritional capacity of a landscape. hobbs and swift (1985) estimated the amount of food required to achieve a diet of specific quality (i.e., food supply), which, when divided by daily dry matter intake rates for the target animal, provides an estimate of animal use days (aud). they used this approach in colorado to evaluate differences in habitat quality between burned and unburned forests for bighorn sheep (ovis canadensis) and mule deer (odocoileus hemionus) in comparison to traditional range supply models (hobbs and swift 1985). hanley et al. (2012) expanded the aud approach with their forage resource evaluation system for habitat (fresh) with sitka black-tailed deer (odocoileus hemionus sitkensis) by using a linear-programming model to estimate the maximum amount of forage biomass that could be pooled from available forage types, while meeting specified nutritional requirements under specific constraints including foraging time, bite size, and diet composition. importantly, such estimates only provide a “snapshot” carrying capacity of a given range. by assuming all available forage was harvested from a given area at a given point in time, the model approximates the days a captive animal is maintained at a desired nutritional plane. such instantaneous estimates ignore plant-herbivore interactions, plant phenology, and dynamic metabolic requirements. as such, they are best interpreted as an index of habitat quality and for comparing the quality of different forage types, cover types, treatments, or areas at a given point in time (cook et al. 2016). though current models are best used as indices, incorporating additional constraints such as resource competition among forage species and accounting for uncertainty may yield estimates more realistic of field conditions (i.e., a multiplier effect; white 1983). one often ignored aspect of estimating range quality is the compounding of alces vol. 58, 2022 adirondack moose biomass – peterson et al. 3 uncertainty as estimates of available biomass are scaled up from individual plants to local plot or transect measures, and again to the level of specific cover types or study areas. dismissal of this attribute of sampling range quality can lead to biased or overly confident estimates of differences in range capacity across heterogeneous landscapes. precision may be grossly overestimated when considered at a single foraging level only, as when accounting for inter-plot variation in biomass while ignoring the precision associated with allometric predictions of the biomass available on a given plant (mcwilliam et al. 1993). monte carlo (mc) simulations can provide a useful approach to account for error propagation across multiple scales or processes (harmon et al. 2007), and have been successfully used to estimate co2 uptake in pine forests (bowler et al. 2012), carbon pools in subtropical forests (conti et al. 2014), and nitrogen density in northeastern hardwood forests (yanai et al. 2010). we applied monte carlo simulations to 1) scale-up estimates of available browse biomass for moose from individual plants to major cover types, and 2) estimated range capacity for moose that accounted for uncertainty in forage quality and foraging constraints during summer and winter. our approach identified the extent to which each component, whether measured empirically in this study or drawn from the literature, contributed to potential bias in and variance around range capacity estimates. we focused on digestible energy and protein as the two most limiting nutritional factors (moen 1995). ultimately, we compared the value of different plants and cover types to identify potentially limiting factors of habitat and management of moose within the adirondack park and forest preserve. study area the study area was delineated as the 23,500 km2 adirondack park and forest preserve of new york state, hereafter referred to as “park” (43°57’08.9”n 74°16’57.5”w), of which ~45% is publicly managed forest preserves interspersed with private inholdings (~55% of the landscape). the majority of public land is protected by article xiv of the new york state constitution as “forever wild” which precludes resource extraction or development of any kind. approximately 25% of privately owned lands are designated for resource management including timber harvest through state-regulated conservation easements. forest canopies in the region are dominated by american beech (fagus grandifolia), red maple (acer rubrum), sugar maple (a. saccharum), yellow birch (betula alleghaniensis), and paper birch (b. papyrifera). common conifer species include white pine (pinus strobus), eastern hemlock (tsuga canadensis), and balsam fir (abies balsamea). elevation in the park ranges from <50 m on the shore of lake champlain to >1600 m in the high peaks (lake placid) area. during data collection, monthly precipitation averaged 84.6 mm in mayseptember 2016, 84.2 mm in december 2016 – march 2017, and 128.3 mm in mayseptember 2017 (n = 17 weather stations; noaa national centers for environmental information). methods plant sampling we sampled woody species along stratified transects (n = 104; fig. 1) proportional to the coverage of upland mixed forest (≥ 497 m elevation, n = 38), lowland mixed forest (< 497 m, n = 34), conifer forest (n = 13), and wetlands (n = 19) within the park using the generalized classification of the terrestrial habitat map produced by the nature adirondack moose biomass – peterson et al. alces vol. 58, 2022 4 fig. 1. study area in northeastern new york state showing public lands (light gray; wild forest, wilderness, primitive use and canoe areas), private lands (white; rural, low, moderate, industrial and intensive use, hamlets and resource management areas), and water bodies (dark gray). also indicated are locations where browse biomass was sampled to build allometric equations (stars), transects where browse components were measured in the field and biomass was predicted using allometric equations (black circles), and locations where browse nutritional samples were collected (white circles). alces vol. 58, 2022 adirondack moose biomass – peterson et al. 5 conservancy for the northeast us and atlantic canada (ferree and anderson 2013; appendix 1). we sub-stratified wetland transects into open wetland (n = 15) and forested wetland (n = 4) classes to capture vegetation patterns based on presence of mature trees within a given wetland. we placed 13 additional transects in the chateaugay woodlands area in the northern region of the park to capture timber harvests of known age (6–8 years old; unpublished data, j. santamour, landvest inc.). transect start points were random locations (create random points in arcgis v.10.4) within each cover type; three 2 × 4-m plots were spaced 50 m apart along a transect. to ensure that we adequately sampled the compositional and productive variation within a landcover type, random locations were spaced a minimum of 500 m from neighboring points. the direction of each transect was randomized using a random number generator (1–360) and modified as needed to ensure all sample plots fell within the same cover type. if we were unable to sample the same cover type within a given plot, we used another random point. we collected summer samples in august-september 2016 and 2017 to assess peak biomass, and winter samples in december 2016 – january 2017. individual plants were sampled to represent the full range of size observed along the transects (peterson et al. 2020). we measured the basal diameter of the main stem (10 cm above the substrate) for tree growth, and the tallest height, longest width, and perpendicular width (to calculate volume) for shrub growth per individual plant (peterson et al. 2020). we clipped all twigs that fell within a 0.5–3.0 m height stratum to 8-mm diameter, a cutoff expected to provide a liberal estimate of available biomass for moose by including both saplings and mature trees with biomass within the stratum (seaton et al. 2002). clippings collected in summer included leaves and twigs, and winter clippings included twigs only for deciduous species and twigs plus needles for balsam fir. we separated and dried leaves and woody mass to a constant mass in a forced air oven at 90 ˚c for 24 h and weighed dried biomass to the nearest gram. allometric equations we developed allometric equations for the 14 species that comprised ≥95% of the woody diet in winter and summer within the adirondack region to more efficiently quantify available moose browse (mcinnes et al. 1992, visscher et al. 2006, peterson et al. 2020). we developed a series of candidate models to evaluate environmental and topographical impacts on dried biomass per species. the dried mass (g) of each individual clipping was ln-transformed to reduce the spread of error and to achieve a normally distributed error structure. we included the environmental site covariates of canopy cover, elevation, percent slope, and aspect. we calculated canopy cover as the average proportion of overstory covering a sample plot from 4 measurements at each of the cardinal directions with a convex densiometer. elevation, percent slope, and aspect were derived from a 90-m resolution digital elevation model (us geological survey) using arcgis (esri, redlands, california, usa). aspect was represented as a binary covariate with northwest values (0–40 degrees and 221–359 degrees) assigned as 0 and southeast values (41–220 degrees) as 1; southeastern facing slopes represent optimal plant growth conditions and increased species richness (olivero and hix 1998). after initial inspection of the relationship between stem size and biomass, we collapsed species (and size classes within species) into 11 groups prior to fitting the final models rather than analyzing the models per individual species based on adirondack moose biomass – peterson et al. alces vol. 58, 2022 6 taxonomic similarities (peterson et al. 2020; table 1). all groups were analyzed for the summer and winter seasons except balsam fir which moose consume in winter and avoid in summer (peterson et al. 2020). we evaluated candidate biomass models using all possible subsets with stepwise regression (whittingham et al. 2006) and used akaike’s information criterion (aicc) for final model selection (burnham and anderson 1998). all model covariates were centered by the mean and standardized prior to model fitting using a z-transformation (schielzeth 2010). we calculated pearson’s correlation coefficient for our predictors and developed candidate models with all plausible covariate pairs having r < 0.7 (dormann et al. 2012), as well as interactions between size and environmental site covariates. additionally, we included quadratic covariates for basal diameter and volume to account for parabolic relationships in plant growth. in the case of uncertainty (i.e., where ∆aicc < 2.0), we predicted biomass with the table 1. allometric models predicting the browsable biomass available to moose in two consecutive winter and summer, 2016–17 as a function of lnbasal diameter (bd; for tree species only), volume (v; for bush species only), percent canopy cover (c), elevation in m (e), percent slope (s), and aspect (a). n indicates the number of samples within each group. interactions among covariates are indicated by “:”. group species included n season model adj. r2 balsam fir abies balsamea 17 winter +8.39 + 0.88bd – 0.01e + 1.04a 0.88 maples sma acer rubrum, a. saccharum 15 summer –2.09 + 2.05bd 0.62 winter –1.43 + 1.89bd – 0.13s 0.71 maples lgb a. rubrum, a. saccharum 17 summer +14.53 – 0.02e 0.29 winter +14.92 – 0.02e 0.32 maples 3 a. pennsylvanicum, a. spicatum 16 summer –8.25 + 6.75bd – 0.85bd2 – 3.51a + 0.31(bd2:a) 0.88 winter –0.01 + 1.17bd – 13.79a + 4.12(bd:a) 0.85 birches betula populifolia, b. papyrifera, 27 summer +0.7 + 3.66bd – 0.52bd2 – 0.02c – 0.79s + 0.17(bd:s) 0.84 b. alleghaniensis, ostrya virginiana winter –26.44 + 15.49bd – 1.85 bd2 + 0.2c – 0.11(bd:c) + 0.01(bd2:c) 0.83 beech smc fagus grandifolia 9 summer +0.84 + 1.53bd – 0.36s 0.69 winter +0.46 + 1.57bd – 0.40s 0.69 beech lgd f. grandifolia 2 summer 5.04 na winter 4.57 na hobblebush viburnum lantanoides 19 summer +4.44 + 0.89v + 0.05v2 0.97 winter +2.96 + 0.96v 0.94 poplars populus grandidentata, 21 summer –8.07 + 6.22bd – 0.71bd2 0.83 p. tremuloides winter –8.98 + 6.26bd – 0.71bd2 0.86 cherries prunus serotina, p. pensylvanica 24 summer +11.49 + 2.19bd – 0.04bd2 – 2.02s + 1.36(bd:s) – 0.23(bd2:s) 0.87 winter –16.07 + 11.42bd – 1.47bd2 – 0.13s 0.79 wild raisin viburnum nudum cassinoides 21 summer +9.11 + 0.99v – 0.01e – 0.02s 0.95 winter +8.97 + 1.00v – 0.01e – 0.03s 0.94 abasal diameter < 55 mm; b basal diameter ≥ 55 mm; cbasal diameter < 60 mm; dbasal diameter ≥ 60 m. alces vol. 58, 2022 adirondack moose biomass – peterson et al. 7 highest ranked model with lowest aicc score that satisfied model assumptions. model assumptions were examined using the shapiro-wilk test for normality, durbinwatson test for autocorrelation, and breuschpagan test for heteroskedasticity. estimating range capacity we estimated available browse biomass for each cover type with monte carlo (mc) sampling (1000 iterations per strata) to account for variation in predictions at each scaling strata (individual plants, plots, transects, and cover types) using r statistical software version 4.0.1 (r core team 2020). we were able to propagate uncertainties at multiple spatial scales by obtaining confidence intervals from the 1000 iterations per each scale. first, we applied allometric equations to predict the available browse biomass on each individual plant by species. we multiplied that prediction by the proportion of the browse on each individual plant species that fell within the defined sampling plot to account for incidences where the entire sampled plant did not fall within the plot. we then resampled predicted values using mc methods to create a grand mean for each individual plant. we summed the predicted biomass across all individuals of a species within each plot. drawing from these values, a second mc sampling generated new values of plot-level biomass for each species, from which we derived a mean biomass per species per plot. across each transect, we summed values of biomass per species/plot and conducted a third mc sampling to generate new values of transect-level biomass for each species. we summed predicted biomass values along transects per cover type, with a final mc sampling generating new values of cover type-level biomass for each species. we converted the final biomass estimates to kg/ha per species for inclusion in range capacity estimates for each cover type. we used the fresh modeling technique (hanley et al. 2012) to quantify the range capacity of each forage and cover type for moose. we specified the model based on the energy and crude protein requirements of an adult female pregnant in winter and lactating in summer. to determine energetic requirements, the mean body weight (bw) for an adult female moose in this region was set to 350 kg based on conversion of fielddressed (carcass) bw reported in new hampshire, vermont, and maine (unpublished data, k. rines, new hampshire fish and game department) with the equation: live weight = carcass weight × 1.46 (crichton 1997) (1) we derived values for daily weight loss, dry matter intake, metabolizable energy requirements (me), foraging time, and bite size from the literature (table 2). estimates of crude protein (cp) and digestible energy (de) available in principle forage species were available from a related study of moose diets in the park (peterson et al. 2020). the fresh model used an energy constraint based on me requirements, but required de values for the input of different forage species; the relationship was assumed as me ≈ de × 0.82 (d. spalinger, author). the fresh model included an input called max that allowed the user to define the maximum amount of biomass of a browsed species that could be included in an individual’s diet (felton et al. 2020). we defined this from the field observed utilization rate of each species from previously conducted browse selection surveys (peterson et al. 2020). we calculated a relative utilization index for each i species with the following equation (gallant et al. 2004, raffel et al. 2009, harrison 2011): ui = # browsed twigs / # available twigs (2) adirondack moose biomass – peterson et al. alces vol. 58, 2022 8 ui was quantified at the transect level and averaged for each species within each cover type. we further adjusted the available browse biomass value predicted for each cover type to account for changes in local moose browsing intensity (fc) due to interfering vegetation (i.e., beech and conifer cover; peterson et al. 2020). to do so, we calculated fc at the cover type level by quantifying the average total biomass of principal browse species, percent cover of beech, and percent cover of conifer within each cover type, and determined the impacts on browse intensity with and without the effects of beech and conifer (peterson et al. 2020). fc was applied as a weight (summer = 0.81– 0.99, winter = 0.94–0.98) to the biomass value for each forage species by cover type in the fresh model. the estimates of animal use days (aud) supported by available browse were ultimately compared with and without the adjustment of fc. applying point estimates for each component in the fresh model, initial estimates of aud were achieved under 2 alternative constraints: 1) a me constraint set at 55,400 kj/day and 27,827 kj/day for summer and winter, respectively, and 2) a cp constraint set at 9.12 and 6.86% for summer and winter, respectively (regelin et al. 1985, reese and robbins 1994). for each scenario, we calculated total range capacity across the park by multiplying the aud/ha of each cover type by the areal extent of each type on the landscape. comprehensive mapping of regenerating forest was lacking for the park. therefore, we estimated its coverage from field surveys that measured the proportion of regenerating forest within private inholdings that were classified a priori as conifer forest, upland deciduous/mixed forest, and lowland deciduous/mixed forest (ferree and anderson 2013); the proportions were 20, 21, and 4%, respectively. we divided the mean aud by 180 days (length of the summer or winter season), multiplied this value by 0.2 to apply a cropping rate ensuring sufficient regeneration without table 2. nutritional requirements and foraging constraints set for a 350 kg, pregnant or nursing, female moose in the fresh cervid model used to estimate animal use days per hectare (aud/ha) by cover type in the adirondack park, new york. constraint season value formula sources daily weight loss both 0.4 kg/day renecker and hudson 1989 dry matter intake summer 11,570 g/day 143 × bw0.75 mcart et al. 2010, renecker and hudson 1989 winter 3075 g/day 38 × bw0.75 metabolizeable energy summer 55,400 kj/day 0.82 × (835 kj × bw0.75) schwartz et al. 1988, renecker and hudson 1989, dungan et al. 2010 winter 27,827 kj/day 0.82 × (124 kcal × bw0.75) × 0.29 kcal/kj crude protein summer 9.12% schwartz et al. 1987, vanballenberghe and miquelle 1990 winter 6.86% foraging time summer 534 min/day risenhoover 1986, dungan et al. 2010 winter 347 min/day bite size summer 1.5 g dry mass moen 1995, moen et al. 1997 winter 1.0 g dry mass alces vol. 58, 2022 adirondack moose biomass – peterson et al. 9 long-term damage to the range (allen et al. 1987), and ultimately considered the season, constraint, and cover types producing the lowest mean aud as most limiting to moose in the system. spatially-explicit estimates of aud were determined by multiplying the estimated aud/ha of a given cover type by the local proportion of that cover type within a defined area. spatially-explicit estimates of moose density were available within a sample of 3 × 10 km blocks surveyed across the park in winter 2016 (j. hinton, suny-esf, unpublished data). survey blocks were excluded from consideration where reliable, on-theground information on regenerating timber cuts was unavailable, yielding 18 blocks for this comparison. we calculated aud using the scenario indicated as most limiting for moose in each season. we also calculated total browse biomass available in each season within these 18 blocks. to quantify the strength of relationship between local aud (or browse biomass) and moose density, we fit a linear model for each scenario (winter versus summer, protein versus energy constraint) and compared models by the variance explained. finally, we applied the fresh model under the most limiting conditions to predict potential changes in aud for moose under 3 scenarios of plausible future landscape changes in the region: 1) warmer and drier conditions leading to 10% conversion of wetland areas (open and wooded wetland types; 33,147 ha) to lowland mixed forest, 2) increased development leading to 10% conversion of coniferous and deciduous forests (176,193 ha) to non-habitat on private lands, and 3) increased timber harvest on private lands converting 2% of deciduous/mixed forest (30,276 ha) and 1% of conifer forest (2481 ha) to regenerating forest, a reasonable estimate based on consultation with local forest managers. we ran 1000 mc simulations in each scenario. results allometric models we sampled 11–32 (ave. = 21) individual plants per species in each season and fit allometric models to 11 distinct plant groupings (appendix 2 and 3). with the exception of large beeches (n = 2, intercept only model) and large maples (n = 17, elevation as sole predictor), basal diameter or volume metrics (with linear or nonlinear terms) were important predictors of available browse biomass of individual plants. we did not detect a difference in allometric relationships for any species among sampling years (models including year effects ∆aicc > 2.0). all selected models met the assumptions of normally-distributed (sw = 0.90 – 0.98), independent (dw = 1.40 – 3.87), and homoscedastic errors (bp = 0.15 – 10.60; all p > 0.05), with the exception of the northern wild raisin (viburnum nudum cassanoides) model in summer which indicated dependent errors (dw = 1.40, p = 0.04). in both seasons, all competing best models (∆aicc < 2.0) violated assumptions of either normally distributed, independent, or homoscedastic error. the first model that met all model assumptions was less supported (∆aicc = 2.20) than those with violations, but that model predicted values that differed from the best-supported model by ~1% only. therefore, we chose to use the top model despite violating model assumptions. available browse biomass four species that were widespread in each cover type represented ~70% of available browse biomass in summer (fig. 2); hobblebush (viburnum lantanoides, 27.7%) and striped maple (acer pensylvanicum, 23.8%) provided ~50% of biomass with red maple (11.4%) and yellow birch (9.2%) combining for ~20%. northern wild raisin (10.2%) and sugar maple (9.7%) also provided ~20% combined but were not distributed evenly adirondack moose biomass – peterson et al. alces vol. 58, 2022 10 fig. 2. estimated density of browsable biomass (kg/ha) for moose during summer in the adirondack park, 2016–17. estimates correspond to principal browse species only, those making up 95% of moose seasonal diets, with mean values reported by primary cover type. alces vol. 58, 2022 adirondack moose biomass – peterson et al. 11 across all cover types. in winter, balsam fir dominated browse availability (57.4%; fig. 3) followed by striped maple (14.9%) and yellow birch (8.7%). across cover types, the estimated reduction in browse intensity ranged from 2.3% (open wetland) to 18.5% (upland and lowland deciduous/mixed forest) in summer, and 1.1% (conifer forest) to 5.5% (open wetland) in winter. by cover type, available browse biomass ranged from 87.0 (lowland deciduous/mixed forest) to 556.5 kg/ha (regenerating forest) during summer, and 146.7 (upland deciduous/mixed forest) to 376.0 kg/ha (wooded wetlands) during winter. regenerating forest provided 2–6 × more available browse in summer than other cover types, and yielded 64% more browse than the dominant cover type (upland and lowland deciduous/mixed forest covering 63% of the landscape). in contrast, availability of winter browse biomass was more evenly distributed with wooded wetland (23.1%), open wetland (21.3%), and regenerating forest (18.5%) providing ~60% combined. available browse within regenerating forest was dominated by yellow birch (38.4%) and several maple species (35.3%), whereas balsam fir (58.5– 93.7%) was dominant in wetland types. in both seasons, available browse biomass was more homogeneous in regenerating forest than in other cover types (cv = 0.42 in summer and 0.23 in winter), as well as among transects in that type (cv = 0.13 in summer and 0.089 in winter). predicted range capacity the distribution of cover types across the park was dominated by deciduous/mixed forest (64% combined): 33.3% lowland deciduous/mixed forest (737,710 ha), 30.9% upland deciduous/mixed forest (685,396 ha), 10.4% conifer forest (230,043 ha), 10.2% forested wetland (225,280 ha), 4.6% regenerating forest (102,214 ha), and 4.4% open wetland (97,557 ha). the summer range supported fewer adult female moose (0.00–5.97 aud/ha across cover types) than the winter range (0.23–26.54 aud/ha) (table 3). only upland deciduous/mixed forest and regenerating forest provided browse sufficient to meet the protein needs of lactating moose in summer. regenerating forest supported 6 × more aud/ha than upland deciduous/mixed forest based on protein requirements. moreover, browse within regenerating forest and wooded wetland provided the most abundant sources of available energy in summer. every habitat type produced sufficient browse in winter to meet the protein and energy requirements of pregnant moose due to the high use of balsam fir and the lower daily energetic requirement in winter relative to summer. in winter, the highest aud was predicted within wooded and open wetlands because of the abundance of balsam fir. we predicted the total available browse (under a protein constraint) in summer to support 457 ± 240 sd reproductive females when accounting for uncertainty (table 4). spatially-explicit estimates (within individual aerial survey blocks, n = 18) of summer aud/ha under the crude protein constraint explained a moderate and significant amount of variation in the observed moose density via aerial surveys (r2 = 0.75, p < 0.01; fig. 4). block-level density estimates for moose were less strongly related to aud estimates under the summer energy constraint (r2 = 0.45, p < 0.01), winter energy constraint (r2 = 0.36, p < 0.01), and winter protein constraint (r2 = 0.12, p = 0.15). likewise, available browse biomass within each sampling block during summer showed a moderately strong relationship with estimated moose density in winter (r2 = 0.64, p < 0.01), whereas browse biomass in winter was a poor predictor of winter moose density (r2 = –0.06, p = 0.90). adirondack moose biomass – peterson et al. alces vol. 58, 2022 12 fig. 3. estimated density of browsable biomass (kg/ha) for moose during winter in the adirondack park, 2016–17. estimates correspond to principal browse species only, those making up 95% of moose seasonal diets, with mean values reported by primary cover type. alces vol. 58, 2022 adirondack moose biomass – peterson et al. 13 landscape changes scenarios we used summer protein constraint for our landscape change scenarios because it was identified as the most limiting factor for moose across season (summer and winter) and resource (crude protein and energy). in terms of model sensitivity, and in the absence of diet selectivity and ignoring model uncertainty, the potential reduction in browse intensity due to interfering vegetation alone yielded a point estimate for summer range capacity of 587 reproductive females. accounting for diet selection, the estimates of summer aud were reduced 82.2%, and accounting for interfering vegetation further reduced the estimates by 11.4%; combined, these factors reduced aud 84.3%. ultimately, uncertainty in diet selectivity explained the greatest proportion of total variance in the final estimates (ave. cv = 22% across types and scenarios). uncertainty in the estimated crude protein content of species or in predicted biomass availability by cover type contributed minimally to the variance in final estimates (<1% each). removal of mature forest yielded the largest change in range capacity either by table 3. the estimated number of seasonal animal use days (aud) per hectare (with standard deviation) for moose in 6 different cover types within the adirondack park, new york. error from biomass availability, nutritional content of browse and diet selection habits of moose (peterson et al. 2020) were incorporated through a series of monte carlo simulations. cover type summer aud winter aud protein energy protein energy conifer forest – 0.48 (0.52) 13.81 (11.14) 13.51 (11.17) upland decid/mixed forest 0.28 (0.23) 0.70 (0.34) 5.74 (3.30) 3.16 (2.69) lowland decid/mixed forest – 0.71 (0.34) 16.18 (14.47) 15.84 (14.03) wooded wetland – 5.97 (5.05) 26.02 (21.43) 15.93 (15.10) open wetland – 0.57 (0.46) 23.74 (21.04) 26.54 (22.90) regenerating forest 1.95 (1.69) 5.61 (2.87) 11.28 (9.23) 0.23 (0.22) table 4. estimates of park-wide nutritional carrying capacity of reproductive female moose in adirondack park, new york by (a) total abundance and (b) density. values shown incorporate error from biomass availability, nutritional quality of forage and diet selection habits (ui ) of moose. browse intensity values (fc) are also included in these estimates. a. total moose summer winter crude protein energy crude protein energy mean 456.5 3453.20 32897.70 26998.70 sd 239.6 1267.10 12341.80 11612.10 b. moose per km2 mean 0.02 0.15 1.4 1.15 sd 0.01 0.05 0.53 0.49 fig. 4. predicted number of reproductive female moose supported by available summer browse (animal use days) within 18, 3 × 10 km2 survey blocks (based on protein requirement) compared to the empirically-estimated density of moose in those blocks during winter 2016. adirondack moose biomass – peterson et al. alces vol. 58, 2022 14 reducing capacity through conversion to non-habitat (development scenario) or increasing capacity through conversion to regenerating forest (timber harvest scenario; fig. 5). in contrast, loss of wetlands (drier conditions scenario) or conversion of wetlands to lowland deciduous/mixed forest (the only scenario that adds additional forest) yielded minor impact (<5%) on range capacity. models predicted that range capacity might increase by 30% in summer through harvest of 30,000 ha of mixed forest and 2400 ha of conifer forest. all landscape change scenarios predicted minimal reduction in winter range capacity (< ~5%). discussion our models indicate that the current and future moose population within the park is constrained by summer protein availability associated with lack of early successional/ regenerating forest. the summer population estimate of ~455 reproductive females (under a protein constraint) was surprisingly similar to the total population estimate from winter aerial surveys (~700 moose; j. hinton, suny college of environmental science and forestry, unpublished data); more importantly, spatially-explicit estimates explained a substantial amount of variation in local moose density. we recognize that while leaves and stems of woody species dominate moose diets (belovsky 1981), aquatic vegetation also provides measurable summer-fall forage when available (crete and jordan 1981). because we did not account for aquatic vegetation in the models and wetlands are widespread and common in the park (14.6% was classified as open or forested wetland), we presume that our estimates of summer range capacity are somewhat low. conversely, our unreasonably large estimate of winter range capacity (>25,000 animals or >1 animal/km2) was biased high because we did not account for selectivity (pastor and danell 2003) within species and age classes of browse (e.g., balsam fir foliage), effects of secondary fig. 5. predicted changes in the total number of reproductive female moose supported by available browse in the adirondack park under 3 landscape change scenarios. scenarios represent drying conditions leading to a 10% conversion of wooded and open wetland habitat to lowland deciduous/ mixed forest, increased development leading to 10% conversion of conifer and deciduous/mixed forest to non-habitat, and increased timber harvest leading to 1% conversion of conifer forest and 2% conversion of deciduous/mixed forest to regenerating forest. values indicate the percent change in the number of female moose relative to contemporary range conditions. alces vol. 58, 2022 adirondack moose biomass – peterson et al. 15 compounds on limiting consumption of balsam fir (parikh et al. 2017, nosko et al. 2020), and the influence of snow depth on browse availability (visscher et al. 2006). however, that winter browse would need to be reduced 99% to match the limitation on summer range arguably reflects the simple interaction of an abundance of balsam fir with low nutrient requirements in the winter model. although sample size per modeled vegetative class was smaller than desirable, we did assess variation in available biomass production and palatability across both species and size classes by separating our samples by size and species when estimating aud (peterson et al. 2020). despite certain limitations in our assumptions and sample sizes, we believe our data and analyses were sufficient to conclude that summer protein limitation is and will be the primary determinant of the stability and growth of the park moose population. where this moose population stands with respect to potential nutritional carrying capacity remains an open question beyond the lack of regenerating forest habitat. the persistent, low density population is consistent with its stable, yet limited availability of summer forage protein, and not necessarily reflective of low recruitment and survival. ungulate populations near carrying capacity trend toward reduced productivity (couturier et al 2009, wam et al. 2010), and the 3-year calf:cow ratio (0.5 ± 0.9 sd) from limited aerial sampling in the park (nys dec, unpublished data) is considered moderate productivity (kuzyk et al. 2018). this population is currently protected from harvest and without a major predator, with brainworm (parelaphostrongylus tenuis) and liver fluke (fascioloides magna) causing 12% of incidental natural mortality (k. schuler, cornell university college of veterinary medicine, unpublished data). brainworm is somewhat of concern given the high deer density around most of the park and the current low-moderate deer density within. the primary objective was to determine if nutritional limitation for moose occurs in the park and our models indicate that the moose population is under summer protein constraint due to lack of optimal foraging habitat associated with regenerating forest. in nearby moose populations inhabiting mostly commercial forestland (maine, new hampshire, vermont), optimal foraging habitat is generated continuously and used year-round (dunfey-ball 2019). numerous regional studies have identified use and preference of regenerating forest in winter, investigated and measured winter browsing damage on valuable, regenerating deciduous species, and documented use of mature coniferous forest when mobility and activity is restricted temporarily in extreme winter conditions (thompson et al. 1995, scarpitti et al. 2005, bergeron et al. 2011, millette et al. 2014, andreozzi et al. 2016). thus, it is not surprising that constrained availability of summer biomass was correlated positively with winter moose density, and conversely, that winter browse biomass dominated by balsam fir was a poor predictor of winter moose density. our inflated estimates of winter adu reflect the high availability and presumed use of balsam fir in the model, and inadvertently minimized the effect of limited deciduous browse in regenerating forest that is preferred winter forage. that we found restricted seasonal forage/nutrition and low moose population density were linked in the park reflects the direct relationship between availability of year-round optimal foraging habitat and population abundance (peek 2007). importantly, conversion of mature forests to regenerating stands would increase optimal and preferred forage in both summer and winter. adirondack moose biomass – peterson et al. alces vol. 58, 2022 16 legal harvest (hunting) is the most common tool for managing ungulate populations within their carrying capacity or to reduce over-populations. instituting harvest of the small park population is not feasible under current statutory constraints. although the recent history of regional moose harvest is short-term (since the 1980s), harvest reductions were implemented by the mid-2000s in new hampshire and vermont despite conservative harvest rates in populations of multiple thousands of animals. in these small states, the interrelationships of moderate-high moose density, winter tick parasitism, and climate change have reduced productivity in primary moose range despite commercial private forests continuously producing optimal foraging habitat (dunfey-ball 2019, jones et al. 2019, pekins 2020, debow et al. 2021). in contrast, the adirondack park has a low density moose population, lack of timber harvest and optimal foraging habitat, and to date, unmeasured population impact from winter tick. because the population is constrained by lack of optimal foraging habitat, strategic forest management (cutting) to address this limitation is desirable, yet that is also prohibited by statute in much of the park. the regional population explosion of moose in the 1970–1990s reflected unprecedented harvest rates of spruce-fir forests in response to a spruce budworm infestation in the 1970–80s. ironically, a trio of forest pests may provide a natural process to improve moose habitat in the park on public lands through canopy openings in response to the current invasion by hemlock woolly adelgid, red pine scale (matsucoccus resinosae), and again, the spruce budworm in the northeastern united states. acknowledgements we would like to thank our partners at suny-esf, nysdec, wcs and cornell university for their assistance. this project would not have been possible without the cooperation of commercial foresters of the adirondacks including lyme adirondacks, the forestland group and molpus woodlands. additionally, we would like to thank k. powers, d. tinklepaugh, r. rich, d. degroff, and r. tam for their assistance with data collection and field support. funding for this project was provided by suny-esf, nys-dec (federal aid in wildlife restoration grant w-173-g), and the american wildlife conservation society. references allen, a. w., p. a. jordan, and j. w. terrell. 1987. habitat suitability index models: moose, lake superior region. united states fish and wildlife service report 82(10.155). united states department of the interior, fish and wildlife service research and development, washington, dc, usa. andreozzi, h. a., p. j. pekins, and l. e. kantar. 2016. using aerial survey observations to identify winter habitat use of moose in northern maine. alces 52: 41–53. doi: 10.1007/bf00346984 belovsky, g. e. 1981. optimal activity times and habitat choice of moose. oecologia 48: 22–30. bergeron, d. h., p. j. pekins, p. j., h. f. jones, and w. b. leak. 2011. moose browsing and forest regeneration: a case study in northern new hampshire. alces 47: 39–51. bontaites, k. m., and k. gustafson. 1993. the history and status of moose and moose management in new hampshire. alces 29: 163–167. bowler, r. a. l., m. freden, m. brown, and t. a. black. 2012. residual vegetation importance to net co2 uptake in pine-dominated stands following mountain pine beetle attack in british columbia, canada. forest ecology and management 269: 82–91. doi: 10.1016/j. foreco.2011.12.011 alces vol. 58, 2022 adirondack moose biomass – peterson et al. 17 burnham, k. p., and d. r. anderson. 1998. model selection and inference: a practical information-theoretic approach. springer-verlag, new york, new york, usa. conti, g., n. perez-harguindeguy, f. quetier, l. d. gorne, p. jaureguiberry, g. a. bertone, l. enrico, a. cuchietti, and s. diaz. 2014. large changes in carbon storage under different land-use regimes in subtropical seasonally dry forests of southern south america. agriculture, ecosystems and environment 197: 68–76. doi: 10.1016/j.agee.2014. 07.025 cook, j. g., r. c. cook, r. w. davis, and l. l. irwin. 2016. nutritional ecology of elk during summer and autumn in the pacific northwest. wildlife monographs 195: 1–81. doi: 10.1002/wmon.1020 couturier, s., s. d. côté, j. huot, and r. d. otto. 2009. body-condition dynamics of a northern ungulate gaining fat in winter. canadian journal of zoology 87: 367–378. doi: 10.1139/z09-020 crete, m., and p. a. jordan. 1981. régime alimentaire des orignaux du sud-ouest québécois pour les mois d’avril à octobre. canadian field-naturalist 95: 50–56. crichton, v. 1997. hunting. pages 617–653 in a. w. franzmann and c. c schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, dc, usa. debow, j., j. blouin, e. rosenblatt, c. alexander, k. gieder, w. cottrell, j. murdoch, and t. donovan. 2021. effects of winter ticks and internal parasites on moose survival in vermont, usa. journal of wildlife management 85: 1423–1439. doi: 10.1002/jwmg.22101 dormann, c. f., j. elith, s. bacheracher, c. buchmann, g. carl, g. carre, j. r. garcia marquez, b. gruber, b. lafourcade, p. j. leitao, t. munkemuller, c. mcclean, p. e. osborne, b. reineking, b. schroder, a. k. skidmore, d. zurell, and s. lautenbach. 2012. collinearity: a review of methods to deal with it and a simulation study evaluating their performance. ecography 36: 27–46. doi: 10.1111/j.1600-0587.2012.07348.x dunfey-ball, k. r. d. 2019. moose density, habitat and winter tick epizootics in a changing climate. ms thesis, university of new hampshire, durham, new hampshire, usa. dungan, j. d., l. a. shipley, and r. g. wright. 2010. activity patterns, foraging ecology, and summer range carrying capacity of moose (alces alces shirasi) in rocky mountain national park, colorado. alces 46: 71–87. felton, a. m., e. holmstrom, j. malmsten, a. felton, j. p. g. m. cromsigt, l. edenius, g. ericsson, f. widemo, and h. k. wam. 2020. varied diets, including broadleaved forage, are important for a large herbivore species inhabiting highly modified landscapes. scientific reports 10: 1904. doi: 10.1038/ s41598020-58673-5 ferree, c., and m. g. anderson. 2013. a map of terrestrial habitats of the northeastern united states: methods and approach. the nature conservancy, boston, massachusetts, usa. gallant, d., c. h. berube, t. termblay, and l. vasseur. 2004. an extensive study of the foraging ecology of beavers (castor canadensis) in relation to habitat quality. canadian journal of zoology 82: 922–933. doi: 10.1139/z04-067 hanley, t. a., d. e. spalinger, k. j. mock, o. l. weaver, and g. m. harris. 2012. forage resource evaluation system for habitat-deer: an interactive deer habitat model. technical report pnwgtr-858. united states department of agriculture, forest service, pacific northwest research station general, portland, oregon, usa. harmon, m. e., d. l. phillips, j. battles, a. rassweiler, r. o. hall, and w. k. lauenroth. 2007. quantifying adirondack moose biomass – peterson et al. alces vol. 58, 2022 18 uncertainty in net primary production measurements. pages 238–260 in t. j. fahey and a. k. knapp, editors. principals and standards for measuring primary production. oxford university press, new york, new york, usa. harrison, a. m. 2011. landscape influences on site occupancy by beaver and resultant foraging impacts on forest composition and structure (adirondack mountains, ny, usa). ms thesis, state university of new york college of environmental science and forestry, syracuse, new york, usa. hicks, a. 1986. history and current status of moose in new york. alces 22: 245–252. hobbs, n. t., and d. m. swift. 1985. estimates of habitat carrying capacity incorporating explicit nutritional constraints. journal of wildlife management 49: 814–822. doi: 10.2307/3801716 illius, a. w., and i. j. gordon. 1987. the allometry of food intake in grazing ruminants. journal of animal ecology 56: 989–99. doi: 10.2307/4961 jensen, w. f., j. r. smith, m. carstensen, c. e. penner, b. m. hosek, and j. j. maskey. 2018. expanding gis analysis to monitor and assess north american moose distribution and density. alces 54: 45–54. jones, h., p. pekins, l. kantar, i. sidor, d. ellingwood, a. lichtenwalner, and m. o’neal. 2019. mortality assessment of moose (alces alces) calves during successive years of winter tick (dermacentor albipictus) epizootics in new hampshire and maine (usa). canadian journal of zoology 97: 22–30. doi: 10.1139/ cjz-2018-0140 kuzyk, g., i. hatter, s. marshall, c. procter, b. cadsand, d. lirette, h. schindler, m. bridger, p. stent, a. walker, and m. klaczek. 2018. moose population dynamics during 20 years of declining harvest in british columbia. alces 54: 101–119. lorimer, c. g., and a. s. white. 2003. scale and frequency of natural disturbances in the northeastern us: implications for early successional forest habitats and regional age distributions. forest and ecology management 185: 41–64. doi: 10.1016/s0378-1127(03)00245-7 mcart, s. h., d. e. spalinger, w. b. collins, e. r. schoen, t. stevenson, and m. bucho. 2009. summer dietary nitrogen availability as a potential bottom-up constraint on moose in south-central alaska. ecology 90: 1400– 1411. doi: 10.1890/08-1435.1 mcinnes, p. f., r. j. naiman, j. pastor, and y. cohen. 1992. effects of moose browsing on vegetation and litter of the boreal forest, isle royale, michigan, usa. ecology 73: 2059–2075. doi: 10.2307/1941455 mcwilliam, a. l. c, j. m. roberts, o. m. r. cabral, m. v. b. r. leitao, a. c. l. decosta, g. t. maitelli, and c. a. g. p. zamparoni. 1993. leaf area index and above-ground biomass of terra firme rain forest and adjacent clearings in amazonia. functional ecology 7: 310–317. doi: 10.2307/2390210 millette, t. l., e. marcano, and d. laflower. 2014. winter distribution of moose at landscape scale in northeastern vermont: a gis analysis. alces 50: 17–26. millward, a. a., and c. e. kraft. 2004. physical influence of landscape on a large-extent ecological disturbance: the northeastern north american ice storm of 1998. landscape ecology 19: 99–111. moen, r. a. 1995. moose energetics, foraging strategies and landscape effects: a spatially explicit simulation model. phd dissertation, university of minnesota, st. paul, minnesota, usa. moen, r. a., j. pastor, and y. cohen. 1997. a spatially explicit model of moose foraging and energetics. ecology 78: 505–521. doi: 10.1890/0012-9658 (1997) 078[0505: asemom]2.0.co;2 nosko, p., k. roberts, r. knight, and a. marcellus. 2020. growth and chemical alces vol. 58, 2022 adirondack moose biomass – peterson et al. 19 response of balsam fir saplings released from intense browsing pressure in the boreal forests of western newfoundland, canada. forest ecology and management 460: 117839. doi: 10.1016/j.foreco.2019. 117839 olivero, a. m., and d. m. hix. 1998. influence of aspect and stand age on ground flora of southeastern ohio forest ecosystems. plant ecology 139: 177–187. doi: 10.1023/a:1009758501201 parikh, g. l., j. s. forbey, b. robb, r. o. peterson, l. m. vucetich, and j. a. vucetich. 2017. the influence of plant defensive chemicals, diet composition, and winter severity on the nutritional condition of a free-ranging, generalist herbivore. oikos 126: 196–203. doi: 10.1111/oik.03359 pastor, j., and k. danell. 2003. moosevegetation-soil interactions: a dynamic system. alces 39: 177–192. peek, j. m. 2007. habitat relationships. pages 351–375 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose, 2nd edition. university press of colorado, boulder, colorado, usa. pekins, p. j. 2020. metabolic and population effects of winter tick infestations on moose: unique evolutionary circumstances? frontiers in ecology and evolution 8:176. doi: 10.3389/fevo. 2020.00176 peterson, s., d. kramer, j. hurst, and j. frair. 2020. browse selection by moose in the adirondack park, new york. alces 56: 107–126. r core team. 2020. r: a language and environment for statistical computing. r foundation for statistical computing, vienna, austria. . raffel, t. r., n. smith, c. cortright, and a. j. gatz. 2009. central place foraging by beavers (castor canadensis) in a complex lake habitat. american midland naturalist 162: 62–73. doi: 10.1674/ 0003-0031-162.1.62 reese, e. o., and c. t. robbins. 1994. characteristics of moose lactation and neonatal growth. canadian journal of zoology 72: 953–957. doi: 10.1139/ z94-130 regelin, w. l., c. c. schwartz, and a. w. franzmann. 1985. seasonal energy metabolism of adult moose. journal of wildlife management 49: 388–393. doi: 10.2307/3801539 renecker, l. a., and r. j. hudson. 1989. ecological metabolism of moose in aspen-dominated boreal forest, central alberta. canadian journal of zoology 67: 1923–1928. doi: 10.1139/z89-275 risenhoover, k. l. 1986. winter activity patterns of moose in interior alaska. journal of wildlife management 50: 727–734. doi: 10.2307/3800990 runkle, j. r. 1982. patterns of disturbance in some old-growth mesic forests of eastern north america. ecology 63: 1533–1546. doi: 10.2307/3800990 scarpitti, d., c. habeck, a. r. musante, and p. j. pekins. 2005. integrating habitat use and population dynamics of moose in northern new hampshire. alces 41: 25–35. schwartz, c. c., m. e. hubbert, and a. w. franzmann. 1988. energy requirements of adult moose for winter maintenance. the journal of wildlife management 52: 26–33. doi: 10.2307/ 3801052 schwartz, c. c., w. l. regelin, and a. w. franzmann. 1987. protein digestion in moose. journal of wildlife management 51: 352–357. doi: 10.2307/3801052 schielzeth, h. 2010. simple means to improve the interpretability of regression coefficients. methods in ecology and evolution 1: 103–113. doi: 10.1111/j.2041-210x.2010.00012.x seaton, c. t. 2002. winter foraging ecology of moose in the tanana flats and alaska range foothills. ms thesis, https://www.r-­project.org/ https://www.r-­project.org/ adirondack moose biomass – peterson et al. alces vol. 58, 2022 20 university of alaska fairbanks, fairbanks, alaska, usa. seymour, r. s., a. s. white, and p. g. demayandier. 2002. natural disturbance regimes in northeastern north americaevaluating silvicultural systems using natural scales and frequencies. forest ecology and management 155: 357–367. doi: 10.1111/j.2041210x. 2010.00012.x shipley, l. a., j. e. gross, d. e. spalinger, n. t. hobbs, and b. a. wunder. 1994. the scaling of intake rate in mammalian herbivores. american naturalist. 143: 1055–1082. doi: 10.1086/285648 thompson, m. e., j. r. gilbert, g. j. matula jr., and k. i. morris. 1995. seasonal habitat use by moose on managed forest lands in northern maine. alces 31: 233–245. visscher, d. r., e. h. merrill, d. fortin, and j. l. frair. 2006. estimating woody browse availability for ungulates at increasing snow depths. forest ecology and management 222: 348–354. doi: 10.1016/j.foreco. 2005.10.035 wam, h. k., o. hofstad, and e. j. solberg. 2010. differential forage use makes carrying capacity equivocal on ranges of scandinavian moose (alces alces). canadian journal of zoology 88: 1179– 1191. doi: 10.1139/ z10-084 wattles, d. w., and s. destefano. 2011. status and management of moose in the northeastern united states. alces 47: 53–68. white, r. g. 1983. foraging patterns and their multiplier effects on productivity of northern ungulates. oikos 40: 377–384. doi: 10.2307/3544310 whittingham, m. j., p. a. stephens, r. b. bradbury, and r. p. freckleton. 2006. why do we still use stepwise modelling in ecology and behaviour? journal of animal ecology 75: 1182–1189. doi: 10.1111/j.1365-2656.2006.01141.x yanai, r. d., j. j. battles, a. d. richardson, c. a. blodgett, d. m. wood, and e. b. rastetter. 2010. estimating uncertainty in ecosystem budget calculations. ecosystems 13: 239––248. doi: 10.1007/ s10021-010-9315-8 alces vol. 58, 2022 adirondack moose biomass – peterson et al. 21 appendices appendix 1: reclassification of ecosystems defined by the nature conservancy’s terrestrial habitat map for the northeastern us and atlantic (ferree and anderson 2013). ecosystems were assigned to either conifer, deciduous/mixed, wetland, wooded wetland or no-sampling based on ecosystem descriptions provided by ferree and anderson 2013. the deciduous/mixed class was further separated into upland and lowland forests, using a cutoff of 497 m in elevation. tnc class tnc ecosystem reclassification upland acadian low elevation spruce-fir-hardwood forest conifer upland acadian sub-boreal spruce flat conifer upland acadian-appalachian alpine tundra no sampling upland acadian-appalachian montane spruce-fir-hardwood forest conifer upland agriculture no sampling upland appalachian (hemlock)-northern hardwood forest: drier deciduous/mixed upland appalachian (hemlock)-northern hardwood forest: moist-cool deciduous/mixed upland appalachian (hemlock)-northern hardwood forest: typic deciduous/mixed upland central appalachian dry oak-pine forest deciduous/mixed upland central appalachian pine-oak rocky woodland deciduous/mixed upland developed no sampling upland glacial marine & lake mesic clayplain forest deciduous/mixed upland great lakes alvar no sampling upland laurentian acidic rocky outcrop no sampling upland laurentian-acadian acidic cliff and talus no sampling upland laurentian-acadian calcareous cliff and talus no sampling upland laurentian-acadian calcareous rocky outcrop no sampling upland laurentian-acadian northern hardwood forest: high conifer deciduous/mixed upland laurentian-acadian northern hardwood forest: moist-cool deciduous/mixed upland laurentian-acadian northern hardwood forest: typic deciduous/mixed upland laurentian-acadian northern pine-(oak) forest deciduous/mixed upland laurentian-acadian pine-hemlock-hardwood forest: moist-cool deciduous/mixed upland laurentian-acadian pine-hemlock-hardwood forest: typic deciduous/mixed upland laurentian-acadian red oak-northern hardwood forest deciduous/mixed upland north-central appalachian acidic cliff and talus no sampling upland north-central appalachian circumneutral cliff and talus no sampling upland northeastern interior pine barrens conifer upland northern appalachian-acadian rocky heath outcrop no sampling upland open water no sampling upland shrubland/grassland; mostly ruderal shrublands, regenerating clearcuts no sampling wetland boreal-laurentian bog: isolated/small stream wetland wetland boreal-laurentian-acadian acidic basin fen: undifferentiated wetland wetland glacial marine & lake wet clayplain forest: undifferentiated wetland wetland laurentian-acadian alkaline conifer-hardwood swamp: isolated wooded wetland wetland laurentian-acadian alkaline conifer-hardwood swamp: lake/pond: any size wooded wetland appendix 1 (continued) adirondack moose biomass – peterson et al. alces vol. 58, 2022 22 appendix 1 (continued): reclassification of ecosystems defined by the nature conservancy’s terrestrial habitat map for the northeastern us and atlantic (ferree and anderson 2013). ecosystems were assigned to either conifer, deciduous/mixed, wetland, wooded wetland or no-sampling based on ecosystem descriptions provided by ferree and anderson 2013. the deciduous/mixed class was further separated into upland and lowland forests, using a cutoff of 497 m in elevation. tnc class tnc ecosystem reclassification wetland laurentian-acadian alkaline conifer-hardwood swamp: smaller river floodplain/riparian wooded wetland wetland laurentian-acadian freshwater marsh: isolated wetland wetland laurentian-acadian freshwater marsh: lake/pond: any size wetland wetland laurentian-acadian freshwater marsh: smaller river floodplain/riparian wetland wetland laurentian-acadian large river floodplain: acidic swamp wetland wetland laurentian-acadian large river floodplain: alkaline conifer-hardwood swamp wooded wetland wetland laurentian-acadian large river floodplain: conifer-hardwood acidic swamp wooded wetland wetland laurentian-acadian large river floodplain: floodplain forest wooded wetland wetland laurentian-acadian large river floodplain: freshwater marsh wetland wetland laurentian-acadian large river floodplain: wet meadow-shrub swamp wetland wetland laurentian-acadian wet meadow-shrub swamp: isolated wetland wetland laurentian-acadian wet meadow-shrub swamp: lake/pond: any size wetland wetland laurentian-acadian wet meadow-shrub swamp: smaller river floodplain/ riparian wetland wetland north-central appalachian acidic swamp: isolated wetland wetland north-central appalachian acidic swamp: lake/pond: any size wetland wetland north-central appalachian acidic swamp: smaller river floodplain/riparian wetland wetland north-central appalachian large river floodplain: acidic swamp wetland wetland north-central appalachian large river floodplain: acidic swamp wetland wetland north-central appalachian large river floodplain: freshwater marsh wetland wetland north-central appalachian large river floodplain: rich swamp wetland wetland north-central appalachian large river floodplain: rich swamp wetland wetland north-central appalachian large river floodplain: wet meadow-shrub swamp wetland wetland north-central interior and appalachian acidic peatland: undifferentiated wetland wetland north-central interior and appalachian rich swamp: isolated wetland wetland north-central interior and appalachian rich swamp: lake/pond: any size wetland wetland north-central interior and appalachian rich swamp: smaller river floodplain/riparian wetland wetland north-central interior wet flatwoods: undifferentiated wetland wetland northern appalachian-acadian conifer-hardwood acidic swamp: isolated wooded wetland wetland northern appalachian-acadian conifer-hardwood acidic swamp: lake/pond: any size wooded wetland wetland northern appalachian-acadian conifer-hardwood acidic swamp: smaller river floodplain/riparian wooded wetland alces vol. 58, 2022 adirondack moose biomass – peterson et al. 23 appendix 2: model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during summer. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight bd 3 –14.22 36.62 0.00 0.19 bd2 3 –14.40 36.99 0.38 0.16 bd+d 4 –12.82 37.65 1.03 0.11 bd2+d 4 –13.04 38.07 1.46 0.09 bd+s 4 –13.26 38.52 1.90 0.07 bd2+s 4 –13.36 38.73 2.11 0.07 bd+bd2 4 –14.09 40.18 3.56 0.03 bd+es 4 –14.14 40.27 3.66 0.03 e+bd 4 –14.17 40.34 3.73 0.03 bd2+es 4 –14.30 40.60 3.99 0.03 e+bd2 4 –14.35 40.70 4.08 0.02 e+bd+d 5 –12.08 40.83 4.21 0.02 e+bd2+d 5 –12.38 41.42 4.81 0.02 bd+bd2+d 5 –12.70 42.06 5.44 0.01 bd+d+s 5 –12.71 42.09 5.48 0.01 bd+d+es 5 –12.81 42.29 5.67 0.01 bd2+d+s 5 –12.88 42.43 5.81 0.01 large maples e 3 –24.63 57.10 0.00 0.21 n 3 –25.38 58.61 1.51 0.10 cc+e 4 –23.76 58.85 1.75 0.09 e+bd 4 –24.01 59.36 2.26 0.07 e+n 4 –24.22 59.77 2.67 0.06 e+s 4 –24.59 60.51 3.41 0.04 2 –28.06 60.98 3.88 0.03 bd+n 4 –24.95 61.24 4.14 0.03 bd2+n 4 –24.95 61.24 4.14 0.03 cc+n 4 –25.02 61.36 4.26 0.03 bd2 3 –26.97 61.79 4.69 0.02 bd 3 –26.99 61.83 4.73 0.02 s+n 4 –25.32 61.98 4.88 0.02 s 3 –27.16 62.16 5.06 0.02 cc+e+n 5 –23.48 62.42 5.32 0.01 cc+e+bd2 5 –23.68 62.82 5.72 0.01 cc+e+bd 5 –23.69 62.83 5.73 0.01 cc+e+s 5 –23.71 62.88 5.78 0.01 e+bd2+n 5 –23.74 62.94 5.84 0.01 e+bd+bd2 5 –23.74 62.94 5.84 0.01 e+bd+n 5 –23.77 62.99 5.89 0.01 appendix 2 (continued) adirondack moose biomass – peterson et al. alces vol. 58, 2022 24 appendix 2 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during summer. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight cc 3 –27.67 63.19 6.09 0.01 birches cc+bd+bd2+s bd × s 7 –30.16 80.21 0.00 0.09 cc+bd+bd2+s bd2 × s 7 –30.42 80.73 0.52 0.07 cc+bd+bd2 cc × bd 6 –32.86 81.93 1.72 0.04 cc+bd+bd2 cc × bd+cc × bd2 7 –31.43 82.75 2.54 0.03 cc+bd+bd2+d+s bd × s 8 –29.76 83.52 3.31 0.02 cc+e+bd+bd2 cc × e+cc × bd 8 –29.78 83.56 3.35 0.02 cc+bd+bd2 cc × bd2 6 –33.68 83.57 3.35 0.02 cc+e+bd+bd2+s cc × e+cc × s 9 –27.54 83.66 3.45 0.02 cc+bd+bd2+s+n bd × s 8 –29.88 83.76 3.55 0.02 cc+e+bd+bd2+s bd × s 8 –29.96 83.91 3.70 0.01 cc+bd+bd2+d+s bd2 × s 8 –29.96 83.92 3.70 0.01 cc+bd+bd2+s cc × s 7 –32.03 83.96 3.75 0.01 cc+e+bd+bd2+d+s cc × e+cc × s 10 –25.14 84.03 3.82 0.01 cc+bd+bd2+s bd × s+bd2 × s 8 –30.11 84.21 4.00 0.01 cc+bd+s bd × s 6 –34.01 84.23 4.01 0.01 cc+bd+bd2+s cc × s+bd × s 8 –30.12 84.24 4.02 0.01 cc+bd+bd2+s cc × bd2+bd × s 8 –30.12 84.25 4.04 0.01 cc+bd+bd2+s+n bd2 ×s 8 –30.14 84.29 4.07 0.01 cc+bd+bd2+s cc × bd+bd × s 8 –30.15 84.30 4.09 0.01 cc+e+bd+bd2+s cc × e+bd × s 9 –27.94 84.47 4.26 0.01 bd+bd2+s 5 –35.86 84.58 4.37 0.01 hobblebush e+v+v2 5 1 11.8 0 0.12 e+v+v2 exv 6 0.778801 12.3 0.5 0.09 e+v 4 0.740818 12.4 0.6 0.09 v 3 0.704688 12.5 0.7 0.08 v+v2 4 0.57695 12.9 1.1 0.07 cc+e+v+v2 6 0.367879 13.8 2 0.04 cc+v+v2 5 0.286505 14.3 2.5 0.03 e+v exv 5 0.201897 15 3.2 0.02 cc+v 4 0.201897 15 3.2 0.02 a+a2+h 5 0.19205 15.1 3.3 0.02 a+a2+h 5 0.19205 15.1 3.3 0.02 a+a2+v 5 0.19205 15.1 3.3 0.02 e+v+v2 5 0.182684 15.2 3.4 0.02 h+v 4 0.157237 15.5 3.7 0.02 a+v 4 0.157237 15.5 3.7 0.02 a+h 4 0.157237 15.5 3.7 0.02 a+h 4 0.157237 15.5 3.7 0.02 appendix 2 (continued) alces vol. 58, 2022 adirondack moose biomass – peterson et al. 25 appendix 2 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during summer. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight v+v2 4 0.157237 15.5 3.7 0.02 h+v+v2 5 0.157237 15.5 3.7 0.02 a+v+v2 5 0.157237 15.5 3.7 0.02 cc+e+v 5 0.149569 15.6 3.8 0.02 e+v+v2+d 6 0.142274 15.7 3.9 0.02 e+a+a2+v 6 0.142274 15.7 3.9 0.02 e+a+a2+h 6 0.142274 15.7 3.9 0.02 e+a+a2+h 6 0.142274 15.7 3.9 0.02 cc+e+v+v2 exv 7 0.122456 16 4.2 0.01 e+a+v 5 0.116484 16.1 4.3 0.01 aspens bd+bd2 4 –13.39 37.28 0.00 0.33 cc+bd+bd2 5 –12.53 39.06 1.77 0.14 bd+bd2+s 5 –12.98 39.95 2.67 0.09 e+bd+bd2 5 –13.35 40.70 3.42 0.06 bd+bd2+d 5 –13.39 40.77 3.49 0.06 cc+e+bd+bd2 6 –11.94 41.88 4.60 0.03 cc+bd+bd2+s 6 –12.19 42.39 5.10 0.03 cc+bd+bd2 cc × bd2 6 –12.36 42.72 5.44 0.02 cc+bd+bd2 cc × bd 6 –12.41 42.82 5.54 0.02 cc+bd+bd2+d 6 –12.53 43.05 5.77 0.02 bd+bd2+d+s 6 –12.76 43.52 6.23 0.01 bd+bd2+s bd2 × s 6 –12.97 43.94 6.66 0.01 e+bd+bd2+s 6 –12.97 43.94 6.66 0.01 bd+bd2+s bd × s 6 –12.97 43.95 6.67 0.01 cherries bd+bd2+s+ea bd × s+bd2 × s 8 –24.88 75.36 0.00 0.16 e+bd+bd2+s bd ×s+bd2 × s 8 –25.41 76.42 1.06 0.09 cc+e+bd+bd2+s bd × s+bd2 × s 9 –23.70 78.26 2.90 0.04 e+bd+bd2+s 6 –30.91 78.77 3.41 0.03 e+bd+bd2+s+n bd × s+bd2 × s 9 –23.99 78.83 3.47 0.03 bd+bd2+s 5 –32.84 79.01 3.65 0.03 e+bd+bd2+s+ea bd × s+bd2 × s 9 –24.10 79.05 3.69 0.03 cc+e+bd+bd2+s cc × bd2+bd × s+bd2 × s 10 –21.34 79.60 4.24 0.02 bd+bd2+s bd × s+bd2 × s 7 –29.41 79.81 4.45 0.02 bd+bd2+d+s+ea bd × s+bd2 × s 9 –24.64 80.13 4.77 0.01 bd+bd2+s+ea bd × s+bd × ea+bd2 × s 9 –24.72 80.30 4.94 0.01 bd+bd2+s+ea+n bd × s+bd2 × s 9 –24.73 80.32 4.96 0.01 bd+bd2+s+ea 6 –31.69 80.33 4.97 0.01 bd+bd2+s+ea bd × s+bd2 × s+bd2 × ea 9 –24.75 80.36 5.00 0.01 appendix 2 (continued) adirondack moose biomass – peterson et al. alces vol. 58, 2022 26 appendix 2 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during summer. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight cc+bd+bd2+s+ea bd × s+bd2 × s 9 –24.81 80.48 5.12 0.01 e+bd+bd2+d+s bd × s+bd2 × s 9 –24.86 80.59 5.23 0.01 cc+e+bd+bd2+s cc × bd+cc × bd2 9 –24.87 80.59 5.23 0.01 cc+e+bd+bd2+s cc × bd+bd × s+bd2 × s 10 –21.98 80.88 5.52 0.01 northen wild raisin e+v+s 5 –2.69 19.40 0.00 0.14 e+v+n 5 –3.07 20.10 0.75 0.09 cc+e+v 5 –3.27 20.50 1.15 0.08 cc+v 4 –5.54 21.60 2.20 0.05 cc+e+v+s 6 –1.98 21.90 2.56 0.04 cc+e+v+n 6 –2.00 22.00 2.61 0.04 e+v+s e × v 6 –2.00 22.00 2.62 0.04 e+v+n e × v 6 –2.11 22.20 2.84 0.03 cc+v cc × v 5 –4.18 22.40 2.98 0.03 e+v+s v × s 6 –2.22 22.40 3.06 0.03 e+v e × v 5 –4.26 22.50 3.13 0.03 e+v+s+n 6 –2.39 22.80 3.39 0.03 cc+e+v cc × v 6 –2.56 23.10 3.73 0.02 e+v+n e × v+v × n 7 –0.61 23.80 4.45 0.02 cc+v+s 5 –4.93 23.90 4.47 0.01 e+v+n v × n 6 –3.03 24.10 4.67 0.01 v+s+n v × n 6 –3.03 24.10 4.68 0.01 cc+e+v e × v 6 –3.08 24.20 4.78 0.01 cc+e+v cc × e 6 –3.27 24.50 5.15 0.01 cc+v+n v × n 6 –3.34 24.70 5.29 0.01 small beech bd+s 4 –8.435 34.9 0 0.14 s 3 –12.516 35.8 0.96 0.09 intercept 2 –14.96 35.9 1.05 0.08 e 3 –12.771 36.3 1.47 0.07 bd+e 4 –9.527 37.1 2.18 0.05 bd 3 –13.411 37.6 2.75 0.04 bd+d 3 –14.667 40.1 5.26 0.01 cc 3 –14.827 40.5 5.58 0.01 alces vol. 58, 2022 adirondack moose biomass – peterson et al. 27 appendix 3: model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during winter. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight bd+s 4 –15.11 42.21 0.00 0.39 bd 3 –18.36 44.89 2.68 0.10 bd+s bd × s 5 –14.37 45.40 3.19 0.08 bd+d 4 –16.76 45.51 3.30 0.07 e+bd+s 5 –14.79 46.25 4.04 0.05 bd+d+s 5 –15.03 46.73 4.52 0.04 bd+s+n 5 –15.10 46.88 4.66 0.04 bd+n 4 –17.53 47.06 4.84 0.03 bd+d+n 5 –15.22 47.11 4.90 0.03 bd+n bd × n 5 –15.58 47.83 5.62 0.02 e+bd 4 –18.14 48.27 6.06 0.02 bd+d bd × d 5 –16.30 49.27 7.06 0.01 s 3 –20.56 49.30 7.09 0.01 e+bd+d+n 6 –13.41 49.31 7.10 0.01 large maples e 3 –25.00 57.85 0.00 0.44 e+bd 4 –24.53 60.39 2.54 0.12 e+n 4 –24.94 61.21 3.36 0.08 e+s 4 –24.99 61.31 3.46 0.08 n 3 –26.86 61.56 3.71 0.07 2 –28.86 62.58 4.73 0.04 e+bd 5 –23.98 63.41 5.56 0.03 bd 3 –27.94 63.72 5.87 0.02 s 3 –28.00 63.85 6.00 0.02 bd+n 4 –26.47 64.28 6.43 0.02 e+bd+n 5 –24.52 64.49 6.64 0.02 e+bd+s 5 –24.52 64.49 6.64 0.02 s+n 4 –26.75 64.84 6.99 0.01 e+s+n 5 –24.93 65.32 7.47 0.01 striped maple bd+d bd × d 5 –18.85 53.70 0.00 0.47 bd+bd2+d bd × d 6 –17.60 56.53 2.83 0.11 bd+bd2+d bd2 × d 6 –17.60 56.54 2.84 0.11 bd+d+s bd × d 6 –18.64 58.60 4.91 0.04 bd2+d+s bd2 × d 6 –19.28 59.90 6.20 0.02 birches cc+bd+bd2 cc × bd+cc × bd2 7 –34.00 87.90 0.00 0.17 cc+bd+bd2 cc × bd 6 –36.43 89.06 1.16 0.09 cc+bd+bd2+s bd × s 7 –35.04 89.98 2.08 0.06 cc+bd+bd2+d cc × bd 7 –35.63 91.15 3.25 0.03 cc+bd+bd2+s bd2 × s 7 –35.65 91.19 3.29 0.03 appendix 3 (continued) adirondack moose biomass – peterson et al. alces vol. 58, 2022 28 appendix 3 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during winter. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight cc+e+bd+bd2 cc × bd+cc × bd2 8 –33.62 91.24 3.34 0.03 cc+bd+bd2 cc × bd2 6 –37.53 91.26 3.36 0.03 cc+bd+bd2+d cc × bd+cc × bd2 8 –33.70 91.40 3.50 0.03 cc+bd+bd2+n cc × bd+cc ×bd2 8 –33.70 91.41 3.51 0.03 bd+bd2+s bd × s 6 –37.82 91.85 3.95 0.02 cc+bd+bd2+s cc × bd+cc × bd2 8 –33.95 91.90 4.00 0.02 bd+bd2+s bd2 × s 6 –38.12 92.44 4.54 0.02 cc+bd+bd2+n cc × bd 7 –36.29 92.47 4.57 0.02 cc+bd+bd2+s cc × bd 7 –36.36 92.61 4.71 0.02 cc+e+bd+bd2 cc × bd 7 –36.43 92.74 4.84 0.01 cc+bd+bd2+d+s bd × s 8 –34.41 92.82 4.92 0.01 cc+bd+bd2+d cc × bd+bd2 × d 8 –34.42 92.84 4.93 0.01 cc+bd+bd2+s bd × s+bd2 × s 8 –34.46 92.92 5.02 0.01 cc+bd+bd2+s+n bd × s 8 –34.51 93.01 5.11 0.01 cc+bd+bd2+d cc × bd2 7 –36.70 93.30 5.40 0.01 hobblebush v 3 –7.68 22.95 0.00 0.13 cc+e cc × v 6 –2.79 24.57 1.62 0.06 cc+e+v2 cc × v 7 –0.52 25.23 2.28 0.04 h 4 –7.20 25.25 2.30 0.04 e 4 –7.26 25.37 2.42 0.04 e+v2 5 –5.38 25.38 2.43 0.04 cc 4 –7.30 25.45 2.50 0.04 d 4 –7.44 25.74 2.79 0.03 cc cc × v 5 –5.60 25.82 2.87 0.03 cc+v2 cc × v 6 –3.65 26.31 3.35 0.03 h+v2 5 –5.85 26.31 3.36 0.03 v2+d 5 –5.99 26.59 3.64 0.02 e+h 5 –5.99 26.60 3.65 0.02 cc+e+v2 6 –4.22 27.44 4.49 0.01 cc+v2 cc × v2 6 –4.30 27.59 4.64 0.01 e+h+v2 6 –4.46 27.93 4.98 0.01 aspens bd+bd2 4 –12.05 34.60 0.00 0.33 bd+bd2+d 5 –11.32 36.63 2.04 0.12 bd+bd2+s 5 –11.45 36.91 2.31 0.10 cc+bd+bd2 5 –11.82 37.65 3.05 0.07 e+bd+bd2 5 –12.04 38.07 3.48 0.06 bd+bd2+d bd2 × d 6 –10.83 39.67 5.07 0.03 bd+bd2+d bd × d 6 –10.89 39.79 5.19 0.02 appendix 3 (continued) alces vol. 58, 2022 adirondack moose biomass – peterson et al. 29 appendix 3 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during winter. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight cc+bd+bd2+d 6 –11.16 40.32 5.72 0.02 e+bd+bd2+s 6 –11.21 40.41 5.82 0.02 bd+bd2+d+s 6 –11.23 40.46 5.86 0.02 cc+bd+bd2+s 6 –11.28 40.56 5.97 0.02 e+bd+bd2+d 6 –11.31 40.63 6.03 0.02 bd+bd2+s bd × s 6 –11.35 40.69 6.10 0.02 bd+bd2+s bd2 × s 6 –11.35 40.69 6.10 0.02 cherries bd+bd2+s 5 –34.04 81.41 0.00 0.15 e+bd+bd2+s 6 –32.76 82.46 1.05 0.09 bd+bd2+s+e 6 –32.88 82.69 1.28 0.08 cc+e+bd+bd2+s cc × bd2 8 –28.80 83.21 1.79 0.06 bd+bd2+d+s 6 –33.67 84.28 2.87 0.04 cc+bd+bd2+s cc × bd2 7 –31.65 84.31 2.89 0.04 cc+e+bd+bd2+s 7 –31.67 84.35 2.93 0.03 bd+bd2+s bd2 × s 6 –33.82 84.57 3.16 0.03 bd+bd2+s+n 6 –33.87 84.69 3.27 0.03 cc+bd+bd2+s 6 –33.98 84.90 3.48 0.03 e+bd+bd2+s+n 7 –32.23 85.46 4.05 0.02 e+bd+bd2+s bd2 × s 7 –32.27 85.54 4.12 0.02 e+bd+bd2+s+e 7 –32.45 85.90 4.49 0.02 e+bd+bd2+d+s 7 –32.51 86.02 4.60 0.01 e+bd+bd2+s e × bd2 7 –32.53 86.05 4.64 0.01 cc+bd+bd2+s+e cc × bd2 8 –30.36 86.32 4.90 0.01 bd+bd2+s+e+n 7 –32.81 86.62 5.20 0.01 cc+bd+bd2+s+e 7 –32.83 86.66 5.24 0.01 bd+bd2+s+e bd2 × e 7 –32.84 86.67 5.26 0.01 bd+bd2+s+e bd2 × s 7 –32.86 86.72 5.31 0.01 bd+bd2+d+s+e 7 –32.87 86.74 5.33 0.01 northern wild raisin e+v+s 5 –5.57 25.15 0.00 0.20 e+v+n 5 –6.58 27.15 2.01 0.07 cc+e+v 5 –6.85 27.71 2.56 0.06 e+v+s e × v 6 –5.02 28.03 2.89 0.05 cc+v 4 –8.83 28.16 3.01 0.04 cc+e+v+s 6 –5.14 28.28 3.13 0.04 e+v+s v × s 6 –5.18 28.37 3.22 0.04 e+v+s+n 6 –5.46 28.92 3.77 0.03 e+v e × v 5 –7.64 29.27 4.13 0.03 cc+v cc × v 5 –7.71 29.42 4.27 0.02 appendix 3 (continued) adirondack moose biomass – peterson et al. alces vol. 58, 2022 30 appendix 3 (continued): model selection table for allometric equations describing browse biomass availability on individual tree and shrub species for moose in adirondack park, new york during winter. models for large (>60 mm diameter) american beech are not displayed, as the small sample size warranted an intercept only model. only candidate models carrying >1% of cumulative model weight are displayed. small maples main effects interactions df loglik aicc delta weight cc+e+v+n 6 –5.80 29.60 4.46 0.02 cc+v+s 5 –7.90 29.81 4.66 0.02 e+v+n e × v 6 –5.91 29.82 4.68 0.02 v+s+n v × n 6 –6.04 30.09 4.94 0.02 v 3 –11.41 30.22 5.08 0.02 v+s 4 –9.90 30.30 5.16 0.02 cc+e+v cc × v 6 –6.30 30.60 5.46 0.01 e+v+n e × v+v × n 7 –4.13 30.87 5.73 0.01 small beech bd+s 4 –8.914 35.8 0 0.14 s 3 –12.83 36.5 0.64 0.1 intercept 2 –15.5 37 1.17 0.08 cc 3 –13.11 37 1.2 0.08 bd+cc 4 –10.04 38.1 2.25 0.05 bd 3 –14.1 39 3.17 0.03 d 3 –15.19 41.2 5.35 0.01 alces34(2)_385.pdf alces29_125.pdf alces32_131.pdf alces30_101.pdf alces30_153.pdf alces 31_247.pdf alces29_75.pdf alces34(2)_269.pdf alces32_61.pdf alces29_47.pdf alces30_13.pdf alces30_63.pdf alces 31_199.pdf alces29_135.pdf alces30_41.pdf alces29_91.pdf alces vol. 48, 2012 aitken et al. carcass weights of moose 105 age, sex, and seasonal differences of carcass weights of moose from the central interior of british columbia: a comparative analysis daniel a. aitken1, kenneth n. child2, roy v. rea3, and olav g. hjeljord4 1college of new caledonia, 3330 22nd avenue, prince george, british columbia, v2n 1p8; 26372 cornell place, prince george, british columbia, v2n 2n7; 3natural resources and environmental studies institute, university of northern british columbia, 3333 university way, prince george, british columbia, canada v2n 4z9; 4department of ecology and natural resource management, norwegian university of life sciences, po box 5003, no-1432 ås, norway. abstract: carcass weight (4 quarters without head, hide, lower legs, or internal organs) of moose (alces alces) harvested in 1995-2007 in the omineca sub-region of the central interior of british columbia, canada were obtained from meat cutters records submitted to the conservation officer service, prince george, british columbia. mean carcass weight of male calves (<1 year) was 82 ± 16 (sd) kg and was not different (p = 0.629) from that of female calves that was 81 ± 13 kg. mean carcass weight of juvenile bulls (spike/fork antlers) was 162 ± 21 kg. the mean carcass weight (249 ± 37) of adult bulls (larger than spike/fork antlers) was heavier (25%, p <0.001) than that of adult cows (199 ± 29 kg. mean carcass weight of adult bulls was heavier (14 kg or 5.9% of carcass weight, p = 0.002) in the pre-rut (10-25 september) than post-rut period (16-31 october); a similar change did not occur in juvenile bulls (p = 0.244). the mean carcass weights of calves (p = 0.651) and adult cows (p = 0.142) were not different between the october and late november-early december hunting seasons. carcass weights and sexual size dimorphism for moose from the omineca were mostly similar to those from european and north american ranges. we recommend increased collection of biological data at hunter check stations to provide more accurate body weight data and associated relationships. alces vol. 48: 105-122 (2012) key words: alces alces, body mass, carcass, moose, body weight, sexual dimorphism. weights of moose (alces alces) have been studied across their ranges in north america (blood et al. 1967, schladweiler and stevens 1973, peterson 1974, schwartz et al. 1987, quinn and aho 1989, adams and pekins 1995, lynch et al. 1995, review of calf and yearling weights in broadfoot et al. 1996) and europe (sæther 1983, sæther and hagenrud 1983, 1985a, b, sæther and heim 1993, ericsson et al. 2002, solberg et al. 2007). body size (mass or weight) and proportions change together (franzmann et al. 1978, sæther 1983, bartosiewicz 1987, wallin et al. 1996, but see sand et al. 1995), and both are sensitive to the nutritional intake of the individual. whatever an animal’s genetic potential, its body size is much influenced by its environment (klein 1964, geist 1999). body mass and condition of male moose peak just before the breeding season, whereas female moose reach their maximum weight in early winter (franzmann et al. 1978, schwartz et al. 1987, schwartz 1998). maximum weight of calves in their first year occurs at about 5 months of age (franzmann et al. 1978, schwartz et al. 1987), some of which is lost during their first winter (schwartz 1998). female moose attain maximum body weight at 3.5-4.5 years, whereas male moose attain their maximum body weight at 5.5-6.5 years (sand et al. 1995, schwartz 1998). live weights have been measured by suscarcass weights of moose aitken et al. alces vol. 48, 2012 106 pending moose in a sling below a tripod (franzmann et al. 1978, haigh et al. 1980, quinn and aho 1989) or having captive moose stand on a scale (lankester et al. 1993, schwartz et al. 1994). several different measures have been used to describe weights of dead moose. for example, blood et al. (1967) defined whole weight as the weight immediately after death, not accounting for blood or tissue loss, and they defined carcass weight (or dressed weight) as the weight without viscera, head, lower legs, and hide. field-dressed weight (hog-dressed weight or eviscerated weight) refers to weight after removal of all viscera (schladweiler and stevens 1973, peterson 1974). for clarity, we use the terms live, whole, carcass, and eviscerated weights. the purpose of this study was to establish a base-line understanding of carcass weights of hunter-harvested moose (a. a. andersoni) from the omineca sub-region of british columbia relative to moose throughout their north american and european ranges. examining records from the omineca for the period from1995-2007, we documented the carcass weights and calculated sexual size dimorphism (ssd) specific to calves, juvenile bulls (spike/ fork antlers), adult cows (older than calves), and adult bulls (antlers larger than spike/fork). we compared carcass weights of both juvenile bulls and adult bulls before and after the rut. we also compared carcass weights of calves and adult cows harvested throughout october (normal cow/calf season) with those harvested during the last week of november and the first week of december (a special late season that ran from 1977-1997). study area the omineca sub-region as delineated by the british columbia ministry of environment for game management purposes is located in the central interior of british columbia, extending across the province from approximately 52° n, 118° w in the southeast to 57° n, 125° w in the northwest. this region is approximately 122,500 km2 in total area representing about 13% of the total land mass of british columbia (fig. 1). rugged mountainous terrain with deeply incised valleys is typical to the north and east of the sub-region (child 1992). by contrast, the terrain is flat to rolling with hundreds of small lakes and wetlands to the south and west (heard et al. 1997). the sub-region contains extensive areas of important moose habitat in the sub-boreal ecotype. this ecotype is a comparatively homogeneous unit, occurring on a drumlinized till plateau surrounding periglacial lake deposits, and dissected by many rivers, lakes, and wetlands (child 1992). the landscape is dominated by coniferous forests of hybrid white spruce (picea engelmannii x glauca) and subalpine fir (abies lasiocarpa). lodgepole pine (pinus contorta var. latifolia) and trembling aspen (populus tremuloides) pioneer secondary successional sites (meidinger & pojar 1991). the climate is generally wet and cool, with precipitation evenly distributed throughout the year. the mean annual temperature at prince george (54° n, 122° w) in the southern portion of the omineca sub-region is 3.7°c, ranging from a monthly mean minimum of -10.3° c in january to a mean maximum of 15.2° c in july. by contrast, at fort saint james (56° n, 124° w) in the western portion, the mean annual temperature is 2.5°c with mean monthly minimum and maximum of -12.2° c in january and 14.8° c in july, respectively. mean annual precipitation at prince george is 636 mm with 200 mm of snowfall; at fort saint james the mean annual precipitation is 465 mm with 160 mm of snowfall (environment canada 2011). fires, logging, and forest pathogens have had major impacts on the forest landscapes in the region. cut blocks created by commercial logging since the 1960s are common (heard et al. 1997). forest succession is characterized by an early shrub stage of 10-25 years duration providing important foods for moose such as willow (salix spp.) and paper birch alces vol. 48, 2012 aitken et al. carcass weights of moose 107 (betula papyrifera) (child 1992). an outbreak of mountain pine beetle (dendroctonus ponderosae) has killed pine stands from the mid-1990s to the present (2009) throughout the study area and extensive salvage logging of these stands occurs (ritchie 2008). methods information on the carcass weights and sex and age class of hunter-harvested moose from 1995-2007 was obtained from meat cutter records on file at the prince george office of the british columbia conservation officer service. at weigh-in, carcass submissions were recorded as a whole carcass or portion thereof (i.e., ¼, ½, ¾, whole carcass). quantity of meat submitted was recorded as weight on the hook or weight of deboned meat submitted. for this study, we only used records classified as a whole carcass (all 4 quarters): these were without head, hide, lower legs, or internal organs (dressed carcass as per blood et al. 1967). information on sex and maturity class (calf, juvenile, or adult) was recorded by meat cutters during carcass submissions. the maturity class “calf” indicated moose of either sex, less than 6 months of age with carcass weight <115 kg (blood et al. 1967, sæther 1983, cederlund et al. 1991, herfindal et al. 2006a, b, but see tiilikainen 2010) that were harvested by hunters during a general open fig. 1. the omineca sub-region (region 7a) of the british columbia ministry of environment in central british columbia (reprinted from 2008 british columbia hunting and trapping regulations and synopsis). carcass weights of moose aitken et al. alces vol. 48, 2012 108 season in mid-october. the maturity class juvenile indicated bulls with spike/fork antlers harvested during a general open season (early september-early november). adult bulls had antlers larger than spike/fork and were harvested during a limited entry season from early september-early november. adult cows were females older than calves that were harvested during a limited entry season in mid-october. a small number of additional records for adult cows and calves were available from animals harvested during a limited entry season from the last week of november and first week of december in 1995-1997. we only used individual data that were complete (those reporting date of kill within a legal hunting season, management unit (mu), whole carcass, sex, and maturity class) from records submitted by 5 meat cutter establishments. we reclassified the records of male calves (n = 57) weighing >115 kg as juvenile bulls, while juvenile bulls (n = 2) weighing <115 kg were reclassified as male calves, and juvenile bulls (n = 4) weighing >230 kg were reclassified as adult bulls. finally, we reclassified juvenile females (n = 66), a maturity class for which no season was advertised, as either female “calves” (n = 2) with weights <115 kg or female adults (n = 64) with weights >115 kg. carcass weights were described by mean ± standard deviation (sd), range, and sample size. we report these statistics for the harvested sample and for 4 maturity classes: calves, juvenile bulls, adult bulls, and adult cows. carcass weights for both juvenile and adult bulls harvested during the pre(10-25 september) and post-rut (16-31 october) periods were compared by t-test to determine whether either class of bulls lost weight between periods; the rut period (26 september-15 october) was determined from conception dates (british columbia ministry of environment, unpublished data). carcass weights of both calves and adult females were compared by t-test between the october and late november-early december seasons to determine whether their weights changed over the course of the hunting season. because only a limited number of records were available from the late november-early december seasons in 1995-1997, records from those 3 years were pooled and compared with similarly pooled records from the october season. sexual dimorphism of carcass weights of calves and adults was tested using carcass weights of calves and of cows from all seasons, but only carcass weights of adult bulls from the pre-rut period. in each case, equality of variances was tested with levene’s test and then equality of means was compared using independent sample t-tests for equal or unequal variances as appropriate (milliken and johnson 1984). a lack of age information prevented us from identifying all yearlings of either sex. consequently, adult ssd was calculated as the ratio of mean adult bull carcass weight to mean adult cow carcass weight. carcass weights from our study were compared with carcass weights reported in other studies. we assumed carcasses from all studies to be equivalent, even though carcass weight may be affected by loss of blood resulting from bullet wounds (blood et al. 1967), additional losses following hanging and cooling (2.5% in the first 24 h; ledger and smith 1964), trimming of damaged tissues (e.g., blood-shot meat) and fat deposits (e.g., rump fat), as well as the exact location of removal of the head (at the atlas-occipital junction or along the cervical vertebrae) and lower legs prior to butchering. we did not correct for these losses; they may account for some of the differences between the carcass weights reported in this and other studies. comparison of live or whole weights reported by others with carcass weights reported in our study required us to convert their measurements to carcass weights. we assumed whole weights of dead moose to be equivalent to live weights. we used the average carcass yield (50% of whole weight, alces vol. 48, 2012 aitken et al. carcass weights of moose 109 n = 35) for moose (a. a. andersoni) from alberta (blood et al. 1967) to calculate carcass weights from live weights and whole weights given in a number of studies throughout north america for a. a. andersoni (crichton 1980, haigh et al. 1980, lynch et al. 1995), a. a. americana ( quinn and aho 1989, addison et al. 1994), and a. a. gigas (franzmann et al. 1978, schwartz et al. 1994). we assumed the carcass yield reported by blood et al. (1967) was applicable to all north american moose, but this probably requires substantiation. we did not use the carcass yields available in sand et al. (1995), wallin et al. (1996), or solberg et al. (2007) because they were developed from moose (a. a. alces) in sweden and norway and we considered them less applicable to our study. comparison of eviscerated weights with carcass weights also required a conversion. visceral weight reportedly varies with body weight, age, volume of food in the digestive tract, and the amount of visceral fat (peterson 1974). first, we converted eviscerated weights to “whole” weights. eviscerated and whole weights have been reported for a. a. andersoni (crichton 1979, 1980) and a. a. americana (peterson 1974, broadfoot et al. 1996). we used the average visceral weight reported by crichton (1979, 1980) for calves (38%, n = 4), yearlings (31%, n = 4), and adults (31%, n = 28) to convert reported eviscerated weights of a. a. andersoni to whole weights. these visceral weights were also used to convert eviscerated weights of a. a. shirasi to whole weights. similarly, to convert eviscerated weights of calf and yearling a. a. americana to whole weights, we used the average eviscerated weight (68% of live weight) for captive 11-month-old moose (n = 12) from broadfoot et al. (1996). the average visceral weight (28%, n = 9) from peterson (1974) was used to convert eviscerated weights (72% of live weight) to whole weights of adult moose. second, we converted these “whole” weights to carcass weights using the conversion factor from blood et al. (1967). the conversion factors for eviscerated weights reported in peterson (1974) and broadfoot et al. (1996) were determined for a. a. americana, while both the visceral weights reported by crichton (1979, 1980) and the carcass yields reported by blood et al. (1967) were based on measurements of a. a. andersoni; it is unknown whether these conversion factors are applicable to other subspecies of moose. we calculated ssd for moose from various north american and european ranges by dividing reported mean weight of males by reported mean weight of females for calves, yearlings, and adults. statistical procedures were performed with pasw’s spss version 18. significance of all statistical tests was set a priori at p = 0.05. results carcass weights carcass weights derived for 2,050 moose ranged from 36-375 kg (fig. 1) with a mean weight of 188 ± 61 (sd) kg. the mean carcass weight of all calves (n = 236) was 81 ± 15 kg (range = 36-114 kg); the mean weight of males (n = 143) was 82 ± 16 kg (range = 36-114 kg) and that of females (n = 93), 81 ± 13 kg (range = 4-110 kg). the mean carcass weight of juvenile bulls (n = 844) was 162 ± 21 kg (range = 117-214 kg). the mean carcass weight was 249 ± 37 kg (range = 135-375 kg) for adult males (n = 747), and 199 ± 29 kg (range = 118-281 kg) for adult females (n = 223) (fig. 2a, 2b). in-season changes of carcass weight adult bulls were heavier (t = 3.241, df = 181, p = 0.001) in the pre-rut (x = 251 ± 39 kg, n = 204) than post-rut period (x = 237 ± 30 kg, n = 79). mean weight loss for adult bulls was 14 kg, or 5.6% of the mean weight at pre-rut. in contrast, weights of juvenile bulls were not different (t = 1.168, df = 375, p = 0.244) between the pre-rut (x = 162 ± 20 kg, n = 241) and post-rut periods (x = 160 ± carcass weights of moose aitken et al. alces vol. 48, 2012 110 21 kg, n = 136). carcass weights of calves in october (x = 79 ± 16 kg, n = 87) were not different (t = 0.454, df = 89, p = 0.651) from weights of a small sample of calves in late november-early december (x = 75 ± 21 kg, n = 4). similarly, carcass weights of adult cows in october (x = 190 ± 30 kg, n = 48) were not different (t = -1.488, df = 60, p = 0.142) from weights in late november-early december (x = 203 ± 27 kg, n = 14). sexual size dimorphism the ssd for calves was 1.01 (table 1) and the mean carcass weight of male calves was not different from that of female calves (t = 0.484, df = 225, p = 0.629). in contrast, the ssd for adults was 1.25 with the mean carcass weight of adult males heavier than that of adult females (t = 15.464, df = 377, p <0.001). we were unable to calculate the ssd for yearlings. carcass weights and ssd in north america and europe carcass weights of moose of all age and sex classes in the omineca had more variation than in most other studies, with heavier maximum and lighter minimum carcass weights. mean weights of moose of all age and sex classes from the omineca were generally heavier than reported for moose elsewhere (tables 2-4). weights of male and female calves from the omineca were not different (i.e., ssd = 1). similarly, equivalent weights were reported for male and female calves in central alberta (blood et al. 1967). in fig. 2. carcass weights of moose harvested in the omineca region of british columbia, 1995-2007. weights (nearest 10 kg) and sample size are provided for 3 age classes of males (2a.) and 2 age classes of females (2b.). b a male calf spike/fork male adult male female calf adult female n um be r of m oo se n um be r of m oo se carcass weight (kg) carcass weight (kg) alces vol. 48, 2012 aitken et al. carcass weights of moose 111 contrast, ssd of calves favoured males in 13 of 17 studies from various north american and european jurisdictions, although ssd of calves favoured females in only 2 areas in north america (table 1). although we were unable to calculate ssd for yearlings in the omineca, ssd favoured males in 8 of 13 studies across north america and europe (table 1). no ssd was calculated for yearlings from 3 studies, and 2 studies found that ssd favoured female yearlings (table 1). in adult moose, ssd favored males by 25% in the omineca. adult male moose were heavier than adult female moose in 13 of 15 studies across north america and europe, whereas no ssd was found in the other 2 studies (table 1). table 1. sexual size dimorphism (ssd) values calculated for calf, yearling, and adult moose of 5 subspieces found in north america and europe. the ssd for each category was calculated as mean weight of males/mean weight of females; calf = 0.5 years, yearling = 1.5 years, adult = 2.5+ years; “–“ indicates no data available. whole weight is for the entire animal immediately after death before evisceration (blood et al. 1967); eviscerated (field dressed) weight is for an animal with all viscera removed (schladweiler and stevens 1973); carcass weight is for animals lacking viscera, head, lower legs, and hide (blood et al. 1967). subspecies calf yearling adult weight category source, location, time of collection andersoni 1.01 1.25 carcass present study, british columbia, sept.-dec. andersoni 1.01 0.94 1.02 carcass blood et al. 1967, alberta, 24 nov.-6 jan. andersoni 1.25 live haigh et al. 1980, alberta & saskatchewan, oct.-feb. andersoni 0.68 1.14 whole and eviscerated crichton 1979, 1980, manitoba, sept.-dec. andersoni 1.16 1.17 1.11 whole lynch et al. 1995, alberta, 1-12 dec. americana 0.9 1.1 1.17 eviscerated timmermann 1972, ontario, autumn americana 1.1 1 1.25 eviscerated peterson 1974, quebec, mid sept-late oct americana 1.06 0.99 1.37 eviscerated dunn and morris 1981, maine, 22-27 sept. americana 1.18 1.04 live and whole quinn and aho 1989, ontario, winter and summer americana 1.13 live addison et al. 1994, ontario, mid oct. (151 days) americana 1.04 0.92 1.25 eviscerated adams and pekins 1995, new hampshire, autumn gigas 1.18 live and whole franzmann et al. 1978, alaska, inside the mrc, unspecified gigas 1.14 live and whole franzmann et al. 1978, alaska, outside the mrc, unspecified shirasi 1.18 1.05 1.18 eviscerated schladweiler and stevens 1973, montana, oct.nov. alces 1.06 1.05 1.23 carcass sæther 1983, sweden, 10 sept.-20 oct. alces 1.05 carcass cederlund et al. 1991, norway, oct.-nov. alces 1.07 1.07 carcass lykke 2005, norway, sept.-nov. alces 1.06 1.08 carcass herfindal et al. 2006a, norway, sept.-oct. alces 1.06 1.06 carcass herfindal et al. 2006b, norway, adjusted to 1 oct. alces 1.05 1 1.17 carcass tiilikainen 2010, finland, adjusted to 15 oct. alces 1.06 carcass tiilikainen 2010, norway, adjusted to 15 oct. carcass weights of moose aitken et al. alces vol. 48, 2012 112 discussion carcass weights the trimodal distribution of carcass weights of males (fig. 2a) and the bimodal distribution for females (fig. 2b) likely reflects the harvest regime, with open seasons for calves of either sex and juvenile (spike/ fork antlers) bulls complimented by restricted (limited entry) seasons for older bulls and cows; calves were <6 months old when harvested in october. previous studies (hatter 1993, child et al. 2010a, b) showed that spike/fork antlered moose from the omineca were principally yearlings (1.5 yr), and bulls with larger than spike/fork antlers comprised 55% of yearlings and more than 98% of bull moose >2.5 years old. thus, for comparative purposes, we considered juvenile bulls (those with spike/fork antlers) from the omineca to be yearlings (1.5 yr) and adult bulls (those with antlers larger than spike/fork) as >2.5 years old. lifelong growth patterns, monthly changes in body condition, and different methods of estimating weight of individual moose, coupled with population age and sex structure may introduce sources of bias that make it difficult to compare weights of moose between populations or different geographical areas (franzmann et al. 1978). it is possible that the carcass weight statistics of both yearling and adult males were biased because we were unable to determine if yearlings with spike/fork antlers were of similar weight as yearlings with larger antlers. specific aging through measurement of incisor teeth would address this potential bias. consequently, caution should be exercised when comparing our results with other studies. the carcass weights of moose of all age and sex classes from the omineca had more variation than in most other studies (table 2-4). this variation may reflect some combination of the large number of carcasses sampled (n = 2050), the length of the study (15 yr), the wide geographic area (53º n 122º w to 55º n 124º w), the range of habitat types (child et al. 2010a, b) from low elevation riparian (700 m) to montane (2000 m), and differences in local densities and sex ratios (heard et al. 1999a, b, walker et al. 2006a, b). the comparatively heavy mean and maximum weights for moose of all age categories from the omineca are suggestive of a population below carrying capacity, not limited by per capita food availability (heard et al. 1997). weights of moose vary with climate (sæther 1985, solberg et al. 2004), latitude (sæther 1985, sand et al. 1995), altitude (hjeljord and histøl 1999, ericcson et al. 2002), habitat quality (sæther and heim 1993, selas et al. 2003, herfindal et al. 2006a, b), density of moose (sand et al. 1996, hjeljord and histøl 1999, ferguson et al. 2000, solberg et al. 2004), and population sex ratio (garel et al. 2006). also, winter ticks (dermacentor albipictus) occur in north central british columbia (samuel 2004) and high density of winter ticks can cause reduction in weight of moose (glines and samuel 1989, addison et al. 1994). the effect of each of these factors on carcass weight of moose from the omineca is unknown. the minimum weights of calves from the omineca were generally lower than reported elsewhere (tables 2-4). these minimum weights were similar to those from an extensive study in finland and norway (tiilikainen 2010; table 2), and weights of captive 3rd estrous calves in alaska (schwartz et al. 1994). the mean and maximum weights of calves from the omineca were generally higher than reported elsewhere during similar time periods (table 2-4). interestingly, mean and maximum weights of captive 1st and 2nd estrous moose calves on a high quality diet (schwartz et al. 1994) were similar to the weights we report. the mean carcass weights of calves reported by blood et al. (1967; table 2), timmerman (1972; table 4), and lynch et al. (1995; table 3) were heavier than those from the omineca. alces vol. 48, 2012 aitken et al. carcass weights of moose 113 however, blood et al. (1967) and lynch et al. (1995) measured calves harvested later in december and january when heavier weights are expected due to continued growth into early winter (schwartz 1998). the weights reported by timmermann (1972) from calves harvested in september may reflect high quality habitat resulting from scattered logging for pulpwood production. the minimum carcass weight of yearling males was generally lower, whereas both maximum and mean weights were generally higher than reported elsewhere (tables 2-4). these findings were only for yearling bulls with spike/fork antlers and did not include yearlings with larger antlers. if larger antlers are indicative of heavier bodies (stewart et al. 2000), the mean and maximum carcass weights we report are presumably conservative; inclusion of yearlings with larger antlers might elevate our mean yearling weight to that of timmermann (1972). carcass weights of both adult male and female moose included a lighter minimum as well as heavier maximum than reported elsewhere (tables 2-4). the lighter minimum weight likely reflects the inclusion of yearlings in our sample of adult moose. in contrast, the maximum weights and mean weights that we report for both males and females are heavier than reported elsewhere, despite the inclusion of yearlings in both categories. it seems subspecies calf yearling adult source, location, male female male female male female time andersoni 82 ± 2, 143 81 ± 12, 93 162 ± 21, 844 -1 249 ± 37, 747 199 ± 29, 223 1, bc, [36-114] [41-110] [117-214] [135-375] [118-281] sept. 10 dec. 5 andersoni 95 ± -, 21 94 ± -, 27 153 ± -, 34 162 ± -, 28 205 ± -, 72 201 ± -, 79 2, alberta, [70-110] [70-112] [115-193] [128-186] [151-258] [144-241] 24 nov. 6 jan. alces 71 ± 14, 63 67 ± 13, 53 152 ± 20, 211 145 ± 18,123 222 ± 39, 681 181 ± 46, 380 3, norway, [-] [-] [-] [-] [-] [-] 10 sept. 20 oct. alces 70 ± -, 161 67 ± -, 172 4, sweden, [32-93] [32-93] oct. nov. alces 63 ± -, 511 59 ± -, 415 129± -, 461 121 ± -, 250 5, norway, [-] [-] [-] [-] sept. nov. alces 67 ± 6, ? 63 ± 6, ? 140 ± 11, ? 130 ± 9, ? 6, norway, [-] [-] [-] [-] sept. oct. alces 66 ± 13, 8268 62 ± 12, 7680 139 ± 20, 8629 131 ± 20, 5483 7, norway, [-] [-] [-] adjusted 1 oct. alces 67 ± 7, 489 63 ± 7, 488 8, norway, [44-96] [28-91] adjusted 15 oct. alces 81 ± 7, 4264 77 ± 7, 4245 152 ± 19, 1639 151 ± 19, 982 227 ± 28, 374 188 ± 22, 198 8, finland, [25-120] [40-107] [-] [-] [-] [-] adjusted 15 oct. table 2. comparison of carcass weights (kg; x ± sd, n, [range]) of moose from the omineca region of british columbia with other jurisdictions. carcass weight (dressed carcass) refers to animals lacking viscera, head, feet, and hide (blood et al. 1967). the three age classes were calf (0.5 years), yearling (1.5 years), and adult (2.5+ years). data sources were this study (1), blood et al. 1967 (2), saether 1983 (3), cederlund et al. 1991 (4), lykke 2005 (5), herfindal et al. 2006a (6), herfindal et al. 2006b (7), and tiilikainen 2010 (8). 1 “-“ indicates no data available. carcass weights of moose aitken et al. alces vol. 48, 2012 114 reasonable to speculate that the mean weights for adults would have been heavier had we identified and corrected for yearlings. in-season changes of carcass weights weight loss of bull moose over the course of the breeding season has been widely reported (franzmann et al. 1978, schwartz et al. 1987, miquelle 1990, mysterud et al. 2005) and has been used as a measure of reproductive effort (mysterud et al. 2005). yearling males lost little weight during the rut, while weight loss of adults (>2 years old) increased with advancing age, but did not vary with either sex ratio or population density (mysterud et al. 2005). adults are involved in the majority of rutting behaviours including fighting, scent-urination, mounting, and copulation (miquelle 1990, 1991, van ballenberghe and miquelle 1993), fast for about 18 days (schwartz et al. 1987, miquelle 1990), and lose large amounts of body fat (schwartz and renecker 1998); their younger counterparts rarely fast, yet feed at reduced rates (miquelle 1990, mysterud et al. 2005). lipid mobilization occurs simultaneously in the carcass subspecies calf yearling adult source, location, time male female male female male female andersoni -1 264 ± -, 6 211 ± -, 12 9, alta. and sask., [238-285] [163-258] oct. feb. andersoni 72 ± 30, 2 107 ± -, 1 146 ± 7, 3 225 ±23, 12 197 ± 198, 9 10, manitoba, [51-93] [-] [-] [190-240] [163-230] 17 sept. 15 dec. americana 72 ± 8, 8 64 ± 3, 10 11, ontario, [-] [-] mid oct. (151 days) andersoni 99 ± -, 13 85 ± -, 12 163 ± -, 6 139 ± -, 8 221 ± -, 40 200 ± -, 46 12, alberta, [75-114] [66-114] [141-182] [89-166] [178-289] [136-261] 1-12 dec. americana 145 ± -, 4 123 ± -, 8 227± -, 29 218 ± -, 45 13, ontario, [115-180] [100-165] [130-271] [155-265] winter and summer gigas 201 ± -, 21 170 ± -, 81 14, alaska inside mrc, [-] [-] unspecified gigas 227 ± -, 5 200 ± -, 66 14, alaska outside mrc, [-] [-] unspecified gigas 85 ± 15, 12 166 ± 29, 8 15, alaska [-] [-] 1st estrous, oct. gigas 70 ± 15, 12 161 ± 22, 7 15, alaska [-] [-] 2nd estrous, oct. gigas 52 ± 9, 3 15, alaska [-] [-] 3rd estrous, oct. table 3. carcass weights (kg; x ± sd, n, [range]) of moose calculated from live weights or whole carcass weights from north america. reported live weights and whole weights were converted to carcass weights using the conversion of carcass weight = 50% of whole weight where whole weight represents the entire animal immediately after death, before evisceration (blood et al. 1967). the three age classes were calf (0.5 years), yearling (1.5 years), and adult (2.5+ years). data sources were haigh et al. 1980 (9), crichton 1979, 1980 (10), addison et al. 1994 (11), lynch et al. 1995 (12), quinn and aho 1989 (13), franzmann et al. 1978 (14), and schwartz et al. 1994 (15). 1 “-“ indicates no data available. alces vol. 48, 2012 aitken et al. carcass weights of moose 115 (subcutaneous fat, intramuscular fat, and bone marrow) and visceral deposits (stephenson et al. 1993, 1998). weight loss in captive bulls (n = 3) in alaska increased from 12% of pre-rut body weight at age 2.5 years to18-19% at age 4.55.5 years (franzmann et al. 1978, schwartz 1998). by comparison, maximum weight loss of harvested male moose (n = 9,949) in norway averaged 9-11% for bulls 6-12 years of age in several hunted populations (mysterud et al. 2005). weight changes of carcasses should be considered conservative since they reflect loss of carcass fat but not loss of visceral fat; weight loss in live moose is larger because it reflects loss of both. the average carcass weight for adult bulls declined significantly between the preand post-rut periods, with losses averaging 5.6%. this value is lower than reported elsewhere (franzmann et al. 1978, schwartz 1998, mysterud et al. 2005) and may indicate lower levels of rutting activity by adult bulls in our study area, or that yearlings with antlers larger than spike/fork were in the adult segment. our lack of precision in estimating age prevented calculation of age-specific weight losses; subsequent comparison with data of mysterud et al. (2005) was also precluded. it is unknown to what extent yearling bulls participate in the rut in the omineca. yearling males generally invest in growth subspecies calf yearling adult source, location, male female male female male female time americana 103 ± -, 7 115 ± -, 3 187 ± -, 19 170 ± -, 7 248 ± -, 26 212 ± -, 8 16, ontario, [80-125] [102-137] [147-247] [157-200] [178-331] [173-255] autumn shirasi 73 ± -, 9 62 ± -, 14 126 ± -, 28 120 ± -, 15 173 ± -,97 147 ± -,70 17, montana, [60-85] [40-75] [92-153] [93-137] [106-265] [96-230] oct. nov. americana 88 ± -, 19 80 ± -, 26 150 ± -, 51 150 ± -, 34 228 ±-, 300 182 ± -, 194 18, quebec, [-] [-] [-] [-] [-] [-] mid sept. late oct. andersoni -1 151 ± 8, 4 120 ± 0, 2 19, manitoba, [-] [120-120] 17 sept. 15 dec. americana 85 ± -, 15 79 ± -, 20 150 ± -, 48 151 ± -, 21 247 ± -, 342 180 ± -, 123 20, maine, [-] [-] [-] [-] [-] [-] 22-27 sept. americana 78 ± 3, 5 21, ontario, [-] mid-oct. (week 24) americana 82 ± 14, 23 79 ± 17, 23 146 ± 30, 139 159 ± 24, 65 222 ± -, 476 178 ± -, 181 22, new hampshire, [-] [-] [-] [-] [-] [-] autumn americana 156 ± -, 450 223 ± -, 2521 22, maine, [-] [-] autumn table 4. carcass weights (kg; x ± sd, n, [range]) of moose from across north america as calculated by converting eviscerated weights to carcass weights; eviscerated (field dressed) weight is for an animal with all viscera removed (schladweiler and stevens 1973). eviscerated weights of calves and yearlings were converted to whole weights using 32% visceral weight (broadfoot et al. 1996), and eviscerated weights of adults were converted using 28% visceral weight (peterson 1974); whole weights were then converted to carcass weights using 50% carcass weight (blood et al. 1967). the three age classes were calf (0.5 years), yearling (1.5 years), and adult (2.5+ years). data sources were timmermann 1972 (16), schladweiler and stevens 1973 (17), peterson 1974 (18), crichton 1980 (19), dunn and morris 1981 (20), lankester et al. 1993 (21), and adams and pekins 1995 (22). 1 “-“ indicates no data available. carcass weights of moose aitken et al. alces vol. 48, 2012 116 rather than reproductive effort, and lose less weight than older bulls; their weight loss is not influenced by sex ratio, but declines with increasing population density (mysterud et al. 2005). we found that preand post-rut carcass weights were not different for spike/fork bulls, suggesting minimal involvement in the rut by these bulls (child et al. 2010a, b). calves achieve maximum body size at or just after the rut, while cows continue to gain weight until early winter (schwartz 1998, cederlund et al. 1991). we found that carcass weights of calves and adult cows from the omineca did not change between the october and late november-early december seasons; however, these findings were based on only 4 calves and 14 adult cows from the late season, and additional data are necessary to substantiate this finding. sexual size dimorphism male moose at all stages of life are generally heavier (franzmann et al. 1978, schwartz et al. 1987, adams and pekins 1995, lynch et al. 1995, schwartz 1998, loison et al. 1999, mysterud 2000) and grow faster and for several more years than females (schwartz 1998, garel et al. 2006, tiilikainen 2010). as a result, ssd favours males, increases with age (geist 1998), and varies with adult sex ratio and length of growing season (garel et al. 2006), as well as location (tiilikainen 2010). male calves were heavier than female calves in 13 of 17 studies (tables 2-4) from across north america and europe, but the difference was not pronounced. in some european studies with large sample sizes (n >500), the ssd of calves favoured males and ranged from 1.04-1.07 (table 1); however, we documented no ssd for calves from the omineca and blood et al. (1967) found no ssd for calves in central alberta. no difference in weight was found (thus, ssd = 1.00) in wild male and female calves in alaska (franzmann et al.1978), or captive moose calves in alaska (schwartz et al. 1994) and ontario (lankester et al. 1993). the only studies indicating ssd favouring female calves were based on sample sizes of 10 (timmermann 1972) and 3 animals (crichton1979, 1980). the differences in body size of yearlings (tables 2-4) were similar to those of calves; ssd for yearlings favoured males in 8 of 13 studies in north america and europe (table 1). the ssd of yearlings in studies with large sample sizes ranged from 1.05-1.08 (table 1). no ssd was calculated for yearlings from quebec (peterson 1974), maine (dunn and morris 1981), or finland (tiilikainen 2010). similarly, no ssd in yearling moose from alaska was reported by franzmann et al. (1978), but ssd of yearlings favoured females in alberta (blood et al. 1967) and new hampshire (adams and pekins 1995). size differences favored males by an average of 24% for norwegian moose. these ssds were larger in areas of norway with short, intense growing seasons that may provide very high forage quality (garel et al. 2006). by contrast, ssds were smaller in populations with an adult sex ratio biased to females (solberg and sæther 1994, garel et al. 2006) which may reflect males diverting resources from growth to reproduction (stearns 1992 in milner et al. 2007), or an altered age distribution of males from harvest practices. in adult moose, ssd favoured males in 13 of 15 studies in north america and europe (table 4). the largest ssd (1.37, table 4) we calculated was from weights of moose harvested 22-27 september, 1980 from a maine population with a balanced sex ratio, an older age structure, and males in peak condition following a 45 year hunting closure (dunn and morris 1981); most ssd of adults were 1.15-1.25 (table 4). the minimal ssd (1.02) in adult moose from alberta (blood et al. 1967) was attributed to the reduced condition of bulls harvested after the rut, as was the ssd (1.04) in ontario (quinn and aho 1989) that was based on weights collected during summer and winter. no records were found indicating alces vol. 48, 2012 aitken et al. carcass weights of moose 117 ssd favouring adult females. we found that adult bull moose were, on average, 25% heavier (ssd = 1.25) than adult cow moose, which is higher than in most jurisdictions (table 4). our value likely represents a maximum for this population as it was calculated with the presumed annual maximum weight of bulls when in prime condition at pre-rut (schwartz 1998). however, potential bias in these data exists since the mean weight for adult males includes some yearlings (antlers larger than spike/fork), and the mean weight for adult females includes all female yearlings. again, more accurate age information would better clarify differences. also, the relationships between ssd and both density and sex ratio should be investigated since the density and sex ratios of moose varied within the study area (heard et al. 1999a, b, walker et al. 2006a, b). record reliability and data quality our examination of the butcher records revealed some obvious confusion with terminology for recording age categories, especially the terms “juvenile” and, to a lesser extent, “calf.” the records used “juvenile” to indicate a spike/fork antlered bull, whereas the hunting regulations currently advertise open seasons for spike/fork bulls, but have historically used either spike/fork or immature bull. calf seasons, on the other hand, are advertised in mid-october for animals 6 months of age or younger. at this time, calf carcasses reportedly weigh <115 kg (blood et al. 1967, sæther 1983, cederlund et al. 1991, herfindal et al. 2006a, b, but see tiilikainen 2010); therefore, weights of male calves >115 kg, particularly those recorded in september and november, suggest that some (~57 of 844) spike/fork (immature) bulls are probably recorded mistakenly as calves. interestingly, no records for female calves >115 kg were found, and only 2 of 66 records for juvenile females had a carcass <115 kg. together, these observations suggest that hunters and/or butchers were accurately identifying female calves; 64 of 66 records for juvenile females were for animals weighing 115-210 kg. the weight distribution for these 64 juvenile females >115 kg parallels that for juvenile males (spike/fork antlers) in our study (117-214 kg), suggesting that hunters and/or butchers were correctly identifying yearling females and recording them as juveniles. conclusions carcass weights of moose from the omineca region in north central british columbia were similar to carcass weights of moose reported for other jurisdictions. our results were also similar to carcass weights we calculated by converting live weights, total weights, and eviscerated weights, even though we used published conversion factors based on small samples from a variety of locations and subspecies. further research is needed to document the range of conversion factors between carcass weights, eviscerated weights, total weights, and live weights and the applicability of these conversion factors to different subspecies of moose. our data represent body mass measurements of a small portion (approximately 10%) of hunter-harvested moose taken in the study area from 1995-2007. additional records would be beneficial and this could be achieved in two ways. first, the accuracy and reliability of the data obtained from the meat cutter records could be improved by employing consistent terminology in both the hunting regulations synopsis and in butcher records. we recommend retaining the terms “calf”, “spike/fork” and “adult” as currently used in the hunting regulations synopsis, while dropping the terms “immature” and “juvenile” on the butcher forms. second, the use of in-season check stops would facilitate collection of additional carcass weights of harvested moose and provide a larger sample for analysis, particularly for calves, since butchers may preferentially accept adults carcass weights of moose aitken et al. alces vol. 48, 2012 118 for butchering. in-season check stops could also provide the opportunity to collect related information including specific kill location coordinates (utm or latitude and longitude), incisor teeth for aging, and tissue samples for analysis (e.g., dna, toxins). despite the short comings of the data set, our findings suggest that butcher records have value because carcass weights were similar to weight records obtained with other methods. without substantially increasing effort at hunter check stations, such records presumably represent the best high volume data available to managers. these data are important as baselines and in developing relationships between animal condition as measured by body mass (adams and pekins 1995, sensu hjeljord and histol 1999) to climate change (sæther 1985, solberg et al. 2004), habitat relationships (sæther and heim 1993, hjeljord and histol 1999, selas et al. 2003, herfindal et al. 2006a, b), and population structure in locally harvested populations (sand et al. 1996, ferguson et al. 2000, solberg et al. 2004, garel et al. 2006). and, enforcement personnel can use such records as supporting evidence during investigations and prosecutions. information about kill location and habitat would help identify specific relationships between weight and a host of factors including age, geography, climate, forage availability, timing of conception, population health, and density of moose in british columbia. acknowledgements we thank barb leslie of the conservation officer service for permitting us to use meat cutter records, anna dehoop for uploading the records, and the meat cutters who diligently recorded the carcass weight information. we also thank the 2 anonymous reviewers for their helpful comments on an earlier draft of this manuscript. references adams, k. p., and p. j. pekins. 1995. growth patterns of new england moose: yearlings as indicators of population status. alces 31: 53-59. addison, e.m., r. f. mclauglin, and j. d. broadfoot. 1994. growth of moose calves (alces alces americana) infested and uninfested with winter ticks (dermacentor albipictus). canadian journal of zoology 72: 1469-1476. bartosiewicz, l. 1987. metacarpal measurements and carcass weight of moose in central sweden. journal of wildlife management 51: 358-359. blood, d. a., j. r. mcgillis, and a. l. lovaas. 1967. weights and measurements of moose in elk island national park, alberta. canadian field naturalist 81: 263-269. broadfoot, j. d., d. g. joachim, e. m. addison, and k. s. macdonald. 1996. weights and measurements of selected body parts, organs and long bones of 11month-old moose. alces 32: 173-184. cederlund, g. n., h. k. g. sand, and a. pehrson. 1991. body mass dynamics of moose calves in relation to winter severity. journal of wildlife management 55: 675-681. child, k. n. 1992. moose management in the omineca sub-region. proceedings of the 1991 moose harvest management workshop, march 1992. wildlife branch, british columbia environment, victoria, british columbia. _____, d. a. aitken, and r. v. rea. 2010a. morphometry of moose antlers in central british columbia. alces 46: 123-134 _____, _____, _____, and r. demarchi. 2010b. potential vulnerability of bull moose in central british columbia to three antler-based hunting regulations. alces 46: 113-121. crichton, v. f. j. 1979. an experimental moose hunt on hecla island, manitoba. alces 15: 245-279. _____. 1980. manitoba’s second experimental moose hunt on hecla island. alces 16: alces vol. 48, 2012 aitken et al. carcass weights of moose 119 489-525. dunn, f. d., and k. i. morris. 1981. preliminary results of the maine moose season (1980). alces 17: 95-110. environment canada national climate data and information archive. 2011. climate. weatheroffice.gc.ca/climate_normals/ index_e.html . ericsson, g., j. p. ball, and k. danell. 2002. body mass of moose calves along an altitudinal gradient. journal of wildlife management 66: 91-97. ferguson, s. h., a. r. bisset, and f. messier. 2000. the influence of density on growth and reproduction in moose alces alces. wildlife biology 6: 31-39. franzmann, a. w., r. e. leresche, r. a. rausch, and j. l. oldemeyer. 1978. alaskan moose measurements and weights and measurement-weight relationships. canadian journal of zoology 56: 298-306. garel, m., e. j. s. solberg, b. -e. sæther, i. herfindal, and k. -a. høgda. 2006. the length of growing season and adult sex ratio affect sexual size dimorphism in moose. ecology 87: 745-758. geist, v. 1998. deer of the world. stackpole books, mechanicsburg, pennsylvania, usa. _____. 1999. moose: behavior, ecology and conservation. voyageur press, stillwater, minnesota, usa. glines, m. v., and w. m. samuel. 1989. effect of dermacentor albipictus (acari: ixodidae) on blood composition, weight gain and hair coat of moose, alces alces. experimental and applied acarology 6: 197-213. haigh, j. c., r. r. stewart, and w. mytton. 1980. relations among linear measurements and weights for moose (alces alces). alces 16: 1-10. hatter, i. 1993. yearling moose vulnerability to spike-fork antler regulation. british columbia environment wildlife branch memorandum, 26 january 1993. heard, d., s. barry, g. watts, and k. child. 1997. fertility of female moose (alces alces) in relation to age and body composition. alces 33: 165-176. _____, k. l. zimmerman, g. s. watts, and s. p. barry. 1999a. moose density and composition around prince george, british columbia, december 1998. final report for common land information base project no. 99004. british columbia ministry of environment, lands and parks, prince george, british columbia, canada. _____, _____, l. l. yaremko, and g. s. watts. 1999b. moose population estimate for the parsnip river drainage, january 1998. final report for forest renewal british columbia. project no. op96004. british columbia ministry of environment, lands and parks, prince george, british columbia, canada. herfindal, i., b. -e. sæther, e. j. solberg, r. andersen, and k. a. høgda. 2006a. population characteristics predict responses in moose body mass to temporal variation in the environment. journal of animal ecology 75: 1110-1118. _____, e. j. solberg, b. -e. sæther, k. -a høgda, and r. andersen. 2006b. environmental phenology and geographical gradients in moose body mass. oecologia 150: 213-224 hjeljord, o., and t. histøl. 1999. rangebody mass interactions of a northern ungulate – a test of hypothesis. oecologia 119: 326-339. klein, d. r. 1964. range-related differences in growth of deer reflected in skeletal ratios. journal of mammalogy 45: 226-235. lankester, m. w., t. wheeler-smith, and s. dudzinski. 1993. care, growth and cost of captive moose calves. alces 29: 249-262. ledger, h. p., and n. s. smith. 1964. the carcass and body composition of the carcass weights of moose aitken et al. alces vol. 48, 2012 120 uganda kob. journal of wildlife management 28: 827-839. loison, a. j., m. gaillard, c. pélabon, and n. g. yoccoz. 1999. what factors shape sexual size dimorphism in ungulates? evolutionary ecology research 1: 611-633. lykke, j. 2005. selective harvest management of a norwegian moose population. alces 41: 9-24. lynch, g. m., b. lajeunesse, j. willman, and e. s. tefler. 1995. moose weights and measurements from elk island national park, canada. alces 31: 199-207. meidinger, d., and j. pojar. 1991. ecosystems of british columbia. special report no. 6. ministry of forests research branch, victoria, british columbia, canada. milner, j. m., e. b. nilsen, and h. p. andreassen. 2007. demographic side effects of selective hunting in ungulates and carnivores. conservation biology 21: 36-47. milliken, g. a., and d. e. johnson. 1984. analysis of messy data: volume i. designed experiments. van nostrand reinhold, new york, new york, usa. miquelle, d. g. 1990. why don’t bull moose eat during the rut? behavioral ecology and sociobiology 27: 145-151. _____. 1991. are moose mice? the function of scent urination in moose. the american naturalist 138: 460-477. mysterud, a. 2000. the relationship between ecological segregation and sexual body-size dimorphism in large herbivores. oecologia 124: 40-54. _____, e. j. solberg, and n. g. yoccoz. 2005. ageing and reproductive effort in male moose under variable levels of intrasexual competition. ecology 74: 742-754. peterson, r. l. 1974. a review of the general life history of moose. naturaliste canadien 101: 9-21. quinn, n. w. s., and r. w. aho. 1989. whole weights of moose from algonquin park, ontario canada. alces 25: 48-51. ritchie, c. 2008. management and challenges of the mountain pine beetle infestation in british columbia. alces 44: 127-135. sæther, b. -e. 1983. relationship between mandible length and carcass weight of moose in norway. journal of wildlife management 47: 226-1229. _____. 1985. annual variation in carcass weight of norwegian moose in relation to climate along a latitudinal gradient. journal of wildlife management 49: 977-983. _____, and h. haagenrud. 1983. life history of the moose (alces alces): fecundity rates in relation to age and carcass weight. journal of mammalogy 64: 226-232. _____, and _____. 1985a. life history of the moose alces alces: relationship between growth and reproduction. holarctic ecology 8: 100-106. _____, and _____. 1985b. geographical variation in body weight and sexual sizedimorphism of norwegian moose (alces alces). journal of zoology 206: 83-96. _____, and m. heim. 1993. ecological correlates of individual variation in age at maturity in female moose (alces alces): the effects of environmental variability. journal of animal ecology 62: 482489. samuel, w. 2004. white as a ghost: winter ticks and moose. the federation of alberta naturalists, edmonton, alberta, canada. sand, h., r. bergstrom, g. cederlund, m. ostergren, and f. stalfelt. 1996. density-dependent variation in reproduction and body mass in female moose alces alces. wildlife biology 2: 233-245. ______, g. cederlund, and k. danell. 1995. geographical and latitudinal variation in growth patterns and adult body size of swedish moose (alces alces). oecologia 102: 433-442. schladweiler, p., and d. r. stevens. 1973. alces vol. 48, 2012 aitken et al. carcass weights of moose 121 weights of moose in montana. journal of mammalogy 54: 772-775. schwartz, c. c. 1998. reproduction, natality and growth. pages 141-172 in a. w. franzmann and c. c. schwartz , editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. _____, k. j. hundertmark, and e. f. becker. 1994. growth of moose calves conceived during the first versus second estrus. alces 30: 91-100. _____, w. l. regelin, and a. w. franzmann. 1987. seasonal weight dynamics of moose. swedish wildlife research supplement 1: 301-310. _____, and l. a. renecker. 1998. nutrition and energetics. pages 441-478 in a. w. franzmann and c. c. schwartz, editors. ecology and management of the north american moose. smithsonian institution press, washington, d.c., usa. selas, v. g. a. sonerud, t. histol, and o. hjeljord. 2003. synchrony in short term fluctuations of moose calf body mass and bank vole population density supports the mast depression hypothesis. oikos 92: 271-278. solberg, e. j., m. heim, v. grotan, b. -e. sæther, and m. garel. 2007. annual variation in maternal age and calving date generate cohort effects in moose (alces alces) body mass. oecologia 154: 259-271. _____, a. loison, j-m. gaillard, and m. heim. 2004. lasting effects of conditions at birth on moose body mass. ecography 27: 677-687. _____, and b. -e. sæther. 1994. male traits as life-history variables: annual variation in body mass and antler size in moose (alces alces). journal of mammalogy 75: 1069-1079. stearns, s. c. 1992. the evolution of life histories. oxford university press, oxford, united kingdom. stephenson, t. r., k. j. hundertmark, c. c. schwartz, and v. van ballenberghe. 1993. ultrasonic fat measurement of captive yearling bull moose. alces 29: 115-123. _____, _____, _____, and _____. 1998. predicting body fat and body mass in moose with ultrasonography. canadian journal of zoology 76: 717-722. stewart, k. m., r. t. bowyer, j. g. kie, and w. c. gasaway. 2000. antler size relative to body mass in moose: tradeoffs associated with reproduction. alces 36: 77-83. tiilikainen, r. 2010. the effects of selective harvesting on life history traits of moose alces alces. publications of the university of eastern finland, dissertations in forestry and natural sciences no. 4, joensuu, finland. timmermann, h. r. 1972. some observations of the moose hunt in the black sturgeon area of northwestern ontario. alces 8: 223-239. van ballenberghe, v., and d. g. miquelle. 1993. mating in moose: timing, behavior, and male access patterns. canadian journal of zoology 71: 1687-1690. walker, a. b. d., d. c. heard, v. michelfelder, and g. s. watts. 2006a. moose density and composition around prince george, british columbia, january 2006. final report for forests for tomorrow. project no. 2914000. british columbia ministry of environment, lands and parks, prince george, british columbia, canada. _____, _____, _____, and _____. 2006b. moose density and composition in the parsnip river watershed, british columbia, december 2005. final report for the ministry of environment. project no. 2914568. british columbia ministry of environment, lands and parks, prince george, british columbia, canada. wallin, k., g. cederlund, and a. pehron. 1996. predicting body mass from chest carcass weights of moose aitken et al. alces vol. 48, 2012 122 circumference in moose alces alces. wildlife biology 2: 53-58. alces vol. 48, 2012 matthews – moose in oregon 63 history and status of moose in oregon patrick e. matthews oregon department of fish and wildlife, 65495 alder slope road, enterprise, oregon 97828, usa. abstract: moose (alces alces) were not present in oregon at the time of european settlement. five alaskan moose were transplanted to the central oregon coast in 1922, but the effort ended in failure. the first recorded sighting of moose in northeast oregon occurred in 1960, and 20 sightings were reported at various locations throughout northeast oregon from 1960-2000. from 2001-2006 the number of sightings increased substantially in the northern blue mountains of oregon, including the first documented calf in 2005. moose in the northern blue mountains became established through natural dispersal, with most believed to have dispersed southwest across the palouse prairie from the moscow mountain area of idaho. moose habitat in the blue mountains is characterized by managed coniferous forest with a mixture of mature timber and logged areas in various stages of succession. survey efforts were established in 2006, with the present population estimated at 60. in 2008, 6 moose were radio-collared to obtain baseline information on reproduction, habitat use, and seasonal movements. captured moose appeared healthy and were without external parasites, although the carotid worm (elaeophora schneideri) was identified in a bull moose that died accidentally in 2010. the oregon department of fish and wildlife plans to continue monitoring population status, range expansion, and health of moose in oregon. alces vol. 48: 63-66 (2012) key words: alces alces, blue mountains, carotid worm, moose, oregon, range, population, range. there is no historic evidence indicating that moose (alces alces) ever occurred in oregon. verts and carraway (1998) did not include moose in their list of mammals of oregon due to lack of physical evidence of the species. however in his book “the mammals and life zones of oregon”, bailey provides an account of moose on the basis of a report (dice 1919 in bailey 1936) that the species formally occurred in the blue mountains of southeast washington. my review of early literature pertaining to oregon, including diaries of explorers and settlers, provided no reference to moose in oregon at the time of european settlement. the first documented moose in oregon were 5 bottle-fed calves (3 females, 2 males) transplanted from the kenai peninsula, alaska to the south central oregon coast at tahkenitch lake in october 1922 (shay 1976). individuals from the transplant and their offspring were habituated to humans resulting in constant damage complaints from rural residents; this effort ended when the last moose was dispatched in 1931. the first record of moose in northeast oregon occurred in august 1960 when a single cow was observed by residents along the imnaha river in wallowa county (matthews 2006). in 1968 a cow moose was reported by dick rivers, a jet boat operator on the snake river at pittsburg creek (bartels 1968). in august 1971 a photograph of a young bull near battle creek on the snake river was taken by another jet boat operator, james zanelli (coggins 1971). from 1960-2000, 20 different sightings were reported at various locations in northeast oregon (matthews 2006); all were believed to be individual transient moose that eventually disappeared. from 2001-2006, the number of moose sightings increased, particularly north of elgin in the wenaha wildlife management unit in the northern blue mountains (matthews 2006). in 2005 a photograph taken by robert larison in the moose in oregon – matthews alces vol. 48, 2012 64 wenaha unit documented the first calf moose observed in oregon (matthews 2006). moose expanded their range and increased in numbers during the last half of the 20th century in both washington (base et al. 2006) and idaho (compton and oldenburg 1994). moose in the northern blue mountains of oregon became established through natural dispersal, and most are believed to have moved southwest across the palouse prairie from the moscow mountain area of idaho. other moose sightings in northeast oregon occur occasionally in and around the wallowa mountains, indicating that moose are probably still expanding their distribution in northeast oregon. based on information from radio-collared moose and surveys, the current range is principally associated with the blue mountains (fig. 1). management recognizing that oregon likely had a resident population of moose, oregon department of fish and wildlife (odfw) biologists began ground and aerial search efforts to locate and classify individual moose in the wenaha unit in 2006. biologists also contacted members of the public and employees of other resource agencies who had reported visual sightings of moose. photographs were used, when available, to identify and record individual moose. a minimum of 19 moose were classified in 2006 (9 cows, 8 calves, and 2 bulls), and similar classification efforts continued through 2010 (table 1). location of moose tracks was recorded during aerial and ground searches when visual observation was unsuccessful. track information coupled with classification of individual moose was used to derive a conservative population estimate of 60 moose in oregon during winter 2010-11. no formal management plan has yet been established to guide moose management in oregon. however, by statute, the odfw will manage moose to protect and perpetuate moose and their habitats to provide optimum recreational and aesthetic benefits to citizens of the state. in cooperation with the confederated tribes of the umatilla indian reservation, and the umatilla national forest, the odfw initiated research to capture and radio-collar moose in the northern blue mountains to gather baseline information on habitat use, seasonal movements, and recruitment of resident moose. in 2008, 6 adult moose (5 cows and 1 bull) were captured by net-gunning and radio-collared; 1 collar (cow) quit transmitting shortly after capture. in 2009 an additional cow was radio collared, another collar quit operating, and 3 collared moose died of unknown causes. in 2010, radio-collars were deployed on 3 more adult cows. of 10 collared moose, 4 received gps transmitters and 6 vhf transmitters. the gps transmitters were programmed to obtain 8 locations per 24 h, and the vhf moose were ground or aerially located 1-4 times per month. all moose were captured during winter months, and appeared healthy with no observed winter ticks (dermacentor albipictus) or other external parasites. fig. 1. geographic range (cross hatching) of moose in northeast oregon, usa, 2011. alces vol. 48, 2012 matthews – moose in oregon 65 habitat the northern blue mountains of oregon contain coniferous forests managed by private landowners, timber companies, odfw, and the umatilla national forest. much of the blue mountains are bisected by steep rugged drainages. preliminary information from radio tracking indicates moose are utilizing much of the same range during winter and summer months, at elevations of 1100-1700 m. currently 7656 and 109 locations have been generated from the gps and vhf moose, respectively. formal analysis of location data has not been completed; however, visual observations of radioed moose and plotted locations on orthoimagery indicate most habitat use by moose occurs in the upper reaches of drainages where topography is less precipitous. past and current silvicultural practices, wildfires, and controlled burns in the region have resulted in a wide variety of timber stands and openings in various stages of succession. in summer months moose appear to be utilizing older stands of mixed lodgepole pine (pinus contorta), western larch (larix occidentalis), engelmann spruce (picea engelmannii), douglas fir (pseudotsuga menziesii glauca), and true firs (abies spp.) that are adjacent to logged openings. recently logged areas often contain scouler willow (salix scouleriana), ceanothus (ceanothus spp.), serviceberry (amelanchier alnifolia), common chokecherry (prunus virginiana), rocky mountain maple (acer glabrum), sticky current (ribes viscosissimum), and other deciduous species. during winter moose remain in mountainous areas where past logging activity has created dense stands of mixed conifer and deciduous species, until deep snow apparently causes them to move to slightly lower elevations. moose move across deep drainages, but rarely remain on canyon side walls for extended periods. only 2 of 10 radio-collared moose spent a substantial amount of time in riparian habitats along river corridors during summer months; river bottoms were not used during winter months. reproduction and mortality at the time of capture, 6 of 9 captured cow moose were accompanied by 7-9 month old calves. all radio-collared cows were believed to have produced calves each year they were monitored based on observation of calves or typical maternal behavior. one of 4 calves born to radio-collared cows survived to at least 11 months of age in 2008, as did 1 of 2 in 2009. in 2010, 5 cows produced 5 calves that survived to at least 11 months of age, including 1 set of twins. of 3 adult mortalities in 2009, 1 was in late april and 2 were in july; no cause of death was determined for any. the adult cow moose that died in april was captured and collared 2 months earlier, and immediately prior to capture, was observed to be slightly ataxic, but otherwise of normal body condition. when captured, it was noted that 4-6 cm of her left ear tip was sloughed. both moose that died in july appeared to be in normal body condition at the time of death; 1 was observed to have necrosis of the left ear tip. ear tip sloughing, observed in 2 of the 3 deaths, can be a clinical sign of the presence of carotid worm in moose, as is loss of muscular coordination (madden et al. 1991) such as observed in the april death. however, none of the 3 carcasses year observed cows calves bulls calves: 100 cows bulls: 100 cows 2006 19 9 8 2 89 22 2007 21 11 6 4 55 36 2008 28 16 7 5 44 31 2009 31 18 7 6 39 33 2010 29 15 7 7 47 47 table 1. gender, age, and composition ratios of moose observed in ground and aerial surveys during fall and winter in the northern blue mountains of oregon, 2006-2010. moose in oregon – matthews alces vol. 48, 2012 66 were examined for presence of carotid worm. during the 2010 captures, a 3-year old bull was accidentally killed, and 3 adult carotid worms were found during gross examination of its carotid arteries. several other cow moose have been observed with ear tip necrosis in the northern blue mountains of oregon, yet appear to be healthy normal individuals. future direction although management of moose in oregon is not a high priority, current monitoring of radio-collared moose will continue into 2012 followed by analysis of habitat use. in the near future, management will include continued monitoring of population growth and distribution, particularly north of wallowa in the sled springs wildlife management unit where recent timber harvesting has created extensive forest openings. efforts will be made to establish annual winter aerial surveys as a means to monitor population trends. limited hunting opportunities may be initiated if moose numbers continue to expand. i also recommend that future research and management include monitoring dead and sick animals for presence of carotid worms. acknowledgements i am grateful for the financial support and cooperation provided by the oregon hunters association, oregon wildlife heritage foundation, confederated tribes of the umatilla indian reservation, blue mountain habitat restoration council, and the umatilla national forest which paid for radio-collaring, monitoring, and survey flights. i thank 2 anonymous reviewers, mike hansen, and vic coggins for comments on earlier drafts of this manuscript. i thank greg rimbach for his help preparing fig. 1. references bailey, v. 1936. the mammals and life zones of oregon. bureau of biological survey, north american fauna number 55. bartels, r. r. 1968. wallowa game district august monthly report. oregon department of fish and wildlife, enterprise, oregon, usa. base, d. l., s. zender, and d. martorello. 2006. history, status, and hunter harvest of moose in washington state. alces 42: 111-114. coggins, v. l. 1971. wallowa game district august monthly report. oregon department of fish and wildlife, enterprise, oregon, usa. compton, b. b., and l. e. oldenburg. 1994. the status and management of moose in idaho. alces 30: 57-62. dice, l. r. 1919. the mammals of southeastern washington. journal of mammalogy 1: 10-22. madden, d. j., t. r. spraker, and w. j. adrian. 1991. elaeophora schneideri in moose (alces alces) from colorado. journal of wildlife diseases 27: 340-341. matthews, p. e. 2006. update on oregon moose. oregon department of fish and wildlife, enterprise, oregon, usa. shay, r. 1976. the great moose adventure. oregon wildlife 31: 3-5. verts, b. j., and l. n. carraway. 1998. land mammals of oregon. university of california press, los angeles, california. usa.